1 The Geological Record of Neoproterozoic Glaciations (Geological Society of London Memoir 36) by Emmanuelle Arnaud, Galen P. Halverson, Graham Shields-Zhou

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 747

The Geological Record of

Neoproterozoic Glaciations
Geological Society Memoirs
The Geological Society of London
Books Editorial Committee

Chief Editor
Bob Pankhurst (UK)

Society Books Editors


John Gregory (UK)
Jim Griffiths (UK)
John Howe (UK)
Howard Johnson (UK)
Rick Law (USA)
Phil Leat (UK)
Nick Robins (UK)
Randell Stephenson (UK)

Society Books Advisors


Eric Buffetaut (France)
Jonathan Craig (Italy)
Tom McCann (Germany)
Mario Parise (Italy)
Satish-Kumar (Japan)
Gonzalo Veiga (Argentina)
Maarten de Wit (South Africa)

IUGS/GSL publishing agreement

This volume is published under an agreement between the International Union of Geological Sciences and the Geological Society of London and arises
from IGCP project number 512.
GSL is the publisher of choice for books related to IUGS activities, and the IUGS receives a royalty for all books published under this agreement.
Books published under this agreement are subject to the Society’s standard rigorous proposal and manuscript review procedures.

It is recommended that reference to all or part of this book should be made in one of the following ways:

Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) 2011. The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36.

Hoffman, P. F., Macdonald, F. A. & Halverson, G. P. 2011. Chemical sediments associated with Neoproterozoic glaciation: iron formation, cap
carbonate, barite and phosphorite. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
Glaciations. Geological Society, London, Memoirs, 36, 67–80.
GEOLOGICAL SOCIETY MEMOIR NO. 36

The Geological Record of Neoproterozoic Glaciations

EDITED BY

EMMANUELLE ARNAUD
University of Guelph, Canada

GALEN P. HALVERSON
McGill University, Canada

and

GRAHAM SHIELDS-ZHOU
University College London, UK

2011
Published by
The Geological Society
London
THE GEOLOGICAL SOCIETY

The Geological Society of London (GSL) was founded in 1807. It is the oldest national geological society in the world and the largest in Europe. It was incorporated
under Royal Charter in 1825 and is Registered Charity 210161.
The Society is the UK national learned and professional society for geology with a worldwide Fellowship (FGS) of over 10 000. The Society has the power to
confer Chartered status on suitably qualified Fellows, and about 2000 of the Fellowship carry the title (CGeol). Chartered Geologists may also obtain the equivalent
European title, European Geologist (EurGeol). One fifth of the Society’s fellowship resides outside the UK. To find out more about the Society, log on to
www.geolsoc.org.uk.
The Geological Society Publishing House (Bath, UK) produces the Society’s international journals and books, and acts as European distributor for selected
publications of the American Association of Petroleum Geologists (AAPG), the Indonesian Petroleum Association (IPA), the Geological Society of America (GSA), the
Society for Sedimentary Geology (SEPM) and the Geologists’ Association (GA). Joint marketing agreements ensure that GSL Fellows may purchase these societies’
publications at a discount. The Society’s online bookshop (accessible from www.geolsoc.org.uk) offers secure book purchasing with your credit or debit card.
To find out about joining the Society and benefiting from substantial discounts on publications of GSL and other societies worldwide, consult www.geolsoc.org.uk,
or contact the Fellowship Department at: The Geological Society, Burlington House, Piccadilly, London W1J 0BG: Tel. þ 44 (0)20 7434 9944; Fax þ 44 (0)20 7439
8975; E-mail: enquiries@geolsoc.org.uk.
For information about the Society’s meetings, consult Events on www.geolsoc.org.uk. To find out more about the Society’s Corporate Affiliates Scheme, write to
enquiries@geolsoc.org.uk.

Published by The Geological Society from:


The Geological Society Publishing House, Unit 7, Brassmill Enterprise Centre, Brassmill Lane, Bath BA1 3JN, UK
(Orders: Tel. þ44 (0)1225 445046, Fax þ44 (0)1225 442836)
Online bookshop: www.geolsoc.org.uk/bookshop
The publishers make no representation, express or implied, with regard to the accuracy of the information contained in this book and cannot accept any legal
responsibility for any errors or omissions that may be made.
# The Geological Society of London 2011. All rights reserved. No reproduction, copy or transmission of this publication may be made without written permission. No
paragraph of this publication may be reproduced, copied or transmitted save with the provisions of The Copyright Licensing Agency Ltd, Saffron House, 6 –10 Kirby
Street, London EC1N 8TS, UK. Users registered with the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, USA: the item-fee code for this
publication is 0435-4052/11/$15.00.

British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library.
ISBN 978-1-86239-334-9

Distributors
For details of international agents and distributors see:
www.geolsoc.org.uk/agentsdistributors
Typeset by Techset Composition Ltd, Salisbury, UK
Printed by CPI Antony Rowe, Chippenham, UK
Mexico Hut, 1978. Clockwise from the back left: Nick Cox (Motorboat Skipper and Engineer), Tim Druitt (Geologist),
Robin (“Bruce”) Davies (Motorboat Skipper and Engineer), Brian Harland (Expedition Leader), James Carter
(undergraduate field assistant), Mike Hambrey (Geologist) and Paul Waddams (PhD Student), Ny-Ålesund, Svalbard

The editors dedicate this Memoir to Mike Hambrey and Brian Harland for their
pioneering work in Neoproterozoic glacial geology and for their comprehensive 1981
volume Earth’s Pre-Pleistocene Glacial Record, which inspired the present work.
Contents
List of Reviewers ix

Introductory chapters

ARNAUD, E., HALVERSON, G. P. & SHIELDS-ZHOU, G. The geological record of Neoproterozoic ice ages 1

HOFFMAN, P. F. A history of Neoproterozoic glacial geology, 1871– 1997 17

ARNAUD, E. & ETIENNE, J. L. Recognition of glacial influence in Neoproterozoic sedimentary successions 39

HALVERSON, G. P. & SHIELDS-ZHOU, G. Chemostratigraphy and the Neoproterozoic glaciations 51

HOFFMAN, P. F., MACDONALD, F. A. & HALVERSON, G. P. Chemical sediments associated with Neoproterozoic glaciation: iron formation, cap 67
carbonate, barite and phosphorite

BAHLBURG, H. & DOBRZINSKI, N. A review of the Chemical Index of Alteration (CIA) and its application to the study of Neoproterozoic 81
glacial deposits and climate transitions

EVANS, D. A. D. & RAUB, T. D. Neoproterozoic glacial palaeolatitudes: a global update 93

GREY, K., HILL, A. C. & CALVER, C. Biostratigraphy and stratigraphic subdivision of Cryogenian successions of Australia in a global context 113

CONDON, D. J. & BOWRING, S. A. A user’s guide to Neoproterozoic geochronology 135

GODDÉRIS, Y., LE HIR, G. & DONNADIEU, Y. Modelling the Snowball Earth 151

Africa

SHIELDS-ZHOU, G. A., DEYNOUX, M. & OCH, L. The record of Neoproterozoic glaciation in the Taoudéni Basin, NW Africa 163

MASTER, S. & WENDORFF, M. Neoproterozoic glaciogenic diamictites of the Katanga Supergroup, Central Africa 173

TAIT, J., DELPOMDOR, F., PRÉAT, A., TACK, L., STRAATHOF, G. & NKULA, V. K. Neoproterozoic sequences of the West Congo and 185
Lindi/Ubangi Supergroups in the Congo Craton, Central Africa

HOFFMAN, P. F. Glaciogenic and associated strata of the Otavi carbonate platform and foreslope, northern Namibia: evidence for large 195
base-level and glacioeustatic changes

PRAVE, A. R., HOFFMANN, K.-H., HEGENBERGER, W. & FALLICK, A. E. The Witvlei Group of East-Central Namibia 211

FRIMMEL, H. E. The Chameis Gate Member, Chameis Group, Marmora Terrane, Namibia 217

FRIMMEL, H. E. The Kaigas and Numees formations, Port Nolloth Group, in South Africa and Namibia 223

FRIMMEL, H. E. The Karoetjes Kop and Bloupoort formations, Gifberg Group, South Africa 233

Eurasia –Nubian Shield

ALLEN, P. A., RIEU, R., ETIENNE, J. L., MATTER, A. & COZZI, A. The Ayn Formation of the Mirbat Group, Dhofar, Oman 239

ALLEN, P. A., LEATHER, J., BRASIER, M. D., RIEU, R., MCCARRON, M., LE GUERROUÉ, E., ETIENNE, J. L. & COZZI, A. The Abu Mahara 251
Group (Ghubrah and Fiq formations), Jabal Akhdar, Oman

MILLER, N. R., AVIGAD, D., STERN, R. J. & BEYTH, M. The Tambien Group, Northern Ethiopia (Tigre) 263

STERN, R. J., JOHNSON, P. R., ALI, K. A. & MUKHERJEE, S. K. Evidence for Early and Mid-Cryogenian glaciation in the Northern 277
Arabian –Nubian Shield (Egypt, Sudan, and western Arabia)

CHUMAKOV, N. M. Glacial deposits of the Bokson Group, East Sayan Mountains, Buryatian Republic, Russian Federation 285

CHUMAKOV, N. M. The Neoproterozoic glacial formations of the North and Middle Urals 289
CONTENTS vii

CHUMAKOV, N. M. Glacial deposits of the Nichatka Formation, Chara River basin and review of Upper Precambrian diamictites 297
of Central Siberia

CHUMAKOV, N. M. Glacial deposits of the Baykonur Formation, Kazakhstan and Kyrgyzstan 303

CHUMAKOV, N. M., POKROVSKY, B. G. & MELEZHIK, V. A. The glaciogenic Bol’shoy Patom Formation, Lena River, central Siberia 309

SOVETOV, J. K. Late Cryogenian (Vendian) glaciogenic deposits in the Marnya Formation, Oselok Group, in the foothills of the 317
East Sayan Range, southwestern Siberian Craton

MACDONALD, F. A. The Tsagaan Oloom Formation, southwestern Mongolia 331

MACDONALD, F. A. & JONES, D. S. The Khubsugul Group, Northern Mongolia 339

ETIENNE, J. L., ALLEN, P. A., LE GUERROUÉ, E., HEAMAN, L., GHOSH, S. K. & ISLAM, R. The Blaini Formation of the Lesser Himalaya, NW India 347

ZHANG, Q.-R., CHU, X.-L. & FENG, L.-J. Neoproterozoic glacial records in the Yangtze Region, China 357

ZHU, M. & WANG, H. Neoproterozoic glaciogenic diamictites of the Tarim Block, NW China 367

North America

MACDONALD, F. A. The Hula Hula Diamictite and Katakturuk Dolomite, Arctic Alaska 379

MACDONALD, F. A. & COHEN, P. A. The Tatonduk inlier, Alaska –Yukon border 389

HOFFMAN, P. F. & HALVERSON, G. P. Neoproterozoic glacial record in the Mackenzie Mountains, northern Canadian Cordillera 397

SMITH, M. D., ARNAUD, E., ARNOTT, R. W. C. & ROSS, G. M. The record of Neoproterozoic glaciations in the Windermere 413
Supergroup, southern Canadian Cordillera

LINK, P. K. & CHRISTIE-BLICK, N. Neoproterozoic strata of southeastern Idaho and Utah: record of Cryogenian rifting and glaciation 425

LUND, K., ALEINIKOFF, J. N. & EVANS, K. V. The Edwardsburg Formation and related rocks, Windermere Supergroup, central Idaho, USA 437

MROFKA, D. & KENNEDY, M. The Kingston Peak Formation in the eastern Death Valley region 449

PETTERSON, R., PRAVE, A. R. & WERNICKE, B. P. Glaciogenic and related strata of the Neoproterozoic Kingston Peak Formation in the 459
Panamint Range, Death Valley region, California

CARTO, S. L. & EYLES, N. The deep-marine glaciogenic Gaskiers Formation, Newfoundland, Canada 467

CARTO, S. L. & EYLES, N. The Squantum Member of the Boston Basin, Massachusetts, USA 475

South America

CHEW, D. & KIRKLAND, C. The Chiquerı́o Formation, southern Peru 481

ALVARENGA, C. J. S., BOGGIANI, P. C., BABINSKI, M., DARDENNE, M. A., FIGUEIREDO, M. F., DANTAS, E. L., UHLEIN, A., SANTOS, R. V., 487
SIAL, A. N. & TROMPETTE, R. Glacially influenced sedimentation of the Puga Formation, Cuiabá Group and Jacadigo Group, and
associated carbonates of the Araras and Corumbá groups, Paraguay Belt, Brazil

FIGUEIREDO, M. F., BABINSKI, M. & ALVARENGA, C. J. S. The Serra Azul Formation, Paraguay Belt, Brazil 499

GUIMARÃES, J. T., MISI, A., PEDREIRA, A. J. & DOMINGUEZ, J. M. L. The Bebedouro Formation, Una Group, Bahia (Brazil) 503

MISI, A., KAUFMAN, A. J., AZMY, K., DARDENNE, M. A., SIAL, A. N. & DE OLIVEIRA, T. F. Neoproterozoic successions of the São Francisco 509
Craton, Brazil: the Bambuı́, Una, Vazante and Vaza Barris/Miaba groups and their glaciogenic deposits

PEDROSA-SOARES, A. C., BABINSKI, M., NOCE, C., MARTINS, M., QUEIROGA, G. & VILELA, F. The Neoproterozoic Macaúbas Group, 523
Araçuaı́ orogen, SE Brazil

ROCHA-CAMPOS, A. C., DE BRITO NEVES, B. B., BABINSKI, M., DOS SANTOS, P. R., DE OLIVEIRA, S. M. B. & ROMANO, A. Moema laminites: 535
a newly recognized Neoproterozoic (?) glaciogenic unit, São Francisco Basin, Brazil
viii CONTENTS

UHLEIN, A., ALVARENGA, C. J. S., DARDENNE, M. A. & TROMPETTE, R. R. The glaciogenic Jequitaı́ Formation, southeastern Brazil 541

PAZOS, P. J., RAPALINI, A. E., BETTUCCI, L. S. & TÓFALO, O. R. The Playa Hermosa Formation, Playa Verde Basin, Uruguay 547

PECOITS, E., GINGRAS, M. K. & KONHAUSER, K. O. Las Ventanas and San Carlos formations, Maldonado Group, Uruguay 555

PAZOS, P. J. & RAPALINI, A. The controversial stratigraphy of the glacial deposits in the Tandilia System, Argentina 565

Europe

HALVERSON, G. P. Glacial sediments and associated strata of the Polarisbreen Group, northeastern Svalbard 571

STOUGE, S., CHRISTIANSEN, J. L., HARPER, D. A. T., HOUMARK-NIELSEN, M., KRISTIANSEN, K., MACNIOCAILL, C. & BUCHARDT-WESTERGÅRD, B. 581
Neoproterozoic (Cryogenian–Ediacaran) deposits in East and North-East Greenland

RICE, A. H. N., EDWARDS, M. B., HANSEN, T. A., ARNAUD, E. & HALVERSON, G. P. Glaciogenic rocks of the Neoproterozoic Smalfjord and 593
Mortensnes formations, Vestertana Group, E. Finnmark, Norway

STODT, F., RICE, A. H. N., BJÖRKLUND, L., BAX, G., HALVERSON, G. P. & PHARAOH, T. C. Evidence of late Neoproterozoic glaciation in the 603
Caledonides of NW Scandinavia

NYSTUEN, J. P. & LAMMINEN, J. T. Neoproterozoic glaciation of South Norway: from continental interior to rift and pericratonic basins in 613
western Baltica

KUMPULAINEN, R. A. & GREILING, R. O. Evidence for late Neoproterozoic glaciation in the central Scandinavian Caledonides 623

KUMPULAINEN, R. A. The Neoproterozoic glaciogenic Lillfjället Formation, southern Swedish Caledonides 629

ARNAUD, E. & FAIRCHILD, I. J. The Port Askaig Formation, Dalradian Supergroup, Scotland 635

PRAVE, A. R. & FALLICK, A. E. The Neoproterozoic glaciogenic deposits of Scotland and Ireland 643

Australia

CALVER, C. R. Neoproterozoic glacial deposits of Tasmania 649

CORKERON, M. Neoproterozoic glacial deposits of the Kimberly Region and northwestern Northern Territory, Australia 659

GOSTIN, V. A., MCKIRDY, D. M., WEBSTER, L. J. & WILLIAMS, G. E. Mid-Ediacaran ice-rafting in the Adelaide Geosyncline and 673
Officer Basin, South Australia

HILL, A. C., HAINES, P. W. & GREY, K. Neoproterozoic glacial deposits of central Australia 677

JENKINS, R. J. F. Billy Springs glaciation, South Australia 693

PREISS, W. V., GOSTIN, V. A., MCKIRDY, D. M., ASHLEY, P. M., WILLIAMS, G. E. & SCHMIDT, P. W. The glacial succession of 701
Sturtian age in South Australia: the Yudnamutana Subgroup

WILLIAMS, G. E., GOSTIN, V. A., MCKIRDY, D. M., PREISS, W. V. & SCHMIDT, P. W. The Elatina glaciation (late Cryogenian), 713
South Australia

Index 723
List of Reviewers
Philip Allen Nick Eyles
Imperial College London, UK University of Toronto at Scarborough, Canada
Carlos Alvarenga Ian Fairchild
University of Brasilia, Brazil University of Birmingham, UK
Jose-Javier Alvaro Mikhail Fedonkin
Centro de Astrobiologia, Madrid, Spain Paleontological Institute, Moscow, Russia
John Arthurs Eric Font
Holywood, County Down, Ireland University of Lisbon, Portugal
Dov Avigad Hartwig Frimmel
The Hebrew University of Jerusalem, Israel University of Würzburg, Germany
Marly Babinsky Claudio Gaucher
University of Sao Paulo, Brazil Universidad de la República, Uruguay
Heinrich Bahlburg Vic Gostin
Westfälische Wilhelms-Universität, Muenster, Germany University of Adelaide, Australia
Richard Bailey Dimitriy Grazhdankin
Northeastern University, USA Institute of Petroleum Geology and Geophysics, Novosibirsk,
Dhiraj Banerjee Russia
University of Delhi, India Mike Hambrey
Angelo Miguel Stipp Basei Aberystwyth University, Wales, UK
University of Sao Paulo, Brazil Richard Hanson
Julie Bartley Texas Christian University, USA
State University of West Georgia Christoph Heubeck
Nicholas Beukes Freie Universität Berlin, Germany
University of Johannesburgh, South Africa Andrew Hill
Knut Bjorlykke Centro de Astrobiologia, Madrid, Spain
University of Oslo, Norway Rick Hiscott
Robert Blodgett Memorial University, Canada
United States Geological Survey, USA Paul Hoffman
Nick Butterfield Harvard University, USA/University of Victoria, Canada
Cambridge, UK John Howe
Clive Calver Scottish Marine Institute, UK
Mineral Resources, Tasmania Ganqing Jiang
Nikolay Chumakov University of Las Vegas, USA
Geological Institute, Russian Academy of Sciences Peter Johnson
Allan Collins Washington, DC, USA
University of Adelaide, Australia Sandra Kamo
Maurice Colpron University of Toronto, Canada
Yukon Geological Survey, Canada A. F. Kamona
Frank Corsetti University of Namibia, Namibia
University of Southern California, USA Alan Jay Kaufman
Lucieth Cruz Viera University of Maryland, USA
University of Sao Paulo, Brazil Brian Kendall
Bley Benjamin de Brito Neves Arizona State University, USA
University of Sao Paulo, Brazil Andy Knoll
Max Deynoux Harvard University, USA
CNRS, Strasbourg, France Risto Kumpulainen
Nicole Dobrzinsky Stockholm University, Sweden
Panterra Geoconsultants B. V., The Netherlands Alexandre Kuzmichev
Eugene Domack Geological Institute, Russian Academy of Sciences, Russia
Hamilton College, USA
Anton Kuznetsov
James Etienne Institute of Precambrian Geology and Geochronology,
Neftex, UK St. Petersburg, Russia
David J. A. Evans Erwan Le Guerroué
Durham University, UK Université de Rennes, France
x LIST OF REVIEWERS

Zheng-Xiang Li Victor Ramos


Curtin University of Technology, Australia University of Buenos Aires, Argentina
John Lindsay Pete Reid
Universities Space Research Association, USA Petratherm, Australia
Paul Link Claudio Riccomini
Idaho State University, USA University of Sao Paulo, Brazil
Karen Lund
United States Geological Survey, USA Hugh Rice
University of Vienna, Austria
Francis Macdonald
Harvard University, USA Ruben Rieu
Repsol YPF, Exploration and Production, Spain
Adam Maloof
Princeton University, USA Anthony Rocha-Campos
University of Sao Paulo, Brazil
Marcelo A. Martins Neto
University of Ouro Preto, Brazil Julius Sovetov
Andrey Maslov Institute of Oil and Gas Geology and Geophysics, Novosibirsk,
Institute of Geology and Geochemistry, Ekaterinburg, Russia Russia

Bill McClelland Anthony Spencer


University of Idaho, USA Statoil, Norway
Roy Miller Doug Sprinkel
Windhoek, Namibia Utah Geological Survey, USA
Aroldo Misi Robert Strachan
Federal University of Bahia, Brazil University of Portsmouth, UK
Malgorzata Moczydlowska-Vidal Nick Swanson-Hysell
Uppsala University, Sweden Princeton University, USA
Paul Myrow
Colorado College, USA Eric Thover
University of Western Australia
Afonso Noguiera
Universidade Federal do Pará, Brazil Elizabeth Turner
Laurentian University, Canada
Robert Pankhurst
British Geological Survey, UK Rob Van der Voo
University of Michigan, USA
Sandra Passchier
Montclair University, USA Pat Vickers-Rich
Pablo Pazos Monash University, Australia
University of Buenos Aires, Argentina Hilmar von Eynatten
Vic Pease University of Göttingen, Germany
University of Stockholm, Sweden
Stephen Warren
Sergei Pisarevsky University of Washington, USA
The University of Western Australia, Australia
George Williams
Boris Pokrovsky University of Adelaide, Australia
Geological Institute, Moscow, Russia
Gary Yeo
Michael Pope Denison Mines Corporation, Canada
Washington State University, USA
Susannah Porter Grant Young
University of California-Santa Barbara, USA University of Western Ontario, Canada

Anthony Prave Chuanming Zhou


University of St. Andrews, UK Nanjing Institute of Geology and Paleontology, Chinese Academy
of Science, China
Wolfgang Preiss
Department of Primary Industries and Resources of South Australia, Udo Zimmermann
Australia University of Stavanger, Norway
Chapter 1

The geological record of Neoproterozoic ice ages

EMMANUELLE ARNAUD1*, GALEN P. HALVERSON2,3 & GRAHAM SHIELDS-ZHOU4


1
School of Environmental Sciences, University of Guelph, Guelph, Ontario N1G 2W1, Canada
2
School of Earth and Environmental Sciences, The University of Adelaide, North Terrace, Adelaide, SA 5005, Australia
3
Present address: Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montreal, Quebec,
H3A 2A7, Canada
4
Department of Earth Sciences, University College London, Gower Street, London WC1E 6BT, UK
*Corresponding author (e-mail: earnaud@uoguelph.ca)

The IUGS- and UNESCO-funded International Geoscience Pro- to just before the renewal of interest in their significance after
gramme Project #512 (Neoproterozoic Ice Ages) was conceived 1998. Arnaud & Etienne provide a ‘user’s guide’ to the identifi-
to contribute towards a global synthesis of current geological cation of glacial influence in the rock record, with emphasis on the
data on the number, duration, extent, causes and consequences processes and sedimentary products found in various glaciated
of glacial episodes during the Neoproterozoic Era. IGCP 512 basins as well as some common issues encountered with determin-
attracted more than 200 scientists from over 30 countries, many ing the palaeoclimatic significance of commonly used indicators of
of whom provided their regional and specialist expertise on Neo- glacial palaeoenvironmental conditions. Geochemistry, in particu-
proterozoic successions around the world to the realization of lar, isotope chemostratigraphy, has been key to the recent revival
this volume. IGCP 512 focused on integrating various aspects of of interest in Neoproterozoic climate change and underpins
Neoproterozoic geology: geochronology, geochemistry, sedimen- models of both glaciation and global correlations. Halverson &
tary geology, biostratigraphy, palaeomagnetism and economic Shields-Zhou review Neoproterozoic chemostratigraphic records
geology. At its inaugural meeting on 27 August 2005 during the with a focus on how they have been applied to palaeoenvironmen-
International Association of Sedimentology conference on tal studies and to constrain both the number and relative timing of
glacial processes and products in Aberystwyth, Wales, IGCP 512 glacial events.
members decided to produce a volume that summarized existing Hoffman et al. then review the occurrence of chemical sedi-
data sets in a form similar to Earth’s Pre-Pleistocene Glacial ments and their depositional environments associated with the
Record by Hambrey & Harland (1981). Neoproterozoic glaciations. In particular, the role of the enigmatic,
An enormous amount of work has been carried out in the 12 post-glacial ‘cap carbonates’ and the reoccurrence of iron for-
years since the publication of Hoffman et al.’s (1998) paper on mations after a billion-year hiatus are explored here. Climate
the Snowball Earth hypothesis for Neoproterozoic glaciation affects how rocks are weathered, and this in turn influences the
(Fairchild & Kennedy 2007). The Snowball Earth hypothesis global carbon cycle, a key factor in modulating global climate
and, more generally, Neoproterozoic climate, have been the change. Bahlburg & Dobrzinski offer in this regard a more
topic of numerous special volumes, special sessions, a dedicated specific review of the application of the Chemical Index of Altera-
conference in Ascona (Switzerland) in 2006 (Shields 2006), and tion (CIA) to Neoproterozoic glacial deposits. If the Snowball
numerous documentaries. Motivated by this intense worldwide Earth hypothesis remains contentious, the following chapter
interest in the Neoproterozoic glaciations and an exploding should at least dispel any doubt that Neoproterozoic glaciation
body of research into the topic, this volume synthesizes the reached the tropics. In their review of palaeolatitudes of Neopro-
state-of-the-art in this now highly multidisciplinary research terozoic glacial deposits, Evans & Raub contribute a comprehen-
field. It is intended to facilitate the integration of data sets, sive compilation of locations and available palaeomagnetic
inspire new research projects, and inform ongoing work into the data for Neoproterozoic glacial deposits worldwide. Condon &
definition and subdivision of the Neoproterozoic timescale, Bowring also provide a much needed user’s guide to Neoprotero-
including selection of the Global Stratotype Section and Point zoic geochronology, with emphasis on the most commonly used
(GSSP) for the base of the Cryogenian Period. Despite such techniques, their strengths and weaknesses and the source of uncer-
lofty aims, any book such as this cannot claim to be complete, tainties that should be considered when using these data to con-
and there are indeed many gaps in our knowledge and also in strain the timing of climatic changes. In addition, Condon &
this book’s coverage, some of which are outlined in this Introduc- Bowring provide a summary of the key geochronological con-
tion and throughout the volume. straints on the timing and duration of glaciations and other signifi-
cant Neoproterozoic events related to isotopic excursions and
biological evolution. Grey et al. tackle the Neoproterozoic fossil
Book organization, format and terminology record, and in particular the question of how climate change
affected life on Earth. In their detailed review, they formulate a
This book contains ten introductory overview chapters followed Cryogenian biostratigraphy based on Australian data, which can
by 60 site- or succession-specific chapters. The multidisciplinary potentially be used as a template for such studies, generally still
overview chapters provide reviews of the study and interpretation in their infancy, in other regions and for other Proterozoic time
of Neoproterozoic glaciations. The first chapter by Hoffman periods. In the final overview chapter, Goddéris et al. provide a
reviews the history of research (1871 – 1997) into late Precambrian critical review of the widely differing climate models that have
glaciations from the first recorded discovery of Neoproterozoic been used to investigate the onset and melting of a Snowball
glaciogenic rocks at Port Askaig on Islay, SW Scotland in 1871 Earth and highlight that whereas initiation of a Snowball Earth is

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 1 –16. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.1
2 E. ARNAUD ET AL.

not difficult to accomplish in many models, how the Earth escapes non-glacial units described in the following boundary relations
from the icy grip of a snowball requires more study from a model- section. The chemostratigraphy section was meant to include geo-
ling perspective. chemical data within a stratigraphic context, including CIA values,
The remaining chapters, organized by current geography, d13C, d18O, d34S, 87Sr/86Sr, and other available isotopic data.
address specific sites from around the world where Neoproterozoic Some authors also included analyses of Rare Earth Elements,
deposits have been studied. The purpose of these chapters is to whole rock geochemistry, major and trace elements, Nd data,
provide a summary of available data and key references; they and Ce, Eu and Ir anomalies. The other characteristics section
are not comprehensive reviews. Each chapter was intended to was designed to cover economic deposits, biomarkers or any
follow a consistent format in order to cover concisely the following other notable feature not covered in the other sections. The palaeo-
topics and to facilitate cross chapter comparisons: structural frame- latitude and palaeogeography section was primarily designed to
work, regional stratigraphy, sedimentary characteristics of glacial discuss the palaeomagnetic data available for that succession in
strata and associated deposits, boundary relations with non-glacial order to consider palaeogeographic location of the deposits at
strata, chemostratigraphy, palaeomagnetism and palaeogeo- the time of deposition. Some authors also included provenance
graphy, geochronology and a discussion outlining interpretations data in the context of palaeogeography based on lithology or geo-
that can be inferred from these data. The structural framework chemistry. The geochronological constraints section was designed
section was designed to include a description of the overall struc- to report any available radiometric or biostratigraphic data that
tural and tectonic setting (such as cratons, types of sedimentary constrains the minimum or maximum age of the succession.
basin, regional-scale folds and faults) as well as the degree to These dates could come from the glacial deposits themselves or
which the sections have been modified by post-depositional tecton- from associated strata. Authors were also asked to discuss regional
ism and metamorphism. Authors were also asked to discuss the stratigraphic correlations based on radiometric data, chemostrati-
history of basin development in the region. The stratigraphy graphy, lithostratigraphy and/or biostratigraphy. Global strati-
section was meant to provide an overview of the relevant strati- graphic correlations were discouraged as these are discussed in
graphic units with comments about any lateral variations. The gla- the introductory chapters by Condon & Bowring (geochronology)
ciogenic and associated strata section was designed to include and Halverson & Shields-Zhou (chemostratigraphy). Lastly, the
descriptions of typical sedimentary characteristics of the glacio- discussion section was designed to include interpretations of
genic and associated strata such as ironstones and carbonates, available datasets with respect to palaeoenvironmental conditions,
with the nature of contact with the overlying and underlying timing of climate change and palaeogeography, showing how

Fig. 1.1. Map showing distribution of


Neoproterozoic glaciogenic successions
within Africa, as covered in this volume. (A)
Taoudeni Basin, NW Africa; (B) Katanga
Supergroup, central Africa; (C) West Congo
and Lindi/Ubangi Supergroups, central
Africa; (D) Otavi Group, northern Namibia;
(E) Witvlei Group, East-Central Namibia;
(F) The Chameis Gate Member, Namibia;
(G) The Kaigas and Numees formations,
South Africa and Namibia, (H) Karoetjes
Kop and Bloupoort formations, South
Africa.
Table 1.1. Summary of Neoproterozoic data sets from Africa

Lead author Glaciogenic strata Structural Chemostratigraphy Geobiology Economic Palaeomagnetism Geochronology
framework
d13C d18O d34S 87
Sr/86Sr Other U–Pb U– Pb Other
detrital direct

Shields-Zhou Basal diamictite of the Foreland basin Y Y Y Y Scarce Barite, BIF, Y K– Ar,
triad sequence, petroleum Rb– Sr
Taoudeni Basin, source rocks
diamictite of the triad
sequence, Volta Basin
Master Grand Conglomerat Fm., Synrift, foreland Y Ir Y Fe, Cu– Co, Zn– Y Y Y
Petit Conglomerat Fm. basin Pb–Cu –Ge–
Ga, Cu –Au,

INTRODUCTION
Ba–Fe, Pb –Zn
Tait Lower and Upper Passive margin Y Y Y Y Ar/Ar
Diamictite Formations,
West Congo
Supergroup;
diamictites of the
Lindi/Ubangi
Supergroup
Hoffman Chuos Fm., Ghaub Fm. Passive margin Y Y Y Y [Fe, Mn, Ba], Y
d11B, d44Ca
Prave Blaubeker Fm. Rift to passive Y Y Y Y reset Y
Frimmel Kaigas Fm., Numees Fm. Failed rift; passive; Y Y Y Y Pb, Zn Y Y Pb/Pb,
back-arc? Ar/Ar
Frimmel Chameis Gate Mb., Back-arc? Y Y Y Limited trace Pb/Pb,
Chameis Group elements Ar/Ar
Frimmel Karoetjes Kop Fm.; Rift; pasive margin? Y Y Y Limited trace Calcitic marble;
Bloupoort Fm. elements Cu, Fe, Mn;
Fe oxides

Note: ‘U –Pb direct’ means dates from volcanic intrusions or tuffs anywhere in the succession, not necessarily from the glaciogenic strata itself. ‘U– Pb detrital’ means dates from detrital zircons from within the succession. Data in
presumed correlative successions were not included.

3
4 E. ARNAUD ET AL.

the tectonic, stratigraphic, sedimentological, isotopic, palaeomag- Neoproterozoic successions and the inability to demonstrate seaso-
netic and geochronological data outlined in previous sections nal cyclicity. In the discussion section, authors were asked, where
support such interpretations. Whereas authors clearly favoured possible, to distinguish between the cap carbonate sequence that
certain interpretations, they were encouraged to discuss competing overlies the glacial deposits and the cap dolostone that occupies
interpretations and continued controversies. In an attempt to avoid the basal transgressive tract of the post-glacial sequence.
redundancy and maintain consistency, photographs were omitted Although the focus of the book is on Cryogenian glaciation,
from site-specific chapters, and are made available in a companion these glacial deposits and associated strata are best understood
online photo atlas (http://neoproterozoic-glaciations.weebly. within the context of their overall stratigraphic successions.
com). Although this format was generally followed, some flexi- Therefore, several site chapters also review associated later Neo-
bility was required to accommodate the complexity of certain proterozoic, Ediacaran-age deposits to highlight the carbonate
areas and the widely differing perspectives that reflect the diversity strata associated with the late Cryogenian glacial deposits and
of authors contributing to this volume. the evolution of early animals that followed the late Cryogenian
Authors were asked to avoid interpretive terms, such as tillite, and mid-Ediacaran glaciations.
varvite, and cap carbonate, in preference to descriptive terms, Every effort was made to include all the sites where Neoproter-
except in the discussion section where interpretations are pre- ozoic glaciogenic successions have been studied. In some cases,
sented. More specifically, the non-genetic term diamictite multiple distinct glaciogenic successions within a single region
was meant to be used to describe poorly sorted materials that were grouped into a single chapter, whereas in others, they were
contain a mixture of gravel-, sand- and mud-sized particles. For treated in separate chapters. For the Adelaide rift-basin in South
example, the term tillite was meant to be used in the discussion Australia, where extensive work has been done on each of the
section in referring to poorly-sorted materials that were demonstra- older (Sturtian and equivalent) and younger (Elatina and equival-
bly deposited by ice without subsequent disaggregation and flow. ent) glacial successions, the reviews are split into two chapters –
The use of the term varvites or varves for rhythmically laminated one for each glaciation.
mudstone that record seasonal fluctuations in ice cover was Unfortunately, not every site known to preserve evidence of
discouraged considering the lack of chronological control on Neoproterozoic glaciation is included in this volume. These

50

F I
G 40
J E
L
H
D K 30

O
20
N
C B M
A 10

N
0
0 1000 2000
km
40 50 60 70 80 90 100 110 120 130
Fig. 1.2. Map showing distribution of Neoproterozoic glaciogenic successions within Eurasia and the Nubian Shield, as covered in this volume. (A) Ayn Fm., Oman;
(B) Abu Mahara Group, Oman; (C) Tambien Group, northern Ethiopia; (D) northern Arabian– Nubian Shield (Egypt, Sudan and western Arabia); (E) Bokson Group,
East Sayan Mountains, Russian Federation; (F) North and Middle Urals, Russian Federation; (G) Nichatka Fm., central Siberia; (H) Baykonur Fm., Kazakhstan and
Kyrgystan; (I) Bol’shoy Patom Fm., central Siberia; (J) Marnya Fm., foothills of East Sayan mountains; (K) Tsagaan Oloom Fm., SW Mongolia; (L) Khubsugul Group,
northern Mongolia; (M) Blaini Fm., NW India; (N) Yangtze region, China; (O) Tarim Block, NW China.
Table 1.2. Summary of Neoproterozoic data sets in Eurasia and the Nubian Shield

Lead author Glaciogenic strata Structural Chemostratigraphy Geobiology Economic Palaeomagnetism Geochronology
framework
d13C d18O d34S 87
Sr/86Sr Other U– Pb U– Pb Other
detrital direct

Allen Ayn Fm., Shareef Fm. Rift Y CIA, MIA Y, unreliable Y Rb –Sr, K– Ar
Allen Ghubrah Fm., Fiq Fm. Rift Y CIA, MIA Y, unreliable Y Y
Stern Meriti Gp.?, Mahd Arc–back-arc, REE, Ce, Eu BIF, Au, Cu, Y
Gp.?, Atud & oceanic basin Zn, Ar,
Nuwaybah
diamictites
Miller Didikama Fm., Matheos Arc, intra-oceanic Y Y Y CIA, PIA, Y Y Sm–Nd, Pb/
Fm. (Tambien Gp.) platform d13C_TOC, Pb, Rb– Sr
Sr
Chumakov Zabit Fm., Kushatay Foreland basin Y Y Y Bauxite, Rb –Sr
Fm. (Bokson Gp.) Phosphorite
Chumakov Churochnaya Fm., Passive margin Y Y Y Y K –Ar, Sm–
Tany/ Koyva/Wil’va Nd, Rb –Sr,
formations, Lower Pb/Pb

INTRODUCTION
Starye Pechi
Subformation
Chumakov Nichatka Fm. Passive margin Y Y Y
Chumakov Baykonur Unknown Y Y Vanadium, Y
phosphorite
Chumakov Bol’shoy Patom Fm. Passive margin Y Y Y Y, reset K –Ar,Rb– Sr
Sovetov Marnya Fm. (Oselok Foreland basin Y Y Y Y Y, overprinted Ar/Ar
Gp.)
Macdonald Tsagaan Oloom Fm. Passive margin? Y Y Ce Y Y Y
Macdonald Ongoluk Fm., Khesen Rift? Y Y Phosphorite
Fm. (Khubsugul Gp.)
Etienne Blaini Fm. Passive margin Y Y Y Unreliable Y Y
Zhang Chang’an/Fulu Fm Rift Y Y TOC Y (Biomarker) Manganese, Y Y Nd (tDM)
(Jiangkou Gp); BIF
Nantuo Fm.
Zhu Polong Fm. Yutang/ Foreland basin Y Y Y Y
Yulmeinak diamictite
(Tarim Basin);
Beiyixi, Altungol,
Tereeken and
Hankalchough
diamictites (Tarim
Block); and others

Note: ‘U –Pb direct’ means dates from volcanic intrusions or tuffs anywhere in the succession, not necessarily from the glaciogenic strata itself. ‘U– Pb detrital’ means dates from detrital zircons from within the succession. Data in
presumed correlative successions were not included.

5
6 E. ARNAUD ET AL.

lacunae are due either to the inability to find an appropriate author and 87Sr/86Sr data sets to reconstruct ocean compositions and
or because little or no work has been published in that region since terrestrial weathering processes (Halverson & Shields-Zhou
the publication of Hambrey & Harland (1981). In such cases, 2011), these rocks have also proven fertile grounds for novel iso-
readers are referred to that earlier volume for more information topic approaches such as Ca and B isotopes, and multiple O
or, where appropriate, to more recent literature referenced in isotopes (e.g. Kasemann et al. 2004; Bao et al. 2008). In addition,
chapters covering neighbouring basins. Where possible, these the hypothesis generated renewed interest in the idea of global syn-
overlooked successions were incorporated into chapters covering chronous glaciation, its onset and initiation, which in turn
nearby regions or related deposits. Most are also listed in Evans prompted focused efforts to provide additional geochronological
& Raub’s review of the palaeomagnetic database. constraints. The development of new techniques such as the use
of Re –Os in dating organic shales, the proliferation of SHRIMP
and LA-ICP-MS analyses, and the refinement of the ID-TIMS
The current knowledge base U –Pb dating technique have simultaneously refined some age
constraints and seriously challenged conventional age models for
General trends key Neoproterozoic successions, particularly in Australia. The
glacial deposits themselves, as the most direct record of the
Many of the successions and regions covered in this book were global glaciations, have come under much greater scrutiny, and
reviewed in Hambrey & Harland (1981). Since that time, the great- many units have been re-evaluated using facies and sequence stra-
est advances in our knowledge have come from the fields of tigraphic analysis, incorporating advances in our understanding of
geochemistry, palaeomagnetism and geochronology, although glaciated basins made in the last 30 years. The role of tectonics in
palaeoenvironmental interpretations of many glaciogenic and modulating climate change during this time period and the role
overlying units have also been variably reinforced and challenged of basin setting in controlling the preservation and nature of the
in recent years. One of the novel aspects of the Snowball Earth sedimentary record has also been explored further (e.g. Eyles &
hypothesis was its integration of various geological data sets, Januszczak 2004; Allen 2007; Stern 2008).
specifically bringing attention to the geochemical signatures Some workers have focused on sites that previously had rela-
associated with the Neoproterozoic glacial deposits. As a result, tively little to no data, such as Mongolia, Alaska, Ethiopia, Egypt
geochemical analyses of associated carbonate rocks proliferated and western Arabia and some parts of Australia, Russian Federa-
from the late 1990s onwards. Although the main focus has been tion and South America, while some of the classic Neoproterozoic
in acquiring traditional stable isotope (namely d13C and d34S) sections have been the subject of additional study by several

Fig. 1.3. Map showing distribution of


Neoproterozoic glaciogenic successions
within North America, as covered in this
volume. (A) Hula Hula diamictite, Arctic
Alaska; (B) Tatonduk Inlier, Alaska–Yukon
border; (C) Windermere Supergroup,
Mackenzie Mountains, Canada; (D)
Windermere Supergroup, southern Canadian
Cordillera; (E) SE Idaho and Utah, USA; (F)
Edwardsburg Fm., central Idaho, USA; (G)
Kingston Peak Fm., eastern Death Valley
region, USA; (H) Kingston Peak Fm.,
Panamint Range, USA; (I) Gaskiers Fm.,
Newfoundland, Canada; (J) Squantum
Member, Boston Basin, USA.
Table 1.3. Summary of Neoproterozoic data sets in North America

Lead author Glaciogenic strata Structural Geochemistry Geobiology Economic Palaeomagnetism Geochronology
framework
d13C d18O d34S 87
Sr/86Sr Other U –Pb U– Pb Other
detrital direct

Macdonald Hula Hula Diamictite Unknown Y Y Pb/Pb


(detrital),
Rb– Sr
Macdonald Rapitan Gp.; Hay Rift to passive Y Y Y Y Ar/Ar
Creek Gp.
Hoffman Rapitan Gp.; Ice Rift to passive Y Y Y Y d57Fe, Ce/Ce*, Y Fe Y (Rap. Gp.) Pb/Pb
Brook Fm. d44Ca, Elemental

INTRODUCTION
(Stelfox Mb.) analysis, REE
Smith Toby Fm., Vreeland Rift; controversial Y Y Y TOC, Mo, V/Cr Y Cu Y Re–Os,
Fm (Windermere upper WSG Sm–Nd
Supergroup)
Link Pocatello Fm. Rift to passive Y CIA, TOC, Nd, Sr Y Y Y
(Idaho), Mineral
Fork Fm. (Utah)
Lund Edwardsburg Fm., Extensional Whole-rock Y
Moores Lake Fm. (inferred) geochem.
Mrofka Kingston Peak Fm. Extensional Y Y Y Y
(Eastern Death
Valley)
Petterson Kingston Peak Fm. Rift Y U, U– Th,
(Panamint Range) Iron ore
Carto Gaskiers Fm. Arc Y FeHR/FeT Y Y Y Y
Carto Squantum Mb. Extensional-rift, CIA Y Y Y Pb/Pb
intra-arc

Note: ‘U –Pb direct’ means dates from volcanic intrusions or tuffs anywhere in the succession, not necessarily from the glaciogenic strata itself. ‘U– Pb detrital’ means dates from detrital zircons from within the succession. Data in
presumed correlative successions were not included.

7
8 E. ARNAUD ET AL.

research groups, notably Svalbard, northern Norway, Scotland/ glaciations are now commonly identified by their relative timing
Ireland, the North American Cordillera (Death Valley, Idaho and and assumed correlation with well-studied localities. Mid-
western Canada), Namibia, various regions in Australia, and Ediacaran glacial deposits are widely referred to as Gaskiers,
China. As a result, some of these classic localities have become based on the eponymous glaciogenic formation on the Avalon
‘type’ localities for investigating Neoproterozoic glaciations and Peninsula of southeastern Newfoundland, which occurs just
their aftermath. below the first appearance of Ediacaran fossils c. 575 Ma (Nar-
With the increasingly global focus in research into the Neopro- bonne & Gehling 2003). Older, Cryogenian-aged glacial deposits
terozoic and ongoing efforts to define and normalize chronostrati- are commonly referred to as Sturtian (older) and Marinoan
graphic units in the Neoproterozoic, the terminology used to refer (younger) (e.g. Kennedy et al. 1998; Halverson et al. 2005) in an
to Neoproterozoic glaciations and, implicitly, their age and global adaptation of terms originally intended for specific sequences in
correlations has significantly evolved. For example, now largely the Adelaidean rift-basin of South Australia (Preiss 2000).
absent from the international lexicon are such previously main- Despite their proliferation and appeal, the application of these
stream terms as Sinian, Vendian and Varangerian, which refer terms usually hinges on inferred inter-regional correlations that
to overlapping subsets of Neoproterozoic time spanning glacial are commonly controversial and often poorly backed up by radio-
epochs. Because these intervals were defined based on regional metric data. Although biostratigraphy (Grey et al. 2011), chemos-
stratigraphic sequences with little basis other than the occurrence tratigraphy (Halverson & Shields-Zhou 2011) and magnetic
of glacial deposits, their use was inconsistent and often confusing. stratigraphy (Evans & Raub 2011) are increasingly able to
New informal terms have emerged in their place and partly as resolve relative age assignments and correlations, we have
a result of the intense efforts to determine the number of Neo- favoured reference to the emerging Neoproterozoic timescale
proterozoic glacial epochs. Specifically, the Neoproterozoic (i.e. Cryogenian and Ediacaran) over the terms Sturtian, Marinoan

Fig. 1.4. Map showing distribution of


Neoproterozoic glaciogenic successions
within South America, as covered in this
volume. (A) Chiquerio Fm., Peru; (B) Puga
Fm., Paraguay Belt, Brazil; (C) Serra Azul
Fm., Paraguay Belt, Brazil; (D) Bebedouro
Fm., Brazil; (E) São Francisco Craton,
Brazil; (F) The Macaúbas Group, SE Brazil;
(G) Moema laminites, São Francisco basin,
Brazil; (H) Jequitaı́ Fm., SE Brazil; (I) Playa
Hermosa Fm., Uruguay; (J) Last Ventanas
and San Carlos formations, Uruguay;
(K) Tandilia system, Argentina.
Table 1.4. Summary of Neoproterozoic data sets in South America

Lead author Glaciogenic strata Structural Chemostratigraphy Geobiology Economic Palaeomagnetism Geochronology
framework
d13C d18O d34S 87
Sr/86Sr Other U –Pb U–Pb Other
detrital direct

Chew Chiquerı́o Fm. Extensional, Y Y Fe, Cu, Au Y Y


not well
constrained
Alvarenga Puga Fm., Cuiabá Gp, Extensional Y Y Y Y Fe (BIF), Mn Y Y Pb/Pb, Nd
Jacadigo Gp.
Figueiredo Serra Azul Fm. Passive margin Ar/Ar, K– Ar,
Sm–Nd
Guimaraes Bebedouro Fm. Extensional Y Y Barite nodules; Y Pb/Pb
phosphate

INTRODUCTION
Misi Bambuı́ Gp., Una Not well Y Y Y (ltd) Fe– Mn, Y Re– Os, Pb/
Gp., Vazante Gp., constrained Phosphorite, Pb, Sm–Nd
Vaza Barris/Miaba fluorite, barite,
Gp. Pb–Zn, Zn
Pedrosa-Soares Serra do Catuni Fm., Rift to passive Mn, Fe, diamonds Y Y
Nova Aurora Fm., margin
Lower Chapada
Acaua Fm.
Rocha-Campos Moema Laminites Intracratonic basin Major and trace Kaolinite
elements, (ceramic)
REE, Eu, Ce,
Uhlein Jequitaı́ Fm. Cratonic/ Y Y Y Y Rb– Sr, Pb/Pb
intracratonic?
Pazos Playa Hermosa Fm. Not well Preliminary Ar/Ar, K– Ar,
constrained Rb– Sr
Pecoits Las Ventanas Fm., Strike–slip basin Y Y Y K– Ar, Rb –Sr,
San Carlos Fm. Ar/Ar
Pazos Sierra del Volcán Fm. Not well Y Y Y? Y Rb– Sr
constrained

Note: ‘U –Pb direct’ means dates from volcanic intrusions or tuffs anywhere in the succession, not necessarily from the glaciogenic strata itself. ‘U– Pb detrital’ means dates from detrital zircons from within the succession. Data in
presumed correlative successions were not included.

9
10 E. ARNAUD ET AL.

and Gaskiers in this volume in recognition that ages and corre- data (Condon & Bowring 2011) confirm the division of Cryogen-
lations of the Neoproterozoic glacial deposits remain controver- ian glaciations into two main episodes (c. 720–c. 660 Ma and
sial. This said, the fact that the lower boundary of the newly c. 650 –635 Ma), although older diamictites of unconfirmed
defined Ediacaran Period is defined as the contact between the gla- origin may hint at localized earlier Cryogenian glaciation. In con-
ciogenic Elatina Fm. (¼Marinoan) and the overlying Nuccaleena trast to the now reasonably well constrained late Cryogenian gla-
cap carbonate (Knoll et al. 2006) is suggestive of the chronostrati- ciation, the wide range of dates obtained from older Cryogenian
graphic importance of at least one of these Neoproterozoic glacial deposits long thought to be correlative (mainly in Australia
glaciations. and the North American cordillera), have called into question their
The definition of this boundary was bolstered by several new, equivalence and have been used to argue against global glaciation.
well placed and precise radiometric ages on basal Ediacaran cap Whereas the available data allow for a single, long-lived earlier
dolostones (i.e. the lower, transgressive, dolomite facies of the Cryogenian glaciation, it is clear that the implications and inter-
cap carbonates; Hoffman et al. 2011) that strengthen the case pretations of these many new ages are contentious and rapidly
for their global age equivalence. The trickle of new radiometric evolving. Therefore, references throughout the book to early,
ages is generally tightening the age constraints on this late earlier or middle Cryogenian glaciation should not be assumed
Cryogenian glaciation and the other glacial epochs and at the to indicate synchronous glaciation.
same time calibrating and helping to refine the secular evolution In recent years, with resource exploitation reaching unprece-
of seawater geochemical proxies (Halverson & Shields-Zhou dented peaks, there has also been increasing interest in the econ-
2011). New ages have demonstrated that some putative ‘Cryogen- omic potential of Neoproterozoic strata. Ediacaran rocks comprise
ian’ glacial deposits are in fact post-Cryogenian in age, such as the petroleum source rocks in China and Oman, and this has triggered
Gaskiers unit in Newfoundland, while the purported ‘Cambrian’ interest in correlative post-glacial successions in Africa, India,
glaciation in the Taoudeni Basin is now regarded unequivocally Australia, Brazil and Russia. Phosphorus resources have also
as Cryogenian in age. Recent U –Pb and increasingly Re –Os become a vulnerable commodity worldwide as Moroccan deposits

Fig. 1.5. Map showing distribution of


Neoproterozoic glaciogenic successions
within Europe, as covered in this volume.
(A) Polarisbreen Group, NE Svalbard; (B)
NE Greenland; (C) Smalfjord and
Mortensnes formations, Finnmark, Norway;
(D) Caledonides, NW Scandinavia; (E)
Moelv and Koppang formations, southern
Norway; (F) Caledonides, central
Scandinavia; (G) Lillfjället Fm., southern
Swedish Caledonides; (H) Port Askaig Fm.,
Scotland; (I) Scotland and Ireland.
Table 1.5. Summary of Neoproterozoic data sets in Europe

Lead author Glaciogenic strata Structural Chemostratigraphy Geobiology Economic Palaeomagnetism Geochronology
framework
d13C d18O d34S 87
Sr/86Sr Other U–Pb U –Pb Other
detrital direct

Halverson Petrovbreen Mb., Thermally Y Y Y D17O Y


Wilsonbreen Fm. subsiding
(Polarisbreen Group) margin
Stouge Ulvesø Fm., Storeelv Fm. Rift to passive Y TC, TOC, TS Y Y/unreliable?
and Støvfanget Fm.
Rice Smalfjord Fm., Rift to passive Y Y Y Rb –Sr
Mortensnes Fm.

INTRODUCTION
Stodt Possible correlatives of Passive margin Y (ltd) Y
the Mortensnes Fm.
Nystuen Moelv Fm., Koppang Rift; pericratonic Y (ltd) Y (ltd) Lu–Hf Y Y Rb –Sr, Re– Os
Fm. shelf
Kumpulainen Längmarkberg Fm., Rift to passive
Risback Gp. margin
Diamictite?
Kumpulainen Lillfjället Fm. Rift to passive Ar/Ar
margin?
Arnaud Port Askaig Fm. Extensional Y Y CIA Y Pb– Zn, magnetite Y Pb/Pb, Sm– Nd
Prave Dalradian Supergroup Rift Y Y Y Y
overview-Port
Askaig Fm.;
Stralinchy-Reelan;
MacDuff/Loch na
Cille/Inishowen
boulder beds

Note: ‘U –Pb direct’ means dates from volcanic intrusions or tuffs anywhere in the succession, not necessarily from the glaciogenic strata itself. ‘U– Pb detrital’ means dates from detrital zircons from within the succession. Data in
presumed correlative successions were not included.

11
12 E. ARNAUD ET AL.

become the only widely traded source of phosphorus for fertilizer. over the Snowball Earth hypothesis. Prave et al. (2011) have pro-
Ediacaran-age phosphorites of China are the only other world- vided a useful review of the hitherto poorly documented Witvlei
class phosphorite resource, sparking interest in possibly correla- Group in east-central Namibia (northern Kalahari craton), which
tive, glacially associated phosphorites in West Africa, South provides a firmer basis for correlation with the numerous Neopro-
America and Australia. In addition to being an oceanographic terozoic glacial deposits elsewhere on the Kalahari craton (see the
curiosity of the Cryogenian Period, the widespread deposition chapters by Frimmel). Whereas few geochronological or geochem-
of iron and manganese deposits may also have economic impli- ical data had been available from the West Congo belt until
cations for those regions where they are found (South America, recently, new research has provided key data that strengthen
northwestern Canada, India, China and Australia; Hoffman glacial interpretations and correlations across the belt and else-
et al. 2011). where on the Congo craton (Tait et al. 2011), such as the Katangan
Significant advances have been made in all regions since the Supergroup, best known for its world class Cu deposits (Master &
publication of Hambrey & Harland (1981). However, with new Wendorff 2011).
data always come new questions and controversies. Some of the
notable advances and contentious issues are highlighted below. Eurasia –Nubian Shield
Readers interested in specific glaciogenic successions are encour-
aged to read chapters from neighbouring sites as controversies are This region encompasses Neoproterozoic glaciogenic successions
sometimes most apparent when viewing the geological record found throughout the Russian Federation, Mongolia, China and
from a regional vantage point. India, as well as those from Oman, the shores of the Red Sea
and Ethiopia (Fig. 1.2, Table 1.2). Excellent preservation of Cryo-
Africa genian glacial deposits in northern Oman have allowed detailed
sedimentological and sequence stratigraphic analysis to be
Neoproterozoic glaciogenic successions reported from Africa carried out over the last several decades. The results have shed
are located primarily in the NW, central and southern regions, light on the environmental conditions during Neoproterozoic gla-
with a cluster of sites in Namibia (Fig. 1.1, Table 1.1). These ciations, with some more recent geochemical and geochronologi-
occur in a wide range of tectonic settings, with few sites contain- cal analyses contributing a relatively comprehensive data set for
ing biostratigraphic or palaeomagnetic data. Geochemical and this region. Unfortunately, palaeomagnetic studies in Oman have
geochronological studies have been carried out at most sites, thus far yielded no reliable palaeolatitudes. In contrast, studies in
with the most complete data sets found in the Taoudeni Basin in Ethiopia and the shores of the Red Sea have focused on geochem-
NW Africa and the Otavi Group of Namibia. The glacial deposits ical and geochronological aspects of Neoproterozoic geology,
in the former (Shields-Zhou et al. 2011) remain significant for making this region an ideal candidate for future work in sedimen-
being among the few to preserve unequivocal terrestrial glacial tology and stratigraphy to confirm glaciogenic conditions existed
deposits, which are now known to be age-equivalent to other end- in these regions and maximize the impact of these new radiometric
Cryogenian glaciogenic units around the world. Robust geochro- ages. The basins of the Arabian –Nubian Shield are unique in that
nological constraints spanning the two glaciogenic intervals of they occur in arc or back-arc basins, a relatively rare tectonic
the Otavi Group are particularly notable. The Otavi Group has setting for glaciated Neoproterozoic basins.
been the subject of intense study over the past two decades, and Glaciogenic successions in the Russian Federation and neigh-
despite lingering controversy over interpretation of glacial depos- bouring Mongolia, Kazakhstan and Kyrgyzstan are found
its in the succession (Hoffman 2011), plays a central role in debates within primarily passive margin and foreland basin settings, with

Fig. 1.6. Map showing distribution of


Neoproterozoic glaciogenic successions
within Australia, as covered in this volume.
(A) Tasmania; (B) the Kimberly region and
NW Northern Territory; (C) Mid-Ediacaran
ice rafted deposits, Adelaide Geosyncline
and Officer Basin; (D) Central Australia; (E)
Billy Springs Fm.; (F) Yudnamutana
Subgroup (Sturtian); (G) Elatina Fm.
Table 1.6. Summary of Neoproterozoic data sets in Australia

Lead author Glaciogenic strata Structural Chemostratigraphy Geobiology Economic Palaeomagnetism Geochronology
framework
d13C d18O d34S 87
Sr/86Sr Other U– Pb detrital U–Pb Other
direct

Calver Cottons Breccia, Julius Rift, epicratonic Y Y Y Y Calcic scheelite Unpublished Y Y Re–Os,
R. Mbr., Croles Hills basin skarns Nd–Sm
Diamictite; Wedge
R. beds, Cotcase Ck. Fm.
Corkeron Walsh Fm., Landrigan/ Unknown Y Y Unreliable Y Y Rb– Sr
Egan Fms., Fargoo/

INTRODUCTION
Moonlight Valley Fm.
Gostin Bunyeroo Fm., Dey Dey Thermally Y Atomic H/C Y Y Y Rb– Sr
Mudstone subsiding basin; ratios
platformal
Hill Glaciogenic deposits of Intracratonic sag Y Y Y Y Y Re–Os
Central Australia basin
(Officer, Amadeus,
Ngalia, Georgina Basins
Jenkins Billy Springs Fm. Rift complex Y Unpub. Y Y (ltd)
Preiss Yudnamutana Subgroup Rift Y Y Y Y Elemental Y Rb– Sr,
(Sturt and correlatives) analysis, REE, Re–Os
Ironstone
geochemistry
Williams Yerelina Subgroup, Rift Y Y Y Y Y Re–Os
Elatina Fm.

Note: ‘U –Pb direct’ means dates from volcanic intrusions or tuffs anywhere in the succession, not necessarily from the glaciogenic strata itself. ‘U– Pb detrital’ means dates from detrital zircons from within the succession. Data in
presumed correlative successions were not included.

13
14 E. ARNAUD ET AL.

geochemical data (primarily carbon, but also oxygen and strontium South America
isotopes in some cases) as well as important biostratigraphic data.
Interestingly, these studies have led to some contradictory strati- Neoproterozoic glaciogenic successions in South America are
graphic correlations across neighbouring basins (see chapters on primarily clustered in the Paraguay Belt and on the São
the Bokson Group and Khubsugul Basin). Chumakov (2011) Franscisco craton of Brazil, with other sections reported from
suggests that two diamictite-bearing horizons within the Bokson Uruguay, Argentina and Peru (Fig. 1.4, Table 1.4). The Araras
Group in the East Sayan Mountains are likely Ediacaran in age Group in the northern Paraguay Belt, in particular, has been the
based on biostratigraphic data, whereas Macdonald & Jones subject of interdisciplinary studies and preserves some of the
(2011) suggest that the two horizons in the neighbouring Mongo- best evidence for Ediacaran glaciation in South America (Figueir-
lian basin correspond to an early and late Cryogenian glacial edo et al. 2011). New entries to the South American database
episode. Dispute also remains over the glacial origin of the include the sections in Peru, Uruguay and Argentina as well as
younger Mongolian diamictite-bearing unit. Nonetheless, the tec- the Moema laminites in Brazil. Despite some new geochronologi-
tonic setting of the region makes these basins suitable targets for cal data, there remain significant challenges in regional strati-
future geochronological studies, which may help to resolve these graphic correlation schemes, limited information regarding
stratigraphic inconsistencies. The chapter on the Marnya Fm. in palaeolatitude, and uncertainty regarding tectonic setting and
the East Sayan Mountains (Sovetov 2011) is notable for its review glacial origin of some of the units described herein. These chal-
of regional stratigraphic correlation and useful comparison of the lenges are manifested in significant controversies over the corre-
Russian stratigraphic scheme with that of the more widely used lation and ages of Proterozoic glacial deposits across Brazil
International Subcommission on Stratigraphy (see also Table (Misi et al. 2011). Despite uncertainties in age, the correlations
27.2). Palaeogeographic reconstructions in northern Eurasia are between the Bambuı́ and Una groups have been strengthened
hampered by the lack of reliable palaeomagnetic data, although (Guimarães et al. 2011).
ongoing studies in Mongolia show some promise.
The Blaini Fm. of India did not receive much attention until
the more recent studies reported in this volume, which include Europe
new sedimentological and geochemical data. In contrast, the gla-
ciogenic successions of the Yangtze region (rift setting) and The Neoproterozoic glaciogenic successions of Europe described
Tarim Block (foreland basin) have a long history of Neo- here are primarily those found in extensional and passive margin
proterozoic research that has resulted in extensive data sets for basins of Scandinavia, Svalbard, Greenland and the British Isles
geochemistry, geobiology and biostratigraphy, palaeomagnetism (Fig. 1.5, Table 1.5). Although most of these sites were described
and geochronology. Notably, a U– Pb zircon age from an ash in Hambrey & Harland (1981), new data have become available in
bed in the Yangtze region has provided a definitive date for the recent decades, including additional geochemical data from most
end of the late Cryogenian glaciation (c. 635 Ma), whereas max- of these successions, additional information on associated carbon-
imum and syn-glacial ages in China and elsewhere have shown ate strata and some additional geochronological constraints. Most
that this late Cryogenian glaciation lasted ,10 Ma (Condon & successions in the region have been significantly impacted by
Bowring 2011). In addition, palaeomagnetic data indicating low Caledonian tectonics and hence lack reliable palaeomagnetic
palaeolatitudes for the Nantuo diamictite have one of the highest data directly related to the glacial deposits. Robust geochronologi-
reliability ratings of available palaeomagnetic data, as discussed cal constraints are also largely lacking, making regional strati-
by Evans & Raub (2011). Lastly, novel geobiological and biostra- graphic correlations relatively tenuous, despite some advances
tigraphic studies have demonstrated the preservation of animal using chemostratigraphic data. Interestingly, ongoing studies in
embryos in the Ediacaran-age Doushantuo Fm. of the Yangtze Greenland have led some researchers to propose that the Greenland
region (Hagadorn et al. 2006). and Svalbard sections are not as intimately connected palaeogeo-
graphically as was previously proposed.

North America
Australia
Many of the Neoproterozoic glaciogenic successions in North
America are found along the length of the North American The Neoproterozoic glaciogenic successions in Australia covered
Cordillera in rift basins and passive margin settings, whereas in this volume are very similar to the entries found in Hambrey
others are found on exotic terranes in Arctic Alaska and along & Harland (1981) over 30 years ago (Fig. 1.6, Table 1.6).
the eastern margin of North America (Fig. 1.3, Table 1.3). The Recent advances include the collection of additional geochemical
Cordilleran sections generally have a long history of research, and biostratigraphic data, continued state-of-the-art palaeomagnet-
but have enjoyed renewed interest and have been the source of ism research with confirmation of low-latitude glaciation, and
much new data. While many researchers have attempted to additional geochronological constraints that have prompted abun-
correlate these widely separated sections, recent studies have dant discussion about the duration and timing of the older Cryo-
highlighted the possible diachroneity of Neoproterozoic glacia- genian glaciation. Entries for South Australia are focused on
tions and rifting related to Rodinian break-up. New additions to individual glacial events (Preiss et al. 2011; Williams et al.
the North American database include the preliminary results of 2011), with new evidence for Ediacaran-aged glaciation within
ongoing multidisciplinary studies in Alaska and the Yukon the Adelaidean rift-basin (Gostin et al. 2011; Jenkins 2011). Cor-
including geochronological and palaeomagnetic data (Macdonald keron, Hill et al. and Calver provide timely and useful syntheses of
2011; Macdonald & Cohen 2011); new geochronological data the separate Neoproterozoic successions in the Kimberly region,
from Idaho (Link & Christie-Blick 2011; Lund et al. 2011) and ‘Centralian Superbasin’ and Tasmania, respectively.
the Canadian Cordillera (Smith et al. 2011; Macdonald et al.
2010); detailed mapping and geochemical data from the Death
Valley region; a better understanding of the tectonic development Calibrating Neoproterozoic change
of the southern Cordillera margin based on work in Idaho,
Utah and Death Valley; the analysis of a deepwater ‘cap’ car- One of the prominent developments in stratigraphy over the last
bonate in the southern Canadian Cordillera; and influential decade has been the attempt to establish consistent, rock-based
studies on the impressive Ediacaran biota and ocean redox in periods of geological time beneath the Cambrian Period. Neopro-
Newfoundland. terozoic glaciation played a pivotal role in this when the Ediacaran
INTRODUCTION 15

Period, which was the first period of the geological timescale to be like to thank all the reviewers (see list of reviewers) who provided meaningful
established for a hundred years, was defined as beginning directly and constructive comments and improved the individual chapters. I. Fairchild
after the widespread, end-Cryogenian (¼ Elatina or Marinoan) gla- kindly provided the cover photo and the historical photograph for the dedication.
ciation. Now, international efforts are under way to further subdi- E. Arnaud would also like to thank those who provided guidance in developing
vide the Ediacaran Period and define the underlying Cryogenian the instructions for authors with suggested content required of specific sections:
Period. In this endeavour, the integrated approach taken by many namely D. Condon, N. Eyles, G. Halverson, B. MacGabhann, P. Martini,
T. Raub, G. Shields-Zhou & R. van der Voo. The following regional coordinators
of the authors in the present volume will be key as a global strati-
also assisted in tracking down authors and names of Neoproterozoic glaciogenic
graphic framework can only be established in the Precambrian by
successions in their regions: P. Allen, C. Calver, N. Chumakov, M. Corkeron,
meshing together sparse fossil information with high-resolution N. Eyles, V. Gostin, P. Hoffman, C. Hoffmann, G. Jiang, K. Kristiansen,
chemostratigraphy, strategically placed age constraints and litho- R. Kumpulainen, P. Link, U. Linnemann, J. Meert, N. Miller, P. Pazos,
stratigraphic markers. T. Prave, Q. R. Zhang, H. Rice, T. Rocha Campos, S. Stouge, M. Wendorff
At present, the base of the Cryogenian Period is defined and S. Xiao. We thank J. Howe for his contribution as the Society Books
chronometrically at 850 Ma, but few age constraints exist to ident- Editor and for his timely review of this chapter. Angharad Hills at the Geological
ify this level in sedimentary successions from around the world. Society of London Publishing House kept us on track and provided editorial
Although, the base of a new chronostratigraphically defined Cryo- assistance throughout the project. This memoir represents a contribution of the
genian Period is yet to be established, any GSSP is likely to be IUGS-UNESCO sponsored International Geoscience Programme (IGCP)
placed ‘beneath the oldest clearly glaciogenic deposits in a Neo- Project # 512.
proterozoic succession’ at an outcrop horizon with ‘proven poten-
tial for global C- and Sr-isotope stratigraphic correlation and
preferably be amenable to microfossil biostratigraphy, isotope References
geochronology and other forms of global correlation such as mag-
netostratigraphy’ (Neoproterozoic Subcommission 2009 Annual Allen, P. A. 2007. The Huqf Supergroup of Oman: basin development
Report). This emphasis on glacial strata will likely lead to a and context for Neoproterozoic glaciation. Earth-Science Reviews,
shorter Cryogenian Period more in line with its Phanerozoic 84, 139– 185.
counterparts, and with a base close in age to the Dunn et al. Bao, H., Lyons, J. R. & Zhou, C. 2008. Triple oxygen isotope evidence
(1971) estimate of c. 750 Ma. The unique Neoproterozoic palaeo- for elevated CO2 levels after a Neoproterozoic glaciation. Nature,
climate, which underpins this volume, has already cemented the 453, 504–506.
place of Neoproterozoic ice ages in the international geological Chumakov, N. M. 2011. Glacial deposits of the Bokson Group, East
timescale (Ogg et al. 2008) and will continue to be central in the Sayan Mountains, Buryatian Republic Russion Federation. In:
Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
subdivision of Proterozoic time.
logical Record of Neoproterozoic Glaciations. Geological Society,
London, Memoirs, 36, 285–288.
Condon, D. J. & Bowring, S. A. 2011. A user’s guide to Neoproterozoic
Future work geochronology. In: Arnaud, E., Halverson, G. P. & Shields-
Zhou, G. (eds) The Geological Record of Neoproterozoic Glacia-
This memoir was inspired by Earth’s Pre-Pleistocene Glacial tions. Geological Society, London, Memoirs, 36, 135– 149.
Record (Hambrey & Harland 1981), and it is hoped that it can Dunn, P. R., Thomson, B. P. & Rankama, K. 1971. Late Pre-Cambrian
serve similarly as a long-lasting reference work that provides glaciation in Australia as a stratigraphic boundary. Nature, 231,
useful summaries of most Neoproterozoic glacial successions, 498– 502.
while also highlighting the state-of-the art in the study of these Evans, D. A. D. & Raub, T. D. 2011. Neoproterozoic glacial palaeo-
rocks. The large number of chapters and impressive list of contri- latitudes: a global update. In: Arnaud, E., Halverson, G. P. &
butors highlight the extraordinary interest in the Neoproterozoic Shields-Zhou, G. (eds) The Geological Record of Neo-
and the vast amount of research carried out over the past 30 proterozoic Glaciations. Geological Society, London, Memoirs, 36,
years. At the same time, the results of this research have revealed 93– 112.
gaping holes in our understanding of the Neoproterozoic and Eyles, N. & Januszczak, N. 2004. Interpreting the Neoproterozoic
aroused compelling new questions. Many glaciogenic units glacial record; the importance of tectonics. In: Jenkins, A.,
described in Hambrey & Harland (1981) have received little or McMenamin, M. A. S., McKey, C. P. & Sohl, L. (eds) The
no attention since that time, such as the widespread diamictites Extreme Proterozoic: Geology, Geochemistry and Climate. Ameri-
of the Sino-Korean craton, which are still of uncertain age and can Geophysical Union, Geophysical Monograph 146, 125–144.
origin. Radiometric age constraints on most glacial units remain Fairchild, I. J. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the
sparse or non-existent, which poses a persistent challenge to inter- Earth System. Journal of the Geological Society, London, 164,
895– 921.
regional correlations. Many important advances have been in the
Figueiredo, M. F., Babinski, M. & Alvarenga, C. J. S. 2011. The Serra
biostratigraphy of Cryogenian and Ediacaran successions, but
Azul Formation, Paraguay Belt, Brazil. In: Arnaud, E., Halverson,
most sections remain poorly studied. In addition, a welter of excit- G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
ing new geochemical methods, such as the D47 palaeothermometer terozoic Glaciations. Geological Society, London, Memoirs, 36,
and transition-metal stable isotopes, await rigorous application to 499– 502.
the Neoproterozoic. Clearly, much remains to be done, but in Gostin, V. A., McKirdy, D. M., Webster, L. J. & Williams, G. E.
almost every region of the world there are now interdisciplinary 2011. Mid-Ediacaran ice-rafting in the Adelaide Geosyncline and
teams of scientists at work in what has become a global endeavour Officer Basin, South Australia. In: Arnaud, E., Halverson, G. P.
not only to explore the extremes of past climate change, but also & Shields-Zhou, G. (eds) The Geological Record of Neopro-
to evaluate the link between eukaryotic evolution and global terozoic Glaciations. Geological Society, London, Memoirs, 36,
glaciations. 673– 676.
Grey, K., Hill, A. C. & Calver, C. 2011. Biostratigraphy and
We wish to thank all the authors who made this collection possible, for agreeing to stratigraphic subdivision of Cryogenian successions of Australia
write these contributions amidst many other writing commitments, and patiently in a global context. In: Arnaud, E., Halverson, G. P. &
awaiting the completion of the volume. We particularly thank N. Chumakov and Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
Q. R. Zhang for promptly contributing to the original template papers that were Glaciations. Geological Society, London, Memoirs, 36, 113– 134.
essential in establishing the format for all subsequent regional chapters. We are Guimarães, J. T., Misi, A., Pedreira, A. J. & Dominguez, J. M. L. 2011.
also especially grateful to P. Hoffman, F. Macdonald, M. Zhu, R. Stern and The Bebedouro Formation, Una Group, Bahia (Brazil). In: Arnaud,
J. Tait for acquiescing to last-minute requests for contributions. We would also E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
16 E. ARNAUD ET AL.

Record of Neoproterozoic Glaciations. Geological Society, London, Macdonald, F. A., Schmitz, M. D. et al. 2010. Calibrating the Cryo-
Memoirs, 36, 503– 508. genian. Science, 327, 1241–1243.
Hagadorn, J. W., Xiao, S. et al. 2006. Integrated X-ray insights into Master, S. & Wendorff, M. 2011. Neoproterozoic glaciogenic diamic-
cellular and subcellular structures of Neoproterozoic animal tites of the Katanga Supergroup, Central Africa. In: Arnaud, E.,
embryos. Science, 314, 291–294. Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
Halverson, G. P. & Shields-Zhou, G. 2011. Chemostratigraphy Record of Neoproterozoic Glaciations. Geological Society, London,
and the Neoproterozoic glaciations. In: Arnaud, E., Halverson, Memoirs, 36, 173–184.
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro- Misi, A., Kaufman, A. J., Azmy, K., Dardenne, M. A., Sial, A. N. &
terozoic Glaciations. Geological Society, London, Memoirs, 36, de Oliveira, T. F. 2011. Neoproterozoic successions of the São
51 –56. Francisco Craton, Brazil: The Bambuı́, Una, Vazante and Vaza
Halverson, G. P., Hoffman, P. F., Schrag, D., Maloof, A. C. & Rice, Barris/Miaba groups and their glaciogenic deposits. In: Arnaud,
A. H. 2005. Towards a Neoproterozoic composite carbon isotope E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
record. Geological Society of America Bulletin, 117, 1181–1207. Record of Neoproterozoic Glaciations. Geological Society, London,
Hambrey, M. J. & Harland, W. B. (eds) 1981. Earth’s Pre-Pleistocene Memoirs, 36, 509–522.
Glacial Record. Cambridge Earth Science series. Cambridge Univer- Narbonne, G. M. & Gehling, J. G. 2003. Life after snowball: the oldest
sity Press, Cambridge. complex Ediacaran fossils. Geology, 31, 27 – 30.
Hoffman, P. F. 2011. A history of Neoproterozoic glacial geology, 1871– Ogg, J. G., Ogg, G. & Gradstein, F. M. 2008. The Concise Geological
1997. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) Time Scale. Cambridge University Press, Cambridge.
The Geological Record of Neoproterozoic Glaciations. Geological Prave, A. R., Hoffmann, K.-H., Hegenberger, W. & Fallick, A. E.
Society, London, Memoirs, 36, 17 –37. 2011. The Witvlei Group of East-Central Namibia. In: Arnaud, E.,
Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P. Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
1998. A Neoproterozoic snowball Earth. Science, 281, 1342– 1346. Record of Neoproterozoic Glaciations. Geological Society, London,
Hoffman, P. F., Macdonald, F. A. & Halverson, G. P. 2011. Chemical Memoirs, 36, 211–216.
sediments associated with Neoproterozoic glaciation: iron formation, Preiss, W. V. 2000. The Adelaide Geosyncline of South Australia and
cap carbonate, barite and phosphorite. In: Arnaud, E., Halverson, its significance in Neoproterozoic continental reconstruction.
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neo- Precambrian Research, 100, 21 – 63.
proterozoic Glaciations. Geological Society, London, Memoirs, 36, Preiss, W. V., Gostin, V. A., McKirdy, D. M., Ashley, P. M., Wil-
67 –80. liams, G. E. & Schmidt, P. W. 2011. The glacial succession of
Jenkins, R. J. F. 2011. Billy Springs Glaciation, South Australia. In: Sturtian age in South Australia: the Yudnamutana Subgroup. In:
Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo- Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
logical Record of Neoproterozoic Glaciations. Geological Society, logical Record of Neoproterozoic Glaciations. Geological Society,
London, Memoirs, 36, 693–699. London, Memoirs, 36, 701–712.
Kasemann, S. A., Hawkesworth, C. J., Prave, A. R., Fallick, A. E. & Shields, G. A. 2006. Snowball Earth is dead! Long live Snowball Earth!
Pearson, P. 2005. Boron and calcium isotope composition in Neo- Episodes, 29, 287–288.
proterozoic carbonate rocks from Namibia: evidence for extreme Shields-Zhou, G. A., Deynoux, M. & Och, L. 2011. The Record of Neo-
environmental change. Earth and Planetary Sciences Letters, 231, proterozoic Glaciation in the Taoudéni Basin, NW Africa. In:
73 –86. Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K. H. & logical Record of Neoproterozoic Glaciations. Geological Society,
Arthur, M. 1998. Two or four Neoproterozoic glaciations? London, Memoirs, 36, 163–171.
Geology, 26, 1059–1063. Smith, M. D., Arnaud, E., Arnott, R. W. C. & Ross, G. M. 2011.
Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blick, N. The record of Neoproterozoic glaciation in the Windermere Super-
2006. The Ediacaran Period: a new addition to the geologic time group, southern Canadian Cordillera. In: Arnaud, E., Halverson,
scale. Lethaia, 39, 13 – 30. G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
Link, P. K. & Christie-Blick, N. 2011. Neoproterozoic strata of South- terozoic Glaciations. Geological Society, London, Memoirs, 36,
eastern Idaho and Utah: record of Cryogenian rifting and glaciation. 413– 423.
In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Sovetov, J. K. 2011. Late Neoproterozoic (Vendian) glaciogenic deposits
Geological Record of Neoproterozoic Glaciations. Geological in the Marnya Formation, Oselok Group, in the foothills of the East
Society, London, Memoirs, 36, 425–436. Sayan Range, Southwestern Siberian craton. In: Arnaud, E., Hal-
Lund, K., Aleinikoff, J. N. & Evans, K. V. 2011. The Edwardsburg For- verson, G. P. & Shields-Zhou, G. (eds) The Geological Record
mation and related rocks, Windermere Supergroup, central Idaho, of Neoproterozoic Glaciations. Geological Society, London,
U.S.A. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. Memoirs, 36, 317–329.
(eds) The Geological Record of Neoproterozoic Glaciations. Geo- Stern, R. J. 2008. Neoproterozoic crustal growth: the solid Earth system
logical Society, London, Memoirs, 36, 437– 447. during a critical episode of Earth history. Gondwana Research, 14,
Macdonald, F. A. 2011. The Hula Hula diamictite and Katakturuk Dolo- 33 – 50.
mite, Arctic Alaska. In: Arnaud, E., Halverson, G. P. & Shields- Tait, J., Delpomdor, F., Préat, A., Tack, L., Straathof, G. & Kanda
Zhou, G. (eds) The Geological Record of Neoproterozoic Glacia- Nkula, V. 2011. Neoproterozoic Sequences of the West Congo and
tions. Geological Society, London, Memoirs, 36, 379–387. Lindi/Ubangi Supergroups in the Congo Craton, central Africa. In:
Macdonald, F. A. & Cohen, P. A. 2011. The Tatonduk inlier, Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
Alaska-Yukon border. In: Arnaud, E., Halverson, G. P. & logical Record of Neoproterozoic Glaciations. Geological Society,
Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic London, Memoirs, 36, 185–194.
Glaciations. Geological Society, London, Memoirs, 36, 389–396. Williams, G. E., Gostin, V. A., McKirdy, D. M., Preiss, W. V. &
Macdonald, F. A. & Jones, D. S. 2011. The Khubsugul Group, northern Schmidt, P. W. 2011. The Elatina glaciation (late Cryogenian),
Mongolia. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. South Australia. In: Arnaud, E., Halverson, G. P. & Shields-
(eds) The Geological Record of Neoproterozoic Glaciations. Zhou, G. (eds) The Geological Record of Neoproterozoic Glacia-
Geological Society, London, Memoirs, 36, 339–345. tions. Geological Society, London, Memoirs, 36, 713–721.
Chapter 2

A history of Neoproterozoic glacial geology, 1871– 1997

PAUL F. HOFFMAN1,2
1
Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA, 02138, USA
2
School of Earth and Ocean Sciences, University of Victoria, Victoria, BC, V8W 2Y2, Canada (e-mail: paulfhoffman@yahoo.com)

Abstract: Neoproterozoic glacial records have been discovered on 23 palaeocontinents, their rate of discovery changing little since
1871. Yet, half of all the resulting publications appeared since 2000. The history of research before 1998 is described in five stages
defined by publication spikes; subsequent work is not covered because historical perspective is lacking. In stage 1 (1871–1907), ‘Cam-
brian’ (now Neoproterozoic) glaciation was recognized successively in Scotland, Australia, India, Norway, Svalbard and China. Criteria
for recognition included faceted and striated pebbles in matrix-supported conglomerates resting on ice-worn bedrock pavements. In
stage 2 (1908–1940), Neoproterozoic glaciation was shown to have been widespread in Africa, Asia and the Americas. Major textbooks
summarized these findings, but the rejection of continental drift (to account for late Palaeozoic glacial dynamics) put a chill on research.
In stage 3 (1942– 1964), the occurrence of glacial deposits within carbonate successions, as well as nascent palaeomagnetic observations,
suggested that Neoproterozoic glaciers reached sea-level at low palaeolatitudes, but the belated recognition of sediment gravity flowage
caused glacial interpretations to be prematurely abandoned in key areas. In stage 4 (1965– 1981), the extent of Neoproterozoic glaciation
was rethought in light of plate tectonics. Distinctive chemical sediments (iron + manganese formations and cap carbonates) were ident-
ified. In basic climate models, modest lowering of solar luminosity resulted in global glaciation due to ice-albedo feedback, and degla-
ciation due to greenhouse forcing resulted from silicate-weathering feedback in the carbon cycle. Neoproterozoic glacial geologists were
blind to these ideas. In stage 5 (1982– 1997), reliable palaeomagnetic data combined with glacial marine sedimentation models confirmed
that Neoproterozoic ice sheets reached sea level close to the palaeoequator.

The first Neoproterozoic glacial deposits were found in the SW of and eastern South America (Fig. 2.1, Table 2.1). It was evident
Scotland in 1871 and the rate of their discovery on 23 palaeoconti- that glaciation ‘at or just before the beginning of the Cambrian’
nents and microcontinents was essentially linear until 1992 had been more extreme than in the Pleistocene, possibly equal to
(Fig. 2.1). However, half of all the papers (n ¼ 811) written on the composite Late Palaeozoic glaciations of Gondwanaland.
them before the end of 2008 were published in the previous 10 These findings were highlighted in a Symposium on Palaeozoic
years (Fig. 2.2). This surge in activity is the justification for the and Pre-Cambrian Climates at the 17th IGC (1937) in Moscow.
present volume, but this history ends at 1997. We lack historical Major textbooks made note of these developments (Köppen &
perspective on the last decade and the endpoint eliminates the Wegener 1924; Brooks 1922, 1926; Coleman 1926).
need to weigh any of the author’s own contributions on this topic. In the third period (1942 – 1964), nascent evidence from
This chapter begins by placing the recognition of the Port palaeomagnetism and from the occurrence of glacial deposits
Askaig boulder beds of Scotland as glaciogenic in its historical atop thick carbonate-dominated successions in the North Atlantic
context, coming in 1871 on the heels of the discoveries of late region and within such successions in central and southern
Palaeozoic glacial deposits in India, Australia and South Africa. Africa implied that Neoproterozoic ice sheets had reached sea
Their inspiration was the Pleistocene glacial controversy of level in low palaeolatitudes. This inference met resistance and,
1837– 1865, which spawned both the orbital and ‘greenhouse’ the- after the discovery of turbidity currents as a cause of graded
ories of bidirectional climate change, and established the principal bedding (1950), some diamictites were reinterpreted as mass-flow
criteria for the recognition of past glaciation. deposits of non-glacial origin. These issues were discussed at
The history is divided into five periods punctuated by spikes major conferences on Palaeoclimates in Newcastle-upon-Tyne,
in publications (Fig. 2.2) associated with the 10th and 17th England (1963), and Köln, Germany (1964). The radiation of
International Geological Congresses in Mexico City (1906) and macrofauna following soon after the ‘infra-Cambrian’ ice age
Moscow (1937), respectively, the twin palaeoclimate con- was often remarked upon.
ferences in Newcastle upon Tyne (1963) and Köln (1964), the The fourth period (1965– 1981) was notable for a detailed
IGCP Project 38 volume on Earth’s Pre-Pleistocene Glacial restudy of the classic Port Askaig Tillite, debate over the roles
Record (Hambrey & Harland 1981), and the onset of the present of true polar wander and plate tectonics in the distribution of
surge triggered in 1998. glaciogenic deposits, and the recognition of widespread post-
The first period (1871 –1908) saw the recognition of ‘Lower glacial ‘cap dolomites’ and syn-glacial iron and manganese
Cambrian’ glaciogenic formations in SW Scotland (1871), South formations, features not known from Phanerozoic glaciations.
Australia (1884), NE Norway (1891), western Svalbard (1898), Climate modelling received a kick-start when simple energy-
South China (1904) and NW India (1908). It ended with overviews balance calculations suggested that the Earth could freeze over
of Late Palaeozoic and Lower Cambrian (‘possibly Pre- from pole to pole due to ice-albedo feedback in response to lower-
Cambrian’) glaciations by Edgeworth David (1907a, b) at the ing of the solar constant by a few percent. Such a ‘white Earth’ dis-
10th IGC, and the winning-over of a skeptical Geological aster was shown to be self-reversing because of negative climate
Society in London by a fellow Australian (Howchin 1908). feedback associated with the geochemical cycle of carbon, but
Faceted and striated clasts in diamictites were the decisive criteria geologists were unaware of these developments.
for glaciogenesis, with a striated subglacial pavement beautifully In the fifth period (1982 –1997), Neoproterozoic diamictites
exposed beneath ‘Reusch’s Moraine’ (Fig. 2.3) in Finnmark, were increasingly interpreted in light of glacial marine sedi-
Norway, for good measure. mentation models and, combined with reliable palaeomagnetic
The second period (1909– 1941) saw Neoproterozoic glacial data, proved that ice sheets had reached sea level close to the
deposits recognized in southern, central and western Africa, palaeoequator. The ‘Snowball Earth’ hypothesis, adapted
central and northern Asia, eastern and western North America from climate and planetary science, was advanced as a

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 17– 37. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.2
18 P. F. HOFFMAN

24 1871 Laurentia (S) overstep the bounds of prudent speculation in suggesting that those erratics
1884 Australia are the reassorted materials of some great Northern Continent that has
22 1887 India yielded to the ceaseless gnawing tooth of time, leaving scattered fragments
1891 Baltica
20 1904 South China
as wreckage of its former greatness, and that the material of which the mass
1910 Avalonia is composed have in time, deeper than we have hitherto suspected, been trans-
18 ported by the agency of ice.’
1913 Laurentia (N)
16 1915 Kalahari
James Thomson, F.G.S. 1871. On the stratified rocks of Islay. Report of the
1929 Congo
14 1930 Siberia 41st Meeting of the British Association for the Advancement of Science,
1930 Sao Francisco Edinburgh, John Murray, London, 110–111.
12 1933 West Africa
1937 Tarim
10 1950 Cadomia
8 1956 Kazakhstan The Port Askaig boulder beds
1959 Amazonia
6 1979 Arequipa
1981 North China The first Neoproterozoic (then ‘Cambrian’) formation to be inter-
4 1982 Arabia
1984 Rio Plata
preted as glaciogenic (see quote above) was the Port Askaig,
2 1990 Mongolia exposed on the east side of Islay (Thomson 1871, 1877), an
0 1992 Iran island in the west of Scotland famous for its single malt whisky.
1880 1900 1920 1940 1960 1980 2000 2003 Nile
2009 Chukotka- The formation is a 750-m-thick succession of 47 mappable
Year of first report Arctic Alaska heterolithic-boulder diamictites with interbeds of crossbedded
sandstone (Kilburn et al. 1965; Spencer 1971; Arnaud 2004). In
Fig. 2.1. Cumulative discovery of Neoproterozoic glaciogenic deposits by his richly detailed sedimentological study, Spencer (1971)
palaeocontinent. See Table 2.1 for locations and references. Note that Laurentia describes the Port Askaig as lying ‘at the same horizon, between
is counted twice because of the large geographical separation between two carbonate-rich formations, for 700 km from NE Scotland to
southern (S) and northern (N) palaeohemisphere deposits of present eastern western Ireland’. It is now thought to be the oldest of three distinct
and western Laurentia, respectively. glaciogenic horizons within the Dalradian Supergroup (McCay
et al. 2006) – older and younger Cryogenian and middle Ediacaran
parsimonious explanation for low-latitude glaciation, associated in age. James Thomson’s speculation (see above) about a ‘great
iron (+manganese) deposits and (later) cap carbonates. Initially Northern Continent’ as the source of the glacial debris turned
ignored, the strong negative reaction to this hypothesis on the out to be true (Cawood et al. 2003), although that continent
part of Neoproterozoic glacial sedimentologists will in future be did survive the ‘gnawing tooth of time’ after all. It is North
seen as the most striking feature of the post-1998 period. America, or more precisely, its pre-Pangaean antecedent Lauren-
tia. It was merely displaced from the Dalradian in the early
Eocene, when Greenland separated from NW Europe. The Port
Prologue: discovery of the Port Askaig Tillite and its Askaig was deposited at palaeolatitudes near 258S according
historical scientific context to reliable palaeomagnetic data from mafic dykes, lavas and
sills of the marginally older (723 –718 Ma) Franklin Igneous
[On the stratified rocks of Islay.] ‘If . . . we compare the embedded boulders of Suite of Arctic Laurentia (Evans 2000). It accumulated on a
granite [in the schist] with the granites found in situ throughout the Highlands, marine shelf, subject to repeated subaerial exposure (Spencer
we feel the necessity of tracing them to another source, and hope we do not 1971; Johnston 1993), situated at the then southeastern margin

1.0
60 60
Cumulative fractional growth AGU
0.8 GM146
55 (n = 811) 55

50 0.6 50

45 0.4 45

40 40
0.2
35 35
0.0 IGCP 38
30 1880 1900 1920 1940 1960 1980 2000 Hambrey & Harland 30
1981
25 25
Newcastle
and Koln

20 20
conferences
1870 1880 1890
17th IGC 1963-64
15 15
10th IGC Moscow
Mexico 1937
10 10
1906
5 5

0 0
1890 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 2010

Fig. 2.2. Growth in the annual number of papers concerning Neoproterozoic glaciation, 1871–2008.
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 19

Late Palaeozoic glaciation and the theory of


continental drift

Thomson’s (1871) interpretation of the Port Askaig came at a time


of growing interest in pre-Pleistocene glaciation. A subglacial
pavement overlain by diamictite in South Australia found in
1859 by A.C. Selwyn proved to be of Permo-Carboniferous age
(Howchin 1912). A Permo-Carboniferous diamictite in the
Talchir coal basin of NE India (Blanford et al. 1859) was shown
to have been deposited by grounded, northward(!)-flowing glaciers
(Fedden 1875; Koken 1907). The late Carboniferous Dwyka Tillite
in Natal, South Africa, rests on a pavement scratched and grooved
by onshore(!)-flowing glaciers (Sutherland 1870). Ironically, these
findings had been inspired by the glacial interpretation of a
Permian-age diamictite in England (Ramsay 1855) that proved
to be non-glacial in origin.
By 1907 it was evident that with continents in fixed relative
positions, no amount of polar wander could prevent early
Permian glaciation in the southern hemisphere from extending
to within 108 latitude of the palaeoequator, while in the northern
hemisphere tropical warmth stretched to the pole (Koken 1907;
Irving 1956). ‘The Permian ice age poses an unsolvable
problem to all models that do not dare to assume horizontal dis-
placements of the continents’, wrote a 32-year-old German
meteorologist (Wegener 1912). With the Atlantic and Indian
oceans closed up, however, all the glacial action could be con-
tained poleward of 458S palaeolatitude, comparable to the
extent of northern hemisphere glaciation in the Pleistocene.
Fig. 2.3. Iconic sketch of the glacially striated pavement beneath the Moreover, a radial pattern of flow could then be inferred,
end-Cryogenian Smalfjord diamictite (‘Reusch’s moraine’) at Bigganjargga, centred on south-central Africa (Martin 1981). ‘This would
Varangerfjord, East Finnmark, northern Norway (from Reusch 1891).
take everything mysterious away from the phenomenon’
(Wegener 1912). The distribution of Late Palaeozoic glacial
of the palaeocontinent. The Port Askaig nicely illustrates the deposits was prime motivation for Wegener’s theory of continen-
basic information needed to interpret Neoproterozoic glaciogenic tal displacement, or drift (Martin 1981). Had he lived, vindication
deposits: their distribution, derivation, depositional environment, in the form of plate tectonics would have come to Wegener at the
palaeogeographic setting and age. age of 87.

Table 2.1. First reported occurrences of Neoproterozoic glaciogenic deposits by palaeocontinent

No. Year Palaeocontinent Continent Ref.

1a 1871 Laurentia (south) Europe Thomson (1871)


2 1884 Australia Australasia Woodward (1884)
3 1891 Baltica Reusch (1891)
4 1904 South China Asia Willis (1904)
5 1908 India Holland (1908)
6 1910 Avalonia North America* Sayles & Laforge (1910)
1b 1913 Laurentia (north) North America Hintze (1913)
7 1915 Kalahari Africa Rogers (1915)
8 1929 Congo Beetz (1929)
9 1930 Siberia Nicolaev (1930)
10 1930 São Francisco† South America Moraes Rego (1930)
11 1933 West Africa Baud (1933), Furon (1933)
12 1937 Tarim Norin (1937)
13 1950 Cadomia Wegmann et al. (1950)
14 1956 Kazakhstan Nalivkin (1956)
15 1959 Amazonia Maciel (1959)
16 1979 Arequipa Caldas (1979)
17 1981 North China Mu (1981), Lu et al. (1985)
18 1982 Arabia Gorin et al. (1982)
19 1984 Rio Plata Spalletti & Del Valle (1984)
20 1990 Mongolia Gibsher & Khomentovsky (1990)
21 1992 Iran Hamdi (1992)
22 2003 Nile‡ Miller et al. (2003)
23 2009 Chukotka-Arctic Alaska Macdonald et al. (2009b)

*Before 1980, the Squantum Tillite was thought to be Late Palaeozoic in age.

Often considered part of the Congo palaeocontinent.

Western margin of East African orogen.
20 P. F. HOFFMAN

The Pleistocene glacial controversy interested in climate, it came as a shock. They had assumed the
the Earth’s climate was slowly cooling over time due to the dissi-
Interest in pre-Quaternary glaciation arose in the wake of the pation of primordial heat in the Solar System according to the
most acrimonious and far-reaching controversy in 19th-century nebular hypothesis. This was consistent with undisputed palaeobo-
geology, the brouhaha over the glacial theory for Pleistocene tanical evidence (e.g. palm fronds in Switzerland) that most
tills, erratic boulders and associated landforms (moraines, drum- regions were warmer in the Cretaceous and early Cenozoic than
lins, eskers, kame-and-kettle, fluted ground, grooved bedrock, they are today. If the climate could change dramatically in either
crag-and-tail, roches moutonnées, whaleback rocks, cirques and direction, previously unknown factors must be at work in the
hanging valleys, horns and arêtes, U-shaped valleys and fjords). climate system.
Glacial theory – that most of northern Europe and North It fell to the Irish physicist, mountaineer and orator John
America had been sculpted by enormous and dynamic ice sheets Tyndall (1861, 1863) to demonstrate experimentally that certain
in the geologically recent past – was independently proposed by gaseous molecules in the atmosphere, notably water vapour and
Esmark (1824) in Norway, Dobson (1925) in the USA, Venetz carbon dioxide (CO2), absorb infrared radiation (‘obscure rays’)
(1830) and Charpentier (1837) in Switzerland and Bernhardi emitted by the Earth but are transparent to sunlight. Fourier
(1832) in Germany. None of these papers – published in major (1824) had previously mentioned the possibility of such a ‘green-
journals by reputable professors (Esmark and Bernhardi), engin- house’ effect, but had rejected it as quantitatively unimportant. For
eers (Venetz and Charpentier) and an industrialist (Dobson) – Tyndall (1863, p. 204), water vapour provides ‘a blanket more
raised a ripple. No geologist ever overturned conventional necessary to the vegetable life of England than clothing is to
wisdom with a single paper. man’. Radiative energy balance is maintained because the atmos-
Conventional theory held that these features were products phere ‘constitutes a local dam, by which the temperature of
of floods – meltwater floods resulting from dam bursts in the the Earth’s surface is deepened: the dam, however, finally over-
Alps (cf. de Saussure, von Humboldt, von Buch, de Beaumont) flows, and we give to space all that we receive from the Sun’
or iceberg-laden floods of Arctic origin in northern Europe and (Tyndall 1863, p. 205).
North America (cf. Sefström, Murchison, Hitchcock, Lyell). The Tyndall understood that water vapour cannot be the ultimate
controversy erupted when glacial theory was taken up by a cause of climate change, because its concentration is itself a
young, ambitious, energetic and multilingual Swiss palaeo- strong function of temperature. It amplifies any given climate
ichthyologist (Agassiz 1837, 1840). A recent convert to glacial change but cannot be its ultimate cause. On the other hand, the
theory, Agassiz linked it to the extinction of the boreal megafauna content of CO2 in the atmosphere, or any of the hydrocarbon
(e.g. mammoth, mastodon, woolly rhinoceros, giant deer), a gases, can vary independently. ‘It is not, therefore, necessary to
linkage easily disproved (Forbes 1846). He tied the end of the assume alterations in the density and height of the atmosphere to
glacial period to the uplift of the Alps, exactly opposite to the account for different amounts of heat being preserved to the
view of Charpentier, his primary glaciological tutor, and laughable Earth at different times’, wrote Tyndall (1861, p. 277), ‘a slight
to those who knew that Alpine orogeny began in the Eocene. Of change in its variable constituents would suffice for this.
the geological establishment, only the former arch-Diluvialist Such changes may in fact have produced all the mutations
William Buckland (Oxford University) was quickly won over to of climate which the researches of geologists have revealed’
glacial theory. [italics added].
The climax of Agassiz’s campaign in support of glacial theory Unknown to Tyndall, a prematurely deceased French ceramicist
came in 1840 (Davies 1968), when he and Buckland toured had already established the basis for the changes in CO2 that he
the British Isles, finding varied and widespread evidence of required, through the geochemical cycle of carbon (Ebelmen
glacial action (Agassiz 1842). The evidence included tills 1845, 1847; see also Hunt 1880; Berner & Maasch 1996).
(boulder clays) with polished, faceted and striated clasts, moraines
marking the limits of former glaciers, streamlined bedrock and
perched terraces resulting from ice-dammed lakes. At first, Agas- 1871 –1908: pioneering discoveries
siz’s findings aroused great interest in the English-speaking world,
but within months of his tour the glacial hypothesis was rejected by Following the first identification of glaciogenic deposits now
most geologists as unworkable (contrast Lyell 1840, 1841; Hitch- known to be Neoproterozoic in age in Scotland (Thomson 1871),
cock 1841). Thereafter, glacial theory would languish until a new similar discoveries were made in Australia (Woodward 1884),
generation of Scottish geologists undertook systematic surveys Norway (Reusch 1891), Svalbard (Garwood & Gregory 1898),
that provided (to their own surprise) overwhelming evidence that South China (Willis 1904) and India (Holland 1908). The deposits
glacial theory was correct after all (Ramsay 1860; Jamieson in the extreme NE of Norway made the biggest initial impact. Hans
1862, 1863, 1865; A. Geikie 1863, 1865; J. Geikie 1874). Reusch (1852– 1922), then Director of the Norwegian Geological
Agassiz had to wait 25 years for vindication. Had he lived, Jens Survey, found them around the head of Varangerfjorden, not far
Esmark would have been 99 years old. from the Russian border. Reusch (1891) describes hills underlain
Lyell, who had ‘an undeniable penchant for the skilful defence by up to 50 m of non-stratified, matrix-supported conglomerate
of lost causes’ (Cunningham 1990, p. 247), never completely gave in which cobbles of Archaean gneiss and granite predominate,
up on his decidedly non-uniformitarian iceberg-drift hypothesis. but which also carry smaller clasts of dolomite with facets and non-
Conversely, it is insufficiently appreciated today that many of parallel striations identical to those found in Quaternary tills
the criteria still used to distinguish ancient glaciogenic deposits throughout Norway. On the coast of the fjord near the lighthouse
were clearly spelled out over 170 years ago by Esmark, Dobson, at Bigganjargga, about 40 minutes’ walk east from Karlebotn,
Venetz, Bernhardi, Charpentier and Agassiz. Reusch (1891) found an isolated ridge of diamictite c. 70 m
long  8 m wide  3 m high (Bjørlykke 1967; Edwards 1975).
The ridge belongs to the older (Smalfjord Formation) of two gla-
The problem of bidirectional climate change and the ciogenic intervals in the region, and is onlapped and draped by
discovery of the ‘greenhouse’ effect shallow-marine sandstone. Just above the waterline of the fjord,
the diamictite has eroded away to reveal a glacial pavement on
The idea of a glacial period, or ‘Ice Age’ (Schimper 1837), was not crossbedded quartzite carrying glacial striations in two main orien-
controversial to geologists per se: characteristic fauna and flora tations (Fig. 2.3). The less distinct (east – west) set of striations
now live c. 1000 km farther north of their occurrences in the preserves lateral ridges and streaks of cataclasite, which suggest
‘Drift’ (Smith 1836, 1839; Forbes 1846). However, to physicists that the overlying diamictite is the melt-out tillite of a stagnant
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 21

ice-cored moraine (Edwards 1975). ‘Reusch’s Moraine’ has been did not become part of Laurentia until the late Silurian (Wilson
restudied by generations of geologists (Strahan 1897; Schiøtz 1966).
1898; Dal 1900; Holtedahl 1918; Rosendahl 1931, 1945; Føyn In southern Africa, glaciogenic diamictites of Neoproterozoic
1937; von Gaertner 1943; Crowell 1964; Reading & Walker age were first recognized in the Numees Formation of northwes-
1966; Bjørlykke 1967; Edwards 1975, 1997; Jensen & Wulff- ternmost South Africa and adjacent Namibia (Rogers 1915).
Pedersen 1996; Rice & Hofmann 2000; Bestmann et al. 2006; ‘Anyone familiar with the southern Dwyka [Tillite] who was
Arnaud 2008) and, despite differences in interpretation, it suddenly put down on the Numees beds would feel sure he was
remains an icon of Neoproterozoic glacial geology. on the Dwyka again’ comments Rogers (1915, p. 89). In fact,
Although they were not the first to recognize glacial action in distal facies of the Dwyka (Carboniferous) rest unconformably
‘Cambrian’ (later Adelaidean) strata of South Australia, Walter on the Numees, ‘which resembles the Dwyka tillite of the southern
Howchin in Adelaide and Edgeworth David in Sydney combined Karoo more closely than the Dwyka beds in the same neighbour-
to bring its extent and importance to wide attention (Cooper hood do’ (Rogers 1915, p. 90).
2009). Howchin (1901, 1903, 1908) traced the older of two There were no new discoveries between 1915 and 1929 – who
glaciogenic intervals from its type section in the Sturt River knows how many careers were cut down in World War I? – but
gorge near Adelaide over an area of nearly 800  400 km2 four book-length syntheses of ancient climates appeared in rapid
(Sprigg 1986), suggesting that ‘the ice gathered on a plateau of succession (Brooks 1922, 1926; Köppen & Wegener 1924;
comparatively low relief’ (David 1907b). The younger (Marinoan) Coleman 1926). In the more theoretical of his books, the English
glaciogenic interval was first recognized by Jack (1913; see also meteorologist Brooks (1926) explicitly recognized and attempted
Mawson 1949b; Preiss 1987). At first, Howchin’s glacial inter- to rationalize the existence of two distinct climate states,
pretation was greeted skeptically in Australia, but his convincing ‘glacial’ (eo-Cambrian, late Palaeozoic, late Cenozoic) and ‘non-
photographs of striated clasts (Howchin 1908) were accepted glacial’ (most of Phanerozoic time). His names are preferable to
without dissent by the Geological Society in London. If the later terms ‘greenhouse’ and ‘icehouse’ used synonymously.
Howchin was responsible for much of the legwork in South Köppen & Wegener (1924) is justly famous – despite having
Australia, David was intrumental in getting the word out overseas never been translated into English – for its detailed treatment of
(Cooper 2009). Reports of the 10th IGC (Mexico City, Phanerozoic palaeoclimates in terms of continental drift, as well
1906) include his 46-page synthesis (David 1907a), as well as as for inserting the first of Milankovic’s graphs of northern-
two shorter papers (one in French) on pre-Cenozoic glacial hemisphere summer insolation calculated for different latitudes
epochs globally, with emphasis on the Australian glacial record. over the past 650 000 years. It devotes three pages to the Neopro-
He also visited India, Great Britain and the USA during his 1906 terozoic (‘Kambrium’) glacial record. Coleman (1926), who had
travels and doubtless raised the awareness of ‘Cambrian’ earlier proposed a glacial origin for the ‘slate conglomerates’ of
glaciation. Huronian age (early Palaeoproterozoic) in Ontario, Canada,
In the lower gorges of the Yangtze River in western Hubei when they were thought to be the oldest known sedimentary
Province, South China, Bailey Willis and Eliot Blackwelder rocks (Coleman 1907), devotes a full chapter to glaciation ‘at
described a body of ‘till’ of probable ‘Lower Cambrian’ age that or just before the beginning of the Cambrian’ (Coleman 1926).
carries heterogeneous boulders exhibiting polish and striae of He concludes that the eo-Cambrian glaciation was more extreme
‘unquestionably glacial’ origin, based on examination of speci- than the Pleistocene and possibly equal to the composite late
mens by Quaternary glacial geologist T.C. Chamberlin (Willis Palaeozoic glaciations, which were known to be diachronous,
1904). The Nantuo tillite (Lee & Chao 1924) extends for younging from west to east (Du Toit 1922). The same conclusion
1200 km from NE to SW across the South China platform (Lee was reached by Kulling (1934), who updated the global
1936; Lee & Lee 1940) and represents the younger of two discrete picture (Table 2.1) at the end of a detailed account of Neoprotero-
Cryogenian glaciations in the region (Wang et al. 1981). Three zoic stratigraphy in the NE of the Svalbard Archipelago of
years later, a similar and likely correlative body (Jiang et al. Arctic Europe.
2003) was described in the Blaini Formation of the Lesser The pace of new discoveries picked up again from 1929 to
Himalaya, northern India (Holland 1908). 1937 with the recognition of glaciogenic formations in the
Katangan Series of southern Zaire (Beetz 1929), beneath the
Bambuı́ Group on the São Francisco craton of central Brazil (de
1909 – 1941: period of globalization Moraes Rego 1930; see also Isotta et al. 1969), on the Yenisey
Ridge at the western margin of the Siberian craton (Nikolaev
The number of palaeocontinents subject to glaciation in what is 1930), in the Taoudeni Basin of the West African craton (Baud
now known to be Neoproterozoic time grew from five in 1908 to 1933; Furon 1933) and in the central Tien Shan of NW China
fourteen in 1937 (Fig. 2.1, Table 2.1), with new finds in eastern (Norin 1937).
and western North America, central and northern Asia, eastern A major symposium on Palaeozoic and Pre-Cambrian climates
South America, and southern, central and western Africa. By the was held at the 17th IGC in Moscow in 1937 (published in 1940
time of the 17th IGC in Moscow in 1937, Neoproterozoic glacia- as Report 6). Ten substantial papers describe Neoproterozoic
tion was known on every continent except Antarctica. glacial deposits in different parts of Asia, Africa, Australia,
In 1913, a boulder slate at Big Cottonwood Canyon in the Europe and North America. Highlights include first-hand accounts
Wasatch Mountains of Utah was interpreted as glacial in origin of various pre-Dwyka glacial deposits throughout southern Africa
(Hintze 1913; see also Blackwelder 1932). Earlier, the Squantum (Gevers & Beetz 1940), the vast extent of the Nantuo Tillite in
‘Tillite’ member of the Roxbury Conglomerate in the Boston South China (Lee & Lee 1940) and the existence of glacial deposits
Basin of eastern Massachusetts was interpreted as glaciogenic in central Africa close to the equator (Davies 1940; Robert 1940).
(Sayles & LaForge 1910; Sayles 1914). This unit has a checquered A consensus emerged that ‘eo-Cambrian’ glaciation was matched
history. As a result of miscorrelation, it was long considered only in the late Palaeozoic, and that invertebrate animals arose fol-
to represent the only Carboniferous glaciation in the northern lowing retreat of the ice sheets.
hemisphere. Its glacial origin was repeatedly questioned (e.g. However, the rejection of continental drift in the late 1920s
Dott 1961) and later its age. Only after 1980 was its mid-Ediacaran following intense debate (Newman 1995; Oreskes 1999) had a
age recognized through geological mapping, geochronology chilling effect on pre-Pleistocene palaeoclimate research. If late
and palynology (Lenk et al. 1982; Thompson & Bowring Palaeozoic glaciation was allowed to remain ‘a hopeless riddle’,
2000). In Ediacaran time, the Boston Basin was part of the as Wegener put it, what chance was there to understand glacial
ribbon continent Avalonia –Cadomia, located off North Africa. It periods twice their age?
22 P. F. HOFFMAN

1942– 1964: rebutting challenges in high latitudes are compensated by reduced weathering rates in
the cooler tropics. The occurrence of glacial marine strata directly
The consensus that emerged from the Moscow Congress of 1937 above or sandwiched between thick carbonate successions in
carried over initially into the postwar period. In a 1948 Royal different areas (e.g. Harland & Wilson 1956; Ziegler 1960; Katz
Society lecture in Sydney, Australia, the normally reserved Sir 1961; Martin 1965a) suggested that, unless the glaciogenic and
Douglas Mawson (Sprigg 1986) proclaimed that ‘the world must carbonate strata are everywhere separated by stratigraphic gaps
then have experienced its greatest Ice Age’ and speculated that of large magnitude, tidewater glaciers existed in the warmest
prolonged genetic isolation during the period of refrigeration con- parts of the surface ocean, consistent with nascent palaeomagnetic
tributed to the subsequent Cambrian radiation of life (Mawson evidence for low palaeolatitudes (Girdler 1964). If glaciers
1949a). However, Mawson was a fixist and therefore unduly occurred in relatively warm areas, including marine platforms
impressed by the evidence for glaciation in equatorial Africa where no mountains existed, higher latitudes and elevations must
(Beetz 1929; Davies 1939, 1940). also have been glaciated. One is compelled to infer glaciation at
But what were mobilists to do? Plotting ‘infra-Cambrian’ all palaeolatitudes, independent of any assumption with respect
glacial deposits on Pangaea reconstructions, assuming them to to correlation (Harland 1964a, b). This is the crux of the Neopro-
be closer to reality than the present geography, still left some terozoic glacial conundrum.
deposits close to the palaeoequator (cf. Harland 1964a, b;
Harland & Rudwick 1964). On the other hand, if the continents
were tightly clustered, they might conceivably have crossed the The Newcastle and Köln palaeoclimate conferences
poles diachronously in infra-Cambrian time due to polar wander- of 1963 – 1964
ing and/or continental drift. In that case, the extent of glacial
deposits could far exceed the extent of ice at any time. What was The spectre of glaciation at low palaeolatitudes caused some to
needed were means of determining the palaeolatitudes of individ- question the glacial origin of Neoproterozoic ‘pebbly mudstones’
ual deposits, so that the ice extent inferred would not depend (Crowell 1957; Schermerhorn & Stanton 1963; Schermerhorn
on correlation. 1974). Since the recognition of turbidity current as a cause of
graded bedding (Kuenen & Migliorini 1950; Natland & Kuenen
1951), mass-flows loomed as alternatives to glaciers in some
Palaeomagnetism and the meridional extent of areas. This would not, however, account for the disproportionate
Neoproterozoic ice sheets occurrence of diamictites in late Neoproterozoic times. Martin
et al. (1985) suggested that mass-flows at low palaeolatitudes
In the early 1950s, statistical corroboration of the geocentric – might have been triggered by glacioeustatic changes driven by
axial –dipole hypothesis by Jan Hospers (Frankel 1987; Irving high-latitude glaciation, but late Palaeozoic glaciation did not
2008) had opened the way for palaeomagnetic and palaeoclimatic result in low-latitude diamictites comparable in setting and
testing of polar wandering and continental drift (Irving 1956, 1959; extent to those of the late Neoproterozoic.
Runcorn 1961). Wegener and Du Toit were soon vindicated, drift These were among the issues discussed at a pair of palaeocli-
was unavoidable, and the revolution in the interpretation of ocean mate conferences that proved to be a watershed for Neoproterozoic
basins followed in 1962–1967. Palaeomagnetism offered a means glacial geology (Fig. 2.2). The NATO Palaeoclimates Conference
by which the palaeolatitudes of ancient glacial deposits could be held at Newcastle-upon-Tyne in January 1963 was the first to high-
determined. Irving (1957a) had shown that the average magnetic light the topic since the Moscow symposium in 1937. Girdler
inclination in Carboniferous glacial varves in Australia is subver- (1964) reviewed the palaeolatitudes of continents during pre-
tical, implying high-latitude deposition, consistent with Wegener’s Mesozoic glacial periods according to existing palaeomagnetic
(1912, 1929) assumption based on their glacial origin. Preservation data. The Permo-Carboniferous and Huronian (early Palaeoproter-
of primary (depositional) natural remnant magnetization in red ozoic) glaciations occurred at high palaeolatitudes, affirming the
clastic sediments of Precambrian age having been demonstrated data, but middle and low palaeolatitudes were registered for the
(Irving 1957b; Irving & Runcorn 1957), it fell to Harland & eo-Cambrian glaciations. There were three possible explanations
Bidgood (1959) to first attempt this method in infra-Cambrian (Girdler 1964): the (eo-Cambrian) palaeomagnetic data were
glacial deposits. Their preliminary results (from a small subset faulty, the rocks are not glacial in origin, or the whole globe was
of samples) implied equatorial palaeolatitudes for glaciogenic for- glaciated. An entire session was devoted to the recognition of
mations in southern Norway and East Greenland. Harland’s glacial sediments and of till-like deposits of non-glacial origin
palaeomagnetic work lacked the sophistication and reach of Run- (Schwarzbach 1964a; Crowell 1964; Heezen & Hollister 1964).
corn’s group three miles to the west (there was little contact In the Discussion, geologist Wally Pitcher and others pushed
between the two Cambridge University laboratories) and it back against non-glacial interpretations, citing detailed strati-
would be 27 years before a truly reliable palaeomagnetic result graphic studies on the Port Askaig boulder beds (Kilburn et al.
was obtained for a Neoproterozoic periglacial marine formation 1965). Harland (1964a) provided a global synthesis of Precam-
(Embleton & Williams 1986). brian glacial deposits and their stratigraphic settings, concluding
Palaeomagnetists have long been interested in the palaeolati- that infra-Cambrian glaciation had extended into the marine
tudes of climate-sensitive sedimentary facies (e.g. thick carbon- environment at low palaeolatitudes, unlike any younger ice age.
ates, evaporites, redbeds, aeolianites, tillites) as a test of the An early tectonic ‘mobilist’, his views were informed by extensive
hypothesis that the time-averaged geomagnetic field closely fieldwork on the diamictite-bearing Hecla Hoek succession in
approximated a geocentric axial dipole over Phanerozoic and Svalbard (e.g. Harland & Wilson 1956; Wilson & Harland
Proterozoic times (Blackett 1961; Opdyke 1962; Briden & Irving 1964). He urged that such a glaciation provided an exceptional
1964; Briden 1970; Evans 2006). Thick shallow-water carbonates basis for global temporal correlation. Rudwick (1964) discussed
are today found only equatorwards of 358 latitude (Rodgers 1957) the glacial aftermath as an ecological cradle for the Cambrian radi-
and their zonal distribution did not change between glacial and ation. Four months after the Newcastle meeting, another global
non-glacial periods of the Phanerozoic (Briden 1970; Ziegler synthesis of Palaeozoic and Precambrian glaciation appeared,
et al. 1984; Opdyke & Wilkinson 1990; Witzke 1990; Kiessling graced with fine images of Neoproterozoic glacial debris from
2001). There are two reasons for this. First, the saturation state central Africa (Cahen 1963).
of seawater with respect to calcium carbonate or dolomite varies In March 1964, Manfred Schwarzbach convened an inter-
with the relative, not the absolute, temperature. Second, in national symposium on palaeoclimates in Köln, simultaneous
glacial periods, higher alkalinity fluxes due to glacial erosion with the publication in English translation of his book Climates
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 23

of the Past (Schwarzbach 1964b). The book has a chapter on Young 1976; Yeo 1981), Namibia (Martin 1965a) and Australia
eo-Cambrian glaciation, the subject of 7 of 37 papers from the (Whitten 1970) turned out to be intimately associated with Cryo-
Köln symposium published in the Geologische Rundschau. genian glaciogenic diamictites. None is closely associated with
Harland (1964b) again summarized the global infra-Cambrian volcanics. They are the only large-scale BIFs in the stratigraphic
glacial record, and Chumakov (1964, 1992) reviewed the record younger than the Palaeoproterozoic (1.9 Ga) deposits
growing number of Precambrian tilloids in the (European) around the Superior craton of North America. Martin (1965b)
USSR. Spjeldnaes (1964, p. 38) affirmed the glacial origin of suggested that ‘this peculiar combination of sediments’ might be
strata in different parts of Norway and Svalbard, noting that the attributed to ‘oxygen deficiency in stagnating bottom waters
eo-Cambrian deposits had been studied ‘with the same methods, caused by an ice cover’ (p. 116).
in the same regions and by the same scientists as have the more The presence of a generally rather thin but remarkably continu-
easily interpreted Quaternary deposits’. ous layer of carbonate, directly above Neoproterozoic diamictites,
Perhaps the defining summation of the infra-Cambrian glacial was occasionally noted by early workers (David 1907b; Norin
problem from this period is Harland & Rudwick (1964) in the 1937; Robert 1940; Mawson 1949b). These descriptions now
popular science magazine The Scientific American – which is become more specific, noting that the ‘capping dolomite’ is a dis-
still worth reading. tinctive pink or less commonly cream colour, invariably laminated
or thin-bedded, and persists over basement highs where the dia-
mictite cuts out (Dow 1965; see also Biju-Duval & Gariel 1969;
1965 – 1981: in the wake of the plate tectonic revolution Dunn et al. 1971; Rankama 1973; Deynoux & Trompette 1976;
Plummer 1978; Williams 1979). Williams (1979) was the first to
Plate tectonics revolutionized the Earth sciences, and Neoprotero- report stable isotope values (measured by Karlis Muehlenbachs)
zoic glacial geology was no exception. It freed the distribution of for ‘cap dolostones’ (and the first to use the term), from the Kim-
ancient glacial deposits from the assumption of fixed relative berleys of Western Australia. He notes that their d18O is similar to
positions of the continents, as foreseen by Harland (1964a) and other Neoproterozoic dolostones, but their d13C is moderately
(earlier) Wegener (1912), just as the concept of polar wandering depleted. Assuming the dolomite to be primary or early diagenetic,
(Gold 1955) had freed them from fixed position with respect to he concludes that ‘abrupt climatic warming at the close of late
the poles. Equally important, plate tectonics provided a basis for Precambrian glacial epochs is implied’, consistent with the d18O
making sense of geochemical cycles, and thereby to understand data (Williams 1979).
the controls on and changes to the chemical compositions of the
oceans and atmosphere (Siever 1968). This opened up the possi-
bility of understanding secular changes in atmospheric CO2 con- Climate models and the ‘white Earth’ instability
centration – the dream of Ebelmen (1845, 1847), Tyndall (1861,
1863) and Chamberlin (1898, 1899) – and consequent changes The geological literature in this period was fixated on the use or
in climate through ‘greenhouse’ radiative forcing. This led to misuse of glaciation for the division of Neoproterozoic time (e.g.
renewed interest during the period 1965 –1981 in certain chemical Harland 1964a; Crawford & Daily 1971; Dunn et al. 1971;
sediments characteristically associated with Neoproterozoic gla- Rankama 1973; Schermerhorn 1977). The cause of glaciation
ciogenic deposits (banded iron- and manganese-formations, and was barely mentioned – Harland (1964a) suggested diminished
post-glacial ‘cap’ carbonates). Finally, the period saw the publi- solar forcing and Williams (1972, 1974, 1975) appealed to large
cation of an influential textbook (Sugden & John 1976) that did variations in the Earth’s orbital obliquity (see below).
much to invigorate the study of processes of glacial erosion, sedi- Climate physicists, however, made a startling discovery when
mentation and landscape development. they used radiative enery-balance equations to calculate surface
temperatures as a function of latitude and variable solar forcing,
with simple parameterizations of meridional heat transport and
Centenary of Thomson’s (1891) study of the Port feedbacks due to snow, ice and clouds. When they reduced solar
Askaig Tillite irradiance by a few percent, surface temperatures fell below
freezing everywhere due to runaway snow and ice-albedo feed-
There could have been no finer centennial celebration than the pub- back (Eriksson 1968; Sellers 1969, 1990; Budyko 1969; see also
lication of Anthony (Tony) Spencer’s Geological Society Memoir North 1990). Although decadal to millenial solar variability is
on Late Pre-Cambrian Glaciation in Scotland (Spencer 1971). It only c. 0.1%, solar luminosity is thought to have slowly increased
remains the gold standard in Neoproterozoic glacial sedimentol- by 25–30% since 4.5 Ga due to the production of helium in the
ogy. The centenary was also marked by papers symbolizing the Sun’s core. The energy-balance calculations were therefore
disunity over the hot-button issue of diachronous or synchronous assumed to be in error because they predicted that the Earth
glaciation (Crawford & Daily 1971; Dunn et al. 1971). If the con- should have been permanently frozen: irradiance roughly 25%
tinents were tightly clustered, it was conceivable that continental higher than present is required to overcome the albedo of an ice-
drift or polar wandering could move them all at different times covered planet, which reflects over 60% of the sunlight it receives.
through the polar region, as was the case for Gondwanaland in This problem was a stimulus for the nascent science of climate
the Palaeozoic (Du Toit 1922; Wegener 1929; Crowell & Frakes modelling. However, the ‘white Earth’ instability (Wetherald &
1970). However, drift alone could not account for Palaeozoic gla- Manabe 1975) turned out to be a robust feature not only of
ciations, because there were long stretches of Palaeozoic time simple energy-balance models but also of most atmospheric
when Gondwanaland was situated over the South Pole yet no general-circulation models.
large ice sheets existed. In 1981, three planetary scientists proposed a solution to the
‘white Earth’ problem (Walker et al. 1981). They appealed to
the geochemical cycle of carbon (which is not accommodated in
Associated chemical sediments: banded iron- and physical climate models because of its timescale of c. 106 years).
manganese-formations and cap carbonates CO2 is supplied to the ocean and atmosphere by metamorphic-
volcanic outgassing and is consumed by silicate rock weathering.
Sedimentary iron- and manganese-formations (BIFs) were the The latter is temperature-dependent because of the direct effect
focus of worldwide mineral exploration for post-World War II of temperature on reaction kinetics and also because moisture
reconstruction. Major deposits in SW Brazil (Dorr 1945; in wet regions, where most weathering occurs, increases
Almeida 1946; Walde et al. 1981), NW Canada (Ziegler 1960; with temperature (Clausius – Clapeyron relationship). The
24 P. F. HOFFMAN

temperature-dependence of weathering provides a negative Glaciomarine deposits are more likely than their terrestrial
climate feedback (abiotic Gaia) that acts as a natural thermostat. counterparts to be preserved in the geological record because
If global temperatures rise (or fall) for any reason, CO2 is con- they are less susceptible to erosion, but ice-proximal processes
sumed at a faster (or slower) rate, thereby limiting the temperature in the marine environment are difficult to study because they
change. In effect, the feedback slowly adjusts the atmospheric CO2 occur underwater and often under ice. Nevertheless, much has
concentration to maintain a balance between CO2 sources and been gleaned from geomorphic and seismic surveys of polar con-
sinks. As solar luminosity increased over geological time, atmos- tinental shelves backed up by drilling programmes (e.g. Elverhøi
pheric CO2 concentration adjusted downwards, keeping surface 1984; Powell 1984, 1990; Alley et al. 1989; Barrett 1989;
temperatures within the range suitable for life. This simple hypoth- Boulton 1990; King et al. 1991; Powell & Domack 1995). A
esis not only went a long way towards solving the ‘faint young growing emphasis on glaciomarine processes and deposits is
Sun’ paradox (Sagan & Mullen 1972), it rationalized long-term reflected in many books written during this period (e.g. Molnia
climate change while explaining why such changes were limited 1983; Drewry 1986; Dowdeswell & Scourse 1990; Anderson &
in magnitude. Ashley 1991; Hambrey 1994; Benn & Evans 1998; Anderson
In the penultimate paragraph of their classic paper, Walker et al. 1999). Neoproterozoic glacial geology benefited enormously
(1981) suggested that silicate weathering feedback had been from these developments, although the complexity of facies archi-
responsible for averting a ‘white Earth’. They added, however, tecture poses a challenge given the limited three-dimensional
that ‘if a global glaciation were to occur, the rate of silicate exposure of ancient deposits and weak or non-existent constraints
weathering should fall nearly to zero, and carbon dioxide should on rates of accumulation.
accumulate in the atmosphere at whatever rate it is released Beginning in the mid-1950s, oxygen isotopes have been used as
from volcanoes. Even the present rate of release would yield 1 bar a basic tool for stratigraphic correlation of Quaternary marine
of carbon dioxide in only 20 Ma. The resultant large greenhouse carbonates and for estimating seawater temperatures and global
effect should melt the ice cover in a geologically short period of ice-sheet volumes. Unfortunately, Neoproterozoic carbonates do
time’ (Walker et al. 1981). If a ‘white Earth’ did occur, it would not preserve their original mineralogy, and their oxygen isotope
be self-reversing. Accordingly, its occurrence in deep time could compositions are seriously compromised by aqueous diagenesis.
not be ruled out a priori. This, in a nutshell, is the climatological However, carbonate rocks buffer the isotopic composition of
concept behind the ‘Snowball Earth’ hypothesis (Kirschvink 1992). the relatively low concentration of carbon in aqueous solution.
Walker et al. (1981) made no reference to ancient glaciation, Consequently, the carbon isotopic compositions of carbonate
Neoproterozoic or otherwise. This was not because they were rocks resist alteration. Beginning in the mid-1980s, carbon iso-
unaware of geology. On the contrary, the ‘faint young Sun’ topes were increasingly used for stratigraphic correlation of Neo-
problem was brought into focus for Walker by his participation proterozoic carbonates and to a lesser extent for organic-rich
in the Precambrian Palaeobiology Research Group (PPRG) shales (Knoll et al. 1986; Knoll & Walter 1992; Kaufman et al.
organized by micropalaeontologist J. William Schopf. (Third 1997). Studies show that Neoproterozoic glaciogenic formations
author Kasting and palaeomagnetist Joe Kirschvink were PPRG are typically bracketed by negative isotopic excursions. The excur-
participants later in the decade.) By 1981, Harland’s great infra- sions are reproduceable and approximately five times larger in
Cambrian glaciation had fallen off the radar screen. Palaeo- magnitude than spatial or depth variations in the modern ocean.
magnetism, on which high hopes were pinned, had encountered Because atmospheric CO2 concentrations were likely higher in
problems. The foremost was the susceptibility of sediments to low- the Neoproterozoic than today because of the dimmer Sun, a
temperature chemical remagnetization (e.g. McCabe & Elmore, larger reservoir of dissolved inorganic carbon should have
1989). Overcoming these problems would require time-consuming damped spatial and depth variations of d13C in Neoproterozoic
stepwise chemical and thermal demagnetization, studies of mag- oceans. Therefore, the observed isotopic excursions are most
netic mineralogy, ‘field tests’ to constrain the age of remnant easily explained as secular variations. Their regular occurrence
magnetic components, and more sensitive magnetometers. directly before and after glaciogenic sedimentation suggests that
glaciation was roughly synchronous in many areas, even if not
every isotopic excursion is associated with glaciation. For those
Earth’s Pre-Pleistocene Glacial Record volume who accepted these basic principles, the arguments for diachro-
nous glaciation receded as the body of Neoproterozoic carbon
In 1981, Cambridge University Press published Earth’s Pre- isotope data grew.
Pleistocene Glacial Record (Hambrey & Harland 1981), a The first robust palaeomagnetic support for glaciation at sea
product of Project 38 (Pre-Pleistocene Tillite Project) of the Inter- level in low palaeolatitudes came from the Elatina Formation in
national Geological Correlation Programme. This sought-after South Australia (Embleton & Williams 1986). The formation is
volume contains 58 unified descriptions of Neoproterozoic glacio- a glaciomarine unit comprised of reddish sandstone, siltstone
genic formations worldwide, and is a tribute to Harland’s vision and diamictite (Lemon & Gostin 1990; Williams et al. 2008),
and Hambrey’s dedication. It was synthesized in Harland (1983) and a 10-m-thick interval of rhythmically laminated siltstone
and Hambrey & Harland (1985). An additional global synthesis has been the subject of repeated palaeomagnetic studies. The
of Neoproterozoic glaciogenic deposits was published by Chuma- palaeopole determined by Embleton & Williams (1986) implies
kov (1981). a palaeolatitude of c. 58 and, although no field test was performed,
the result is considered reliable because the natural remanent
magnetization is stably carried by detrital haematite. At the time
of the study, the laminations were considered to be annual
1982– 1997: the gathering storm varves, and the rhythmic bundles of 12–14 laminae were
thought to represent sunspot cycles (Williams 1981; Williams &
There were many developments during this period – burgeoning Sonett 1985). Later, they were reinterpreted as tidal rhythmites
information regarding glaciomarine processes and deposits, (Williams 1989), supporting a glaciomarine origin for the associ-
widespread use of carbon isotopes as a tool for correlation, ated diamictites. Field tests were subsequently carried out using
acquisition of reliable palaeomagnetic constraints on palaeolati- soft-sediment folds (Schmidt et al. 1991; Schmidt & Willliams
tudes of proximal glaciomarine deposits, increased awareness of 1995) and polarity reversals (Sohl et al. 1999), which confirmed
the sedimentological peculiarites of post-glacial ‘cap’ carbonates, the primary nature of the remanent magnetization and the palaeo-
and growing interest in causal mechanisms for low-latitude latitude of 7.58 or less. A somewhat larger palaeolatitude of c. 158
glaciation. was obtained in a recent palaeomagnetic study (Raub & Evans
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 25

2006) of the cap dolostone (Nuccaleena Formation) above the Kendall & Warren 1987), which result from layer-parallel expan-
Elatina Formation, implying a certain amount of inclination flat- sive growth associated with cementation under alternating vadose
tening during compaction of the Elatina siltstone, but this hardly and phreatic conditions. The cap ‘tepees’ are not associated with
alters the thrust of the result because temperatures are broadly breccias or void-filling cements, as are peritidal tepees, and in
uniform across the tropics. plan view their crestlines are always linear and parallel, not poly-
gonal like true tepees (Assereto & Kendall 1977). Bundles of
laminae may exhibit onlap or offlap relations on ‘tepee’ flanks.
Cap carbonates These metre-scale trochoidal bedforms were later analysed as
giant wave ripples (Allen & Hoffman 2005).
Arguably the first person to fully appreciate that the basal Edia- Large-scale stromatolites, within which are spaced, micrite-
caran (Marinoan) cap carbonate was not only ‘an important strati- or cement-filled tubes or gutters that maintain a palaeovertical
graphic marker’ but also ‘a perplexing and paradoxical (‘geoplumb’) orientation irrespective of the inclination of lami-
lithostratigraphic unit’ (Table 2.2) was Aitken (1991), a leading nations in the host stromatolite, were described from cap dolo-
Rocky Mountain carbonate stratigrapher. Concerning the cap stones in California (Cloud et al. 1974; Wright et al. 1978),
dolostone atop the younger glaciogenic unit (Stelfox Member of Namibia (Hegenberger 1987) and later in Alaska, Brazil, Canada
the Icebrook Formation) in the Mackenzie Mountains of NW and Mongolia (Fig. 2.3).
Canada, Aitken (1991) stressed that the aphanitic microcrystalline Aitken (1991) identified crystal fans and pseudo-stromatolites
dolostone was not deposited as fine-grained mud, but as silt and in limestone overlying the cap dolostone, correctly interpreting
sand-sized peloids. Ubiquitous subparallel lamination is defined them as pseudomorphic after aragonite sea-floor cements. His
by oscillatory variation in peloid size. His photographs show photograph of ‘blades of calcite’ is from a unique layer of sea-floor
well-sorted, coarse-grained peloids in layers only one or two barite (BaSO4) cement, variably calcitized, which occurs in the top
peloids deep. Where coarse peloids form the base of a layer, ,10 cm of the cap dolostone for .150 km along a strike south-
they commonly rest on facets, apparently resulting from abrasion. eastward from the location of Aitken’s (1991) photograph. Thus,
The layers are not laterally continuous, and low-angle cross- cap dolostones and overlying limestones display a panoply of idio-
lamination is ubiquitous. The lamination is clearly mechanical in syncratic features (Fig. 2.3) that generally occur in a consistent
origin, interspersed locally with microbially stabilized masses of vertical sequence (Hoffman et al. 1987).
micropeloidal stromatolite (James et al. 2001). Aitken (1991) offered no genetic explanation for cap carbonates.
Aitken (1991) argued that the intrastratal ‘tepees’ for which Eyles (1993) suggested that they represent detrital rock ‘flour’,
the unit was informally named (Eisbacher 1981) are not true peri- as proposed earlier for carbonate layers within glaciogenic
tidal tepees as defined by Assereto & Kendall (1977; see also diamictites (Fairchild 1983). A detrital origin could hardly

Table 2.2. Idiosyncratic sedimentary and early diagenetic features in cap dolostones

Palaeocontinent Cap dolostone (ref.) SCC TPB DGB PEL LAC GWR TBS SFB
p p p p
Amazonia Mirassol d’Oeste (1) – – – –
Arabia Hadash (2) – – – – – – – –
p p p p
Arctic Alaska Nularvik (3) – – – –
p p p p p
Australia Mount Doreen (4) – – –
p p
Australia Nuccaleena (5) – – – – – –
p p p
Australia Cumberland Creek (6) – – – – –
p p
Australia Lower Ranford (7) – – – – – –
Baltica Lower Nyborg (8) – – – – – – – –
p p p p p
Congo Keilberg (9) – – –
Congo Calcaire Rose (10) – – – – – – – –
Congo C1 (11) – – – – – – – –
India Upper Blaini (12) – – – – – – – –
p p p p
Kalahari Dreigratberg (13) – – – –
p p p p p p
Laurentia Ravensthroat (14) – –
p p p p
Laurentia Noonday (15) – – – –
p
Laurentia Lower Canyon (16) – – – – – – –
p p p
Laurentia Lower Dracoisen (17) – – – – –
p p p
Laurentia Cranford (18) – – – – –
p p
Laurentia Hard Luck (19) – – – – – –
p p p p p
Mongolia Ol (20) – – –
p p
Mongolia Baxha (21) – – – – – –
p p p p
South China Lower Doushantuo (22) – – – –
p
Tarim Lower Zhamoketi (23) – – – – – – –
p p
West Africa Oued Djouf (24) – – – – – –
p p p
West Africa Amogjar (25) – – – – –
p p p p p
West Africa Mid Sud-Banboli (26) – – –

Abbreviations: SCC, sheet-crack cements; TPB, tepee breccia; DGB, diagenetic barite; PEL, peloids; LAC, low-angle cross-laminae; GWR, giant wave ripples; TBS, tube
biostrome; SFB, seafloor barite.
References: (1) Nogueira et al. (2003); (2) Allen et al. (2004); (3) Macdonald et al. (2009b); (4) Kennedy (1996); (5) Plummer (1978); (6) Calver & Walter (2000);
(7) Corkoran (2007); (8) Edwards (1984); (9) Hoffman et al. (2007); (10) Cahen & Lepersonne (1981); (11) Cahen (1950); (12) Kaufman et al. (2006); (13) Macdonald
et al. (2011); (14) James et al. (2001); (15) Corsetti & Grotzinger (2005); Corsetti & Kaufman (2005); Wright et al. (1978); (16) Hambrey & Spencer (1987); (17)
Halverson et al. (2004); (18) McCay et al. (2006); (19) F. A. Macdonald, pers. comm.; (20) Macdonald et al. (2009a); (21) F. A. Macdonald & D.S. Jones, pers. comm.;
(22) Jiang et al. (2006); (23) Xiao et al. (2004); (24) Bertrand-Sarfati et al. (1997); (25) Shields et al. (2007); (26) Nédélec et al. (2007).
26 P. F. HOFFMAN

account for the systematic variations in d13C observed in cap Table 2.3. Causative theories for Neoproterozoic glaciation
dolostones (Kennedy et al. 1998; Hoffman et al. 2007).
Kaufman et al. (1993) and Grotzinger & Knoll (1995) related Astronomical theories
cap carbonates to the overturn of alkalinity-charged, isotopically 1. Solar variation Harland (1964a)
depleted, bottom water after a lengthy period of ocean stagnation. 3. Large obliquity Williams (1972, 1975), Jenkins (2000),
This suggestion requires that primary production be maintained Donnadieu et al. (2002)
in the absence of upwellings, which is the overwhelmingly pre- 5. Ice-ring collapses Sheldon (1984)
dominant nutrient flux in the modern ocean. Upwellings would 13. Impact ejecta Bendtsen & Bjerrum (2002), Fawcett &
not be necessary if organic matter did not settle, but then alkalinity Boslough (2002)
would not build up in the deep nor would a vertical isotopic 4. Giant molecular clouds Pavlov et al. (2005)
gradient develop. This and the physical implausibility of pro- Oceanographic theories
longed (.10 000 years) ocean stagnation (in the absence of an 2. Carbonate burial Roberts (1971)
ice cover) make the overturn hypothesis unconvincing, despite 7. Ocean stagnation Kaufman et al. (1993), Grotzinger & Knoll
its popular appeal. (1995)
Kennedy (1996) proposed that cap carbonates represent non- 10. Hypsometric effect Kennedy (1996), Ridgwell et al. (2003)
skeletal analogues of carbonates deposited (in low latitudes) 11. Organic burial Kaufman et al. (1997)
during times of Quaternary deglaciation according to the ‘coral Geodynamic theories
reef’ hypothesis (Berger 1982; Opdyke & Walker 1992). This 6. Continental Marshall et al. (1988), Worsley & Kidder
results from the hypsometry of ocean basins. When sea level is distribution (1991), Kirschvink (1992), Schrag et al.
lowered due to ice-sheet growth, the area of shallow shelves (2002)
and inland seas is disproprotionately reduced. These areas are 9. Continental break-up Eyles (1993), Young (1995a), Donnadieu et al.
favoured for carbonate burial because the sediment never (2004a)
encounters undersaturated deepwaters. When these areas are lost, 15. Basalt weathering Goddéris et al. (2003)
all carbonate production occurs in the open ocean and the 16. True polar wander Li et al. (2004)
alkalinity of deepwaters increases. Upon deglaciation, alkalinity- Biological theories
rich waters flood the shelves, depositing carbonate and 8. Biocatalysed Carver & Vardavas (1994), Lenton & Watson
releasing CO2 to the atmosphere. The process can account for weathering (2004), Kennedy et al. (2006)
much of the glacial –interglacial variation in pCO2 (Opdyke & 12. Methane destruction Pavlov et al. (2000, 2003), Catling et al. (2001),
Walker 1992) and thus provides a positive climate feedback. Claire et al. (2006)
This mechanism must have contributed to cap carbonates, but 14. Methane substitution Halverson et al. (2002), Schrag et al. (2002)
the estimated average thickness of carbonate produced (Ridgwell
et al. 2003) is only approximately one-tenth of the average
thickness of cap dolostones of 18.5 m (Hoffman et al. 2007, Large orbital obliquity. The Australian sedimentologist George
Table 2.1). The coral reef hypothesis does not explain why cap car- E. Williams always had an eye on ‘the big picture’ and possessed
bonates are only associated with the ultimate deglaciation, and do an early appetite for planetary orbital mechanics. Impressed by the
not accompany glacial –interglacial cycles within the glacial prevalence of varves (i.e. rhythmic lamination, supposedly seaso-
period where these are recognized (Leather et al. 2002; Allen nal) and other seasonal indicators such as polygonal sand wedges
et al. 2004). If cap carbonates were eroded during glacial read- associated with Neoproterozoic glaciogenic strata (Chumakov
vances, they would be present as clasts in glacial deposits, which 1968; Spencer 1971), he proposed that the Earth’s obliquity (i.e.
they are not. the angle between the equatorial and ecliptic planes) oscillated
Nearly 70 papers concerning cap carbonates appeared in the between large (e.g. Neoproterozoic) and small (e.g. Phanerozoic)
decade after 1997. This reflects the central but contentious role values (Williams 1972). When the angle was large, the seasonal
they have assumed in the ongoing controversy over the nature of cycle was strong everywhere; when small, the seasons did not
Cryogenian glaciation. Arguably they are the most richly enig- differ greatly in low latitudes (except in strongly monsoonal
matic horizons in the entire stratigraphic record. areas). In addition, as the angle increased, the meridional insola-
tion gradient declined and actually reversed (i.e. greater annual
insolation at the poles than at the equator) whenever the angle
Causative theories for Neoproterozoic glaciation exceeded 548. Williams (1972, 1974, 1975) inferred that when
the obliquity was very large, low latitudes were subject to glacia-
Walker et al. (1981) provided the conceptual basis for a self- tion preferentially. As low latitudes cover more area than high lati-
reversing global glaciation from the point of the initial ‘white tudes, there should be more ice (and higher planetary albedo)
Earth’ (ice-albedo) instability. They did not speculate on how during periods of low-latitude glaciation than when ice is limited
the climate ever reached that point because they credited silicate- to the poles. Accordingly, glacial periods occur when obliquity
weathering feedback with its avoidance. However, a number is large and when there is strong seasonality everywhere (contrary
of causative theories (not mutually exclusive) were proposed to the Wegener – Milankovic hypothesis that ice-sheet growth is
between 1971 and 1997 to explain Neoproterozoic glaciation favoured by cool summers and low seasonality). Under large obli-
(Table 2.3). quity, glaciation at the poles is precluded by intense summer-
time insolation.
The large obliquity hypothesis was strengthened by the discov-
Carbonate burial. Roberts (1971, 1976) was struck by the fact that ery of 2.5-m-deep sand wedges associated with the Elatina For-
the Cryogenian glaciation was preceded by thick successions of mation (Williams & Tonkin 1985), which, combined with
shallow-water carbonate strata in the North Atlantic (Greenland, palaeomagnetic evidence (Embleton & Williams 1986), indicates
Svalbard and the British Isles), western North America and strong seasonality at low palaeolatitude (Williams 1993; but see
central Africa. He postulated a period of anomalous carbonate Maloof et al. 2002). In the Phanerozoic, deep periglacial sand
burial in which atmospheric CO2 was sequestered, producing an wedges occur only in middle and high latitudes, where seasonal
‘anti-greenhouse’ effect. He did not give any reason why this forcing is strong (Lachenbruch 1962). With large obliquity, sea-
might have occurred. It would require a rise in the global rate of sonality is strong at all latitudes.
silicate weathering relative to the rate of CO2 outgassing (see Like any theory, large obliquity was not without difficulties.
below; Continental distribution and Continental break-up). Although a large (and chaotic) obliquity could have been imparted
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 27

on the early Earth by stochastic accretion (Laskar & Robutel ocean largely covered by pack ice, perhaps with ‘warm tropical
1993), obliquity cannot oscillate between large and small angles “puddles” in the sea of ice, shifting slightly from north to south
because of ‘entrapment’ at small angles by gravitational attraction with the seasons’ (Kirschvink 1992). At the 1989 PPRG meeting
between the Moon and the Earth’s equatorial bulge (Laskar et al. at UCLA (Maugh 1989), he proposed that the ‘white Earth’
1993). Therefore, the theory was revised such that obliquity was disaster had actually occured in the Neoproterozoic, perhaps
persistently large until Ediacaran times, when it was rapidly repeatedly, and that each pan-glacial period was abruptly termi-
reduced to small angles before the Cambrian (Williams 1993, nated when the slow build up of CO2 reached a critical level, as
2000). This proved difficult to replicate in models or justify from deduced by Walker et al. (1981). He attributed the deposition
orbital theory (Néron de Surgy & Laskar 1997; Williams et al. of banded iron-formation during Neoproterozoic glaciation to
1998; Hoffman & Maloof 1999; Pais et al. 1999; Donnadieu ocean stagnation and deepwater anoxia, as previously proposed
et al. 2002; Levrard & Laskar 2003). Moreover, large obliquity by Martin (1964b), and he predicted that the abrupt climate
over most of Precambrian time is at odds with a growing body switches accompanying glaciation and deglaciation should be
of palaeomagnetic evidence that pre-Ediacaran evaporites and represented in widely separate areas by lithologically similar
carbonate-dominated successions formed at latitudes inconsistent strata, a result of the global scale of the climatic fluctuations
with a reversed climatic gradient (Park 1994; Buchan et al. (Kirschvink 1992). He called the glacial state a ‘Snowball’ Earth
2001; Evans 2006; Maloof et al. 2006). However, decisive falsifi- (Maugh 1989) to highlight the central role of planetary albedo in
cation of the large obliquity hypothesis, glaciation at high palaeo- the phenomenon. It is an evocative description of how the late
latitude, has yet to be observed. Cryogenian Earth might have looked from outer space.

Ice-ring collapses. Richard P. Sheldon was a sedimentary phos- Impact ejecta. Impact ejecta figured in discussions of Neoprotero-
phorite specialist with the United States Geological Survey who zoic glaciation in two completely different ways. They were pro-
developed an ‘outlandish hypothesis’ (Sheldon 1984) for Neopro- posed as an alternative emplacement mechanism for diamictite,
terozoic glaciation following a field excursion in northern Mongo- implying that the low-latitude and carbonate-associated glacial
lia, where phosphorites, carbonates and glacial marine strata occur problem was a chimera (Oberbeck et al. 1993; Rampino 1994).
in close stratigraphic proximity. He speculated that ice rings denser This was a small reenactment of the pebbly mudflow challenge
than those of Saturn orbited the Archaean Earth. In the Proterozoic, of 1957 –1974. Impact ejecta aprons do in fact share many
sublimation thinned the rings and they became more discrete. diagnostic features with glaciogenic diamictites (e.g. unsorted
As the Moon pulled away from the Earth, their orbits slowly debris, faceted and striated erratic stones, outsized dropstones).
decayed and they successively entered the atmosphere. As each However, continuous ejecta blankets extend only about one crater-
ring approached the Earth, its shadow caused an ice age. Upon radius from the crater rim, regardless of crater size (Melosh 1989).
entering the atmosphere and before the next ring approached, the Therefore, an improbably large flux of big impacts (roughly one
shadow was lost and greenhouse gases were added, causing a Cretaceous –Tertiary sized impact every 100 ka for 200 Ma)
warm interglacial. Once the final ring had fallen, the tropics would be required to account for the distribution of Neoproterozoic
were shadowless for the first time. Ocean overturning driven by diamictites, given the low probability of their exposure at the
the steepened climatic gradient led to phosphogenesis and Earth’s surface today.
organic diversification. Sheldon (1984) noted that seasonality It was later proposed on the basis of modelling experiments
would be greatest in glacial times, consistent with Williams’ that large impacts could have caused Neoproterozoic glaciation
(1993) observation, because the rings cast shadows only on (Bendtsen & Bjerrum 2002; Fawcett & Boslough 2002). The
the winter hemisphere. Again, this is contrary to the Wegener – models suggest that short-term (,10 year) shielding of sunlight
Milankovic theory of ice ages, which posits that ice sheets grow by ejecta from a Cretaceous –Tertiary sized impact could cause
when summers (not winters) are cool and seasonality is weak the ocean surface to freeze over if seawater was as cold as today,
(Köppen & Wegener 1924; Milankovic 1941). unlike the Cretaceous ocean, which was 8– 128 warmer than
present.
Continental distribution. Atmospheric pCO2 will adjust downwards
in response to an equatorward migration of the continents, because Ocean stagnation. As the patterns of C and Sr isotope variations
weathering rates are greatest in the tropics (Marshall et al. 1988; associated with Neoproterozoic and Phanerozoic glaciations are
Worsley & Kidder 1991). Global cooling is therefore consistent distinct (Kaufman et al. 1993), the cause of glaciation might also
with sedimentological and palaeomagnetic evidence for an be different. Kaufman et al. (1993) and Grotzinger & Knoll
unusual preponderance of low-latitude continents in the Neopro- (1995) attempted to genetically link the isotopic patterns, glacia-
terozoic, ‘a situation that has not been encountered at any sub- tion and the associated Fe- and Mn-ore formations and ‘cap’ car-
sequent time in Earth history’ (Kirschvink 1992). Equatorward bonates in a conceptual model involving prolonged ocean
displacement of the continents also changes the planetary stagnation followed by overturn. They hypothesized that when
albedo: the tropical ocean is a strong absorber of sunlight, unlike the ocean was stagnant, organic export and subsequent respiration
land areas (minimally vegetated in the Neoproterozoic) and the in anoxic deep waters would progressively enrich the surface
fog-bound polar oceans, which are relatively good reflectors waters in 13C, while simultaneously creating an isotopically
(Kirschvink 1992). Moreover, a positive albedo feedback would depleted, bicarbonate-enriched reservoir at depth (Kaufman
result from the enlargement of tropical land area that would et al. 1993; Grotzinger & Knoll 1995). In their model, the attendant
accompany any glacioeustatic fall in sea level (Kirschvink 1992). transfer of CO2 from the atmosphere to the deep ocean con-
A palaeomagnetist, Kirschvink (1992) accepted the low-latitude tributes to global cooling. Ultimately, the growth of sea ice triggers
result for the Elatina glaciation (Embleton & Williams 1986) the formation of cold, saline, deep water, and the meridional
but rejected the large-obliquity hypothesis because of the co- overturning circulation is reestablished. Upwelling of alkalinity-
occurrence of Neoproterozoic glacial deposits and thick carbonate- laden deepwater leads to the precipitation of isotopically depleted
dominated successions. If the meridional climatic gradient were cap carbonates and the simultaneous release of CO2 to the atmos-
reversed, the carbonate belts would move from low latitudes to phere, which melts the ice (Kaufman et al. 1993; Grotzinger &
the poles, ‘where the glaciers (in Williams’ model) should not Knoll 1995).
encounter them’ (Kirschvink 1992). He was therefore compelled The parsimony of the model temporarily outshone its flawed
to assume that if ice sheets reached sea level close to the foundation. Short of covering the ocean with ice, physical
equator, higher latitudes must have been glaciated as well, wher- stagnation (as distinct from dynamic stratification) on geological
ever ablation did not exceed precipitation. He envisaged an timescales is implausible because the energy driving the upwelling
28 P. F. HOFFMAN

flux (the rate-limiting step in the overturning circulation) derives only a handful of the highest peaks (.4.0 km above sea level)
from winds and tides (Wunsch 2002). Winds and tides are not on the Ethiopean Dome were glaciated at the Last Quaternary
so easily turned off. Moreover, if the surface ocean was ever Maximum (Osmaston 2004). For Neoproterozoic ice sheets to
deprived of nutrients upwelled from depth, primary production have reached sea level at the same latitudes implies a drastically
would crash, depriving the deep of its alkalinity pump and the iso- colder climate.
topic gradient of its driver. To its credit, the overturn hypothesis
at least lined up all the key observations in the same viewfinder:
glaciation, isotopes, iron formations and cap carbonates. Epilogue: the long road to consensus

Biocatalysed weathering. Carver & Vardavas (1994) constructed a Figure 2.2 clearly shows that publications concerning Neoprotero-
time-dependent model of the Earth’s mean surface temperature zoic glaciation ‘took off’ in the mid-1990s and possibly peaked
controlled by the geochemical cycle of carbon (Walker et al. in 2007. There is little doubt that the galvanizing factor was the
1981). They parameterized CO2 outgassing as having declined Snowball Earth hypothesis (Kirschvink 1992; Hoffman &
rapidly before and more slowly after 3.5 Ma. The solar flux rose Schrag 2002). In the first four years after its publication,
almost linearly, and the consumption of CO2 through silicate Kirschvink (1992) had been cited only three times, and favourably
weathering was shaped by the rapid growth of 90% of the only once (Klein & Beukes 1993). However, as soon as it was
present continental crust between 3.5 and 2.0 Ga, and by stepwise advocated prominently (Hoffman et al. 1998; Hoffman & Schrag
increases in weathering efficiency (i.e. the rate of CO2 consump- 2000) and in the public arena (Walker 2003), the reaction
tion for any given pCO2) resulting from biological colonizations was swift.
of the land (Schwartzman & Volk 1991). The first colonization This is not the place to review all the papers that have appeared
(microbial) is assumed to have occurred before 3.5 Ga. The since 1998: more time and more space are needed. Opinions about
second (organisms unspecified) supposedly occurred between 1.2 the nature of Neoproterozoic glaciation are now at a stalemate.
and 0.7 Ga, and the third in the middle Phanerozoic in response Most workers agree that one, and possibly two, Cryogenian glacia-
to the rise of vascular plants. The observed enrichment in 13C tions occurred simultaneously on virtually all palaeocontinents
of most pre-Ediacaran carbonates (e.g. Knoll et al. 1986) is cited (Evans 2000; Halverson 2006; Hoffman & Li 2009; but see
in support of the middle (and largest) step in weathering efficiency Eyles & Januszczak 2003 for a contrary view). The basic argument
(Carver & Vardavas 1994), although the isotopic pattern requires is that if ice sheets reached sea level in the warmest areas (i.e. close
an increase in fractional organic burial as opposed to overall to the palaeoequator and where thick non-skeletal carbonates
carbon burial. Increased Neoproterozoic biocatalysed weathering accumulated), then colder areas must have been frozen as well.
receives some support from micropalaeontology (Horodyski & This argument makes no a priori assumption about correlation,
Knauth 1994), molecular divergence analysis (Heckman et al. but correlation follows from the premise.
2001) and clay mineralogy (Kennedy et al. 2006), but remains The stalemate concerns the state of the ocean: was it largely
controversial. The modelled temperature curves (Carver & ice-covered as assumed in the Snowball Earth hypothesis (Kirsch-
Vardavas 1994) feature two relatively cold intervals in Earth vink 1992; Hoffman & Schrag 2002), or substantially ice-free
history, early Palaeoproterozoic in response to continental as posited in the so-called slushball solution (Hyde et al. 2000;
growth and mid-Neoproterozoic in response to the prescribed Peltier et al. 2004; but see Bendtsen 2002; Voigt et al. 2011)?
increase in biocatalysed weathering. Because of subduction, direct evidence from the ocean floor has
been eliminated, or survives only from post-Cryogenian (Edia-
Continental break-up. Supercontinents tend to be dry because most caran), regional-scale glaciation (Kawai et al. 2008). Those who
land is far from the ocean. Upon fragmentation, new continental favour the snowball model point to its ability to account for syngla-
margins are created and land overall is brought closer to the cial Fe- and Fe –Mn deposits, post-glacial cap carbonates, and iso-
source of moisture. As silicate-weathering is catalysed by moisture topic evidence for highly elevated pCO2 during and after glaciation
as well as by temperature (Walker et al. 1981), continental (Bao et al. 2008, 2009). The slushball solution does not account for
break-up should result in lower pCO2 and a colder global these features (Pollard & Kasting 2005). Those who favour the
climate. Cooling will be most severe if rifting occurs in the slushball solution point to evidence that the snowball hypothesis
tropics, and least so for rifting in the polar regions where chemical is incompatible with the fossil record (Knoll 2003; Xiao 2004; Cor-
weathering rates are low. Continental break-up in tropical Pangaea setti et al. 2006; Moczydlowska 2008) and also with the sedimen-
was followed by the coolest period (Late Jurassic –Early Cretac- tology of glacial deposits (e.g. McMechan 2000; Condon et al.
eous) of the Mesozoic era (Frakes et al. 1992); break-up in the 2002; Allen & Étienne 2008). The biotic argument presupposes
polar North Atlantic by the warmest period (Early Eocene) of a knowledge of the limits to survival of species. The argument
the Cenozoic (Zachos et al. 2001). This is consistent with Cryogen- from sedimentology assumes knowledge of when the deposits
ian cooling, related to the break-up and dispersal of equatorial were formed because the snowball hypothesis predicts that glacia-
Rodinia beginning c. 800 Ma (Li et al. 2008). These ideas have tion was ‘polar’ in character at the beginning and ‘temperate’ or
recently been tested with a simplified atmospheric general circula- ‘Alaskan’ at the end. Studies of modern ice-sheet stability show
tion model coupled to a model of the geochemical cycle of carbon that outlet glaciers are buttressed by ice shelves (Dowdeswell
(Donnadieu et al. 2004a, b). et al. 2000; De Angelis & Skvarca 2003; Nick et al. 2009): ice-
It is therefore difficult to accept the argument that Neoprotero- shelf removal triggers ice-sheet deflation, with potential for
zoic glaciation could be explained in terms of a close spatial (as ice-sheet collapse. On a Snowball Earth, the ice shelf is global. Ice-
well as temporal) association with rift valleys and rifted margins, shelf removal, the first step in snowball deglaciation, should
without recourse to extreme climate states (Eyles 1993; Young trigger rapid deflation and collapse of low-latitude ice sheets.
1995a; Eyles & Januszczak 2003). [In Reply to a Comment, Much of the glacial sediment record on continental margins will
Young (1995b) wrote that he did not intend to imply a genetic date from this period. These deposits will bear evidence of open
relationship, but then his paper should have been titled, ‘Was the water (e.g. iceberg rafting and dumping, wave ripples) because
preservation of Neoproterozoic glacial deposits on the margins open water then existed. Sedimentology is thus a clumsy means
of Laurentia related to the fragmentation of two superconti- of testing the snowball hypothesis because of uncertainty over
nents?’.] The Late Quaternary glacial maxima were arguably as which stage of the glacial cycle is stratigraphically preserved
cold as any time in the Phanerozoic, but moraines on the mount- (Hoffman 2005). Moreover, sediment fluxes cannot be estimated
ains of the East African rift system do not extend below because chronology is lacking. Even varve chronology (De Geer
c. 3.5 km above sea level (Osmaston 2004). In the Red Sea rift, 1912), from which the duration of the Holocene was estimated
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 29

1.0

A. Geikie
Fig. 2.4. Bibliographic history of the

Jamieson
Revival 0.8 Pleistocene (‘Newer Pliocene’) glacial
Agassiz
10

Ramsay
controversy. Number of papers (black) and
books (grey) per year concerning the ‘Great
Northern Drift’. Early Martyrs refer to the
Rejection
0.6 essentially correct but ignored glacial
Early Martyrs theories of Esmark (1824), Dobson (1925),

5 0.4 Venetz (1830) and Bernhardi (1832).

Bernhardi
Agassiz refers to the time of tumult
MacCulloch

Esmark associated with the championing of the

Venetz
Playfair
Hutton

glacial theory by Agassiz (1837, 1840,


0.2
Buch
Hall

1842). Rejection was a 20-year period of


eclipse of glacial theory. Revival marks the
widespread acceptance of glacial theory
0 0 following the studies of Ramsay (1860),
1800 1820 1840 1860 1880 Jamieson (1862, 1863, 1865) and Geikie
Year (AD) (1863).

with a fair degree of accuracy, is closed to us because of the rela- age was accompanied by submergence of the land, followed by
tively low palaeolatitudes of most Neoproterozoic deposits, where slow reemergence after the ice disappeared? This problem
seasonality is weak. The decisive falsifying test is chronometric: (based on marine fossils within tills that are now elevated above
any glaciation that is short-lived (,5 Ma) cannot have been a sea level) was central to the interpretation of the Drift (e.g.
snowball glaciation because of the time required to accumulate Smith 1836). Agassiz had offered no explanation for it, there
enough CO2 to overcome the ‘white Earth’s’ albedo. It is the dur- being no marine fossils in the Swiss glacial deposits. Geophysicists
ation of the glaciation, not that of particular glacial deposits, that attributed ice-age submergence to the gravitational ‘pull’ of a
provides the test. The tightest existing constraint on the duration gigantic polar ice cap on the adjacent seas (Adhémar 1842;
of the end-Cryogenian (Marinoan) glaciation is 19.3 + 4.2 Ma Croll 1875), a theory easily accommodated by those like Lyell
(Condon et al. 2005; Zhang et al. 2008). The older Cryogenian and Murchison who believed that the drift came by way of
(Sturtian) glaciation is virtually unconstrained, whereas the mid- iceberg-laden floods of Arctic origin. In contrast, Shaler (1847)
Ediacaran (Gaskiers) glaciation could not have been a snowball and Jamieson (1865) suggested, and later demonstrated with
glaciation because of its short duration (Hoffman & Li 2009). geological evidence (Jamieson 1882), that submergence was
If the snowball hypothesis is not falsified geochronologically, caused by deflection of the lithosphere under the load of the ice
how and when might a consensus on the nature of Cryogenian sheet (glacioisostasy). However, Jamieson’s (1882) theory of
glaciation come about? The history of the Pleistocene glacial submergence came after broad acceptance of the glacial theory
controversy (Fig. 2.4) may offer a preview. Agassiz’s outspoken and cannot therefore have been its cause.
advocacy of the glacial theory of Esmark, Venetz and Bernhardi Consensus on the glacial theory came about because of the over-
engendered much excitement, but failed to achieve consensus. In whelming weight of evidence of the same kind as that obtained in
fact, with few exceptions (James Smith, Charles Maclaren, skeletal form by Agassiz (1842). If consensus on the Snowball
Robert Buckland, Robert Chambers and, on-and-off, Charles Earth hypothesis ultimately rests on the weight of evidence of
Darwin), British geologists soon turned hostile to the glacial the kind identified by Kirschvink (1992) (palaeomagnetic evidence
theory and remained so from 1842 until 1860. One can track the for low-latitude glaciation at sea level, syn-glacial sedimentary
fortunes of the theory through the writings of Charles Lyell Fe2O3 and MnO2 ore formations, and global evidence of abrupt
(1841, 1845, 1851, 1852, 1855, 1857, 1863, 1865), or the Anniver- climate switching (cap carbonates)) a positive outcome appears
sary addresses of the President to the Geological Society, London more and more probable in the end because that evidence has
(see also Woodward 1907; Davies 2007). only become stronger.
Then, in the first half of the 1860s, everything changed. This Must we also wait for two decades or more for consensus
was the result of comprehensive regional studies by a new gener- to emerge? Long delays between initial discussion of radical
ation of Scottish geologists led by Ramsay (1860), Jamieson hypotheses and their broad acceptance is a general phenomenon
(1862, 1863, 1865) and the Geikie brothers (1863, 1874) in what in science (e.g. heliocentrism, biological evolution, continental
may be termed the ‘Scottish glacial revival’. When they began drift, the big bang), so we must suppose there is a reason for the
their studies, they had assumed the glacial theory was false delays. The simplest one is that no consensus can emerge until
because that was the prevailing view. Their own work, systematic after the original antagonists, tethered to hasty judgements, have
and wide-ranging, left them no choice but to conclude that passed from the scene. Given the lengthening of productive
Agassiz had been correct after all. From their time onwards, lifespans, we could be in for a long wait.
critics of the glacial theory became the exception rather than Whatever the final outcome concerning the state of the
the majority. Cryogenian glacial ocean, it is apparent that a third climate state
One is tempted to think that consensus might be reached faster needs to be added to the two first recognized by Brooks (1926):
in the early 21st century than in the mid-19th century. This is ‘non-glacial’ (no ice sheets), ‘glacial’ (polar to temperate ice
doubtful. It is clear from the literature that Victorian geologists sheets) and ‘pan-glacial’ (ice sheets at all latitudes).
actually read each others’ papers. No one could claim that this is
true today. As a comparative newcomer to Neoproterozoic glacial studies, I am grateful
Will consensus be achieved through some conceptual or techni- for reviews of a draft of this paper by M. Hambrey and two anonymous
cal breakthough? I doubt it: there have been several already and reviewers. I also thank Barry Cooper for a preprint of his paper in Earth Sciences
consensus is no closer than at the millenium. Was consensus History on early contributions from South Australia. However, I bear sole res-
on Pleistocene glaciation brought about by Jamieson’s (1882) ponsibility for the views expressed here. This work was supported by the US
solution to the submergence problem, which was that each ice National Science Foundation (grant EAR-0417422), the Earth System Evolution
30 P. F. HOFFMAN

Project of the Canadian Institute for Advanced Research (CIFAR) and Harvard Benn, D. I. & Evans, D. J. A. 1998. Glaciers and Glaciation. Arnold
University. This represents a contribution of the IUGS- and UNESCO-funded Publishers, London.
IGCP (International Geoscience Programme) Project #512. Berger, W. H. 1982. Increase of carbon dioxide in the atmosphere during
deglaciation: The coral reef hypothesis. Naturwissenschaften, 69,
87 – 88.
References Berner, R. A. & Maasch, K. A. 1996. Chemical weathering and controls
on atmospheric O2 and CO2: fundamental principles were enunciated
Adhémar, J. A. 1842. Révolutions de la Mer. Fain et Thunot, Paris. by J. J. Ebelmen in 1845. Geochimica et Cosmochimica Acta, 60,
Agassiz, L. 1837. Discours prononcé à l’ouverture des séances de la 1633– 1637.
Société Helvétique des Sciences Naturelles, à Neuchâtel, le 24 Bertrand-Sarfati, J., Flicoteaux, R., Moussine-Pouchkine, A. &
juillet 1837. Actes de la Société Helvétique des Sciences Naturelles Aı̈t Kaci Ahmed, A. 1997. Lower Cambrian apatitic stromatolites
(22ème session, Neuchâtel, 24 – 26 juillet 1837), 5– 32. (Reprinted and phospharenites related to the glacio-eustatic cratonic rebound
in English: Upon glaciers, moraines, and erratic blocks. Edinburgh (Sahara, Algeria). Journal of Sedimentary Research, 67, 957– 974.
New Philosophical Journal, 24, 364– 383, 1838.) Bernhardi, A. 1832. Wie kamen die aus dem Norden stammenden
Agassiz, L. 1840. Études sur les glaciers. Neuchâtel et Soleure. (Re- Felsbruchstücke und Geshiebe, welche man in Norddeutschland
printed in English: Studies on glaciers, preceded by the Discourse und den benachbarten Ländern findet, an ihre gegenwärtigen
of Neuchâtel. Hafner, New York, 1967.) Fundorte? (How did the rock fragments and boulders of northern
Agassiz, L. 1842. The glacial theory and its recent progress. New origin found in northern Germany and neighboring countries get to
Edinburgh Philosophical Journal, 33, 217–283. their present positions?) Jahrbuch für Mineralogie, Geognosie,
Aitken, J. D. 1991. The Ice Brook Formation and post-Rapitan, Late Geologie, und Petrefaktenkunde, 3, 257– 267.
Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. Bestmann, M., Rice, A. H. N., Langenhorst, F., Grasemann, B. &
Geological Survey of Canada, Bulletin, 404, 43. Heidelbach, F. 2006. Subglacial bedrock welding associated with
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves glacial earthquakes. Journal of the Geological Society, London,
in the aftermath of a Neoproterozoic glaciation. Nature, 433, 163, 417– 420.
123– 127. Biju-Duval, B. & Gariel, O. 1969. Nouvelles observations sur les
Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to Snowball phénomènes glaciaires ‘Éocambriens’ de la bordure nord de la
Earth. Nature Geoscience, 1, 817–825. synéclise de Taoudeni, entre le Hank et le Tanezrouft, Sahara occi-
Allen, P., Leather, J. & Brasier, M. D. 2004. The Neoproterozoic Fiq dental. Palaeogeography, Palaeoclimatology, Palaeoecology, 6,
glaciation and its aftermath, Huqf Supergroup of Oman. Basin 283– 315.
Research, 16, 507 – 534. Blackett, P. M. S. 1961. Comparison of ancient climates with the ancient
Alley, R. B., Blankenship, D. D., Rooney, S. T. & Bentley, C. R. latitudes deduced from rock magnetic measurements. Proceedings of
1989. Sedimentation beneath ice shelves – the view from ice the Royal Society of London, Series A, 263, 1 –30.
stream B. Marine Geology, 85, 101– 120. Blackwelder, E. 1932. An ancient glacial formation in Utah. Journal of
Almeida, F. F. M. 1946. Origem dos minérios de ferro y manganes de Geology, 40, 289–304.
Urucum (Corumbá, Estado Mato Grosso). Boletim de Divisão de Blanford, W. T., Blanford, H. F. & Theobald, W. 1859. On the
Geologia e Mineralogia, 119, 58. geological structure and relations of the Talcheer Coal Field, in the
Anderson, J. B. 1999. Antarctic Marine Geology. Cambridge University District of Cuttack. Memoirs of the Geological Survey of India, 1,
Press, Cambridge. 33 – 89.
Anderson, J. B. & Ashley, G. M. (eds) 1991. Glacial Marine Sedimen- Boulton, G. S. 1990. Sedimentary and sea level changes during glacial
tation: Paleoclimatic Significance. Geological Society of America cycles and their control on glacimarine facies architecture. In: Dow-
Special Paper, 261, Boulder, CO. deswell, J. A. & Scourse, J. D. (eds) Glacimarine Evironments:
Arnaud, E. 2004. Giant cross-beds in the Neoproterozoic Port Askaig Processes and Sediments. Geological Society, London, Special Pub-
Formation, Scotland: implications for snowball Earth. Sedimentary lications, 53, 15– 52.
Geology, 165, 155–174. Briden, J. C. 1970. Palaeolatitude distribution of precipitated sediments.
Arnaud, E. 2008. Deformation in the Neoproterozoic Smalfjord In: Runcorn, S. K. (ed.) Palaeogeophysics. Academic Press,
Formation, northern Norway: an indicator of glacial depositional London, 437– 444.
conditions? Sedimentology, 55, 335– 356. Briden, J. C. & Irving, A. 1964. Palaeolatitude spectra of sedimentary
Assereto, R. L. A. M. & Kendall, C. G. St. C. 1977. Nature, origin palaeoclimatic indicators. In: Nairn, A. E. M. (ed.) Problems in
and classification of peritidal tepee structures and related breccias. Palaeoclimatology. John Wiley & Sons, New York, 199–224.
Sedimentology, 24, 153–210. Brooks, C. E. P. 1922. The Evolution of Climate. Ernest Benn, London.
Bao, H., Lyons, J. R. & Zhou, C. 2008. Triple oxygen isotope evidence Brooks, C. E. P. 1926. Climate Through the Ages. R.V. Coleman,
for elevated CO2 levels after a Neoproterozoic glaciation. Nature, New York.
452, 504– 506. Buchan, K. L., Ernst, R. E., Hamilton, M. A., Mertanen, S.,
Bao, H., Fairchild, I. J., Wynn, P. M. & Spötl, C. 2009. Stretching the Pesonen, L. J. & Elming, S.-Å. 2001. Rodinia: the evidence from
envelope of past surface environments: Neoproterozoic glacial lakes integrated palaeomagnetism and U–Pb geochronology. Precambrian
from Svalbard. Science, 323, 119–122. Research, 110, 9– 32.
Barrett, P. J. (ed.) 1989. Antarctic Cenozoic History from the CIROS-1 Budyko, M. I. 1969. The effect of solar radiation variations on the climate
Drillhole, McMurdo Sound. DSIR Bulletin, 245, Science Information of the Earth. Tellus, 21, 611–619.
Publishing Center, Wellington, New Zealand. Bjørlykke, K. 1967. The Eocambrian Reusch moraine at Bigganjargga
Baud, L. 1933. Le conglomérat argilo-calcareux dans la région de Kayes and the geology of Varangerfjord, northern Norway. Norsk Geolo-
et de Bafoulabé et sa position stratigraphique. Compte Rendu, 197, giske Undersøkelse, 251, 18 –44.
172– 173. Cahen, L. 1950. Le Calcaire de Sekelolo, le complexe tillitique et la
Beetz, W. 1929. Über das Wahrscheinlich altcambrische oder jungproter- dolomie rose C1 dans l’anticlinal de Congo de Kati (Bas-Congo).
ozoische Alter der Glazialschichten an der Basis des Kundelungu- Annales de la Musée Royale Congo Belge, Sciences Géologiques,
Systems in Katanga und am unteren Kongo. Neues Jahrbuch für 7, 1 –55.
Mineralogie, Geologie und Paläontologie, Beilagebände, Abteiling Cahen, L. 1963. Glaciations anciennes et dérive des continents (Ancient
B, 61, 61– 82. glaciations and continental drift). Annales de la Société Géologique
Bendtsen, J. 2002. Climate sensitivity to changes in solar insolation in a de Belgique, 86, 19– 83.
simple coupled climate model. Climate Dynamics, 18, 595–609. Cahen, L. & Lepersonne, J. 1981. Proterozoic diamictites of Lower
Bendtsen, J. & Bjerrum, C. J. 2002. Vulnerability of climate on Earth Zaire. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-
to sudden changes in insolation. Geophysical Research Letters, 29, Pleistocene Glacial Record. Cambridge University Press, Cam-
doi: 10.1029/2002GL014829. bridge, 153– 157.
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 31

Caldas, J. 1979. Evidencias de una glaciación Precambriana en la costa Crowell, J. C. 1957. Origin of pebbly mudstones. Geological Society of
sur del Perú. Segundo Congreso Geológico Chileno, Arica, Vol. J, America Bulletin, 68, 993–1010.
29 – 37. Crowell, J. C. 1964. Climatic significance of sedimentary deposits con-
Calver, C. R. & Walter, M. R. 2000. The late Neoproterozoic Grassy taining dispersed megaclasts. In: Nairn, A. E. M. (ed.) Problems in
Group of King Island, Tasmania: correlation and palaeogeographic Palaeoclimatology. Interscience, London, 86– 99.
significance. Precambrian Research, 100, 299– 312. Crowell, J. C. & Frakes, L. A. 1970. Phanerozoic glaciations
Carver, J. H. & Vardavas, I. M. 1994. Precambrian glaciations and the causes of ice ages. American Journal of Science, 268,
and the evolution of the atmosphere. Annales Geophysicae, 12, 193– 224.
674– 682. Cunningham, F. F. 1990. James David Forbes, Pioneer Scottish Glaciol-
Catling, D. C., Zahnle, K. J. & McKay, C. P. 2001. Biogenic methane, ogist. Scottish Academic Press, Edinburgh.
hydrogen escape, and the irreversible oxidation of early Earth. Dal, A. 1900. Geologiske iagttagselser omkring Varangerfjorden. Norges
Science, 293, 839– 843. Geologiske Undersøkelse, 28, 1– 16.
Cawood, P. A., Nemchin, A. A., Smith, M. & Loewy, S. 2003. Source of David, T. W. E. 1907a. Conditions of climate at different geological
the Dalradian Supergroup constrained by U– Pb dating of detrital epochs, with special reference to glacial epochs. 10th International
zircon and implications for the East Laurentian margin. Journal of Geological Congress, Mexico City, 1906, Compte Rendu, 437–482.
the Geological Society, London, 160, 231–246. David, T. E. W. 1907b. Some problems of Australian glaciations. Report
Chamberlin, T. C. 1898. The ulterior basis of time divisions and the of the Australasian Association for the Advancement of Science,
classification of geological history. Journal of Geology, 6, 449– 462. 11, 457– 465.
Chamberlin, T. C. 1899. An attempt to frame a working hypothesis of Davies, G. L. 1968. The tour of the British Isles made by Louis Agassiz in
the cause of glacial periods on an atmospheric basis. Journal of 1840. Annals of Science, 24, 131– 146.
Geology, 7, 545–584. Davies, G. L. H. 2007. Whatever is under the Earth: the Geological
Charpentier, J. de. 1837. Some conjectures regarding the great revolu- Society of London, 1807– 2007. Geological Society, London.
tions which have so changed the surface of Switzerland, and particu- Davies, K. A. 1939. The glacial sediments of Bunyoro, N.W. Uganda.
larly that of the Canton of Vaud, as to give rise to its present aspect. Bulletin of the Geological Survey of Uganda, 3, 20 –37.
Edinburgh New Philosophical Journal, 22, 27 –36. Davies, K. A. 1940. The glacial series of Bunyoro, north Uganda. 17th
Chumakov, N. M. 1964. Präkambrische tillit-ähnliche Gesteine der International Geological Congress, Moscow, 1937, 6, 115– 119.
Sowjetunion. Geologische Rundschau, 54, 83 –102. De Angelis, H. & Skvarca, P. 2003. Glacier surge after ice shelf
Chumakov, N. M. 1968. On the character of the Late Precambrian glacia- collapse. Science, 299, 1560– 1562.
tion of Spitsbergen (in Russian). Doklady Akademiya Nauk USSR, De Geer, G. 1912. A geochronology of the last 12,000 years. 11th Inter-
Geological Series, 180, 1446–1449. national Geological Congress, Stockholm, 1910, 1, 241– 253.
Chumakov, N. M. 1981. Upper Proterozoic glaciogenic rocks and their Deynoux, M. & Trompette, R. 1976. Discussion: Late Precambrian mix-
stratigraphic significance. Precambrian Research, 15, 373–395. tites: glacial and/or nonglacial? Dealing especially with the mixtites
Chumakov, N. M. 1992. The problems of old glaciations (pre-Pleistocene of West Africa. American Journal of Science, 276, 1302– 1315.
glaciogeology in the USSR). Soviet Scientific Reviews, Section G Dobson, P. 1925. Remarks on Bowlders. American Journal of Science and
Geology, 1, 1– 208. Arts, 10, 217.
Claire, M. W., Catling, D. C. & Zahnle, K. J. 2006. Biogeochemical Donnadieu, Y., Ramstein, G., Fluteau, F., Besse, J. & Meert, J. 2002.
modelling of the rise in atmospheric oxygen. Geobiology, 4, Is high obliquity a plausible cause for Neoproterozoic glaciations?
239– 269. Geophysical Research Letters, 29, doi: 10.1029/2002GL015902.
Cloud, P., Wright, L. A., Williams, E. G., Diehl, P. & Walter, M. R. Donnadieu, Y., Goddéris, Y., Ramstein, G., Nédélec, A. & Meert, J.
1974. Giant stromatolites and associated vertical tubes from the 2004a. A ‘snowball Earth’ climate triggered by continental break-up
upper Proterozoic Noonday Dolomite, Death Valley region, eastern through changes in runoff. Nature, 428, 303– 306.
California. Geological Society of America Bulletin, 85, 1869– 1882. Donnadieu, Y., Ramstein, G., Fluteau, F., Roche, D. & Gonopolski,
Coleman, A. P. 1907. A Lower Huronian ice age. American Journal of A. 2004b. The impact of atmospheric and oceanic heat transport on
Science, 23, 187– 192. the sea-ice instability during the Neoproterozoic. Climate Dynamics,
Coleman, A. P. 1926. Ice Ages: Recent and Ancient. MacMillan, 22, 293– 306.
New York. Dorr, J. V. H. 1945. Manganese and iron deposits of Morro do Urucum,
Condon, D. J., Prave, A. R. & Benn, D. I. 2002. Neoproterozoic glacial- Mato Grosso, Brazil. United States Geological Survey Bulletin,
rainout intervals: observations and implications. Geology, 30, 35– 38. 946-A, 1 –47.
Condon, D., Zhu, M., Bowring, S. A., Wang, W., Yang, A. & Jin, Y. Dott, R. H. Jr. 1961. Squantum ‘tillite’, Massachusetts – evidence of
2005. U– Pb ages from the Neoproterozoic Doushantuo Formation, glaciation or subaqueous mass movements? Geological Society of
China. Science, 308, 95– 98. America Bulletin, 72, 1289– 1305.
Cooper, B. 2009. Snowball Earth: the early contribution from South Dow, D. B. 1965. Evidence of a Late Pre-Cambrian glaciation in the Kim-
Australia. Earth Sciences History, 29, 121– 145. berley Region of Western Australia. Geological Magazine, 102,
Corkoran, M. 2007. ‘Cap carbonates’ and Neoproterozoic glacigenic 407– 419.
successions from the Kimberley region, north-west Australia. Sedi- Dowdeswell, J. A. & Scourse, J. D. (eds) 1990. Glacimarine Environ-
mentology, 54, 871– 903. ments: Processes and Sediments. Geological Society, London,
Corsetti, F. A. & Grotzinger, J. P. 2005. Origin and significance of tube Special Publications, 53.
structures in Neoproterozoic post-glacial cap carbonates: example Dowdeswell, J. A., Whittington, J. A., Jennings, A. E., Andrews, J.
from Noonday Dolomite, Death Valley, United States. Palaios, 20, T., Mackensen, A. & Marienfield, P. 2000. An origin for lami-
348– 363. nated glacimarine sediments through sea-ice build-up and supressed
Corsetti, F. A. & Kaufman, A. J. 2005. The relationship between the iceberg rafting. Sedimentology, 47, 557– 576.
Neoproterozoic Noonday Dolomite and the Ibex Formation: new Drewry, D. 1986. Glacial Geologic Processes. Edward Arnold, London.
observations and their bearing on ‘snowball Earth’. Earth-Science Dunn, P. R., Thomson, B. P. & Rankama, K. 1971. Late Pre-Cambrian
Reviews, 73, 63 –78. glaciation in Australia as a stratigraphic boundary. Nature, 231,
Corsetti, F. A., Olcott, A. N. & Bakermans, C. 2006. The biotic 498– 502.
response to Neoproterozoic snowball Earth. Palaeogeography, Du Toit, A. L. 1922. The Carboniferous glaciation of South Africa.
Palaeoclimatology, Palaeoecology, 232, 114– 130. Transactions of the Geological Society of South Africa, 24, 188– 227.
Crawford, A. R. & Daily, B. 1971. Probable non-synchroneity of Late Ebelmen, J.-J. 1845. Sur les produits de la décomposition des especes
Precambrian glaciations. Nature, 230, 111– 112. minérales de la famille silicates. (On the products of the weathering
Croll, J. 1875. Climate and Time in Their Geological Relations: A Theory of silicate minerals.) Annales des Mines, 7, 3– 66.
of Secular Changes of the Earth’s Climate. Appleton & Co., Ebelmen, J.-J. 1847. Sur la décomposition des roches. (On the weathering
New York. of rocks.) Annales des Mines, 12, 627– 654.
32 P. F. HOFFMAN

Edwards, M. B. 1975. Glacial retreat sedimentation in the Smalfjord Geikie, A. 1865. The Scenery of Scotland, Viewed in Connexion with its
Formation, Late Precambrian, north Norway. Sedimentology, 22, Physical Geology. MacMillan, London.
75 –94. Geikie, J. 1874. The Great Ice Age. D. Appleton & Co., New York, 545.
Edwards, M. B. 1984. Sedimentology of the Upper Proterozoic glacial Gevers, T. W. & Beetz, W. 1940. Pre-Dwyka glacial periods in southern
record, Vestertana Group, Finnmark, North Norway. Norges Geolo- Africa. 17th International Geological Congress, Moscow, 1937, 6,
giske Undersøkelse Bulletin, 394, 76. 65 – 98.
Edwards, M. B. 1997. Discussion of glacial or non-glacial origin of the Gibsher, A. S. & Khomentovsky, V. V. 1990. The section of the Tsagaan
Bigganjargga tillite, Finnmark, northern Norway. Geological Maga- Olum and Bayan Gol Formations of the Vendian – Lower Cambrian
zine, 134, 873– 876. in the Dzabkhan zone of Mongolia. In: Khomentovsky, V. V.,
Eisbacher, G. H. 1981. Sedimentary tectonics and glacial record in the Gibsher, A. S. & Karlova, G. A. (eds) The Late Precambrian and
Windermere Supergroup, Mackenzie Mountains, northwestern Early Paleozoic of Siberia. Institut Geologii I Geofiziki, Sibirskoe
Canada. Geological Survey of Canada Paper, 80-27, 40. Otdelenie, Akademiya Nauk SSSR, Novosibirsk, 79 –91.
Elverhoi, A. 1984. Glacigenic and associated marine sediments in the Girdler, R. W. 1964. The palaeomagnetic latitudes of possible ancient
Weddell Sea, Fjords of Spitsbergen and the Barents Sea: a review. glaciations. In: Nairn, A. E. M. (ed.) Problems in Palaeoclimatology.
Marine Geology, 57, 53– 88. Interscience, London, 119–149.
Embleton, B. J. J. & Williams, G. E. 1986. Low latitude of deposition for Goddéris, Y., Donnabieu, Y. et al. 2003. The Sturtian ‘snowball’ gla-
late Precambrian periglacial varvites in South Australia: implications ciation: fire and ice. Earth and Planetary Science Letters, 211, 1– 12.
for palaeoclimatology. Earth and Planetary Science Letters, 79, Gold, T. 1955. Instability of the Earth’s axis of rotation. Nature, 175,
419– 430. 526– 529.
Eriksson, E. 1968. Air– ocean – icecap interactions in relation to climatic Gorin, G. E., Racz, L. G. & Walter, M. R. 1982. Late Precambrian –
fluctuations and glaciation cycles. Meteorological Monographs, 8, Cambrian sediments of Huqf Group, Sultanate of Oman. American
68 –92. Association of Petroleum Geologists Bulletin, 66, 2609– 2627.
Esmark, J. 1824. Bidrag til vor jordklodes historie. (Contribution to the Grotzinger, J. P. & Knoll, A. H. 1995. Anomalous carbonate
history of our Earth.) Magazin for Naturvidenskaberne, 2(1), precipitates: is the Precambrian the key to the Permian? Palaios, 10,
29 –54. (Reprinted in English: Remarks tending to explain the geo- 578– 596.
logical history of the Earth. Edinburgh New Philosophical Journal, Halverson, G. P. 2006. A Neoproterozoic chronology. In: Xiao, S. &
2, 107–121, 1827.) Kaufman, A. J. (eds) Neoproterozoic Geobiology and Paleobiology.
Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic Springer, Dordrecht, 231– 271.
constraints upon the Neoproterozoic climatic paradox. American Halverson, G. P., Hoffman, P. F., Schrag, D. P. & Kaufman, J. A.
Journal of Science, 300, 347– 433. 2002. A major perturbation of the carbon cycle before the Ghaub
Evans, D. A. D. 2006. Proterozoic low orbital obliquity and axial-dipolar glaciation (Neoproterozoic) in Namibia: prelude to snowball
geomagnetic field from evaporite palaeolatitudes. Nature, 444, Earth? Geophysics, Geochemistry, Geosystems, 3, doi: 10.1029/
51 –55, doi: 10.1038/nature05203. 2001GC000244.
Eyles, N. 1993. Earth’s glacial record and its tectonic setting. Earth- Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
Science Reviews, 35, 1– 248. glaciation (Neoproterozoic) in northeast Svalbard. Basin Research,
Eyles, N. & Januszczak, N. 2003. ‘Zipper-rift’: a tectonic model for 16, 297– 324.
Neoproterozoic glaciations during the breakup of Rodinia after 750 Hambrey, M. J. 1994. Glacial Environments. CRC Press, Boca Raton, FL.
Ma. Earth-Science Reviews, 65, 1– 73. Hambrey, M. J. & Harland, W. B. 1981. Earth’s Pre-Pleistocene
Fairchild, I. J. 1983. Effects of glacial transport and neomorphism on Glacial Record. Cambridge University Press, Cambridge.
Precambrian dolomite crystal sizes. Nature, 304, 714– 716. Hambrey, M. J. & Harland, W. B. 1985. The Late Proterozoic glacial
Fawcett, P. J. & Boslough, M. B. E. 2002. Climatic effects of an era. Palaeogeography, Palaeoclimatology, Palaeoecology, 51,
impact-induced equatorial debris ring. Journal of Geophysical 255– 272.
Research, 107(D15), doi: 10.1029/2001JD001230. Hambrey, M. J. & Spencer, A. M. 1987. Late Precambrian glaciation of
Fedden, F. G. S. 1875. On the evidence of ‘ground-ice’ in tropical India, central East Greenland. Meddelelser om Grønland, Geoscience, 19,
during the Tálchı́r period. Records of the Geological Survey of India, 1 –50.
8, 16 –18. Hamdi, B. 1992. Late Precambrian glacial deposits in central Iran. 29th
Forbes, E. 1846. On the connexion between the distribution of the existing International Geological Congress, Kyoto, 1992, Abstracts, 2, 263.
fauna and flora of the British Isles, and the geological changes which Harland, W. B. 1964a. Evidence of late Precambrian glaciation and its
have affected their area, especially during the epoch of the Northern significance. In: Nairn, A. E. M. (ed.) Problems in Palaeoclimatol-
Drift. Memoirs of the Geological Survey of Great Britain, 1, ogy. Interscience, London, 119– 149.
336– 403. Harland, W. B. 1964b. Critical evidence for a great infra-Cambrian
Fourier, J. 1824. Remarques générales sur les températures du globe ter- glaciation. Geologische Rundschau, 54, 45 –61.
restre et des espaces planétaires. Annales de Chimie et de Physique Harland, W. B. 1983. The Proterozoic glacial record. In: Medaris, L. G.,
(2nd series), 27, 136–167. Translated as, General remarks on the Jr., Byers, W. C., Mickelson, D. M. & Shanks, W. C. (eds) Proter-
temperature of the terrestrial globe and the planetary spaces. Ameri- ozoic Geology. Geological Society of America, Memoir, 161,
can Journal of Science and Arts, 32, 1– 20, 1837. 279– 288.
Føyn, S. 1937. The Eocambrian series of the Tana district, northern Harland, W. B. & Bidgood, E. T. 1959. Palaeomagnetism in some
Norway. Norsk Geologisk Tidsskrift, 17, 65– 164, 4 plates, Norwegian Sparagmites and the Late Pre-Cambrian ice age.
1:250,000 scale map. Nature, 184, 1860– 1862.
Frakes, L. A., Francis, J. E. & Syktus, J. I. 1992. Climate Modes of the Harland, W. B. & Rudwick, M. J. S. 1964. The great infra-Cambrian ice
Phanerozoic. Cambridge University Press, Cambridge, 274. age. Scientific American, August, 42 –49.
Frankel, H. 1987. Jan Hospers and the rise of paleomagnetism. Eos, Harland, W. B. & Wilson, C. B. 1956. The Hecla Hoek Succession in
Transactions of the American Geophysical Union, 68, 577, 579– 581. Ny Friesland, Spitsbergen. Geological Magazine, 93, 2265– 2286.
Furon, R. 1933. Observations sur la stratigraphie de l’ouest africain Heckman, D. S., Geiser, D. M., Eldell, B. R., Stauffer, R. L., Kardos,
(Mauritanie et Soudan). Compte Rendu, 196, 1905–1906. N. L. & Hedges, S. B. 2001. Molecular evidence for the early colo-
Gaertner, H. R. von. 1943. Bemerkungen über den Tillit von Biggan- nization of land by fungi and plants. Science, 293, 1129–1133.
jargga am Varangerfjord. Geologische Rundschau, 34, 226–231. Hegenberger, W. 1987. Gas escape structures in Precambrian peritidal
Garwood, E. J. & Gregory, J. W. 1898. Contribution to the glacial carbonate rocks. Communications of the Geological Survey of
geology of Spitsbergen. Quarterly Journal of the Geological Namibia, 3, 49 –55.
Society, London, 54, 197–225. Heezen, B. C. & Hollister, C. 1964. Turbidity currents and glaciation.
Geikie, A. 1863. On the phenomena of the glacial drift of Scotland. Trans- In: Nairn, A. E. M. (ed.) Problems in Palaeoclimatology. Inter-
actions of the Geological Society of Glasgow, 1, 190. science, London, 99– 108.
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 33

Hintze, F. F. 1913. A contribution to the geology of the Wasatch Moun- Jamieson, T. F. 1862. On the ice-worn rocks of Scotland. Quarterly
tains, Utah. Annals of the New York Academy of Sciences, 23, Journal of the Geological Society of London, 18, 164– 184.
85 – 143. Jamieson, T. F. 1863. On the parallel roads of Glen Roy, and their place in
Hitchcock, E. 1841. Diluvium, or drift. In: Final Report on the Geology the history of the glacial period. Quarterly Journal of the Geological
of Massachusetts. J.S. & C. Adams, Amherst, MA, 350– 406; post- Society of London, 19, 235– 259.
script, 3a –11a. Jamieson, T. F. 1865. On the history of the last geological changes in Scot-
Hoffman, P. F. & Maloof, A. C. 1999. Glaciation: the snowball theory land. Quarterly Journal of the Geological Society of London, 21,
still holds water. Nature, 397, 384. 161– 203.
Hoffman, P. F. & Schrag, D. P. 2000. Snowball Earth. Scientific Amer- Jamieson, T. F. 1882. On the cause of the depression and re-elevation of
ican, 282, 68 –75. the land during the glacial period. Geological Magazine, 9, 400– 407,
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: 457– 466.
testing the limits of global change. Terra Nova, 14, 129– 155. Jenkins, G. S. 2000. Global climate model high-obliquity solutions to the
Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P. ancient climate puzzles of the faint-young-Sun paradox and low-
1998. A Neoproterozoic snowball Earth. Science, 281, 1342–1346. latitude Proterozoic glaciation. Journal of Geophysical Research,
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., 105, 7357–7370.
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635 Jensen, P. A. & Wulff-Pedersen, E. 1996. Glacial or non-glacial origin
Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary for the Bigganjargga tillite, Finnmark, northern Norway. Geological
Science Letters, 258, 114– 131. Magazine, 133, 137– 145.
Holland, T. H. 1908. On the occurrence of striated boulders in the Jiang, G., Sohl, L. E. & Christie-Blick, N. 2003. Neoproterozoic
Blaini Formation of Simla, with a discussion of the geological stratigraphic comparison of the Lesser Himalaya (India) and Yangtse
age of the beds. Records of the Geological Survey of India, 37, clock (south China): paleogeographic implications. Geology, 31,
129– 135. 917– 920.
Holtedahl, O. 1918. Varangerhalvøn: Strøet omkring bunden av Varan- Jiang, G., Kennedy, M. J., Christie-Blick, N., Wu, H. & Zhang, S.
gerfjord. Norsk Geologiske Undersøkelse, 84, 148– 173. 2006. Stratigraphy, sedimentary structures, and textures of the late
Hoffman, P. F. 2005. 28th DeBeers Alex. Du Toit Memorial Lecture: on Neoproterozoic Doushantuo cap carbonate in South China. Journal
Cryogenian (Neoproterozoic) ice-sheet dynamics and the limitations of Sedimentary Research, 76, 978– 995.
of the glacial sedimentary record. South African Journal of Geology, Johnston, J. D. 1993. Ice wedge casts in the Dalradian of South Donegal
108, 557– 576. – evidence for subaerial exposure of the Boulder Bed. Irish Journal
Hoffman, P. F. & Li, Z. X. 2009. A palaeogeographic context for Neopro- of Earth Sciences, 12, 13– 26.
terozoic glaciation. Palaeogeography, Palaeoclimatology, Palaeo- Katz, H. R. 1961. Late Precambrian to Cambrian stratigraphy in East
ecology, 277, 158– 172. Greenland. In: Raasch, G. O. (ed.) Geology of the Arctic, Vol 1.
Horodyski, R. J. & Kauth, L. P. 1994. Life on land in the Precambrian. University of Toronto Press, Toronto, 299– 328.
Science, 263, 494– 498. Kaufman, A. J., Jacobsen, S. B. & Knoll, A. H. 1993. The Vendian
Howchin, W. 1901. Preliminary note on the existence of glacial beds of record of Sr and C isotopic variations in seawater: implications for
Cambrian age in South Australia. Transactions of the Royal Society tectonics and paleoclimate. Earth and Planetary Science Letters,
of South Australia, 21, 74 –86. 120, 409–430.
Howchin, W. 1903. Report on South Australian Glacial Investigation Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes, ice
Committee. Report of the Australasian Association for the Advance- ages, and terminal Proterozoic Earth history. Proceedings of the
ment of Science, 9, 194– 200. National Academy of Sciences USA, 94, 6600–6605.
Howchin, W. 1908. Glacial beds of Cambrian age in South Kaufman, A. J., Jiang, G., Christie-Blick, N., Banerjee, D. M. & Rai,
Australia. Quarterly Journal of the Geological Society, London, 64, V. 2006. Stable isotope record of the terminal Neoproterozoic Krol
234– 259. platform in the Lesser Himalayas of northern India. Precambrian
Howchin, W. 1912. Australian glaciations. Journal of Geology, 20, Research, 147, 156–185.
193– 227. Kawai, T., Windley, B. F., Terabayashi, M., Yamamoto, H., Isozaki,
Hunt, T. S. 1880. The chemical and geological relations of the atmos- Y. & Maruyama, S. 2008. Neoproterozoic glaciation in the
phere. American Journal of Science, 19, 349– 363. mid-oceanic realm: an example from hemi-pelagic mudstones on
Hyde, W. T., Crowley, T. J., Baum, S. K. & Peltier, W. R. 2000. Neo- Llanddwyn Island, Anglesey, UK. Gondwana Research, 14,
proterozoic ‘snowball Earth’ simulations with a coupled climate/ice- 105– 114.
sheet model. Nature, 405, 425–429. Kendall, C. G. St. C. & Warren, J. 1987. A review of the origin and
Irving, E. 1956. Palaeomagnetic and palaeoclimatological aspects of setting of tepees and their associated fabrics. Sedimentology, 34,
polar wandering. Geofisica Pura e Applicada, 33, 23 – 41. 1007–1027.
Irving, E. 1957a. Directions of magnetization in the Carboniferous glacial Kennedy, M. J. 1996. Stratigraphy, sedimentology, and isotopic geo-
varves of Australia. Nature, 180, 280– 281. chemistry of Australian Neoproterozoic postglacial cap dolostones:
Irving, E. 1957b. Analysis of the palaeomagnetism of the Torridonian deglaciation, d13C excursions, and carbonate precipitation. Journal
sandstone series of north-west Scotland. Philosophical Transactions of Sedimentary Research, 66, 1050– 1064.
of the Royal Society, London, Series A, 250, 83 –89. Kennedy, M., Droser, M., Mayer, L. M., Pevear, D. & Mrofka, D.
Irving, E. 1959. Palaeomagnetic pole positions: a survey and analysis. 2006. Late Precambrian oxygenation: inception of the clay mineral
Geophysical Journal of the Royal Astronomical Society, 2, 51 –79. factory. Science, 311, 1446–1449.
Irving, E. 2008. Jan Hospers’s key contributions to geomagnetism. Eos, Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K.-H. &
Transactions, American Geophysical Union, 89, 457–458. Arthur, M. A. 1998. Two or four Neoproterozoic glaciations?
Irving, E. & Runcorn, S. K. 1957. The origin of the palaeomagnetism of Geology, 26, 1059– 1063.
the Torridonian sandstones of north-west Scotland. Philosophical Kilburn, C., Pitcher, W. S. & Shackleton, R. M. 1965. The stratigra-
Transactions of the Royal Society, London, Series A, 250, 100– 110. phy and origin of the Portaskaig Boulder Bed series (Dalradian). Geo-
Isotta, C. A., Rocha-Campos, A. C. & Yoshida, R. 1969. Striated pave- logical Journal, 4, 343– 360.
ment of the Upper Pre-Cambrian glaciation in Brazil. Nature, 222, Kiessling, W. 2001. Paleoclimatic significance of Phanerozoic reefs.
466– 468. Geology, 29, 751– 754.
Jack, R. L. 1913. The Mount Grainger Goldfield. Report of the Geological King, L. H., Rokoengen, K., Fader, G. B. J. & Gunleiksrud, T. 1991.
Survey of South Australia, 1, 1 –24. Till-tongue stratigraphy. Geological Society of America Bulletin,
James, N. P., Narbonne, G. M. & Kyser, T. K. 2001. Late Neoprotero- 103, 637–659.
zoic cap carbonates: Mackenzie Mountains, northwestern Canada: Kirschvink, J. L. 1992. Late Proterozoic low-latitude glaciation: the
precipitation and global glacial meltdown. Canadian Journal of snowball Earth. In: Schopf, J. W. & Klein, C. (eds) The Proterozoic
Earth Sciences, 38, 1229–1262. Biosphere. Cambridge University Press, Cambridge, 51 –52.
34 P. F. HOFFMAN

Klein, C. & Beukes, N. J. 1993. Sedimentology and geochemistry of the Lyell, C. 1845. Travels in North America, 1841– 2, with Geological
glacigenic Late Proterozoic Rapitan iron-formation in Canada. Econ- Observations on the United States, Canada, and Nova Scotia
omic Geology, 88, 542– 565. (2 vol.). Wiley & Putnam, New York, 231. (Reprinted in 1978 by
Knoll, A. H. 2003. Life on a Young Planet, the First Three Billion Arno Press, New York.)
Years of Evolution on Earth. Princeton University Press, Princeton. Lyell, C. 1851. Manual of Elementary Geology, 3rd edn. John Murray,
Knoll, A. H. & Walter, M. R. 1992. Latest Proterozoic stratigraphy and London.
Earth history. Nature, 356, 673–678. Lyell, C. 1852. Manual of Elementary Geology, 4th edn. John Murray,
Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K. & Lambert, I. B. London.
1986. Secular variation in carbon isotope ratios from Upper Protero- Lyell, C. 1855. Manual of Elementary Geology, 5th edn. John Murray,
zoic successions of Svalbard and East Greenland. Nature, 321, London.
831– 838. Lyell, C. 1857. Manual of Elementary Geology, 6th edn. John Murray,
Koken, 1907. Indisches Perm und die Permische Eiszeit. (Indian Permian London.
and the Permian ice age.) Neues Jahrbuch für Mineralogie, Geologie, Lyell, C. 1863. The Geological Evidence of the Antiquity of Man.
und Paläontologie, Festband 1907, 446–545, 19 plates (including Reprinted unabridged in 2004 by Dover Publications, Mineola, NY.
map). Lyell, C. 1865. Elements of Geology, 6th enlarged edn. John Murray,
Köppen, W. & Wegener, A. 1924. Die Klimate der geologischen London.
Vorzeit. (Climates of the geological past.) Gebrüder Borntraeger, Macdonald, F. A., Jones, D. S. & Schrag, D. P. 2009a. Stratigraphic
Berlin. and tectonic implications of a newly discovered glacial diamictite-
Kuenen, P. H. & Migliorini, C. I. 1950. Turbidity currents as a cause of cap carbonate couplet in southwestern Mongolia. Geology, 37,
graded bedding. Journal of Geology, 58, 91 –127. 123– 126.
Kulling, O. 1934. The Hecla Hoek Formation round Hinlopenstredet. Macdonald, F. A., McClelland, W. C., Schrag, D. P. & Macondald,
Geografiska Annaler, 14, 161–253. W. P. 2009b. Neoproterozoic glaciation on a carbonate platform
Lachenbruch, A. 1962. Mechanics of thermal contraction cracks and margin in Arctic Alaska and the origin of the North Slope subterrane.
ice-wedge polygons in permafrost. Special Paper, Geological Geological Society of America Bulletin, 121, 448–473.
Society of America, 70, 1 –69. Macdonald, F. A., Strauss, J. V., Rose, C. V., Dudás, F. Ö. & Schrag,
Laskar, J. & Robutel, P. 1993. The chaotic obliquity of the planets. D. P. 2011. Stratigraphy of the Port Nolloth Group of Namibia and
Nature, 361, 608–612. South Africa and implications for the age of Neoproterozoic iron for-
Laskar, J., Joutel, F. & Robutel, P. 1993. Stabilization of the Earth’s mations. American Journal of Science, 310, 862– 888.
obliquity by the Moon. Nature, 361, 615– 617. Maciel, P. 1959. Tilito Cambriano(?) no Estado de Mato Grosso. Socie-
Leather, J., Allen, P. A., Brasier, M. D. & Cozzi, A. 2002. Neoproter- dad Brasilieras Geologia Boletino, 8, 3– 49.
ozoic snowball Earth under scrutiny: evidence from the Fiq glaciation Maloof, A. C., Kellogg, J. B. & Anders, A. M. 2002. Neoproterozoic
of Oman. Geology, 30, 891–894. sand wedges: crack formation in frozen soils under diurnal forcing
Lee, Y. Y. 1936. The Sinian glaciation in the lower Yangtze valley. Bul- during a snowball Earth. Earth and Planetary Science Letters, 204,
letin of the Geological Society of China, 15, 131–134. 1 –15.
Lee, J. S. & Chao, Y. T. 1924. Geology of the gorge district of the Maloof, A. C., Halverson, G. P., Kirschvink, J. L., Schrag, D. P.,
Yangtze (from Ichang to Tzekuei) with special reference to the Weiss, B. P. & Hoffman, P. F. 2006. Combined paleomagnetic,
development of the gorges. Bulletin of the Geological Society of isotopic and stratigraphic evidence for true polar wander from the
China, 3, 351– 391. Neoproterozoic Akademikerbreen Group, Svalbard. Geological
Lee, J. S. & Lee, Y. Y. 1940. Sinian glaciation of China. 17th Inter- Society of America Bulletin, 118, 1099– 1124.
national Geological Congress, Moscow, 1937, Report, 6, 33 – 41. Marshall, H. G., Walker, J. C. G. & Kuhn, W. R. 1988. Long-term
Lemon, N. M. & Gostin, V. A. 1990. Glacigenic sediments of the late climate change and the geochemical cycle of carbon. Journal of
Proterozoic Elatina Formation and equivalents, Adelaide Geosyn- Geophysical Research, 93, 791– 801.
cline, South Australia. In: Jago, J. B. & Moore, P. S. (eds) The Martin, H. 1965a. The Precambrian Geology of South West Africa and
Evolution of a Late Precambrian–Early Paleozoic Rift Complex: Namaqualand. Precambrian Research Unit Bulletin, 1, University
the Adelaide Geosyncline. Geological Society of South Australia of Cape Town, South Africa.
Special Publication, Adelaide, 16, 149–163. Martin, H. 1965b. Beobachtungen zum Problem der jung-
Lenk, C., Strother, P. K., Kaye, C. A. & Barghoorn, E. S. 1982. Pre- präkambrischen Glazialen Ablagerungen in Südwestafrika
cambrian age of the Boston Basin: new evidence from microfossils. (Observations concerning the problem of the late Precambrian
Science, 216, 619–620. glacial deposits in South West Africa). Geologische Rundschau, 54,
Lenton, T. M. & Watson, A. J. 2004. Biotic enhancement of weathering, 115– 127.
atmospheric oxygen and carbon dioxide in the Neoproterozoic. Geo- Martin, H. 1981. The late Palaeozoic Gondwana glaciation. Geologische
physical Research Letters, 31, L05202, doi:10.1029/2003GL018802. Rundschau, 70, 480 – 496.
Levrard, B. & Laskar, J. 2003. Climate friction and the Earth’s obli- Martin, H., Porada, H. & Walliser, O. H. 1985. Mixtite deposits of the
quity. Geophysical Journal International, 154, 970–990. Damara sequence, Namibia, problems of interpretation. Palaeogeo-
Li, Z. X., Evans, D. A. D. & Zhang, S. 2004. A 908 spin on Rodinia: poss- graphy, Palaeoclimatology, Palaeoecology, 51, 159–196.
ible causal links between the Neoproterozoic supercontinent, super- Maugh, T. H. II. 1989. Super ice age gave life on Earth growing pains.
plume, true polar wander and low-latitude glaciation. Earth and Los Angeles Times, September 7, 1, 3, 28.
Planetary Science Letters, 220, 409–421. Mawson, D. 1949a. The Late Precambrian ice age and glacial record of
Li, Z. X., Bogdanova, S. V. et al. 2008. Assembly, configuration, and the Bibliando dome. Journal and Proceedings of the Royal Society
break-up history of Rodinia: a synthesis. Precambrian Research, of New South Wales, 82, 150– 174.
160, 179– 210. Mawson, D. 1949b. The Elatina glaciation: a third recurrence of glacia-
Lu, S. N., Ma, G. G., Gao, Z. J. & Lin, W. X. 1985. Primary research on tion evidenced in the Adelaide system. Transactions of the Royal
the glaciogenic rocks of the Late Precambrian in China. In: Precam- Society of South Australia, 73, 117–121.
brian Geology Committee (eds) Precambrian Geology, No. 1, The McCabe, C. & Elmore, R. D. 1989. The occurrence and origin of Late
Collected Works on the Late Precambrian Glaciogenic Rocks of Paleozoic remagnetization in the sedimentary rocks of North
China. Geology Publication House, Beijing, 1– 86. America. Reviews of Geophysics, 27, 471–494.
Lyell, C. 1840. On the geological evidence of the former existence of McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial
glaciers in Forfarshire. Edinburgh New Philosophical Journal, trinity: Neoproterozoic Earth history within the British –Irish Caledo-
30, 199–202. Reprinted in Philosophical Magazine (2nd series), nides. Geology, 34, 909–912, doi: 10.1130/G22694A.1
18, 579– 591, 1841. McMechan, M. E. 2000. Vreeland Diamictites – Neoproterozoic glacio-
Lyell, C. 1841. Elements of Geology, 2nd edn. Hilliard, Gray, and Co., genic slope deposits, Rocky Mountains, northeast British Columbia.
Boston. Bulletin of Canadian Petroleum Geology, 48, 246– 261.
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 35

Melosh, H. J. 1989. Impact Cratering: A Geologic Process. Oxford Uni- discussion of dynamical constraints. Earth and Planetary Sciences,
versity Press, New York. 174, 155–171.
Milankovic, M. 1941. Canon of Insolation and the Ice-Age Problem. Park, J. K. 1994. Palaeomagnetic constraints on the position of Laurentia
(English translation published by Zavod za Udzbenike i Nastavna from middle Neoproterozoic to Early Cambrian times. Precambrian
Sredstva, Belgrade, 634, 1998.) Research, 69, 95– 112.
Miller, N. R., Alene, M. et al. 2003. Significance of the Tambien Pavlov, A. A., Kasting, J. F., Brown, L. L., Rages, K. A. & Freedman,
Group (Tigrai, N. Ethiopia) for Snowball Earth events in the R. 2000. Greenhouse warming by CH4 in the atmosphere of early
Arabian-Nubian Shield. Precambrian Research, 121, 263–283. Earth. Journal of Geophysical Research, 105, 11 981– 11 990.
Moczydlowska, M. 2008. The Ediacaran microbiota and the survival of Pavlov, A. A., Hurtgen, M. T., Kasting, J. F. & Arthur, M. A. 2003.
Snowball Earth conditions. Precambrian Research, 167, 1 –15. Methane-rich Proterozoic atmosphere? Geology, 31, 87– 90.
Molnia, B. F. 1983. Glacial– Marine Sedimentation. Plenum Press, Pavlov, A. A., Toon, O. B., Pavlov, A. K., Bally, J. & Pollard, D.
New York. 2005. Passing through a giant molecular cloud: ‘snowball’ glaciations
Moraes Rego, L. F. de. 1930. Glaciação eopaleozóica no centro do produced by interstellar dust. Geophysical Research Letters, 32,
Brazil (Glaciation in the earliest Palaeozoic of central Brazil). L03705, doi: 10.1029/2004GL021890.
Anais Academia Brasileira de Ciências, 2, 109–112. Peltier, W. R., Tarasov, L., Vettoretti, G. & Solheim, L. P. 2004.
Mu, Y. 1981. Luoquan Tillite of the Sinian System in China. In: Climate dynamics in deep time: modeling the ‘snowball bifurcation’
Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene and assessing the plausibility of its occurrence. In: Jenkins, G. S.,
Glacial Record. Cambridge University Press, Cambridge, 402– 413. McMenamin, M. A. S., McKey, C. P. & Sohl, L. (eds) The
Nalivkin, D. V. 1956. Ucheniye o fatsiyak. Geograficheskiye usloviya Extreme Proterozoic: Geology, Geochemistry, and Climate. Geophy-
obrazovaniya osadkov, 2 (Science of facies. Geographical environ- sical Monograph, American Geophysical Union, Washington, DC,
ment of sedimentation). Izvestiya Akademia Nauk SSSR, Moscow. 146, 107–124.
Natland, M. L. & Kuenen, P. H. 1951. Sedimentary history of the Plummer, P. S. 1978. Note on the palaeoenvironmental significance of the
Ventura Basin, California and the action of turbidity currents. In: Tur- Nuccaleena Formation (upper Precambrian), central Flinders Ranges,
bidity Currents and the Transportation of Coarse Sediments into South Australia. Journal of the Geological Society of Australia, 25,
Deep Water. Special Publication 2, Society of Economic Paleontolo- 395– 402.
gists and Mineralogists, Tulsa, OK, 76 –107. Pollard, D. & Kasting, J. F. 2005. Snowball Earth: a thin-ice solution
Nédélec, A., Affaton, P., France-Lanord, C., Charrière, A. & with flowing glaciers. Journal of Geophysical Research, 110,
Alvaro, J. 2007. Sedimentology and chemostratigraphy of C07010, doi: 10.1029/2004JC002525.
the Bwipe Neoproterozoic cap dolostones (Ghana, Volta Basin): Powell, R. D. 1984. Glacimarine processes and inductive lithofacies
a record of microbial activity in a peritidal environment. Comptes modelling of ice shelf and tidewater glacier sediments based on
Rendus Geoscience, 339, 223– 239. Quaternary examples. Marine Geology, 57, 1– 52.
Néron de Surgy, O. & Laskar, J 1997. On the long term evolution of the Powell, R. D. 1990. Glacimarine processes at grounding-line fans and
spin of the Earth. Astronomy and Astrophysics, 318, 975–989. their growth to ice-contact deltas. In: Dowdeswell, J. A. &
Newman, R. P. 1995. American instransigence: the rejection of continen- Scourse, J. D. (eds) Glacimarine Environments: Processes and Sedi-
tal drift in the great debates of the 1920’s. Earth Sciences History, 14, ments. Geological Society, London, Special Publication, 53, 53 –73.
62 – 83. Powell, R. D. & Domack, E. W. 1995. Modern glaciomarine environ-
Nick, F. M., Vieli, A., Howat, I. M. & Joughin, I. 2009. Large-scale ments. In: Menzies, J. (ed.) Glacial Environments, Vol. 1. Modern
changes in Greenland outlet glacier dynamics triggered at the termi- Glacial Environments: Processes, Dynamics and Sediments.
nus. Nature Geoscience, 2, 110– 114. Butterworth-Heinemann, Oxford, 445– 486.
Nikolaev, J. 1930. The glacial deposits (tillites) of Lower Cambrian age Preiss, W. V. 1987. The Adelaide Geosyncline: Late Proterozoic stratigra-
in the Yenissei Range. Bulletin of the Geological and Prospecting phy, sedimentation, palaeontology and tectonics. Geological Survey
Service of the U.S.S.R., 49(7), 1 – 15. of South Australia Bulletin, 53, 438.
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V. & Fair- Rampino, M. R. 1994. Tillites, diamictites, and the ballistic ejecta of large
child, T. R. 2003. Soft-sediment deformation at the base of the Neo- impacts. Journal of Geology, 102, 439– 456.
proterozoic Puga cap carbonate (southwestern Amazon craton, Ramsay, A. C. 1855. Angular, subangular, polished and striated fragments
Brazil): confirmation of rapid icehouse to greenhouse transition in and boulders in the Permian breccia of Shropshire, Worcestershire,
snowball Earth. Geology, 31, 613– 616. etc.; and on the probable existence of glaciers and icebergs in the
Norin, E. 1937. Geology of the western Quruq Tagh, eastern Tien Shan. Permian epoch. Quarterly Journal of the Geological Society of
Reports of the Sino-Swedish Expedition III. Geology. Bokförlags London, 11, 185– 205.
Aktiebolaget Thule, Stockholm. Ramsay, A. C. 1860. The Old Glaciers of Switzerland and North Wales.
North, G. R. 1990. Multiple solutions in energy balance climate models. Longman, Green, Longman & Roberts, London.
Palaeogeography, Palaeoecology, Palaeoclimatology (Global and Rankama, K. 1973. The Late Precambrian glaciation, with particular
Planetary Change Section), 82, 225– 235. reference to the Southern Hemisphere. Journal and Proceedings,
Oberbeck, V. R., Marshall, J. R. & Aggarwal, H. 1993. Impacts, til- Royal Society of New South Wales, 106, 89 –97.
lites, and the breakup of Gondwanaland. Journal of Geology, 101, Raub, T. D. & Evans, D. A. D. 2006. Magnetic reversals in basal
1 – 19. Ediacaran cap carbonates: a critical review. Eos, Transactions of
Opdyke, B. N. & Walker, J. C. G. 1992. Return of the coral reef hypoth- the American Geophysical Union, Joint Assembly Supplement, 87,
esis: basin to shelf partitioning of CaCO3 and its effect on atmos- Abstract GP41– 02.
pheric CO2. Geology, 20, 733–736. Reading, H. G. & Walker, R. G. 1966. Sedimentation of Eocambrian
Opdyke, B. N. & Wilkinson, B. H. 1990. Palaeolatitude distribution of tillites and associated sediments in Finnmark, northern Norway.
Phanerozoic marine ooids and cements. Palaeogeography, Palaeocli- Palaeogeography, Palaeoclimatology, Palaeoecology, 2, 177– 212.
matology, Palaeoecology, 78, 135–148. Reusch, H. 1891. Skuringmærker og morængrus eftervist i Finnmarken
Opdyke, N. D. 1962. Palaeoclimatology and continental drift. In: fra en periode meget ældre end ‘istiden’ (Glacial striae and boulder-
Runcorn, S. K. (ed.) Continental Drift. Academic Press, clay in Norwegian Lapponie from a period much older than the last
New York, 41– 65. ice age). Norges Geologiske Undersøkelse, 1, 78 –85 (English
Oreskes, N. 1999. The Rejection of Continental Drift. Oxford University summary, 97– 100).
Press, New York. Rice, A. H. N. & Hofmann, C.-C. 2000. Evidence for a glacial origin of
Osmaston, H. 2004. Quaternary glaciations in the East African moun- Neoproterozoic III striations at Oaibaccannjar’ga, Finnmark, north-
tains. In: Ehlers, J. & Gibbard, P. L. (eds) Quaternary Glaciations ern Norway. Geological Magazine, 137, 355– 366.
– Extent and Chronology, Part III. Elsevier, Amsterdam, 139–150. Ridgwell, A. J., Kennedy, M. J. & Caldeira, K. 2003. Carbonate depo-
Pais, M. A., Le Mouël, J. L., Lambeck, K. & Poirier, J. P. 1999. Late sition, climate stability, and Neoproterozoic ice ages. Science, 302,
Precambrian paradoxical glaciation and obliquity of the Earth – a 859– 862.
36 P. F. HOFFMAN

Robert, M. 1940. La glaciation du Kundelungu au Katanga (Congo other similar changes. Memoirs of the Boston Society of Natural
Belge). 17th International Geological Congress, Moscow, 1937, History, 2, 321–323, 335– 340.
Report 6, 99 – 113. Sheldon, R. P. 1984. Ice-ring origin of the Earth’s atmosphere and
Roberts, J. D. 1971. Late Precambrian glaciation: an anti-greenhouse hydrosphere and late Proterozoic –Cambrian phosphogenesis.
effect? Nature, 234, 216. Special Publication, Geological Survey of India, 17, 17 – 21.
Roberts, J. D. 1976. Late Precambrian dolomites, Vendian glaciation, and Shields, G. A., Deynoux, M., Strauss, H., Paquet, H. & Nahon, D.
synchroneity of Vendian glaciations. Journal of Geology, 84, 47 – 63. 2007. Barite-bearing cap dolostone of the Taoudéni Basin, northwest
Rodgers, J. 1957. The distribution of marine carbonate sediments: a Africa: sedimentary and isotopic evidence for methane seepage after
review. In: Le Blanc, R. J. & Breeding, J. G. (eds) Regional a Neoproterozoic glaciation. Precambrian Research, 154, 209– 235.
Aspects of Carbonate Deposition. Society of Economic Paleontolio- Siever, R. 1968. Sedimentological consequences of a steady-state
gists and Mineralogists (SEPM) Special Publication No. 5, 2 – 14, ocean-atmosphere. Sedimentology, 11, 5 –29.
Tulsa, Oklahoma. Smith, J. 1836. On indications of changes in the relative levels of sea and
Rogers, A. W. 1915. The geology of part of Namaqualand. Transactions land in the West of Scotland. Proceedings of the Geological Society,
of the Geological Society of South Africa, 18, 72 –101, 14 plates. London, 2, 427–429.
Rosendahl, H. 1931. Bidrag til Varangernesets geologi. Norsk Geologisk Smith, J. 1839. On the climate of the newer pliocene tertiary period.
Tidsskrift, 25, 327– 349. Proceedings of the Geological Society of London, 3, 118– 119.
Rosendahl, H. 1945. Prekambrium-Eokambrium i Finnmark. Norsk Sohl, L. E., Christie-Blick, N. & Kent, D. V. 1999. Paleomagnetic
Geologisk Tidsskrift, 25, 327–349. polarity reversals in Marinoan (ca 600 Ma) glacial deposits of
Rudwick, M. J. S. 1964. The infra-Cambrian glaciation and the origin of Australia: implications for the duration of low-latitude glaciation in
the Cambrian fauna. In: Nairn, A. E. M. (ed.) Problems in Palaeo- Neoproterozoic time. Geological Society of America Bulletin, 111,
climatology. Interscience, London, 150–155, 184– 185. 1120– 1139.
Runcorn, S. K. 1961. Climatic change through geological time in light of Spalletti, L. & Del Valle, A. 1984. Las diamictitas del sector oriental
the palaeomagnetic evidence for polar wandering and continental de Tandilia: characteres sedimentológicos y origen. Revista de la
drift. Quarterly Journal of the Royal Meteorological Society, 87, Asociación Geológica Argentina, 39, 188– 206.
282– 313. Spencer, A. M. 1971. Late Pre-Cambrian glaciation in Scotland. Geologi-
Sagan, C. & Mullen, G. 1972. Earth and Mars: evolution of atmospheres cal Society of London Memoir, 6, 99.
and surface temperatures. Science 177, 52 –26. Spjeldnaes, N. 1964. The Eocambrian glaciation in Norway. Geologische
Sayles, R. W. 1914. The Squantum Tillite. Bulletin of the Museum of Rundschau, 54, 24 –45.
Comparative Zoology at Harvard College, 56, 141– 175, 12 plates. Sprigg, R. C. 1986. The Adelaide Geosyncline: a century of controversy.
Sayles, R. W. & LaForge, L. 1910. The glacial origin of the Roxbury Earth Sciences History, 5, 66– 83.
Conglomerate. Science, 32, 723–724. Strahan, A. 1897. On glacial phenomena of Palaeozoic age in the Varan-
Schermerhorn, L. J. G. 1974. Late Precambrian mixtites: glacial and/or ger Fiord. Quarterly Journal of the Geological Society, London, 53,
non-glacial? American Journal of Science, 274, 673– 824. 137– 146, Discussion, 153–156.
Schermerhorn, L. J. G. 1977. Late Precambrian dolomites, Vendian Sugden, D. E. & John, B. S. 1976. Glaciers and Landscape: A Geomor-
glaciation, and synchroneity of Vendian glaciations: a discussion. phological Approach. Edward Arnold, London.
Journal of Geology, 85, 247– 250. Sutherland, P. C. 1870. Notes on an ancient boulder-clay of
Schermerhorn, L. J. G. & Stanton, W. I. 1963. Tilloids in the West Natal. Quarterly Journal of the Geological Society of London, 26,
Congo geosyncline. Quarterly Journal of the Geological Society of 514– 517.
London, 119, 201–241. Thompson, M. D. & Bowring, S. A. 2000. Age of the Squantum ‘tillite’,
Schimper, K. 1837. Über die Eiszeit. (On the Ice Age.) Actes Société Boston Basin, Massachusetts: U– Pb zircon constraints on terminal
Helvetique des Sciences Naturelles, Neuchâtel, 38– 51. Neoproterozoic glaciation. American Journal of Science, 300,
Schmidt, P. W. & Williams, G. E. 1995. The Neoproterozoic climatic 630– 655.
paradox: equatorial palaeolatitude for Marinoan glaciation near sea Thomson, J. 1871. On the stratified rocks of Islay. Report of the 41st
level in South Australia. Earth and Planetary Science Letters, 134, Meeting of the British Association for the Advancement of Science,
107– 124. Edinburgh. John Murray, London, 110– 111.
Schmidt, P. W., Williams, G. E. & Embleton, B. J. J. 1991. Low palaeo- Thomson, J. 1877. On the geology of the island of Islay. Transactions of
latitude of Late Proterozoic glaciation: early timing of remnance the Geological Society of Glasgow, 5, 200–222.
in Haematite of the Elatina Formation, South Australia. Earth and Tyndall, J. 1861. On the absorbtion and radiation of heat by gases and
Planetary Science Letters, 105, 355–367. vapours, and on the physical connexion of radiation, absorbtion,
Schiøtz, O. E. 1898. Om Dr Reusch’s präglaciale skuringdmerker. Nyt and conduction – the Bakerian Lecture. Philosophical Magazine
Magazin Naturvissenschaften, 36, 1 –10. (series 4), 22, 169– 194, 273–285.
Schrag, D. P., Berner, R. A., Hoffman, P. F. & Halverson, G. P. 2002. Tyndall, J. 1863. On the radiation through the Earth’s atmosphere. Phi-
On the initiation of a snowball Earth. Geophysics, Geochemistry, losophical Magazine (series 4), 25, 200–206.
Geosystems, 3, doi: 10.1029/2001GC000219. Venetz, I. 1830. Sur l’extension qu’il présume que les glaciers avaient
Schwartzman, D. W. & Volk, T. 1991. Biotic enhancement of weather- autrefois, et sur leur retraits dans leurs limites actuelles. Actes
ing and surface temperatures since the origin of life. Palaeogeogra- Société Helvetique des Sciences Naturelles, Lausanne, 31.
phy, Palaeoecology, Palaeoclimatology, 90, 357– 374. Voigt, A., Abbot, D. S., Pierrehumbert, R. T. & Marotzke, J. 2011.
Schwarzbach, M. 1964a. Criteria for the recognition of ancient glacia- Initiation of a Marinoan Snowball Earth in a state-of-the-art atmos-
tions. In: Nairn, A. E. M. (ed.) Problems in Palaeoclimatology. phere-ocean general circulation model. Climate of the Past, 7, 1– 15.
Interscience, London, 81 –85. Walde, D. H. G., Gierth, E. & Leonardos, O. H. 1981. Stratigraphy and
Schwarzbach, M. 1964b. Climates of the Past. D. Van Nostrand Co., mineralogy of the manganese ores of Urucum, Mato Grosso, Brazil.
London. (English translation of the 2nd revised edn of Das Clima Geologische Rundschau, 70, 1077– 1085.
der Vorzeit. Ferdinand Enke Verlag, Stuttgart, 1961.) Walker, G. 2003. Snowball Earth: The Story of the Great Global Cata-
Sellers, W. D. 1969. A global climatic model based on the energy strophe that Spawned Life As We Know It. Crown Publishers,
balance of the Earth-atmosphere system. Journal of Applied Meteor- New York, 269.
ology, 8, 392–400. Walker, J. C. G., Hays, P. B. & Kasting, J. F. 1981. A negative feedback
Sellers, W. D. 1990. The genesis of energy balance modeling and mechanism for the long-term stabilization of Earth’s surface tempera-
the cool Sun paradox. Palaeogeography, Palaeoecology, Palaeocli- ture. Journal of Geophysical Research, 86(C10), 9776– 9782.
matology (Global and Planetary Change Section), 82, 217–224. Wang, Y., Lu, S. & Gao, Z. 1981. Sinian tillites of China. In: Hambrey,
Shaler, N. S. 1847. Preliminary report on the secent changes of level M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial
on the coast of Maine: with reference to their origin and relation to Record. Cambridge University Press, Cambridge, 386–401.
HISTORY OF NEOPROTEROZOIC GLACIAL GEOLOGY 37

Wegener, A. 1912. Die Entstehung der Kontinente. Petermanns Geogra- Wilson, C. B. & Harland, W. B. 1964. The Polarisbreen Series and other
phische Mitteilungen, 58, 185–195, 253– 256, 305–309. (Reprinted evidences of late pre-Cambrian ice ages in Spitsbergen. Geological
in English translation: The origins of the continents. Journal of Magazine, 101, 198– 219.
Geodynamics, 32, 29– 63, 2001.) Wilson, J. T. 1966. Did the Atlantic close and then reopen? Nature, 211,
Wegener, A. 1929. Die Enstehung der Kontinente und Ozeane, 4th edn. 676– 681.
Friedrich Vieweg & Sohn, Braunschweig. (Reprinted in English Witzke, B. J. 1990. Palaeoclimatic constraints for Palaeozoic palaeolati-
translation by John Biram: The Origin of Continents and Oceans. tudes of Laurentia and Euramerica. In: McKerrow, W. S. &
Dover Publications, New York, 246, 1966.) Scotese, C. R. (eds) Palaeozoic Palaeogeography and Biogeogra-
Wegmann, C. E., Dangeard, L. & Graindor, M. J. 1950. Sue quelques phy. Geological Society of London Memoirs, 12, 57 –73.
caractères remarquables de la formation pré-cambrienne connue Woodward, H. B. 1907. The History of the Geological Society of
sous le nom de Poudinage de Granville. Compte Rendus, 230, London. Geological Society, London, 336. Reprinted in 1978
979– 980. by Arno Press, New York.
Wetherald, R. T. & Manabe, S. 1975. The effects of changing the Solar Woodward, H. P. 1884. Report on the range to the east of Farina.
constant on the climate of a general circulation model. Journal of the Parliamentary Papers for 1884. Government of South Australia,
Atmospheric Sciences, 32, 2044–2059. 40, 1– 5.
Whitten, G. F. 1970. The investigation and exploitation of the Razorback Worsley, T. R. & Kidder, D. L. 1991. First-order coupling of paleogeo-
Ridge iron deposit. Geological Survey of South Australia Reports of graphy and CO2, with global surface temperature and its latitudinal
Investigations, 33, 165. contrast. Geology, 19, 1161– 1164.
Williams, D. M., Kasting, J. F. & Frakes, L. A. 1998. Low-latitude Wright, L., Williams, E. G. & Cloud, P. 1978. Algal and cryptalgal
glaciation and rapid changes in the Earth’s obliquity explained by structures and platform environments of the late pre-Phanerozoic
obliquity – oblateness feedback. Nature, 396, 453–455. Noonday Dolomite, eastern California. Geological Society of
Williams, G. E. 1972. Geological evidence relating to the origin America Bulletin, 89, 321–333.
and secular rotation of the Solar system. Modern Geology, 3, Wunsch, C. 2002. What is the thermohaline circulation? Science, 298,
165– 181. 1179–1180.
Williams, G. E. 1974. Discussion of Late Precambrian glacial climate and Xiao, S. 2004. Neoproterozoic glaciations and the fossil record. In:
the Earth’s obliquity. Journal of the Geological Society, London, 130, Jenkins, G. S., McMenamin, M. A. S., McKey, C. P. & Sohl, L.
599– 601. (eds) The Extreme Proterozoic: Geology, Geochemistry, and
Williams, G. E. 1975. Late Precambrian glacial climate and the Earth’s Climate. Geophysical Monograph, American Geophysical Union,
obliquity. Geological Magazine, 112, 441–544. Washington, DC, 146, 199–214.
Williams, G. E. 1979. Sedimentology, stable-isotope geochemistry and Xiao, S., Bao, H. et al. 2004. The Neoproterozoic Quruqtagh Group in
palaeoenvironment of dolostones capping late Precambrian glacial eastern Chinese Tianshan: evidence for a post-Marinoan glaciation.
sequences in Australia. Journal of the Geological Society of Austra- Precambrian Research, 130, 1– 26.
lia, 26, 377–386. Yeo, G. M. 1981. The Late Proterozoic Rapitan glaciation in the northern
Williams, G. E. 1981. Sunspot periods in the late Precambrian glacial Cordillera. In: Campbel, F. H. A. (ed.) Proterozoic Basins of Canada.
climate and solar – planetary relations. Nature, 291, 624–628. Geological Survey of Canada Paper, 81– 10, 25 –46.
Williams, G. E. 1989. Late Precambrian tidal rhythmites in South Austra- Young, G. M. 1976. Iron-formation and glaciogenic rocks of the Rapitan
lia and the history of the Earth’s rotation. Journal of the Geological Group, Northwest Territories, Canada. Precambrian Research, 3,
Society, London, 146, 97 –111. 137– 158.
Williams, G. E. 1993. History of the Earth’s obliquity. Earth-Science Young, G. M. 1995a. Are Neoproterozoic glacial deposits preserved on
Reviews, 34, 1 –45. the margins of Laurentia related to the fragmentation of two super-
Williams, G. E. 2000. Geological constraints on the Precambrian history continents? Geology, 23, 153–156.
of Earth’s rotation and the Moon’s orbit. Reviews of Geophysics, 38, Young, G. M. 1995b. Reply: Are Neoproterozoic glacial deposits pre-
37 – 59. served on the margins of Laurentia related to the fragmentation of
Williams, G. E. & Sonett, C. P. 1985. Solar signature in sedimentary two supercontinents? Geology, 23, 1054– 1055.
cycles from the late Precambrian Elatina Formation, Australia. Zachos, J., Pagani, M., Sloan, L. & Thomas, E. 2001. Trends, rhythms,
Nature, 318, 523– 527. and aberrations in global climate 65 Ma to present. Science, 292,
Williams, G. E. & Tonkin, D. G. 1985. Periglacial structures and palaeo- 686– 693.
climatic significance of a late Precambrian block field in the Cattle Zhang, S., Jiang, G. & Han, Y. 2008. The age of the Nantuo
Grid copper mine, Mount Gunson, South Australia. Australian Formation and Nantuo glaciation in South China. Terra Nova, 20,
Journal of Earth Sciences, 32, 287–300. 289– 294.
Williams, G. E., Gostin, V. A., McKirdy, D. M. & Preiss, W. V. 2008. Ziegler, P. A. 1960. Frühpaläozoische Tillite in östlichen Yukon-
The Elatina glaciation, late Cryogenian (Marinoan Epoch), South Territorium (Kanada). Eclogae Geologicae Helveticae, 52, 735– 741.
Australia: sedimentary facies and palaeoenvironments. Precambrian Ziegler, A. M., Hulver, M. L., Lottes, A. L. & Schmachtenberg, W.
Research, 163, 307– 331. F. 1984. Uniformitarianism and paleoclimates: inferences from the
Willis, B. 1904. Geological research in eastern Asia. Carnegie Institution distribution of carbonate rocks. In: Brenchley, P. J. (ed.) Fossils
of Washington, Yearbook, 3, Washington, DC, 275–291. and Climate. John Wiley & Sons, New York, 3 –25.
Chapter 3

Recognition of glacial influence in Neoproterozoic sedimentary successions

EMMANUELLE ARNAUD1* & JAMES L. ETIENNE2


1
School of Environmental Sciences, University of Guelph, Guelph, Ontario, N1G 2W1, Canada
2
Neftex Petroleum Consultants Ltd, 97 Milton Park, Abingdon, Oxfordshire OX14 4RY, UK
*Corresponding author (e-mail: earnaud@uoguelph.ca)

Abstract: This chapter provides an overview and key references of glacial processes and resulting sedimentary products in subglacial,
terrestrial proglacial and glaciomarine or glaciolacustrine settings. These settings are characterized by a wide variety of processes ranging
from subglacial lodgement and deformation, ice-push and sediment remobilization, which in turn result in a wide range of products such
as diamictite, conglomerate, sandstone, siltstone and mudstone. The sedimentary record of proglacial settings exhibits the most lateral
and vertical variability due to the dynamic nature of ice margins and the most direct record of climatic fluctuations. Many Neoproterozoic
successions, however, preserve glaciomarine deposits that can provide a more continuous and high-resolution (though indirect) record of
change. This chapter will enable the reader to identify features that may be used to infer a glacial influence on the formation of ancient
deposits. The chapter also outlines some of the important issues that require consideration when evaluating palaeoclimatic models for
Neoproterozoic sedimentary successions. These include the equivocal significance of most commonly used proxies such as occurrence
of diamictite, outsized clasts in laminated sediments, clast characteristics, lithostratigraphic trends and sequence boundaries. Careful
analysis of multiple lines of sedimentary evidence, together with other proxies of climatic changes, can yield meaningful reconstructions
and provide a basis for testing palaeoclimate models for this time period. A summary table outlining the characteristics of diamictite with
different depositional origins is also included in order to assist with the interpretation of the Neoproterozoic sedimentary record.

The ability to recognize glacial deposits, or a glacial influence on Flint 1975; Hambrey & Harland 1978; Boulton & Deynoux 1981;
sedimentation, has long been a subject of debate (Schermerhorn Dreimanis & Schluchter 1985; Chumakov 1992; Eyles 1993;
1974; Deynoux & Trompette 1976; Jensen & Wulff-Pedersen Crowell 1999). This improved understanding has occurred as a
1996), and yet is critical for the accurate reconstruction of result of the emergence of modern sedimentological techniques,
palaeoenvironmental conditions during past ice ages. This which focus on process and products, and the application of
chapter summarizes the historical development of the terminology facies and basin analysis (Reading 1978), which have enhanced
and approach used in identifying deposits of glacial origin, the our ability to reconstruct palaeoenvironmental conditions from
characteristics of glacial environments in terms of processes, the ancient sedimentary successions (see Eyles & Januszczak 2004
resulting sedimentary facies and stratal geometries, as well as for a historical review).
some key points to consider when reconstructing palaeoenviron- For the purpose of this book, we encouraged authors describing
mental conditions in ancient successions. The chapter is far from deposits from around the world to use the non-genetic term diamic-
exhaustive, and many others have previously covered these tite to describe poorly sorted materials that contain a mixture of
topics in more detail elsewhere. Rather, it is meant to provide an gravel-, sand- and mud-sized particles (Fig. 3.1). Terms such as
overview with key references for workers who focus on other ‘tillite’ were to be reserved for the interpretation section where
aspects of Neoproterozoic geology and may not necessarily be poorly sorted materials could be shown to have been deposited
familiar with glacial deposits. As the focus of this book is on directly by ice, without significant subsequent disaggregation
Neoproterozoic ice ages, characteristics related to micro- or and flow. In describing reworked glacial deposits, the interpretive
macrofauna and biogenic indicators associated with glacial set- term ‘glaciogenic’ was recommended as a modifier (e.g. glacio-
tings are not discussed. Similarly, geomorphic features are only genic debris flow). Authors were also encouraged to use ‘outsized
briefly mentioned, as most Neoproterozoic successions do not clast’ or ‘lonestone’ rather than the term ‘dropstone’ in their
have the prerequisite outcrop exposure or preservation potential descriptions, leaving the latter word to the discussion where a
of landforms found in more recent glacial settings. dropstone origin could be interpreted based on deformation of
underlying laminae and onlapping overlying laminae. The use of
terms ‘varvite’ or ‘varve’ to describe laminated fine-grained sedi-
Historical development ment in glaciogenic successions was discouraged, considering the
lack of radiometric control required to confirm annual depositional
The terminology used to refer to glacial deposits has evolved over cycles. These are best described as ‘rhythmically laminated’ when
the years, with an increasing appreciation of the processes occur- a cyclical nature can be identified or simply ‘laminated’ when the
ring in glaciated basins and a desire to refine palaeoenvironmental stacking pattern appears random. Readers are referred to Hambrey
interpretations to include more information about the severity of & Glasser (2003) for an up-to-date discussion of terminology
climatic conditions, the nature of the ice mass, the record of associ- related to glacial sediments and Evans et al. (2006) for a more
ated sea-level changes and depositional setting (Flint et al. 1960; specific discussion of subglacial till classification.
Carey & Ahmad 1961; Dreimanis 1978; Hambrey & Harland
1978; Boulton & Deynoux 1981; Boulton 1990; Eyles 1993;
Hambrey & Glasser 2003). The criteria used to identify glacial Characteristics of glacial environments
deposits in ancient sedimentary successions has also evolved
with this refined understanding of processes and a recognition of Deposition occurs within a number of different settings in gla-
the non-uniqueness of many of the sedimentary characteristics ciated basins, including subglacial, proglacial, glaciolacustrine
that were once thought diagnostic of glacial deposits (Crowell and glaciomarine environments. Each of these depositional set-
1964; Harland 1964; Harland et al. 1966; Schermerhorn 1974; tings has characteristic processes, deposits and landforms

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 39– 50. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.3
40 E. ARNAUD & J. L. ETIENNE

Fig. 3.1. Classification of poorly sorted sediments according to Hambrey & Glasser (2003). Reproduced with permission from Springer.

(Menzies 1995, 1996; Benn & Evans 1998, 2010; Evans 2005). Subglacial settings
Sedimentary analysis of Neoproterozoic successions can thus
provide much-needed palaeoenvironmental information when Conditions in subglacial settings are primarily controlled by the
this variability is taken into consideration. In contrast to more relationship between pressure and temperature, which ultimately
recent glacial deposits, landforms tend not to be preserved or are controls the transition between solid and liquid phases and the
poorly exposed in ancient successions, although there are excep- alternation between melting and regelation in the basal traction
tions. An understanding of the processes, together with the result- zone. In other instances, ice will melt and refreeze in response to
ing deposits and their three-dimensional distribution in these broader temperature fluctuations. These subglacial conditions in
settings, is thus critical for palaeoenvironmental reconstruction. turn control the ability of the ice to erode, transport and deposit
Glaciomarine successions tend to be the most common in the debris. The thermal characteristics of ice masses thus have a
Neoproterozoic geological record, as marine basins provide direct impact upon the accumulation of subglacial debris, with
accommodation space away from subsequent ice advances and polythermal (variable distribution of warm and cold-based ice)
erosion, allowing preservation of thick packages of glaciogenic and temperate glaciers (basal ice close to pressure melting point)
sediments (Carey & Ahmad 1961; Nystuen 1985; Eyles 1993). accumulating the most significant volumes of debris. Although
Although these settings tend to provide the longest and highest res- cold-based glaciers have been shown to entrain, transport and
olution records of Neoproterozoic climate change, the depocentres deposit sediment, they generally accumulate significantly less
tend to be dislocated from the glaciated shelves, which acted as debris. In some cases, subglacial lakes may form, as shown by
the sediment supply pathway, complicating efforts to reconstruct recent inventories of subglacial lakes in Antarctica (Siegert et al.
the distribution of palaeo-ice masses. An appreciation of other 2005). Satellite data have demonstrated significant ice elevation
glacial settings is important as some successions will record con- changes associated with these subglacial lakes, and these have
ditions at the basin margin, close to land-based glaciers (Rieu been attributed to major discharge events and reorganization of
et al. 2006; Arnaud 2008). Basin margin successions such as subglacial meltwater (Wingham et al. 2006; Fricker et al. 2007).
these are particularly useful in terms of palaeoclimate reconstruc- Subglacial processes that are most relevant for palaeoenviron-
tion as they can provide information regarding the nature of the ice mental reconstructions include lodgement or plastering of debris
mass (stagnant, active or fast-moving) and allow better control on onto the substrate, deformation of soft-sediment substrates, hydro-
the relative timing of the changes associated with that ice margin. fracturing of substrate (Boulton & Caban 1995), plucking of
They may also provide a better sense of the temporal relationship angular material from hardrock substrates as a result of freeze
between glacial conditions and palaeoenvironmental perturbations and thaw, abrasion of both the debris within the ice and the under-
inferred from geochemical data that have recently been so central lying substrate, and meltwater erosion. Meltwater processes are
to Neoproterozoic research. particularly significant in temperate glacial settings. The meltwater
A discussion of the various settings in glaciated basins follows can be found in subglacial distributary channels, which comprise
with particular emphasis on the nature of these environments, part of the drainage system of the glacier, or as a thin film at the
the processes that operate within them and the resulting sedimen- base of the glacier, which helps facilitate movement.
tary products. References provided herein are by no means exhaus- These processes are responsible for subglacial deposits such as
tive, but were selected as they focus on sedimentary characteristics lodgement and deformation tills (Boyce & Eyles 2000; Evans et al.
and emphasize their connection to processes, thus facilitating the 2006), commonly characterized by diamicton (the unlithified
reconstruction of palaeoenvironmental conditions from the sedi- equivalent of diamictite), in addition to gravel and sand deposited
mentary record. Readers are referred to the ‘glaciers online’ site in subglacial meltwater tunnels (Fig. 3.2; Table 3.1). Whereas
of Alean & Hambrey (2008) as well as Eyles (1993), Crowell deposition related to meltout or undermelting of ice in a subglacial
(1999), Hambrey & Glasser (2003) and Etienne et al. (2007) for setting is theoretically possible, studies have shown that it is very
helpful photographs and additional general discussions of criteria difficult to distinguish these deposits from other subglacial depos-
used in the recognition of ancient glacial deposits. its considering the high likelihood of subsequent overprinting by
RECOGNITION OF GLACIAL DEPOSITS 41

Fig. 3.2. Photographs of typical glacial sedimentary indicators in Neoproterozoic successions. (a) Mud-rich diamictite with extra-basinal clasts; (b) outsized clast
showing clear piercing and onlap relationship typical of dropstones; (c) striated pavement, Varangerfjorden, Norway; (d) striated clast; (e) bullet-shaped clast; (f) faceted
clast. It should be noted that none of these features are exclusive to glacial conditions, so careful analysis of other features is necessary to confirm a glacial origin. See text
for discussion.

deformation and lodgement processes (Evans et al. 2006). Subgla- & Boulton 1991; Hart & Roberts 1994; McCarroll & Rijsdijk
cial processes also result in striated pavements, and clasts that 2003; Roberts & Hart 2005; Arnaud 2008). Clastic dykes may
show evidence of transport and erosion in the basal traction zone also develop in areas where a frozen substrate and the overlying
such as striations, faceting, chattermarks and a bullet shape (also ice mass result in high pore water pressures (Boulton & Caban
referred to as stoss and lee clasts, flat-iron clasts or bullet-shaped 1995; Le Heron & Etienne 2005). Preferred orientation of the
boulders; Fig. 3.2, Table 3.1). Care must be taken to distinguish a-axes of clasts (fabric) has also been used to discriminate
between striations arising from subglacial erosion and those that between various depositional origins of diamictite (Domack &
are related to slickenside development associated with faulting Lawson 1985; Hart 1994; Benn & Evans 1996); however, others
or pressure solution. Diamictite units may exhibit a gradational have argued that clast fabric is not always indicative of a particular
change from deformed and sheared at the base through chaotically depositional setting and that other indicators of glacial conditions
stratified to massive at the top, indicative of the increase in stress are needed to confirm interpretations (Bennett et al. 1999;
with proximity to the ice – sediment interface (Boulton & Hind- Hambrey & McKelvey 2000). Preferred clast orientations can be
marsh 1987; Hart & Boulton 1991; Boulton 1996; Benn & very useful in determining direction of transport, which may in
Evans 1996). Other sedimentary characteristics that can be indic- turn facilitate palaeoenvironmental discrimination (Table 3.1).
ative of subglacial settings are boulder pavements (boulders Clast fabric analyses are somewhat more difficult for Neoprotero-
preferentially aligned one behind the other) and soft-sediment zoic successions that have undergone multiple phases of post-
deformation features such as boudinage, tectonic lamination, depositional deformation. Finally, it is difficult to say what
rooted recumbent folding and inclusions of underlying material sedimentary expression subglacial lakes will have in the rock
that exhibit shear stress in a preferred direction (Table 3.1; Hart record and how their existence may be inferred without the
42
Table 3.1. Characteristics of diamictite units of different origin

Types of diamictite Textural features Compositional Sedimentary Deformation Basal contact Geometry Associated deposits Useful references
features structures

Lodgment & Poorly sorted Mixed provenance Massive or Glaciotectonic Sharp Thin (,m to m) Well bedded conglomerate Benn & Evans (1996)
deformation tillite* Boulder pavements (intra- and chaotically deformation such as Erosional or and tabular and sandstone, commonly Boulton (1972)
Evidence of glacial extrabasinal clasts) stratified shear structures (rooted deformed Laterally cross-stratified and Boyce & Eyles
transport on particle Compositional Gradational upward folds, boudinage, Basal shear plane extensive exhibiting cut and fill (2000)
surface† immaturity change from highly attenuated bedding, Grooves or (hundreds of geometry (glaciofluvial McCarroll & Rijsdijk
Textural immaturity Inclusions/ intraclasts deformed to tectonic laminae), deformation in km2) outwash) (2003)
Wide range of stone or rafts of other massive nappes, imbricate sub-till Discontinuous if Thick deposits of windblown Hart & Roberts
roundness sediment types that Preferred orientation structures, thrust or sediments close to ice silt (loess), sandstone (1994)

E. ARNAUD & J. L. ETIENNE


Micromorphological exhibit ductile of shear structures reverse faulting, shear margin wedges, convoluted Menzies (2000)
characteristics deformation planes and decollement bedding from Evans et al. (2006)
associated with high surfaces. cryoturbation and in situ
stress conditions Clastic dykes breccia‡ (periglacial)
Strong a-axis fabrics Laminated or massive
(lodgement) mudstone with/without
outsized clasts (lacustrine)
Meltout tillite § Poorly sorted Mixed provenance Crudely stratified Extensional faulting from Conformable Rapid lateral Well bedded conglomerate Boulton (1972)
Localized sorting from (intra- and (discontinuous melting of buried ice Sharp facies changes and sandstone, commonly Lawson (1982)
meltwater reworking extrabasinal clasts) layers of sorted Convoluted bedding and Undisturbed cross-stratified and McDonald & Shilts
Evidence of glacial Compositional sediments folding substrate exhibiting cut and fill (1975)
transport on particle immaturity or draped layers geometry (glaciofluvial Evans et al. (2006)
surface such as Soft sediment or soft over large clasts) outwash)
striations bedrock inclusions Laminated or massive
Textural immaturity or intraclasts that mudstone with or without
Wide range of stone exhibit no outsized clasts from
roundness deformation ponded water or lake
deposits
‘Rainout’ deposits Poorly sorted Mixed provenance Massive to crudely Grounding ice berg Gradational Variable Lodgement till from ice Eyles et al. (1985)
Evidence of glacial (intra- and stratified structures and ice-keel thickness advance onto shelf Thomas & Connell
transport on particle extrabasinal clasts) Discontinuous furrows Tabular to Sediment gravity flow (1985)
surface† Compositional stringers of sand blanket-like deposits
Textural immaturity immaturity and gravel (boulder geometry, Laminated mudstone with/
Wide range of stone Clast clusters from pavements) from draping without outsized clasts
roundness iceberg dumping winnowing underlying Subaquatic fan deposits in
Diamict pellets from Variable clast content topography proximal settings
frozen debris on from changing Laterally Pelagic clastic or carbonate
icebergs sedimentation rate continuous sediments
Dropstones or supply of (distal) to rapid
ice-rafted debris lateral facies
changes
(proximal)
Terrestrial debris flowsPoorly sorted Local (non-glacial) or Massive to Flow nose Erosional with Extent (m– km) Lodgement or meltout tills Lawson (1982)
Evidence of glacial mixed (glaciogenic chaotically Basal shearing structures rip-up clasts Benvenuti & Martini
transport on particle source) provenance stratified Convoluted bedding or (2002)
surface* if glaciogenic Compositional folding Colella & Prior
source maturity depending (1990)
Textural maturity on source
depending on source Inclusions or rafts of
Wide range of stone other sediment
roundness depending types that exhibit
on source ductile deformation
Subaqueous debris flows Poorly sorted Local (non-glacial) or Massive to Flow noses Erosional with Moderate to thick Lodgement tills Nardin et al. (1979),
Evidence of glacial mixed (glaciogenic chaotically Basal shear structures rip-up clasts Tabular to ‘Rainout’ deposits Visser (1983),
transport on particle source) provenance stratified Folding and convoluted Conformable lenticular Laminated mudstone with/ Prior et al. (1984),
surface† Compositional Graded bedding bedding when Channelized without outsized clasts Postma et al.
Textural immaturity maturity depending Coarse-tail grading Load casts, water escape hydroplaning Extent (m to 10s Subaquatic fan deposits (1988), Hart &
Wide range of stone on source structures has occurred of km) Pelagic clastic or carbonate Roberts (1994),
roundness Inclusions or rafts of sediments Mohrig et al.

RECOGNITION OF GLACIAL DEPOSITS


other sediment Other sediment gravity flow (1999), Eyles &
types that exhibit deposits (e.g. turbidites) Eyles (2000),
ductile Hambrey &
deformation, some McKelvey (2000),
can be very large Laberg & Vorren
(10s of m in size) (2000), Mulder &
Alexander (2001),
McCarroll &
Rijsdijk (2003)

Note: No one of these features are sufficient for identification of a glacial or glacially influenced origin and any one of the features may be absent.
*We combine these two types of till, as lodgement and deformation processes will often both affect individual deposits (see discussion in Evans et al. 2006); Evans et al. (2006) argued that meltout till generated in subglacial
settings could similarly not be easily distinguished).

Evidence of glacial transport: striations, chattermarks, polish, grooves, facets, bullet shape (also referred to as ‘stoss and lee’ or ‘flat iron’ clasts).

These sedimentary characteristics are not unique to periglacial settings; their origin may be difficult to confirm even with careful stratigraphic and sedimentological consideration.
§
Although meltout of debris is theoretically possible in subglacial settings, their identification in the rock record seems unlikely (Evans et al. 2006). Here, we only consider meltout tillite in the context of proglacial terrestrial settings
(see text for discussion).

43
44 E. ARNAUD & J. L. ETIENNE

benefit of the geomorphic context that is available in modern envi- The wide range in depositional conditions associated with
ronments. Studies to date have been based on remote sensing data, terrestrial proglacial environments often leads to an inherently
so very little is known about the types of deposits within these complex distribution of sediment types and three-dimensional geo-
lakes and their longer-term preservation potential. Their unequivo- metries that vary from irregular to channelized and lenticular with
cal identification in the rock record would appear difficult, as common intraformational unconformities and sudden bed termin-
deposits would look very similar to glaciomarine successions ation. Basal contacts are variable from sharp to irregular and
recording advances and recessions of ice. In Neoproterozoic suc- deformed or erosional to conformable and draping over irregular
cessions, biota are not available to distinguish between the lacus- topography. Associated sediments are likely to be subglacial and
trine and marine settings. glaciofluvial outwash deposits.
In terms of three-dimensional geometry, lodgement and defor-
mation tills tend to be tabular, relatively thin (on the scale of less
than a metre to several metres) and can cover extensive areas (typi- Glaciolacustrine and glaciomarine settings
cally hundreds of km2). Basal contacts can be erosional or show
evidence of deformation and incorporation of materials from the Conditions in glacially influenced water bodies are primarily con-
underlying substrate. Conglomerate and sandstone from subglacial trolled by their proximity to the ice margin, whether or not the
tunnels will have sharp outer contacts and a pipe-like geometry margin terminates in the water body, the nature of the ice margin
in cross-section. The nature of the associated sediments will (i.e. tidewater glacier or floating ice shelf), the thermal regime of
depend on whether or not this ice mass overrode land or a the ice mass (e.g. temperate v. polar) and the characteristics of
marine shelf. Terrestrial subglacial deposits may be associated the water body (lake v. ocean; deep sea, shelf, fjord or shoreline;
with coarse-grained glaciofluvial outwash sediments and other Smith & Ashley 1985; Eyles et al. 1985; Powell & Elverhoi
terrestrial proglacial sediments such as glaciolacustrine facies 1989; Dowdeswell & Scourse 1990; Anderson & Ashley 1991;
(Boulton 1972), whereas glaciomarine subglacial deposits will Dowdeswell et al. 1998; Evans & Pudsey 2002). Lakes that are
be associated with subaqueous proglacial sediments, tidal rhyth- dammed by ice or sediment (such as moraines) are also regulated
mites, calcareous, biogenic or clastic fine-grained sediments by dam stability (Tweed & Russell 1999). Glacial processes dom-
deposited by suspension settling and/or a range of fine- to coarse- inate in settings that are close to the ice margin (,kms), whereas
grained sediment gravity flow deposits (e.g. turbidites and hyper- normal lacustrine/marine processes dominate in settings that are
concentrated flows; Domack 1988; Cowan & Powell 1990; distal to the ice margin (.tens of km). Where ice terminates in a
Anderson & Ashley 1991; Fairchild 1993; Dowdeswell et al. water body (ice-contact lakes or tidewater glacier marine settings),
1996, 1998; Anderson 1999; Taylor et al. 2002). depositional settings are affected by melt-out of subglacial and
englacial debris, slumping off the ice margin, iceberg calving,
efflux from subglacial and englacial meltwater conduits and sedi-
Terrestrial proglacial settings ment input from surface meltwater drainage. These processes
supply sediment to grounding line fan complexes and more
Conditions in proglacial settings are primarily controlled by the distal areas. In contrast, water bodies fed by glacial meltwater
proximity to the ice margin and whether or not the ice margin is are primarily influenced by diurnal and seasonal fluctuations in
stagnant, actively advancing or surging. Proglacial processes meltwater discharge and seasonal ice cover. Meltwater processes
include in situ meltout of debris, remobilization of debris down- are very important in settings influenced by temperate glaciers
slope, reworking by meltwater and melting of buried stagnant and limited to non-existent in polar settings. The configuration
ice (Boulton 1972; Lawson 1982). These processes typically of the basin will influence the types of non-glacial processes
result in highly irregular topography, ponding of meltwater and that also contribute to the distribution and nature of deposits
development of proglacial lakes. The entire proglacial succession found in these settings. For example, lakes may experience
may be deformed and cannibalized by subsequent ice advances thermal stratification and seasonal mixing, whereas deposition in
that bulldoze, thrust and re-work the deposits. Such events are oceans will be affected by tides and salinity. Sediment gravity
known from modern environments (Huddart & Hambrey 1996) flows resulting from high sedimentation rates are common in
and the ancient glacial geological record (Le Heron et al. 2005). most water bodies but may be more prevalent in steep-sided
Stagnating ice margins will often have multiple recessional mor- lakes or fjords. However, it is important to note that subaqueous
aines and abundant supraglacial debris with buried ice. In contrast, glaciogenic debris flows may have run-out distances of hundreds
active ice margins will be characterized by multiple push moraines of kilometres, even on shallow slopes (Dowdeswell et al. 1996,
and ridges associated with the fluctuating ice margin (Boulton 1998; Taylor et al. 2002). Oscillatory waves will affect shallow
et al. 1999; Bennett 2001; Bennett et al. 2004). water settings, while storm waves will affect both shallow and
Highly variable depositional conditions in terrestrial proglacial shelf settings.
environments (ponding, meltwater flow, bulldozing and remobi- Processes include rapid deposition of coarse-grained sediments
lization of sediments) tend to result in complex lithofacies assem- at the mouth of subglacial tunnels, density underflows and
blages with a high degree of lateral variability. These assemblages suspension-settling of fine-grained sediments related to buoyant
may include diamictite (derived from meltout, subaerial debris meltwater plumes and sub-ice shelf melting, winnowing or
flow and bulldozing and deformation of glaciogenic debris; reworking by sub-ice shelf or open water currents as well as bull-
Table 3.1), conglomerate, sandstone, siltstone and mudstone dozing, lodgement and deformation of sediments related to ice
(meltwater processes) and laminated mudstone and siltstone advances (Lønne 1995; Ó Cofaigh & Dowdeswell 2001; Evans
(ponding) in these settings. Clasts may show evidence of recent et al. 2005; McKay et al. 2008; Naish et al. 2009). Remobilization
transport in the basal traction zone such as faceting, striations, of debris by gravity flows is common in these settings and results
chattermarks and bullet-shape, but surface features rarely survive from high sedimentation rates, oversteepening along the edges of
fluvial transport. Deformation structures ranging in scale from grounding-line fans, and normal sediment-transfer processes at
centimetres to hundreds of metres can be quite prevalent in these the shelf-slope break or along steep-sided fjords. Gravity flows
proglacial settings (Benn & Evans 1996; Phillips et al. 2002). may also be initiated directly by subaqueous iceberg calving, or
Extensional faulting and ductile folding of sediments associated as a result of fluctuating water depths associated with calving or
with in situ melting of ice can be found in all facies types. In abrupt lake drainage events. Icebergs can have an impact on
addition, reverse or thrust faults, shear planes and low-angle décol- both proximal and distal settings through iceberg roll-over and
lement surfaces, nappes and complex folding and shear structures dumping, meltout of entrained debris and scouring and bulldozing
can be found resulting from bulldozing by ice. when grounded in shallow water (Thomas & Connell 1985; Gilbert
RECOGNITION OF GLACIAL DEPOSITS 45

1990; Woodworth-Lynas & Guigne 1990). Deposition of cold- be found in ice-proximal grounding line fan deposits as a result of
water carbonates is also common in glacially influenced polar slumping and bulldozing by ice, in remobilized glaciogenic debris-
settings. Ikaite (hydrated calcium carbonate) is one such mineral flow deposits along the continental slope and in other deposits
that forms in near-freezing water temperatures. Unstable at where iceberg grounding has occurred (Table 3.1; Woodworth-
higher temperatures, ikaite has poor preservation potential and Lynas & Guigne 1990; Hart & Roberts 1994).
typically undergoes replacement by calcite. Where the original
crystal structure is preserved, these pseudomorphs are called
glendonites and are an important palaeoenvironmental indicator The record of climatic changes over time in glacial
of cold-water conditions (James et al. 2005; Domack et al. 2007; environments
Halverson 2011).
Processes such as those identified above often lead to a predomi- Ancient glaciogenic sedimentary successions contain a record of
nance of diamictite and mudstone in glacially influenced lacus- fluctuating climatic conditions over time (Boulton 1990). Ice
trine/marine settings, although conglomerate and sandstone can ages typically consist of interstadials (relatively warmer) and sta-
also be found associated with grounding-line fans and subglacial dials (relatively colder) periods, which are recorded in the partial
tunnel outflow as well as stratified facies related to sub-ice shelf advance and recession of ice and changes in the lateral and vertical
sedimentation or reworking by sub-ice shelf or open water currents distribution of sediments associated with these fluctuating environ-
(Eyles et al. 1985; Lønne 1995; Nemec et al. 1999; Barrett 2007; ments. Such fluctuations have been documented in several Neopro-
McKay et al. 2008). In open water conditions where icebergs are terozoic successions (Allen et al. 2004; Arnaud & Eyles 2006), as
calving or in sub ice-shelf settings, suspension settling of fine- well as in more recent glacial successions (Domack 1983; Barrett
grained sediment and rainout of ice-rafted debris can lead to the 2007; Naish et al. 2009). Changes from glacial to non-glacial
formation of diamictite (Table 3.1) or mudstone with outsized periods will similarly be recorded at the margins of marine
clasts. Where outsized clasts rafted from icebergs deform under- basins. Here, glacioisostacy (crustal movements associated with
lying laminated sediments (truncating laminae, loading sediment ice loading and unloading) and glacioeustasy (changes in global
to develop compaction-related folds and/or water escape struc- ocean volume associated with changes in terrestrial ice volume)
tures) they may be referred to as dropstones (Fig. 3.2; Thomas & will combine with local tectonics and sediment supply to influence
Connell 1985). Material of pebble size or greater is generally con- relative sea level and ultimately the nature of stratal geometries
sidered as a good proxy for ice-rafting, although some care is (Ravnås & Steel 1998; Powell & Cooper 2002; Allen 2007).
needed in interpretation (see discussion below). In some cases, Glacio-eustatic variability can lead to significant correlative
frozen masses of ice-rafted debris are released from icebergs into global cycles of deposition.
water bodies and result in till pellets or inclusions of diamict Glaciations affect various parts of the world and can be
within either fine-grained sediments or a diamict with slightly expressed quite differently depending on the thermal character-
different characteristics (Gilbert 1990). These can be distinguished istics of the ice mass mentioned above. These characteristics can
from intraclasts resulting from erosion at the base of sediment vary spatially and temporally. Today, temperate glaciers (i.e.
gravity flows by closely examining the type of materials and warm ice close to melting point) are found in Alaska, polythermal
their relationship to underlying materials, the nature of the outer glaciers (mixture of warm and cold ice) in Greenland and polar gla-
boundary of these inclusions (sharp and irregular for erosional ciers (cold ice well below melting point) in Antarctica. However,
intraclast v. diffuse for till pellets) as well as the basal contact of these thermal characteristics can change over time, as shown by the
the unit in which they are found. Ice advance onto the continental sedimentary record of the Ross Ice Shelf, which shows temperate
shelf can also result in the deposition of lodgement or deformation conditions in the past compared to the polar conditions of today
tillite (Table 3.1), as demonstrated by overconsolidated diamicton (Naish et al. 2009).
displaying shear structures, clastic dykes and associated subglacial
landforms such as grooves, lineations and drumlins in recent con-
tinental shelf sediments of Antarctica (Camerlenghi et al. 2001; Reconstructing Neoproterozoic palaeoenvironmental
Ó Cofaigh et al. 2002; Evans et al. 2005; Naish et al. 2009). Dia- conditions: key considerations
mictite may of course also form as a result of sediment gravity
flows (Table 3.1; Wright & Anderson 1980; Eyles & Eyles 2000; The sedimentary record is at times difficult to interpret as many, if
Mulder & Alexander 2001). Rhythmically laminated mudstone not most, sedimentary characteristics are not unique to specific
may result from tidal pumping in ice-proximal marine settings environments. The following discussion highlights key points to
(Smith et al. 1990) or from seasonal discharge fluctuations in consider when reconstructing palaeoenvironmental conditions
non-ice contact lakes (Leonard 1986). The latter are referred to from Neoproterozoic successions, focusing on indicators typically
as varves in more recent deposits, but the use of this term is dis- used to infer glacial conditions.
couraged in Neoproterozoic successions as a seasonal control
cannot be clearly demonstrated due to limited geochronological
constraints. Although Evans & Pudsey (2002) suggested that ice- The palaeoclimatic significance of diamictite
shelf deposits could be distinguished from tidewater glacier depos-
its by the absence of meltwater-derived facies, similar stratified Diamictite can result from both glacial and non-glacial processes
sediments have been reported from ice-shelf settings in Antarctica (Table 3.1). In glaciated settings, diamictite results from lodge-
and attributed to grounding line zones (Evans et al. 2005; McKay ment and deformation of subglacial sediments, in situ meltout of
et al. 2008; Naish et al. 2009), making it unlikely that the two will ice and release of poorly sorted basal, englacial or supraglacial
be easy to distinguish in ancient glaciogenic successions in the debris, slumping and reworking of glaciogenic debris in proglacial,
absence of biogenic facies, which are sensitive to ice cover, or con- glaciolacustrine or glaciomarine settings and from ‘rainout’ of
textual knowledge of the ice margin available in more recent gla- fine-grained sediment and ice-rafted debris. In non-glaciated set-
ciomarine environments. tings, diamictite can form as a result of terrestrial or subaqueous
Laminated or massive fine-grained sediments with rare clasts debris flows (hyperconcentrated flows) or as a result of impact
dominate in ice distal settings where typical marine processes ejecta (Fisher 1971; Nardin et al. 1979; Rampino 1994; Martins-
dominate and suspension settling from icebergs provide a marginal Neto 1996; Sohn 2000; Weaver et al. 2000; Mulder & Alexander
contribution to sedimentation. Boulder pavements may develop on 2001; Benvenuti & Martini 2002). Studies of modern glaciated set-
marine shelves as a result of winnowing of diamict and overriding tings and of sediment gravity flows and their deposits have greatly
by ice during sea-level lowstands (Eyles 1994). Deformation may improved our ability to link depositional processes with their
46 E. ARNAUD & J. L. ETIENNE

sedimentary products. However, careful consideration of sedi- our depositional models by distinguishing between different
mentary characteristics, geometry, as well as lateral and vertical types of depositional process, glacial setting (e.g. terrestrial
distribution of diamictite and associated deposits, is needed to dis- proglacial v. glaciomarine/glaciolacustrine) and thermal regime,
tinguish between these various depositional origins (Arnaud & which can in turn contribute much needed constraints on evaluating
Eyles 2006; Table 3.1). palaeoclimate models that have been proposed for this time period
For example, a diamictite formed as a result of ‘rainout’ may (Condon et al. 2002; Leather et al. 2002; Allen et al. 2004; Arnaud
have a gradational basal contact, whereas those related to sedi- & Eyles 2006; Rieu et al. 2006; Allen & Etienne 2008).
ment gravity flows or subglacial lodgement tend to have sharp con-
formable or erosional basal contacts (Table 3.1). In a marine
The palaeoclimatic significance of outsized clasts in
setting, diamictite showing draped upper contact may be attributed
to a sediment gravity flow rather than subglacial processes. bedded sediments
Diamictite formed as a result of subglacial lodgement may
exhibit upwardly increasing deformation (Benn & Evans 1996), Outsized clasts in laminated sediments have often been interpreted
in contrast to deposits of sediment gravity flows where the as ice-rafted debris or ‘dropstones’ and a clear indication of glacial
maximum deformation will occur at the base of the deposit conditions. Piercing or disturbance of underlying laminae and
(Nardin et al. 1979; Mulder & Alexander 2001). Associated draping of sediment on top of the outsized clasts are probably
facies play a particularly key role in distinguishing between terres- the best criteria to imply that the clast was dropped from floating
trial or marine settings in these ancient successions where faunal ice (Thomas & Connell 1985). Without this deformation and
indicators are lacking. onlap relationship, the outsized clast may in fact be an outrunner
Careful analysis is particularly important in considering in a succession of turbidites and other sediment gravity flows
ancient glaciogenic successions, as many of them were deposited (Postma et al. 1988) or a clast surrounded by lamination resulting
in tectonically active settings where diamictite may record tectoni- from post-depositional compaction or deformation.
cally induced sediment instability (Krzyszkowski 1993; Tanaka & Once a ‘dropstone’ origin has been established, care must still
Maejima 1995) rather than renewed glacial conditions. Sediment– be taken in terms of inferring glacial conditions from the presence
landform associations may not be available in these basins to better of these outsized clasts. Ice-rafted debris may be derived from
constrain a glacial influence. seasonal ice unrelated to any ice margin, proximal icebergs or
Even with the most careful examination, the depositional origin of far-travelled icebergs (Gilbert 1990). Distinction of seasonal ice,
diamictite will sometimes be difficult to determine unequivocally as proximal icebergs and far-travelled icebergs is very unlikely in
indicators are sometimes absent or obscured by post- depositional ancient successions. It is therefore important to recognize that
tectonism or weathering. This is particularly significant considering the occurrence of dropstones can have varying palaeoclimatic
that Neoproterozoic deposits will have been subjected to a long significance (cold climate if resulting from seasonal ice, locally gla-
history of post-depositional modification with sometimes very ciated or glaciated some distance away if resulting from icebergs).
limited and poor preservation in structurally complex settings. This is an important consideration when using the sedimentary
Weathering is problematic when looking for surface markings on record to test competing palaeoclimate models for this time period.
carbonate clasts (e.g. striations, faceting or bullet-shape) within Neo-
proterozoic diamictite, a common problem considering that many The palaeoclimatic significance of clast characteristics
Neoproterozoic glaciogenic successions overlie carbonate strata
and have incorporated these softer lithologies. Clast characteristics suggesting transport in the basal traction zone of
In other situations, interpretation will be confounded by the glaciers are also often used to infer glacial conditions. These include
fact that the poorly sorted debris has been affected by a number the presence of striations, chattermarks, faceting, bullet-shaped
of different processes, and therefore retains a composite sedi- boulders or flat iron clasts and extra-basinal provenance. It is impor-
mentary signature of these processes. In their comprehensive tant to note however, that although these confirm a glaciogenic influ-
review of subglacial processes and products, Evans et al. (2006) ence on transport history, contextual analysis is required to ensure
suggested that it will be difficult to distinguish between different that the clasts have not been re-worked into younger deposits
types of tillites in ancient glacial deposits, as diamictite will lacking a direct glacial influence. Although it is true that such
ultimately record a composite of these subglacial processes and surface markings are unlikely to survive with significant transport
the dynamic nature of subglacial systems as they change over in meltwater glaciofluvial systems, many glaciomarine deposits
time and space. In another example, glaciomarine deposits could (which tend to be the most common in ancient successions) may
be subsequently over-ridden by ice, preserving some character- not have undergone significant fluvial reworking. This is particularly
istics of both glaciomarine and subglacial processes and thus relevant when considering the amount of time elapsed between the
making it very difficult to identify whether these represent open deposition of a diamictite thought to record glacial conditions and
glaciomarine or subglacial conditions. Another equally difficult its associated ‘cap’ carbonate thought to record warmer post-glacial
situation is where large-scale controls on the sedimentary suc- conditions in Neoproterozoic successions.
cession can leave similar signatures in the rock record such as
diamictites formed in offshore Alaska during the late Miocene
(Eyles et al. 1991), which were affected by both climatic and The palaeoclimatic significance of stratigraphic trends and
tectonic controls and Neoproterozoic volcaniclastic deposits in sequence boundaries
the Arabian –Nubian shield (Miller et al. 2009), which may be
related to glaciogenic and/or volcaniclastic processes. Debates Many studies of glacially influenced successions have equated
over the palaeoclimatic significance of the Neoproterozoic alternating beds of diamictite and other sediments as the product
diamictite in Namibia, for example, revolve in part around the of multiple advances and recessional phases of ice margins and
relative importance of tectonic and climatic controls on sedimen- multiple diamictite-bearing intervals of successions separated by
tation and the different processes ascribed to these deposits. thick sections of non-diamictite bearing sediments as representing
It is important to acknowledge these uncertainties where rel- discrete ice ages. Although this may turn out to be true (Naish et al.
evant, because very different palaeoenvironmental conditions are 2009), such interpretations must be made with caution considering
implied depending on the depositional process responsible for the various depositional origins that can be ascribed to diamictite,
the formation of diamictite. In places where it is possible, careful the possibility that the succession at times is too distal to the ice
consideration of the sedimentary characteristics and spatial geome- margin to preserve any clear indicator of glacial conditions (e.g.
try of the diamictite and its associated deposits can greatly refine relatively high-order transgressive or highstand systems tracts
RECOGNITION OF GLACIAL DEPOSITS 47

characterized by ice-distal mudstones in glaciomarine settings evaluation of sedimentary characteristics, geometry of deposits,
with limited or no icebergs), and the lack of geochronological deformation styles and associated facies are needed to identify a
control to constrain the amount of time represented by those glaciogenic source and reconstruct the extent and nature of
thick non-diamictite bearing successions. glacial conditions at the time of deposition.
The lateral and vertical heterogeneity of lithofacies within
glacially influenced successions can also be highly complex. In E. Arnaud’s research in glacial sedimentology is funded by the Natural Sciences
both terrestrial and shallow marine (neritic) environments, and Engineering Research Council. J. L. Etienne’s work follows research experi-
regional sequence boundaries may be developed as a result of sub- ence on modern, Pleistocene and Neoproterozoic glacial deposits funded under
glacial erosion either directly by ice via processes of regelation, NERC Studentship NER/S/A/2000/03690/ and Schweizerischer National-
abrasion from subglacially entrained debris, or by meltwater drai- Fonds Grant 103502. E. Arnaud thanks Carolyn Eyles (McMaster University).
nage under normal or elevated hydrostatic pressures. Multiple J.L. Etienne thanks Mike Hambrey, Neil Glasser (University of Aberystwyth),
glacial advance-recessional cycles across marine shelf areas may Dan Le Heron (Royal Holloway) and Philip Allen (Imperial College London).
Constructive reviews from David Evans and Eugene Domack greatly improved
result in stacked sequence boundaries, with lowstand packages
the manuscript and are much appreciated. This work represents a contribution
of sediment in slope to bathyal environments providing the only
of the IUGS- and UNESCO-funded IGCP (International Geoscience Programme)
continuous record of glacial – interglacial cyclicity. A cryptic
Project #512.
shallow marine record provided by multistorey stacking of sub-
glacial drainage conduits such as tunnel valleys may also be
preserved, although these are not well documented from Neopro-
terozoic successions (see Ordovician examples in Le Heron
References
et al. 2009). More localized disconformities may develop in Alean, J. & Hambrey, M. J. 2008. ‘Earth’s glacial record’ in Glaciers
marine environments as a result of iceberg keel-scour and pro- Online. (Available at http://www.swisseduc.ch/glaciers/earth_icy_
cesses of submarine mass-wasting related to high glaciomarine planet/glaciers15-en.html.)
sedimentation rates or local tectonic activity. In terrestrial progla- Allen, P. A. 2007. The Huqf Supergroup of Oman: basin development
cial environments, the suite of erosion processes associated with and context for Neoproterozoic glaciation. Earth-Science Reviews,
braided fluvial systems, glaciotectonic deformation, and incision 84, 139– 185.
generated by glacial lake outburst floods or jökulhlaups resulting Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to snowball
from subglacial eruptions can all lead to local or regional sequence Earth. Nature Geoscience, 1, 817– 825.
boundary development. This picture can become even more com- Allen, P. A., Leather, J. & Brasier, M. D. 2004. The Neoproterozoic
plicated where thin cold-based ice occurs in accumulation areas Fiq glaciation and its aftermath, Huqf Supergroup of Oman. Basin
and pre-existing sediments such as periglacial deposits can be Research, 16, 507– 534.
preserved with little or no modification, effectively leading to a Anderson, J. B. 1999. Antarctic Marine Geology. Cambridge University
correlative surface lacking erosion. Press, Cambridge.
In addition, as ancient glaciogenic successions tend to be prefer- Anderson, J. B. & Ashley, G. M. (eds) 1991. Glacial marine sedimen-
entially preserved in basins that were tectonically active, the evol- tation; paleoclimatic significance. Geological Society of America
ution of the basin must be well constrained in order to assess the Special Paper, 261, 232.
impacts on accommodation space, sediment supply, gross-scale Arnaud, E. 2008. Deformation in the Neoproterozoic Smalfjord
stratigraphic architecture and characteristics of sequence bound- Formation, northern Norway: an indicator of glacial depositional
conditions? Sedimentology, 55, 335–356.
aries (Allen 2007).
Arnaud, E. & Eyles, C. H. 2006. Neoproterozoic environmental change
The points above highlight some of the considerations when
recorded in the Port Askaig Formation, Scotland: climatic vs tectonic
applying a lithostratigraphic or sequence stratigraphic analysis controls. Sedimentary Geology, 183, 99 –124.
within glacially influenced successions. Barrett, P. J. 2007. Cenozoic climate and sea level history from glaci-
marine strata off the Victoria Land coast, Cape Roberts Project,
Antarctica. In: Hambrey, M. J., Christoffersen, P., Glasser, N.
Conclusions F. & Hubbard, B. (eds) Glacial Sedimentary Processes and Pro-
ducts. International Association of Sedimentologists Special Paper,
Criteria for the recognition of glacial deposits in Neoproterozoic 39, 259– 287.
sedimentary successions have developed over the last 50 years Benn, D. I. & Evans, D. J. A. 1996. The interpretation and classification
with an increasing appreciation for the different kinds of settings of subglacially-deformed materials. Quaternary Science Reviews, 15,
and processes that occur within glaciated basins. The analysis of 23– 52.
modern glacial settings and processes has allowed us to better Benn, D. I. & Evans, D. J. A. 1998. Glaciers and Glaciation. Arnold
recognize the complexity of glaciated basins and the various Publishers, London.
challenges in confirming a glacial influence on deposition. This Benn, D. I. & Evans, D. J. A. 2010. Glaciers and Glaciation, 2nd edn.
includes a better appreciation of the processes occurring in glacio- Hodder, London.
marine settings as well as the difficulty in distinguishing the nature Bennett, M. R. 2001. The morphology, structural evolution and signifi-
of floating ice producing dropstones (local iceberg, distant iceberg cance of push moraines. Earth Science Reviews, 53, 197–236.
or sea ice), the depositional origin of diamictite (glacial, glacially Bennett, M. R., Waller, R. I., Glasser, N. F., Hambrey, M. J. &
influenced or non-glacial) and the controls on large-scale Huddart, D. 1999. Clast fabrics: genetic fingerprint or wishful
thinking? Journal of Quaternary Science, 14, 125–135.
stratigraphic trends in facies successions (alternating lithofacies
Bennett, M. R., Huddart, D. et al. 2004. Sedimentary and tectonic
assemblages associated with autocyclicity or eustatic trends of
architecture of a large push moraine: a case study from Hagafjellsjo-
transgression and regression). kull-Eystri, Iceland. Sedimentary Geology, 172, 269– 292.
Poorly sorted deposits (diamictite), outsized clasts in laminated Benvenuti, M. & Martini, I. P. 2002. Analysis of terrestrial hypercon-
lithofacies (dropstones), high lateral facies variability in proximal centrated flows and their deposits. Special Publications International
settings and the presence of clasts that show evidence of subglacial Association of Sedimentologists, 32, 167– 193.
transport are common features in glacial settings and these remain Boulton, G. S. 1972. Modern arctic glaciers as depositional models for
common indicators used in the study of ancient glaciogenic succes- former ice sheets. Journal of the Geological Society, London, 128,
sions. However, recent studies have shown that such sedimentary 361– 393.
characteristics can have widely different climatic implications Boulton, G. S. 1990. Sedimentary and sea-level changes during
such that an understanding of the processes and products of glacial cycles and their control on glacimarine facies architecture.
glacial settings and sediment gravity flows as well as careful In: Dowdeswell, J. A. & Scourse, J. D. (eds) Glacimarine
48 E. ARNAUD & J. L. ETIENNE

Environments: Processes and Sediments. Geological Society, Dowdeswell, J. A., Kenyon, N. H., Elverhøi, A., Laberg, J. S.,
London, Special Publications, 53, 15– 53. Hollender, F.-J., Mienert, J. & Siegert, M. J. 1996. Large-scale
Boulton, G. S. 1996. Theory of glacial erosion, transport and deposition sedimentation on the glacier-influenced Polar North Atlantic margins:
as a consequence of subglacial sediment deformation. Journal of Long-range side-scan sonar evidence. Geophysical Research Letters,
Glaciology, 140, 43 –62. 23, 3535– 3538.
Boulton, G. S. & Caban, P. 1995. Groundwater flow beneath ice sheets: Dreimanis, A. 1978. The problems of waterlain tills. In: Schluchter, C.
Part II — its impact on glacier tectonic structure and moraine for- (ed.) Moraines and Varves. A. A. Balkema Publishers, Rotterdam,
mation. Quaternary Science Reviews, 14, 563– 587. 167– 177.
Boulton, G. S. & Deynoux, M. 1981. Sedimentation in glacial environ- Dreimanis, A. & Schluchter, C. 1985. Field criteria for the recognition
ments and the identification of tills and tillites in ancient sedimentary of till and tillite. Palaeogeography, Palaeoclimatology, Palaeoecol-
sequences. Precambrian Research, 15, 397– 422. ogy, 51, 7– 14.
Boulton, G. S. & Hindmarsh, R. C. A. 1987. Sediment deformation Etienne, J. L., Allen, P. A., Rieu, R. & Le Guerroué, E. 2007.
beneath glaciers: rheology and sedimentological consequences. Neoproterozoic glaciated basins: a critical review of the Snowball
Journal of Sedimentary Research, 92, 9059–9082. Earth hypothesis by comparison with Phanerozoic basins. In:
Boulton, G. S., Van Der Meer, J. J. M., Beets, D. J., Hart, J. K. & Hambrey, M. J., Christoffersen, P., Glasser, N. F. & Hubbard,
Ruegg, G. H. J. 1999. The sedimentary and structural evolution of B. (eds) Glacial Sedimentary Processes and Products. International
a recent push moraine complex: Holmstrombreen, Spitsbergen. Association of Sedimentologists Special Publication, 39, 343–399.
Quaternary Science Reviews, 18, 339–371. Evans, D. J. A. (ed.) 2005. Glacial Landsystems. Oxford University Press.
Boyce, J. I. & Eyles, N. 2000. Architectural element analysis applied to Evans, J. & Pudsey, C. J. 2002. Sedimentation associated with Antarctic
glacial deposits: internal geometry of a late Pleistocene till sheet, Peninsula ice shelves: implications for palaeoenvironmental recon-
Ontario, Canada. Geological Society of America Bulletin, 112, structions of glaciomarine sediments. Journal of the Geological
98 –118. Society, London, 159, 233–237.
Camerlenghi, A., Domack, E. et al. 2001. Glacial morphology and Evans, D. J. A., Phillips, E. R., Hiemstra, J. F. & Auton, C. A. 2006.
post glacial contourites in northern Prince Gustav Channel (NW Subglacial till: Formation, sedimentary characteristics and classifi-
Weddell Sea, Antartica). Marine Geophysical Researches, 22, cation. Earth Science Reviews, 78, 115– 176.
417– 443. Evans, J., Pudsey, C. J., O’Cofaigh, C., Morris, P. & Domack, E. 2005.
Carey, S. W. & Ahmad, N. 1961. Glacial marine sedimentation. In: Late Quaternary glacial history, flow dynamics and sedimenta-
Raasch, G. O. (ed.) Geology of the Arctic. Volume II. University tion along the eastern margin of the Antarctic Peninsula Ice Sheet.
of Toronto Press, Toronto, 865– 894. Quaternary Science Reviews, 24, 741– 774.
Chumakov, N. M. 1992. The Problems of Old Glaciations (Pre- Eyles, C. H. 1994. Intertidal boulder pavements in the northeastern Gulf
Pleistocene Glaciogeology in the USSR). Harwood Academic Pub- of Alaska and their geological significance. Sedimentary Geology,
lishers, New York. 88, 161– 173.
Colella, A. & Prior, D. B. (eds) 1990. Coarse-Grained Deltas. Inter- Eyles, N. 1993. Earth’s glacial record and its tectonic setting. Earth-
national Association of Sedimentologists Special Publication, 10. Science Reviews, 35, 1 –248.
Blackwell Scientific Publications, Oxford. Eyles, C. H. & Eyles, N. 2000. Subaqueous mass flow origin for Lower
Condon, D. J., Prave, A. R. & Benn, D. J. 2002. Neoproterozoic glacial- Permian diamictites and associated facies of the Grant Group, Barb-
rainout intervals: Observations and implications. Geology, 30, wire Terrace, Canning Basin, Western Australia. Sedimentology, 47,
35 –38. 343– 356.
Cowan, E. A. & Powell, R. D. 1990. Suspended sediment transport and Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for
deposition of cyclically interlaminated sediment in a temperate Neoproterozoic glaciations during the breakup of Rodina after 750
glacial fjord, Alaska, U.S.A. In: Dowdeswell, J. A. & Scourse, Ma. Earth-Science Reviews, 65, 1– 73.
J. D. (eds) Glacimarine Environments: Processes and Sediments. Eyles, C. H., Eyles, N. & Lagoe, M. B. 1991. The Yakataga Formation: a
Geological Society, London, Special Publications, 53, 75– 89. six million year record of temperate glacial marine sedimentation
Crowell, J. C. 1964. Climatic significance of sedimentary deposits con- in the Gulf of Alaska. In: Anderson, J. B. & Ashley, G. M. (eds)
taining dispersed megaclasts. In: Nairn, A. E. M. (ed.) Problems in Glacial Marine Sedimentation: Paleoclimatic Significance. Geologi-
Palaeoclimatology. Interscience Publishers, London, 86 – 99. cal Society of America Special Paper, 261, 159– 180.
Crowell, J. C. 1999. Pre-Mesozoic Ice Ages: their Bearing on Under- Eyles, C. H., Eyles, N. & Miall, A. D. 1985. Models of glaciomarine
standing the Climate System, 192. Geological Society of America, sedimentation and their application to the interpretation of ancient
Memoir, 106. glacial sequences. Palaeogeography, Palaeoclimatology, Palaeo-
Deynoux, M. & Trompette, R. 1976. Late Precambrian mixtites: glacial ecology, 51, 15 –84.
and/or non-glacial? Dealing especially with the mixtites of West Fairchild, I. 1993. Balmy shores and icy wastes: The paradox of carbon-
Africa. American Journal of Science, 276, 1302–1324. ates associated with glacial deposits in Neoproterozoic times. Sedi-
Domack, E. W. 1983. Facies of late Pleistocene glacial-marine sediments mentology Review, 1, 1– 16.
on Whidbey Island, Washington: an istostatic glacial marine Fisher, R. V. 1971. Features of coarse-grained, high concentration
sequence. In: Molnia, B. F. (ed.) Glacial-Marine Sedimentation. fluids and their deposits. Journal of Sedimentary Petrology, 41,
Plenum, New York, 535–570. 916– 927.
Domack, E. W. 1988. Biogenic facies in the Antarctic glaciomarine Flint, R. F. 1975. Features other than diamicts as evidence of ancient gla-
environment: basis for a polar glaciomarine summary. Palaeogeogra- ciations. In: Wright, A. E. & Morsley, F. (eds) Ice Ages: Ancient
phy, Palaeoclimatology, Palaeoecology, 63, 357– 372. and Modern. Seelhouse Press, Liverpool, 121–135.
Domack, E. W. & Lawson, D. E. 1985. Pebble fabric in an ice-rafted Flint, R. F., Sanders, J. E. & Rodgers, J. 1960. Diamictite, a substitute
diamicton. Journal of Geology, 93, 577– 591. term for symmictite. Bulletin of the Geological Society of America,
Domack, E. W., Halverson, G. P., Willmott, V., Leventer, A., 71, 1809– 1810.
Brachfield, S. & Ishman, S. 2007. Spatial and temporal distri- Fricker, H. A., Scambos, T., Bindschadler, R. & Padman, L. 2007. An
bution of ikaite crystals in Antarctic glacial marine sediments. active subglacial water system in West Antarctica mapped from
United States Geological Survey and The National Academies, space. Science, 315, 1544–1548.
USGS OF-2007-1047, Extended Abstract 015, 5. Gilbert, R. 1990. Rafting in glacimarine environments. In: Dowdes-
Dowdeswell, J. A. & Scourse, J. D. (eds) 1990. Glaciomarine Environ- well, J. A. & Scourse, J. D. (eds) Glacimarine Environments:
ments: Processes and Sediments. Geological Society, London, Processes and Sediments. Geological Society, London, Special Pub-
Special Publications, 53. lications, 53, 105– 121.
Dowdeswell, J. A., Elverhøi, A. & Spielhagen, R. 1998. Glacimarine Halverson, G. P. 2011. Glacial sediments and associated strata of the
sedimentary processes and facies on the polar North Atlantic Polarisbreen Group, Northeastern Svalbard. In: Arnaud, E.,
Margins. Quaternary Science Reviews, 17, 243–272. Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
RECOGNITION OF GLACIAL DEPOSITS 49

Record of Neoproterozoic Glaciations. Geological Society, London, McDonald, B. C. & Shilts, W. W. 1975. Interpretation of faults in gla-
Memoirs, 36, 571–579. ciofluvial sediments. In: Jopling, A. V. & McDonald, B. C. (eds)
Hambrey, M. J. 1994. Glacial Environments. UCL Press, London. Glaciofluvial and Glaciolacustrine Sedimentation. SEPM Special
Hambrey, M. J. & Glasser, N. F. 2003. Glacial sediments: processes, Publication, 123–131.
environments and facies. In: Middleton, G. V. (ed.) Encyclopedia McKay, R. M., Dunbar, G. B., Naish, T. R., Barrett, P. J., Carter, L.
of Sediments and Sedimentary Rocks. Kluwer Academic Publishers, & Harper, M. 2008. Retreat history of the Ross Ice Sheet (Shelf)
Dordrecht, 316–331. since the last glacial maximum from deep-basin sediment cores
Hambrey, M. J. & Harland, W. B. 1978. Analysis of Pre-Pleistocene around Ross Island. Palaeogeography, Palaeoclimatology, Palaeo-
glacigenic rocks: aims and problems. In: Schluchter, C. (ed.) ecology, 260, 245– 261.
Moraines and Varves. A. A. Balkema Publishers, Rotterdam, Menzies, J. (ed.) 1995. Modern Glacial Environments: Processes,
271– 275. Dynamics and Sediments. Butterworth-Heinemann, Oxford.
Hambrey, M. J. & McKelvey, B. 2000. Neogene fjordal sedimentation Menzies, J. (ed.) 1996. Past Glacial Environments: Sediments, Forms and
on the western margin of the Lambert Graben, East Antarctica. Sedi- Techniques. Butterworth-Heinemann, Oxford.
mentology, 47, 577– 607. Menzies, J. 2000. Microstructures within diamictites of the Lower Gow-
Harland, W. B. 1964. Critical evidence for a great infra-cambrian glacia- ganda Formation (Huronian) near Elliot Lake, Ontario. Journal of
tion. Geologisches Rundschau, 54, 45– 61. Sedimentary Research, 35, 210– 216.
Harland, W. B., Herod, K. N. & Krinsley, D. H. 1966. The definition Miller, N. R., Stern, R. J., Avigad, D., Beyth, M. & Schilman, B.
and identification of tills and tillites. Earth-Science Reviews, 2, 2009. Cryogenian slate-carbonate sequences of the Tambien Group,
225– 256. northern Ethiopia (I): pre-sturtian chemostratigraphy and corre-
Hart, J. K. 1994. Till fabric associated with deformable beds. Earth lations. Precambrian Research, 170, 129– 156.
Surface Processes and Landforms, 19, 15 – 32. Mohrig, D., Elverhoi, A. & Parker, G. 1999. Experiments on the rela-
Hart, J. K. & Boulton, G. S. 1991. The interrelation of glaciotectonic tive mobility of muddy subaqueous and subaerial debris flows and
and glaciodepositional processes within the glacial environment. their capacity to remobilize antecedent deposits. Marine Geology,
Quaternary Science Reviews, 10, 335– 350. 154, 117–129.
Hart, J. K. & Roberts, D. H. 1994. Criteria to distinguish between Moncrieff, A. M. 1989. Classification of poorly sorted sedimentary
subglacial glaciotectonic and glaciomarine sedimentation, I. rocks. Sedimentary Geology, 65, 191– 194.
Deformation styles and sedimentology. Sedimentary Geology, 91, Mulder, T. & Alexander, J. 2001. The physical character of subaqueous
191– 213. sedimentary density flows and their deposits. Sedimentology, 48,
Huddart, D. & Hambrey, M. J. 1996. Sedimentary and tectonic develop- 269– 299.
ment of a high-arctic, thrust-moraine complex: Comfortlessbreen, Naish, T., Powell, R. et al. 2009. Obliquity-paced Pliocene West
Svalbard. Boreas, 25, 227– 243. Antarctic ice sheet oscillations. Nature, 458, 322–328.
James, N. P., Narbonne, G. M., Dalrymple, R. W. & Kyser, C. 2005. Nardin, T. R., Hein, F. J., Gorsline, D. S. & Edwards, B. D. 1979.
Glendonites in Neoproterozoic low-latitude interglacial sedimentary A review of mass movement processes, sediment and acoustic
rocks, northwest Canada: insights on the Cryogenian ocean and characteristics and contrasts in slope and base of slope
Precambrian cold carbonates. Geology, 33, 9 – 12. systems versus canyon-fan basin floor systems. In: Doyle, L. J. &
Jensen, P. A. & Wulff-Pedersen, E. 1996. Glacial or non-glacial origin Pilkey, O. H. (eds) Geology of Continental Slopes. Society
of the Bigganjargga tillite, Finnmark, northern Norway. Geological of Economic Paleontologists and Mineralogists, Special Publication,
Magazine, 133, 137– 145. 61– 73.
Krzyszkowski, D. 1993. Pleistocene glaciolacustrine sedimentation in Nemec, W., Lønne, I. & Blikra, L. H. 1999. The Kregnes moraine in
a tectonically-active zone, Kleszczow Graben, central Poland. Gauldalen, west central Norway: anatomy of a Younger Dryas pro-
Sedimentology, 40, 623– 644. glacial delta in a palaeofjord basin. Boreas, 28, 254–476.
Laberg, J. S. & Vorren, T. O. 2000. Flow behaviour of the submarine Nystuen, J. P. 1985. Facies and preservation of glaciogenic sequences
glacigenic debris flows on the Bear Island Trough Mouth Fan, from the Varanger ice age in Scandinavia and other parts of the
western Barents Sea. Sedimentology, 47, 1105–1117. North Atlantic Region. Palaeogeography, Palaeoclimatology,
Lawson, D. E. 1982. Mobilization, movement and deposition of active Palaeoecology, 51, 209–229.
subaerial sediment flows, Matanuska Glacier, Alaska. Journal of ´ Cofaigh, C. & Dowdeswell, J. A. 2001. Laminated sediments in gla-
O
Geology, 90, 279–300. cimarine environments: diagnostic criteria for their interpretation.
Leather, J., Allen, P. A., Brasier, M. D. & Cozzi, A. 2002. Neoproter- Quaternary Science Reviews, 20, 1411– 1436.
ozoic snowball Earth under scrutiny: evidence from the Fiq glaciation ´ Cofaigh, C., Pudsey, C. J., Dowdeswell, J. A. & Morris, P. 2002.
O
of Oman. Geology, 30, 891– 894. Evolution of subglacial bedforms along a paleo-ice stream, Antarctic
Le Heron, D. P. & Etienne, J. L. 2005. A complex subglacial clastic Peninsula continental shelf. Geophysical Research Letters, 29,
dyke swarm, Sólheimajökull, southern Iceland. Sedimentary 1199–1203.
Geology, 181, 25 –37. Phillips, E. R., Evans, D. A. J. & Auton, C. A. 2002. Polyphase
Le Heron, D. P., Craig, J. & Etienne, J. L. 2009. Ancient glaciations and deformation at an oscillating ice margin following the Loch
hydrocarbon accumulations in North Africa and the Middle East. Lomond Readvance, central Scotland, U.K. Sedimentary Geology,
Earth Science Reviews, 93, 47 –76. 149, 157–182.
Le Heron, D. P., Sutcliffe, O. E., Whittington, R. J. & Craig, J. Postma, G., Nemec, W. & Kleinspehn, K. L. 1988. Large floating clasts
2005. The origins of glacially-related soft-sediment deformation in turbidites: a mechanism for their emplacement. Sedimentary
structures in Upper Ordovician glaciogenic rocks: implication for Geology, 58, 47 –61.
ice sheet dynamics. Palaeogeography, Palaeoclimatology, Palaeo- Powell, R. D. & Cooper, J. M. 2002. A glacial sequence stratigraphic
ecology, 218, 75 – 103. model for temperate glaciated continental shelves. In: Dowdeswell,
Leonard, E. M. 1986. Varve studies at Hector Lake, Alberta, Canada, and J. A. & O ´ Cofaigh, C. (eds) Glacier-Influenced Sedimentation on
the relationship between glacial activity and sedimentation. Quatern- High Latitude Continental Margins. Geological Society, London,
ary Research, 25, 199– 214. Special Publications, 203, 215–244.
Lønne, I. 1995. Sedimentary facies and depositional architecture of ice- Powell, R. D. & Elverhoi, A. (eds) 1989. Modern glacimarine environ-
contact glaciomarine systems. Sedimentary Geology, 98, 13 –43. ments: Glacial and marine controls of modern lithofacies and bio-
Martins-Neto, M. A. 1996. Lacustrine fan-deltaic sedimentation in a facies. Marine Geology, 85, 416.
Proterozoic rift basin: the Sopa-Brumadinho Tectonosequence, Prior, D. B., Bornhold, B. D. & Johns, M. W. 1984. Depositional
southeastern Brazil. Sedimentary Geology, 106, 65 –96. characteristics of a submarine debris flow. Journal of Geology, 92,
McCarroll, D. & Rijsdijk, K. F. 2003. Deformation styles as a key for 707– 727.
interpreting glacial depositional environments. Journal of Quatern- Rampino, M. R. 1994. Tillites, diamictites and ballistic ejecta of large
ary Science, 18, 473–489. impacts. The Journal of Geology, 102, 439–456.
50 E. ARNAUD & J. L. ETIENNE

Ravnås, R. & Steel, R. J. 1998. Architecture of marine rift-basin succes- Taylor, J., Dowdeswell, J. A., Kenyon, N. H. & O’Cofaigh, C. 2002.
sions. American Association of Petroleum Geologists Bulletin, 82, Late Quaternary architecture of trough-mouth fans: debris flows and
110– 146. suspended sediments on the Norwegian margin. In: Dowdeswell,
Reading, H. G. 1978. Sedimentary Environments and Facies, 1st edn. J. A. & O ´ Cofaigh, C. (eds) Glacier-Influenced Sedimentation on
Blackwell Scientific Publications, Oxford. High Latitude Continental Margins. Geological Society, London,
Rieu, R., Allen, P. A., Etienne, J. L., Cozzi, A. & Weichert, U. 2006. Special Publications, 203, 55– 71.
A Neoproterozoic glacially influenced basin margin succession and Thomas, G. S. P. & Connell, R. J. 1985. Iceberg drop, dump and ground-
‘atypical’ cap carbonate associated with bedrock paleovalleys, ing structures from Pleistocene glacio-lacustrine sediments, Scotland.
Mirbat area, southern Oman. Basin Research, 18, 471– 496. Journal of Sedimentary Petrology, 55, 243– 249.
Roberts, D. H. & Hart, J. K. 2005. The deforming bed characteristics of Tweed, F. S. & Russell, A. J. 1999. Controls on the formation and
a stratified till assemblage in north East Anglia, U.K.: investigating sudden drainage of glacier-impounded lakes: implications for
controls on sediment rheology and strain signatures. Quaternary jökulhaup characteristics. Progress in Physical Geography, 23,
Science Reviews, 24, 123–140. 79 – 110.
Schermerhorn, L. J. G. 1974. Late Precambrian mixtites: glacial and/or Visser, J. N. J. 1983. The problems of recognizing ancient subaqueous
non-glacial. American Journal of Science, 274, 673–824. debris flow deposits in glacial sequences. Transactions of the Geo-
Siegert, M. J., Carter, S., Tabacco, I., Popov, S. & Blankenship, D. D. logical Society of South Africa, 86, 127– 135.
2005. A revised inventory of Antarctic subglacial lakes. Antarctic Weaver, P. P. E., Wynn, R. B., Kenyon, N. H. & Evans, J. 2000. Con-
Science, 17, 453–460. tinental margin sedimentation, with special reference to the north-east
Smith, N. D. & Ashley, G. M. 1985. Proglacial lacustrine environments. Atlantic margin. Sedimentology, 47(Suppl. 1), 239– 256.
In: Ashley, G. M., Shaw, J. & Smith, N. D. (eds) Glacial Sedimen- Wingham, D. J., Siegert, M. J., Sheperd, A. & Muir, A. S. 2006. Rapid
tary Environments. SEPM Short Course Notes, 16, 135– 216. discharge connects Antarctic subglacial lakes. Nature, 440,
Smith, N. D., Phillips, A. C. & Powell, R. D. 1990. Tidal drawdown: 1033– 1036.
A mechanism for producing cyclic sediment laminations in glacio- Woodworth-Lynas, C. M. T. & Guigne, J. Y. 1990. Iceberg scours in
marine deltas. Geology, 18, 10 –13. the geological record: examples from glacial Lake Agassiz. In:
Sohn, Y. K. 2000. Depositional process of submarine debris flows in the Dowdeswell, J. A. & Scourse, J. D. (eds) Glaciomarine Environ-
Miocene fan deltas, Pohang Basin, S.E. Korea, with special reference ments: Processes and Sediments. Geological Society, London,
to flow transformation. Journal of Sedimentary Research, 70, Special Publications, 53, 217– 223.
491– 503. Wright, R. & Anderson, J. B. 1980. The importance of sediment gravity
Tanaka, J. & Maejima, W. 1995. Fan-delta sedimentation on the basin flow to sediment transport and sorting in a glacial marine environ-
margin slope of the Cretaceous, strike-slip Izumi Basin, southwestern ment: Eastern Weddell Sea, Antarctica. Geological Society of
Japan. Sedimentary Geology, 98, 205– 213. America Bulletin, 93, 951–963.
Chapter 4

Chemostratigraphy and the Neoproterozoic glaciations

GALEN P. HALVERSON1,3* & GRAHAM SHIELDS-ZHOU2


1
School of Earth and Environmental Sciences, The University of Adelaide, North Terrace, Adelaide, SA 5005, Australia
2
Department of Earth Sciences, University College London, Gower Street, London, WD1E 6BT, UK
3
Present address: Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montreal,
QC, H3A 2A7, Canada
*Corresponding author (e-mail: galen.halverson@mcgill.ca)

Abstract: Although the pre-glacial Proterozoic isotopic record is poorly constrained, it is apparent that the chemical and isotopic
composition of the oceans began to change during the early to mid-Neoproterozoic and experienced considerable fluctuations alongside
climatic instability during much of the subsequent Cryogenian and Ediacaran periods. The earliest known large negative d13C excursion
appears to post-date 811 Ma and fluctuations became progressively more extreme, culminating in the late-Ediacaran ‘Shuram– Wonoka’
anomaly. The negative excursions are commonly associated with pre-glacial and post-glacial times, while extremely high d13C values are
characteristic of strata between glaciations. The precise causal mechanism for these excursions is subject to debate. Seawater 87Sr/86Sr
rose during the Neoproterozoic, with abrupt increases following deglaciation consistent with enhanced weathering rates. Reported
marine sulphate and pyrite d34S data exhibit marked variation through this interval, although the changes are not always consistent
within or between sedimentary successions of equivalent age. Iron-speciation studies indicate that much of this variation was caused
by fluctuating and low sulphate concentrations in seawater, which at times led to the build-up of ferruginous conditions in the
ocean. The application of chemostratigraphy to understanding and correlating the Neoproterozoic glaciations evokes considerable
controversy, and many questions persist regarding the reliability and calibration of the d13C, 87Sr/86Sr and d34S record. Nevertheless,
the individual glaciations appear to be characterized by distinct combined chemostratigraphic signatures, in large part due to the generally
increasing strontium isotope composition of seawater through the Neoproterozoic Era.

Chemical stratigraphy has enjoyed widespread application to the increasing) use (Knoll & Walter 1992; Knoll 2000). Multiple nega-
study of the Neoproterozoic sedimentary record, in particular tive carbon-isotope excursions and intervening intervals of sus-
with regard to the number, correlation, causes and consequences tained high d13Ccarb appear to correlate globally (e.g. Kaufman
of glaciations during this era. Owing to the scarcity of biostratigra- et al. 1997; Halverson et al. 2005). However, because there are
phically useful fossils, it has become the tool of choice for global also multiple glaciations, each apparently associated with negative
stratigraphic correlation of Neoproterozoic strata (Kaufman & carbon-isotope anomalies, these excursions are non-unique and
Knoll 1995; Walter et al. 2000). In their broad survey of carbon- robust correlations, in the absence of firm radiometric ages,
ates, Schidlowski et al. (1975) pioneered the application of require additional data.
carbon-isotope stratigraphy to the Precambrian sedimentary Like carbon isotopes, sulphur-isotope data on both pyrite and
record. Williams (1979) subsequently documented negative d13C sulphate show impressive variability spanning the Neoproterozoic
signatures associated with post-glacial carbonates in Australia. glaciations (Gorjan et al. 2000; Hurtgen et al. 2002). However, the
Knoll et al. (1986) later focused attention on the Neoproterozoic strontium-isotope record is the most useful in discriminating
Era with coupled carbonate (d13Ccarb) and total organic carbon between the various negative carbon-isotope anomalies because
(d13Corg) carbon-isotope records from carbonate-dominated suc- 87
Sr/86Sr increases throughout most of the Neoproterozoic
cessions in Svalbard and East Greenland that showed large and (Shields 1999; Melezhik et al. 2001; Halverson et al. 2007) such
in-phase variations ostensibly linked to glacial episodes. This that each of the prominent carbon-isotope anomalies appears to
study complemented contemporaneous investigations that simi- be associated with distinct strontium-isotope signatures.
larly showed large fluctuations in the d13C composition of Many other chemostratigraphic methods have been applied to
ancient marine carbonates spanning the Precambrian –Cambrian the Neoproterozoic sedimentary record. Iron speciation data
boundary (Hsu et al. 1985; Tucker 1986; Magaritz et al. 1986). have revealed remarkable fluctuations in the redox state of
These studies bolstered the argument that ancient carbonates and seawater throughout the Neoproterozoic, in particular spanning
sedimentary organic matter could be used as proxies for the glaciations and leading up to the first appearance of the Ediacaran
carbon-isotopic composition of the seawater in which they fauna (Canfield et al. 2007, 2008). Post-glacial carbonates have
formed. Inasmuch, they opened the door to carbon-isotope che- also proven to be exceptional archives of the highly unusual
mostratigraphy as a commonplace tool for basin- and global-scale environmental conditions during recovery from the Cryogenian
correlation and as a measure of the behaviour of the exogenic glaciations, preserving in some cases extraordinary geochemical
carbon cycle, in particular as it related to extreme climate vari- anomalies in non-traditional geochemical proxies, such as
ability and biospheric change (e.g. Derry et al. 1992; Kaufman calcium and boron isotopes (Kasemann et al. 2005) and D17O
& Knoll 1995). (Bao et al. 2008). For a more extensive review of these and
Notwithstanding challenges to the integrity of carbon-isotope other chemostratigraphic methods applied to the Neoproterozoic
signatures as proxies for seawater composition (e.g. Frimmel sedimentary record, the reader is referred to Halverson et al.
2009; Knauth & Kennedy 2009; Derry 2010) and poor radiometric (2010). A detailed account of the application of the chemical
age control on many Neoproterozoic successions, carbon-isotope index of alteration (CIA) to the Neoproterozoic glacial record is
stratigraphy is a powerful tool for establishing a chronological presented in this volume by Bahlburg & Dobrzinski (2011), and
framework for this interval, during which extensive carbonates chemostratigraphic and geochronological data from individual
were deposited but for which biostratigraphy is of limited (but sedimentary successions are summarized in reviews throughout

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 51– 66. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.4
52 G. P. HALVERSON & G. SHIELDS-ZHOU

this volume. The aim of this chapter is to provide a brief review of plausibility of sufficient oxidants being available to generate the
the carbon-, sulphur- and strontium-isotope systems and the iron- accompanying highly negative d13Ccarb anomalies (Bristow &
speciation proxy and a synthesis of published data bracketing the Kennedy 2008), the undeniable lack of correlation in certain data
Neoproterozoic glaciations, with a focus on the structure of these sets requires a mechanistic explanation.
records and their temporal relation to glaciation.

Sulphur isotopes
Carbon isotopes
Sedimentary sulphur-isotope ratios are measured on sedimentary
Under normal conditions and at the scale of resolution appropriate sulphates or sulphides. Sulphur-isotope data on sulphates can be
in Precambrian chemostratigraphy, the carbon-isotope compo- recovered, with varying degrees of reliability, from evaporites,
sitions of unaltered marine carbonates (d13Ccarb) precipitated in barite, phosphorites and carbonates (as carbonate-associated sul-
equilibrium with seawater closely approximate the composition phate, CAS). Sulphate-bearing marine minerals commonly do
of the total dissolved inorganic carbon (DIC) pool (Hayes et al. not form in isotopic equilibrium with seawater, so additional argu-
1999). If it is assumed that the carbonates retain their primary ments need to be marshalled to support the presumption that they
signatures after burial, then carbon-isotope chemostratigraphy record seawater sulphate isotopic compositions. The d34S values of
has diverse applications, such as establishing global correlations marine evaporite minerals may be higher or lower than contem-
(e.g. Kaufman et al. 1993, 1997), regional studies of basin archi- poraneous seawater depending on the extent of basin restriction
tecture and carbonate sequence stratigraphy (Halverson et al. and connection to meteoric influence. Stratiform barite d34S
2002; Cozzi et al. 2004; Jiang et al. 2007), modelling of the values are also prone to deviate from seawater, although lowermost
global carbon cycle (Derry et al. 1992), and as a constraint on mech- values are generally close approximations to ocean composition
anisms for the causes of Neoproterozoic glaciation (Kaufman et al. (Shields 2007). Phosphorite can also show higher d34S values
1997; Halverson et al. 2002; Schrag et al. 2002) and post-glacial than seawater due to Rayleigh fractionation during bacterial sul-
carbonate deposition (Kaufman et al. 1997; Hoffman et al. 1998; phate reduction (e.g. Hough et al. 2006), and the same may also
Kennedy et al. 2001; Higgins & Schrag 2003; Shields 2005). be true of some diagenetic carbonate rocks (e.g. Marenco et al.
Debate continues as to whether Neoproterozoic marine carbonates 2008a). An additional problem with carbonate- (and phosphate-)
faithfully preserve their primary depositional signatures, particu- associated sulphate sulphur-isotopic analyses derives from the
larly where they are highly 13C-depleted (Bristow & Kennedy inadvertent incorporation of contaminant sulphate during the dis-
2008; Frimmel 2009; Knauth & Kennedy 2009; Derry 2010). solution process before analysis (Marenco et al. 2008b).
Nevertheless, the general consensus is that, by and large, ancient Unlike marine sulphate minerals, sulphides (typically pyrite)
marine carbonates do preserve seawater compositions, so secular record the net fractionation imparted during bacterial sulphate
trends in d13Ccarb can be interpreted in terms of inputs of carbon reduction (BSR), plus additional fractionation effects contributed
to and outputs from the global ocean (Holser 1997). by disproportionation reactions during oxidative recycling of sul-
The record of the carbon-isotope composition of sedimentary phide (Canfield & Teske 1996; Detmers et al. 2001). Sulphate
organic matter (d13Corg), while much less applied in the Neoproter- and pyrite d34S trends broadly mirror one another, but pyrite is gen-
ozoic, is often used where carbonates are not sufficiently abundant erally much more variable, attributable to the fact that reduction,
to construct a d13Ccarb profile. The analysis of d13Corg has recently oxidation and disproportionation reactions commonly occur
surged due to the hypothesis that some of the extreme negative during early diagenesis, where distinct local effects, such as pore
d13Ccarb anomalies might be the result of partial oxidation of an water sulphate concentrations, availability of appropriate sub-
enormous reactive organic carbon reservoir in the deep ocean strates for BSR, and physiology of the BSR population (Detmers
(Rothman et al. 2003). Organic carbon data are intrinsically et al. 2001) strongly influence the final sulphur-isotope signature.
more difficult to acquire than inorganic carbon data, and the Sulphur isotopes have been widely applied to problems in Pre-
connection between extant kerogen (or total organic carbon, cambrian Earth history. For example, a gradual increase in the
TOC) and original biomass is inevitably blurred by diagenesis, spread of d34Spyrite in shales (d34Ssulphate – d34Spyrite isotopic dis-
thermal alteration of the organic matter (Hayes et al. 1983; crimination) broadly coincides with separate geological and geo-
Kaufman et al. 1991; Hayes et al. 1999) and input of detrital chemical evidence for a Neoproterozoic rise in the oxygenation
kerogen. Nevertheless, early stratigraphically constrained Neopro- of the Earth’s environment (Canfield & Teske 1996; Fike et al.
terozoic d13Corg data sets (e.g. Knoll et al. 1986; Kaufman et al. 2006). Originally, this isotopic shift was linked to the evolutionary
1997) showed broad correlation with contemporaneous d13Ccarb, radiation of sulphide-oxidizing bacteria (Canfield & Teske 1996),
suggesting that it could be used in parallel with or in place of but recent evidence puts this evolutionary event much earlier
(where primary carbonates are unavailable) inorganic carbon (Johnston et al. 2005). Instead, this isotopic shift could have
data as a proxy for seawater compositions, albeit with the caveat been caused by a rise in the importance of sulphide oxidation
that the net fractionation between the original DIC reservoir and and disproportionation reactions during early diagenesis, which
extracted kerogen is inherently variable. A compilation of may be indirectly linked to oxygenation through seafloor redox
organic carbon data by Hayes et al. (1999) suggested an average changes. The d34S pyrite compilation is also the basis for the
difference between contemporaneous carbonates and organic model that the global deep ocean became dominantly euxinic,
matter (1TOC) of c. 30‰ in the Neoproterozoic, but with large fluc- rather than oxygenated, as massive banded iron-formation depo-
tuations (+10‰) associated with d13Ccarb anomalies, in particular sition came to a close in the late Palaeoproterozoic (Poulton
those spanning glaciations. et al. 2004). The Neoproterozoic is characterized by extraordinary
More recent and detailed data sets from Ediacaran-aged sedi- fluctuations in both d34Spyrite and d34Ssulphate in the Neoprotero-
ments have revealed sustained intervals where d13Ccarb and zoic, which are closely coupled to glacial events (Halverson &
d13Corg are decoupled or even anti-correlated (Calver 2000; Fike Hurtgen 2007). However, the extent to which these variations
et al. 2006; McFadden et al. 2008). Swanson-Hysell et al. truly reflect global seawater sulphate compositions and environ-
(2010) recently argued, based on new, paired data from mental conditions is poorly established.
Australia, that decoupling of these two proxies began sometime Highly precise measurement of all four sulphur isotopes
in the Cryogenian. Decoupling between these two proxies is a (32S, 33S, 34S, 36S) has shown resolvable and consistent variations
prediction of the Rothman et al. (2003) hypothesis, because the in the parameters D33S and D36S, which can reflect both mass-
implied oceanic organic-carbon reservoir would be much larger independent (MIF) and mass-dependent fractionation (MDF) pro-
than the DIC reservoir. While questions remain as to the cesses. MIF of sulphur isotopes has been applied to reconstructing
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 53

the evolution of atmospheric oxygen levels (Farquhar et al. 2000), because of its long ocean residence time. Although isotopic analy-
but has not been systematically applied to the Neoproterozoic, sis is now standard procedure, samples need to be carefully
when atmospheric O2 concentrations likely exceeded the low screened because of the risk of contaminating the sample with
level (,1025 present levels) required to suppress MIF fraction- diagenetically altered or detrital materials. This risk is reduced
ation in the atmosphere (Pavlov & Kasting 2002). On the other for Phanerozoic studies for which microscopically well-preserved
hand, variations in D33S and D36S resulting from MDF record low-Mg calcite tests (e.g. foraminifera, brachiopods and belem-
information about specific sulphur transformations and pathways nites) can be analysed. However, for much of the Precambrian, it
that cannot be gleaned from traditional measurements of is necessary to identify least altered components of the bulk car-
2
S/4S alone (Johnston et al. 2005; Ono et al. 2006). Although bonate rock (e.g. Derry et al. 1992; Kaufman et al. 1993).
not reviewed here, the technique of measuring quadruple sulphur Strontium isotope analysis comprises the following six steps, of
isotopes promises to yield significant insight into behaviour of which the first four vary considerably between research groups: (i)
the sulphur cycle leading up to and following the Neoproterozoic sample selection using petrographic and geochemical criteria; (ii)
glaciations. physical extraction from the host sample; (iii) chemical pre-
treatment to remove contaminant Sr ions; (iv) sample dissolution;
(v) chemical separation of the element Sr; (vi) mass spectrometric
Iron speciation analysis. Studies show that least altered ratios can most easily be
obtained by targeted microsampling of fine-grained calcite
The impressive range of d34Spyrite and d34Ssulphate values that are (micrite), which is free of later recrystallization and dolomitiza-
typical of the Neoproterozoic are commonly interpreted in terms tion, although early diagenetic cements may also be reliable
of a low sulphate ocean (Halverson & Hurtgen 2007). This tracers of ocean composition (e.g. Kaufman et al. 1993; Fairchild
interpretation is consistent with the observation that seawater et al. 2000). Chemical pre-leaching helps remove contaminant Rb
during the Neoproterozoic was occasionally rich in dissolved and radiogenic Sr (Gorokhov et al. 1995), while subsequent
ferrous iron, based on the occurrence of iron formation. Iron- incomplete dissolution in weak acids such as acetic acid has
speciation studies shed further light on the redox state of the been shown to limit inadvertent dissolution of clay mineral-bound
oceans and are being carried out increasingly on Neoproterozoic Sr (Bailey et al. 2000).
successions (Canfield et al. 2008). Redox conditions in the water The Neoproterozoic strontium-isotope record is typically
column affect the speciation of iron in marine sediments. The reconstructed from calcite that has relatively high strontium and
highly reactive iron pool (FeHR) is that which is geochemically low rubidium concentrations or otherwise exhibits evidence of
available during diagenesis and includes iron bound in carbonates, limited exchange with diagenetic fluids, as measured, for
iron oxides and oxyhydroxides, and iron sulphides (Canfield example, in Mn/Sr ratios (Brand & Veizer 1981; Jacobsen &
1989). Sediments deposited under an oxic water column tend to Kaufman 1999). Although carbonate originally precipitated as
contain less reactive iron as a fraction of total sedimentary iron aragonite is intrinsically more suitable due to its high original Sr
(FeHR/FeT) than those deposited under an anoxic water column, content (Kulp et al. 1952), the abundance of metastable carbonate
with a proposed cut-off at 0.38 (Raiswell et al. 1988; Lyons & precipitates in the Precambrian indicates that some apparently
Severmann 2006). Canfield et al. (2008) further proposed that pristine components could be the result of diagenetic replacement
the measure of pyrite-bound iron v. total reactive iron (FePy/ (Fairchild et al. 2000). Because alteration usually increases the
87
FeHR) can be used to distinguish iron-rich anoxic v. euxinic Sr/86Sr of carbonates (Banner & Hanson 1990), the lowest
water columns, with sediments exhibiting FePy/FeHR . 0.8 measured values within a suite of rocks are typically inferred to
being classified as euxinic. most closely approximate the composition of the seawater from
Neoproterozoic sediments deposited beneath a storm wave which it precipitated. However, this ought not merely to be
base typically exhibit high FeHR/FeT ratios (Fig. 4.1), indicating assumed, as post-depositional fluids are not always more radio-
that anoxia was common in the deeper marine environment. genic than coeval seawater. With this in mind, least-altered ratios
However, sulphidic conditions are so far only indicated during are best extrapolated from diagenetic trends, for example by
the pre-glacial Neoproterozoic and early Cambrian (Canfield plotting Mn/Sr, Mg/Ca, 1/Sr or d18O against measured 87Sr/86Sr.
et al. 2008), suggesting ferruginous rather than sulphidic Recent compilations of the 87Sr/86Sr evolution of Neoprotero-
(euxinic) anoxia during the interval of climatic fluctuations. zoic oceans share the common, dominant feature of ratios rising
Because ferruginous conditions can only arise when the molar from as low as c. 0.7055 in the early Neoproterozoic to as high
flux of FeHR to the deep ocean is greater than half the flux of as 0.7085 in the Ediacaran Period (Jacobsen & Kaufman 1999;
sulphide, their reappearance during the Neoproterozoic after an Walter et al. 2000; Melezhik et al. 2001; Halverson et al. 2007a;
absence of more than a billion years is consistent with a low- Shields 2007). This feature alone makes strontium-isotope
sulphate ocean reservoir. chemostratigraphy a key tool in establishing correlations and rela-
tive age in otherwise poorly radiometrically dated Neoproterozoic
successions. Additional detail in the strontium-isotope record
Strontium isotopes remains controversial due to the relatively poor resolution of the
Neoproterozoic record and variable quality of samples used in
Over timescales of .106 years, the 87Sr/86Sr ratio of seawater compilations (Melezhik et al. 2001). However, an apparently
records the relative strontium fluxes from continental weathering reproducible finer-scale structure in the record is emerging with
and hydrothermal input, superimposed on the long-term increase in the promise of improving the Neoproterozoic chronology and the
87
Sr/86Sr as a result of radioactive decay of 87Rb (Edmond 1992). link between coupled tectonic– climatic processes. Importantly,
As such, marine 87Sr/86Sr is a useful measure of evolving tectonic and as will be discussed further below, the seawater strontium-
and long-term climatic regimes. Nevertheless, the ensemble of isotope record appears to be distinct, spanning each of the Neo-
driving mechanisms for the variations in seawater 87Sr/86Sr proterozoic glacial epochs (Table 4.1).
spanning the Neoproterozoic remains controversial. It is doubtful
that seawater 87Sr/86Sr can be simplified to a simple balance
between the hydrothermal and continental inputs, but rather The pre-glacial Neoproterozoic
must also be influenced by continental configuration and CO2 out-
gassing rates (Halverson et al. 2007; Shields 2007). The most characteristic features of the Neoproterozoic chemostra-
Strontium-isotope stratigraphy is based on the presumption that tigraphic record are the generally high d13Ccarb values that prevail
seawater was always isotopically homogeneous with respect to Sr throughout most of the era and the episodic negative anomalies
54 G. P. HALVERSON & G. SHIELDS-ZHOU

Fig. 4.1. Working chemostratigraphic compilations of d13Ccarb, 87Sr/86Sr, d34Ssulphate, d34Spyrite and iron-speciation (FeHR/FeT) data for the Neoproterozoic (modified
after Halverson et al. 2005, 2007a, 2010; Halverson 2006; Canfield et al. 2008). This compilation does not include all available high-quality data due to the difficulty
of integrating data that have poor and variable age constraints (see text for discussion). The names shown vertically in the top of the figure are the informal names of the
principal negative anomalies, as well as the distinct positive shift in d13Ccarb recorded in the lower Little Dal Group (Halverson 2006) and the ‘Keele Peak’ positive
anomaly (Kaufman et al. 2007). Note that there are conflicting sulphur-isotope data, in particular for the late Ediacaran Period (e.g. Ries et al. 2009). A low of 87Sr/86Sr in
the earliest Neoproterozoic (dashed oval) is based on data from Gorokhov et al. (1995). Iron-speciation data (FeHR/FeT) are from Canfield et al. (2008), replotted as
five-point running averages, and from Nagy et al. (2009). Open boxes and diamonds are additional d34Ssulphate and iron-speciation data from the Chuar Group as published
in Johnston et al. (2010). Principle U–Pb geochronological constraints on the d13Ccarb record are shown in triangles at the top of the compilation: (1) an 811.5 + 0.3 Ma
tuff in the Fifteenmile Group, Yukon Territory (Macdonald et al. 2010b); (2) a 760 + 1 Ma tuff in the Ombombo Subgroup, NW Namibia (Halverson et al. 2005); (3) a
716.5 + 0.2 Ma tuff in the lower Rapitan Group, Yukon Territory (Macdonald et al. 2010b); (4) an ash bed in the Ghaub Formation in northern Namibia dated at
635 + 0.6 Ma (Hofmann et al. 2005) and an ash bed in the lower Duoshantuo Formation in South China dated at 635.2 + 0.6 Ma (Condon et al. 2005) bracket the
age of the Cryogenian –Ediacaran boundary; (5) a 542.0 + 0.6 Ma ash bed at the Precambrian –Cambrian boundary in Oman (Amthor et al. 2003); and several
slightly older ages (overlapping triangles) from the latest Ediacaran of the Nama Group, S Namibia (Grotzinger et al. 1995). Shaded boxes and lines show the inferred
range of glacial events, with the gradational shading in the middle Cryogenian box reflecting the wide range of ages for middle Cryogenian glaciation (e.g. Allen &
Etienne 2008).
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 55

that punctuate it (Knoll et al. 1986; Kaufman & Knoll 1995; providing a more robust basis for global correlations. At the
Shields & Veizer 2002). These basic observations raise several upper boundary of the Bitter Springs anomaly, d13Ccarb shifts
questions. First, when did the average values near 0‰ typical of back to values exceeding 5‰.
the Mesoproterozoic (Brasier & Lindsay 1995) give way to the The d13Ccarb of seawater at the time of the putative and possibly
average values of þ5‰ (Halverson et al. 2005) characteristic of regional Kaigas (Bayisi) glaciation at c. 750– 740 Ma (Frimmel
the Neoproterozoic? Second, did the increased variability in et al. 1996; Xu et al. 2009) is not well known. Negative d13C
d13Ccarb accompany the higher average values and, if not, when values occur in the lower Wallekraal Formation, which overlies
did this begin? Finally, what drove these fundamental shifts in the Kaigas Formation (Fölling & Frimmel 2002) but no large
the behaviour of the exogenic carbon cycle? d13C anomaly occurs between the Bitter Springs anomaly and
These questions are difficult to answer because of the poor the Islay anomaly in Svalbard (Fig. 4.1). Instead, d13Ccarb values
carbon-isotope coverage and radiometric ages spanning the remain generally high until the onset of the Islay anomaly (Prave
Mesoproterozoic– Neoproterozoic transition. Bartley et al. et al. 2009), which appears to precede the first major Cryogenian
(2001) and Frank et al. (2003) argued that the increase in the glaciation (Fig. 4.1) and is discussed in more detail in the
amplitude and frequency of fluctuations in d13Ccarb began some- following section.
time in the middle to late Mesoproterozoic, based on data from The strontium-isotope record through the early Neoproterozoic
Siberia and NE Canada, respectively. Halverson (2006) argued shown in Figure 4.1 is largely derived from data from Svalbard and
that sustained d13Ccarb  5‰ did not start until the early Neo- northwestern Canada, even though other presumably primary
proterozoic based on data from the Platform Assemblage of the values have been obtained from other successions such as the
Little Dal Group in NW Canada, where the lowermost data show Bitter Springs Formation. As noted by Bartley et al. (2001),
a distinct ramp from 0 to 5‰ (Fig. 4.1), followed by generally seawater 87Sr/86Sr values reach a nadir around the Mesoproter-
high (5‰) values up to the onset of the Bitter Springs anomaly ozoic –Neoproterozoic transition, presumably related to the final
(see below). However, the age and correlation of the lowermost assembly of Rodinia. The least radiogenic latest Mesoproterozoic
87
Little Dal Group and its equivalence to other rocks within Succes- Sr/86Sr values reported by Bartley et al. (2001) hover around
sion B in NW Canada is not known and only constrained to be 0.7055, which is virtually the same as a value of 0.7056 from
between c. 1004 (Rainbird et al. 1996) and 811 Ma (Macdonald limestones of the Atar Group, Taoudeni Basin (Veizer et al.
et al. 2010b). It is therefore unclear whether the lower Little Dal 1983), now dated at 1.1 Ga (Rooney et al. 2010) and the lowest
positive d13C shift (Fig. 4.1) truly does record the first highly values in the lower Little Dal Group (Halverson et al. 2007a).
13
C-enriched seawater in the Neoproterozoic, or what the precise However, Gorokhov et al. (1995) reported values from pristine
age of this event might be. limestones of the presumed early Neoproterozoic Burovaya For-
Much better constrained now is the timing of presumably the mation (Turukhansk uplift, NW Siberia) as low as 0.7052. Thus,
first major negative d13Ccarb excursion in the Neoproterozoic: the 0.7055 is a plausible approximation for seawater 87Sr/86Sr at the
Bitter Springs anomaly (Table 4.1; Halverson et al. 2005). Mesoproterozoic –Neoproterozoic boundary, but probably does
Although older negative d13Ccarb excursions have been reported not represent the nadir before the protracted Neoproterozoic rise
(Bartley et al. 2001; Shields 2007), these are generally from in 87Sr/86Sr.
poorly dated successions and may be the result of alteration. For Strontium-isotope ratios had risen to 0.7063 by the onset of the
example, 13C-depleted I6 unit rocks of the Atar Group, Mauritania Bitter Springs anomaly (Fig. 4.1). Although a more discreet and
(Shields 2007), although now known to be c. 1.1 Ga in age smaller shift to slightly higher ratios accompanies the negative
(Rooney et al. 2010), are likely to have undergone significant iso- d13C shift at the base of the anomaly, 87Sr/86Sr returns
topic alteration (Fairchild et al. 1990). The Bitter Springs negative to c. 0.7063 at the end of the Bitter Springs anomaly, before appar-
anomaly was first fully documented in the Bitter Springs ently increasing to as high as 0.7070 in the Backlundtoppen
Formation in the Amadeus basin of central Australia (Hill et al. Formation in Svalbard (prior to the Islay anomaly). No intervening
2000). It was subsequently recorded in more detail in Svalbard data from this transition have been documented. As will be
(Halverson et al. 2005), where it is clearly defined by reciprocal discussed below, this middle Neoproterozoic 87Sr/86Sr peak is fol-
d13C shifts of .8‰ that perfectly coincide with rare exposure sur- lowed by a temporary return to values as low as 0.7063.
faces in the Akademikerbreen Group. The anomaly also appears to The sulphur-isotope record for the early Neoproterozoic is rela-
occur in the upper Little Dal Group (Upper Carbonate Formation) tively sparse, despite the fact that voluminous evaporites were
in the Mackenzie Mountains (Halverson 2006). These global cor- deposited in Australia and NW Canada c. 830 to 800 Ma (Evans
relations are reinforced by additional geochemical data (Halverson 2006). Indeed, the compilation in Figure 4.1 does not include any
et al. 2007a), namely 87Sr/86Sr (see below) and an updated and data older than c. 820 Ma, although this lacuna is in part an artefact
high-resolution d13Ccarb record through the Bitter Springs For- of poor control on ages and correlations, which limit the data that
mation (Swanson-Hysell et al. 2010), as well as the coincidence can be used. d34Ssulphate from evaporites in the Gillen Member of
of the shifts with prominent sequence boundaries. Both features the Bitter Springs Formation (i.e. before the Bitter Springs
have been interpreted to relate to large true polar wander episodes anomaly) ranges from 15 to 21‰ (Gorjan et al. 2000; Hill et al.
(Maloof et al. 2006). 2000). From the available data, it appears that early Neoproterozoic
The Bitter Springs anomaly (Table 4.1; Fig. 4.1) has now been seawater d34Ssulphate values do not depart greatly from the Gillen
reproduced across northwestern Canada: within the middle Member values (Fig. 4.1), averaging c. 20‰ through the middle
Lower Tindir Group (Tatunduk Inlier), upper Fifteenmile Group Neoproterozoic, close to average values for the Cretaceous and
(Coal Creek Inlier), and the Wynniatt Formation of the Shaler Cenozoic (Paytan et al. 1998, 2004; Kampschulte & Strauss
Supergroup (Victoria Island; Macdonald et al. 2010b; Jones 2004), but distinct from the highly 34S-enriched values that
et al. 2010), although in this latter case only the negative shift is characterize the early and late Ediacaran and early Palaeozoic
preserved and this is recorded in sediments intruded by mafic (Shields et al. 2004; Kampschulte & Strauss 2004; Halverson &
sills, which means it may be an overprint related to contact meta- Hurtgen 2007).
morphism rather than a primary signal. A likely correlative has Although most d34Ssulphate data of the composite curve for this
also been documented in the Tambien Group of northern Ethiopia time interval are derived from CAS measurements on Svalbard
(Alene et al. 2006; Miller et al. 2009). Macdonald et al. (2010b) samples (Halverson et al. 2010), available data from evaporites
produced a precise U – Pb zircon age of 811.51 + 0.25 Ma on a spanning the Bitter Springs anomaly in the Amadeus Basin
tuff c. 50 m beneath the negative shift presumed to define the (Gorjan et al. 2000; Hill et al. 2000) yield similar values. Pyrite
base of the anomaly in upper Fifteenmile Group, thus constrain- data in this record are also sparse and similarly derived mainly
ing the onset of the Bitter Springs anomaly to c. 810 Ma and from Svalbard (Halverson et al. 2010) and the Bitter Springs
56 G. P. HALVERSON & G. SHIELDS-ZHOU

Table 4.1. Summary of the d13Ccarb and 87Sr/86Sr signatures associated with the major Neoproterozoic negative carbon isotope anomalies (Fig. 4.1) and an abridged
compilation of locations where they are well preserved (for the Maieberg and Rasthof anomalies, only key locations where the anomalies are preserved in great detail
are given). PB, Paraguai Belt; UTG, Upper Tindir Group

Shuram-Wonoka anomaly
d13Ccarb Minimum: , –10‰; protracted recovery
87
Sr/86Sr 0.7081– 0.7084

Region Unit Reference


S Namibia Nama Gp Workman et al. (2002)
Oman Shuram/Kufai Fm Burns & Matter (1993), Le Guerroué et al. (2006)
S China Doushantuo Fm Jiang et al. (2007), McFadden et al. (2008)
S Australia Wonoka Fm Pell et al. (1993), Calver (2000)
N Norway Nyborg Fm Halverson et al. (2005)
SW USA Johnnie Fm Corsetti & Kaufman (2003)
SE Siberia Nikol’skaya/Chenchinskaya Pokrovskii et al. (2006), Melezhik et al. (2009)

Maieberg anomaly
d13Ccarb Minimum: –4 to – 6‰, gradual to abrupt return to near 0‰ values
87
Sr/86Sr 0.7072 at base, increases upsection to 0.7078– 0.7080

Region Unit Reference

NW Namibia Maieberg Fm Hoffman et al. (1998)


S Namibia Bloeddrif Mb Fölling & Frimmel (2002)
Oman Hadash/Masirah Bay Le Guerroué et al. (2006)
NW China Tureeken Fm Xiao et al. (2004)
S China Doushantuo Fm Jiang et al. (2007), McFadden et al. (2008) and others
SW Mongolia Oi Mb Macdonald et al. (2009a)
E Svalbard Dracoisen Fm Kaufman et al. (1997), Halverson et al. (2004)
Scotland Cranford Ls McCay et al. (2006)
NE Alaska Katakturuk Dolomite (K2) Macdonald et al. (2009b)
SW USA Noonday Dolomite Prave (1999)
Brazil (PB) Araras Fm Nogueira et al. (2007)
SE Siberia Barakunskaya Fm Pokrovskii et al. (2006), Melezhik et al. (2009)

Trezona anomaly
d13Ccarb Minimum: –6 to – 8‰, increases before glaciation
87
Sr/86Sr 0.7072– 0.7074

Region Unit Reference

NW Namibia Ombaatjie Fm Halverson et al. (2002)


South Australia Trezona Fm McKirdy et al. (2001)
N Norway Grasdal Fm Halverson et al. (2005)
Scotland Craignish/Ardrishaig Prave et al. (2009)
NE Alaska Katakturuk Dolomite (K1) Macdonald et al. (2009b)
Mackenzie Mts Keele Fm Halverson et al. (2005)

Rasthof anomaly
d13Ccarb Minimum: –4 to 0‰, rises rapidly
87
Sr/86Sr 0.7067– 0.7069 at base, increases up section to 0.7072– 0.7073

Region Unit Reference

NW Namibia Rashtof Fm Hoffman et al. (1998), Yoshioka et al. (2003)


SW Mongolia Base Tayshir Mb Shields et al. (1997), Macdonald et al. (2009a)
S Australia Tapley Hill Fm McKirdy et al. (2001)
Scotland Bonahaven Fm* Brasier & Shields (2000), McCay et al. (2006), Prave et al. (2009)
NE Alaska Katakturuk Dolomite (K1) Macdonald et al. (2009b)
Mackenzie Mts Twitya Fm. Hoffman & Schrag (2002)
SW USA base Beck Spring Fm Corsetti & Kaufman (2003)

Islay anomaly
d13Ccarb Minimum: –5 to – 9‰, return to .0‰ before glaciation
87
Sr/86Sr 0.7065– 0.7067, declines upsection to 0.7064

Region Unit Reference

Tasmania Black River Fm Calver (1998)


E Svalbard Elbobreen Fm Halverson et al. (2004)

(Continued)
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 57

Table 4.1. Continued

Scotland Islay/Lossit Fm Brasier & Shields (2000), Prave et al. (2009), McCay et al.
(2006), Sawaki et al. (2010b)
E Greenland Bed Group 19 Fairchild et al. (2000)
Victoria Island Killian Formation Macdonald et al. (2010b)
Mackenzie Mts Copper Cap Fm Halverson (2006)
Alaska-Yukon Upper Dolomite (LTG) Macdonald et al. (2010a)
SW USA Upper Beck Spring Fm Corsetti & Kaufman (2003)

Bitter Springs anomaly


d13Ccarb Minimum: –5 to 0‰, bracketed by sharp positive and negative shift
87
Sr/86Sr 0.7063–0.7065

Region Unit Reference

N Ethiopia Tambien Group Alene et al. (2006)


Central Australia Bitter Springs Formation Hill et al. (2000)
Western Australia Hussar Formation Hill (2005)
E Svalbard Grusdievbreen/Svanbergfjellet Halverson et al. (2005)
Scotland Ballachulish Sbgp Prave et al. (2009)
Mackenzie Mts Upper Carbonate Fm Halverson (2006)
Yukon Upper Fifteenmile Gp (PF1) Macdonald et al. (2010b)
Alaska-Yukon Upper Shale (LTG) Macdonald et al. (2010a)

*Prave et al. (2009) argued that the post-Port Askaig negative d13C anomaly is missing, in which case the Bonahaven anomaly, which is stratigraphically above the
glacigenic Port Askaig Formation, might be equivalent to the Tayshir anomaly in Mongolia.

Formation (Hill et al. 2000). The most salient pattern in the d34S impossible to discriminate at this point between the model that
data is a coupled increase in d34Ssulphate and decrease in d34Spyrite there were multiple middle Cryogenian glaciations or a single,
coincident with the Bitter Springs anomaly (Fig. 4.1), resulting long-lived glacial epoch that spanned from c. 716 (Macdonald
in a transient increase in D34S. Other constraints on the d34Ssulphate et al. 2010b) to 660 Ma (Fanning & Link 2008).
of early Neoproterozoic seawater come from studies on succes- These deficiencies in the Cryogenian chronology are exacer-
sions from Arctic Canada. d34Ssulphate values there range from 14 bated by presumably larges hiatuses in the records on continental
to 18‰ in the pre-Bitter Springs anomaly (.810 Ma) Gypsum shelves (from whence most of the chemostratigraphic records are
Formation of the Little Dal Group (Turner 2009), while similar derived) resulting from glacioeustatic fall in sea level and glacial
values (16 – 19‰) have been reported from the post-Bitter erosion. These complications in the Cryogenian record make it
Springs anomaly (,780 Ma) Redstone River Formation of the impossible to produce an accurate chemostratigraphic record for
overlying Coates Lake Group (Strauss 1993). the Cryogenian period and the compilation in Figure 4.1 can
The sulphur-isotope record throughout the remainder of the only be regarded as a work in progress. Nevertheless, it is
early –middle Neoproterozoic is thin, but suggests a return notable that independent compilations emerging from widely sep-
to reduced D34S resulting from higher average d34Spyrite sometime arate sedimentary basins do seem to be converging on a similar
prior to the middle Cryogenian glaciation. overall chemostratigraphic structure for the interval spanning the
Cryogenian glaciations (e.g. Halverson et al. 2005; Prave et al.
2009; Macdonald et al. 2010b).
Spanning the Cryogenian glaciations The negative d13Ccarb anomalies that coincide with deglaciation
and are preserved in transgressive –regressive sequences overly-
Significant discussion remains as to the number, timing and dur- ing glacial deposits and their equivalent glacial unconformities
ation of the Cryogenian glaciations, and the base of the Cryogenian (i.e. cap carbonate sequences; Hoffman & Schrag 2002) are well
period has yet to be formally defined. Two particular persistent documented (Table 4.1; Williams 1979; Kaufman et al. 1997;
sources of controversy and confusion are the possible occurrence Kennedy et al. 1998; Halverson et al. 2005; Hoffman et al.
of an early (c. 745–750 Ma; Borg et al. 2003) ‘Kaigas’ glaciation 2007). Therefore, only the most important features are summarized
(Frimmel et al. 1996) that may not have been global in extent and here. Importantly, no more than two distinct, post-glacial negative
the seemingly contradictory age constraints on the so-called ‘Stur- d13Ccarb anomalies linked to Cryogenian glaciation are known to
tian’ or early Cryogenian glaciation. There is no incontrovertible occur in a single succession (Kennedy et al. 1998). Where there
evidence that the Kaigas diamictite in the Gariep belt of southern are two such post-glacial anomalies within one succession or,
Namibia (see Macdonald et al. (2011) for a stratigraphic and where there is just one of known age, consistent distinctions can
sedimentological reappraisal of the Kaigas diamictites) and other be drawn between the two anomalies (Table 4.1; Kennedy et al.
older diamictites in the Katanga Supergroup in Zambia (Key 1998; Hoffman & Schrag 2002).
et al. 2001) and NW China (Xu et al. 2009) are glacial in origin As a general rule, the older of the two post-glacial sequences
(Hoffman & Li 2009). Chemostratigraphic data spanning these lacks the basal transgressive portion, meaning the maximum flood-
isolated diamictites are also sparse. Therefore, although the ing surface is the base of the sequence (Hoffman & Schrag 2002;
putative Kaigas glaciation is plotted in Figure 4.1, we do not Hoffman et al. 2011). d13Ccarb typically begins negative and trends
review the chemostratigraphic framework for this possible middle- positive, returning to values .0‰ within a few metres (Kennedy
Neoproterozoic glaciation. As for the sticky question of the age of et al. 1998; Halverson et al. 2005). For example, in the Rasthof
early Cryogenian glaciation (i.e. glaciation between the older Formation of NW Namibia, d13Ccarb begins at around –4.3‰
Kaigas event and the end-Cryogenian ‘Marinoan’ event), reviewed (herein referred to as the Rasthof anomaly) and crosses over
recently by Hoffman & Li (2009), we acknowledge that it is abruptly to positive values c. 10 m up-section, at the contact
58 G. P. HALVERSON & G. SHIELDS-ZHOU

between a basal rhythmite member and an overlying chaotic equivalent to the Trezona anomaly and the return to positive
microbialite member (Hoffman et al. 1998; Yoshioka et al. d13Ccarb values above the anomaly has been removed by erosion
2003). The thickness of this anomaly in Namibia is in large part as it was in Namibia (Macdonald et al. 2009a).
a function of the amount of allodapic dolomite within the lower Although the homologous pattern of the Islay and Trezona
member (Halverson et al. 2005). Only in rare instances is a nega- anomalies suggests a similar origin, it diminishes the utility of
tive trend preserved in the lowermost part of the sequence (e.g. in these anomalies for correlation purposes. Fortunately, the
the presumed Rasthof-equivalent basal Tayshir Member of the anomalies are clearly distinguished from one another chemostrati-
Tsagaan Oloom Formation, Mongolia; Macdonald et al. 2009a). graphically by their strontium-isotope signatures. Whereas
87
In contrast, the basal transgressive portion of the sequence over- Sr/86Sr ratios of 0.7072–0.7073 coincide with the Trezona
lying the end-Cryogenian glacials and defining the base of the anomaly (Halverson et al. 2007a), 87Sr/86Sr through the older
Ediacaran Period (Knoll et al. 2006) commonly preserves an Islay anomaly distinctly declines from c. 0.7068 to 0.7064
upward negative trend, with values typically in the range of –2 (Fig. 4.2; Table 4.1), with the lowest values in the Islay Formation
to –4‰ (Hoffman & Schrag 2002; Hoffman et al. 2007; Rose & corresponding to the return towards positive d13Ccarb values
Maloof 2010). In the carbonate-dominated post-glacial platformal (Sawaki et al. 2010a). Similar 87Sr/86Sr values have been reported
sequence in NW Namibia (Maieberg Formation) where the Maie- from the likely equivalent interval in the Coates Lake Group
berg anomaly has been most extensively studied, d13Ccarb con- (0.7066; Halverson et al. 2007a) and Bed Group 20 (0.7063;
tinues to decline to nearly –6‰ at the maximum flooding Fairchild et al. 2000). Thus, it appears that the onset of the
surface before gradually returning to higher values over hundreds middle-Cryogenian glaciation is preceded by a 0.0005 decline in
of metres of the section (Hoffman et al. 1998). However, this the 87Sr/86Sr of seawater following the c. 765 Ma peak of
apparent protracted recovery is not typical of the basal Ediacaran c. 0.7070 (Fig. 4.1) documented in the Backlundtoppen Form-
negative d13Ccarb anomaly, and in many other carbonate- ation. In contrast to this negative slope leading up to the middle
dominated locations, including the slope equivalents of the Cryogenian glaciation, the 87Sr/86Sr record in the following
Maieberg Formation in Namibia (Halverson et al. 2005), the interglacial interval has a positive slope (Fig. 4.1), with a smooth
Doushantuo Formation of South China (e.g. McFadden et al. and seemingly rapid increase in 87Sr/86Sr from 0.7067 to 0.7073
2008), and the Barakunskaya Formation in southern Siberia (Pok- recorded in the Tsagaan Oloom Formation in Mongolia (Shields
rovskii et al. 2006), the anomaly is relatively condensed, spanning et al. 1997).
only several metres to a few tens of metres. Thus, relative sedimen- Strontium-rich aragonite and barite cements of the Hayhook
tation rates must be considered when correlating chemostrati- Formation in the upper part of the transgressive portion of the
graphic signals. post-Ice Brook (base Ediacaran) sequence in the Mackenzie
Large negative d13Ccarb anomalies also precede two Cryogenian Mountains, NW Canada, have 87Sr/86Sr ratios of c. 0.70715
glaciations (Table 4.1; Fig. 4.2). The so-called Trezona anomaly, (Halverson et al. 2007a; Hoffman & Halverson 2011). These
named after the Trezona Formation in South Australia where values are effectively indistinguishable from a ratio of 0.70718
the anomaly was first fully documented (McKirdy et al. 2001), measured within the pre-glacial Keele Peak in the same section
occurs prior to the end-Cryogenian (i.e. the ‘Marinoan’ or (Halverson et al. 2007a) and therefore indicate no net change
‘Elatina’ glaciation in South Australia) glaciation and following in 87Sr/87Sr spanning the end-Cryogenian glaciation.
the so-called Keele Peak (Kaufman et al. 1997), where d13Ccarb The sulphur-isotope record spanning the Cryogenian glaciations
reaches values þ9‰ (Figs 4.1 & 4.2). This negative anomaly is almost entirely based on pyrite and CAS (there being few eva-
is best documented in the Ombaatjie Formation in the Otavi porites during this time span) and is intensely variable (Fig. 4.1).
Group, NW Namibia, where it is defined by a gradual, facies- This variability reflects a combination of low seawater sulphate
independent decline from c. þ7 to –5‰ through up to 50 m of concentrations and strongly fluctuating environmental conditions
dominantly shallow-water carbonates, followed by a positive tied to global glaciation (Hurtgen et al. 2002), as well as possible
shift of a few ‰ (Halverson et al. 2002). The extent of truncation post-depositional effects. Sulphur-isotope data (Fig. 4.1) from the
of the anomaly can be used as a general indicator of the depth of Chuar Group in southwestern USA show a return to d34Spyrite
erosion on the glacial surface. values clustering around 0‰ and low d34S values around
A negative anomaly of similar magnitude, referred to as the 750 Ma, coinciding with at least local development of euxinic
Islay anomaly (Prave et al. 2009) after the formation in Scotland deep waters as determined from iron-speciation data on mid-shelf
(Figs 4.1 & 4.2) where it was first documented (Brasier & black shales (Johnston et al. 2010). Sulphur-isotope data are sparse
Shields 2000; McCay et al. 2006; Sawaki et al. 2010a), occurs for the subsequent interval leading up to the middle Cryogenian
prior to the older Cryogenian glaciation in several successions. glaciation (i.e. the c. 717 Ma Rapitan glaciation; Macdonald
In NE Svalbard, this anomaly is virtually identical in shape, mag- et al. 2010b). However, abundant pyrite and CAS data have
nitude and stratigraphic context to the Trezona anomaly in been produced from the interglacial interval, in particular from
Namibia (Fig. 4.2) and is similarly variably erosionally truncated Namibia (Hurtgen et al. 2002; Gorjan et al. 2003) and Australia
(Halverson et al. 2004). However, in Scotland, d13C returns to (Gorjan et al. 2000; McKirdy et al. 2001). Whereas d34Spyrite
positive values prior to glaciation (Prave et al. 2009). In NW data from within the Sturt and equivalent glacial units in Australia
Canada and East Greenland, the presumably (but not definitively) range from 0 to –20‰, they are strongly positive, with values
equivalent anomalies in the Coates Lake Group (Halverson 2006) commonly 20– 60‰ in post-glacial rocks, such as the Tapley
and Bed Groups 19–20 (Fairchild et al. 2000), respectively, Hill, Aralka, Datangpo, Rasthof and lower Court forma-
feature an even more prominent positive shift, possibly providing tions (Fig. 4.1). This extreme, immediately post-glacial trend in
an important distinction from the Trezona anomaly. pyrite values is matched by highly 34S-enriched d34Ssulphate
A potentially different late Cryogenian negative d13Ccarb values from CAS, such that D34S  0‰. d34Spyrite values then
anomaly of similar magnitude has been found in the Tayshir decline precipitously in the middle interglacial period, paralleling
Member of the lower Tsagaan Oloom Formation in Mongolia a smaller decline in d34Ssulphate, resulting in D34S values approach-
(Macdonald et al. 2009a). This anomaly has been tentatively cor- ing 30‰. D34S reduces again just before the onset of the end-
related with the Bonahaven anomaly in Scotland (Prave et al. Cryogenian glaciation, coinciding with the Trezona negative
2009) and an anomaly of much smaller magnitude in the Gruis For- d13C anomaly (Fig. 4.1; Table 4.1).
mation of NW Namibia, between the Rasthof and Trezona d34S values are again erratic immediately following the
anomalies (Fig. 4.2). A similar large negative anomaly in this stra- end-Cryogenian glaciation: a trend of rapidly increasing
tigraphic position has not been documented elsewhere. Therefore, d34Ssulphate that coincides with a wide range in d34Spyrite (– 30 to
it remains plausible that the anomaly in the Tayshir Member is þ20‰) may be sufficient to distinguish this post-glacial (i.e.
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 59

Fig. 4.2. Summary stratigraphic plots and


accompanying d13Ccarb profiles and
87
Sr/86Sr data for key sections spanning the
Trezona and Islay pre-glacial (Cryogenian)
negative d13C anomalies. Note the variable
scales across the non-scaled glaciogenic
intervals in each column. Nucc, Nuccaleena
Formation; Brach, Brachina Formation;
Macdon., Macdonaldryggen Member.
Australia data from McKirdy et al. (2001);
Namibia data from Halverson et al. (2002,
2007a), which includes data from both the
inner and outer continental shelf of the
Ombaatjie Formation; Scotland data from
Brasier & Shields (2000), McCay et al.
(2006) and Prave et al. (2009); Svalbard
data from Halverson et al. (2004, 2007a).

basal Ediacaran) sulphur-isotope record from the middle Cryogen- The Ediacaran Period
ian post-glacial record. However, coeval d34S profiles that vary
from more proximal to more open ocean post-glacial sections in The beginning of the Ediacaran Period coincides with a major
Namibia (Hurtgen et al. 2006) and South China (Li et al. 2010) negative d13Ccarb anomaly (the Maieberg anomaly; Fig. 4.1) that
suggest significant heterogeneity in the concentration and punctuates the final global glaciation of the Neoproterozoic
sulphur-isotope composition of seawater sulphate and generally (Knoll et al. 2006), as discussed above. Although d13Ccarb values
low sulphate concentrations in the earliest Ediacaran Period. return to near 0‰ within about 3 million years of the end of glacia-
Therefore, sulphur-isotope chemostratigraphy may be useful in tion (Condon et al. 2005), they then remain relatively low for much
establishing an early Ediacaran age, but would not be reliable for of the early Ediacaran Period, with the exception of a positive spike
detailed correlations for sedimentary rocks of this age. to þ6 –10‰ seen in some sections, such as NW Namibia and NE
60 G. P. HALVERSON & G. SHIELDS-ZHOU

Svalbard (Halverson et al. 2005). This pattern contrasts with the anomalies and glaciation is exacerbated by the possibility that
sustained high d13Ccarb values typical of much of the earlier whereas the anomaly should be globally synchronous, the glacia-
Neoproterozoic. tion may be diachronous, or there may be multiple discrete glacia-
Superimposed on the lower average d13Ccarb values in the Edia- tions (e.g. Chumakov 2010). Indeed, d13Ccarb values are highly
caran Period, the dominant feature in the record is the so-called variable in relation to other reported Ediacaran glacial deposits,
Shuram (or Shuram-Wonoka) anomaly, named after the Shuram such as in the Kimberley of northern Australia, where positive
Formation in the Huqf Supergroup of Oman (Burns & Matter d13C values characterize carbonates overlying the Egan Formation
1993). The Shuram anomaly features a precipitous drop to (Corkeron 2007, 2011). Other contrasting d13Ccarb values are
d13Ccarb values , – 10‰, followed by a protracted recovery back found in the Hankalchough Formation in the Tarim Basin, NW
to values near 0‰ (Fig. 4.1). This extraordinary anomaly has China (Xiao et al. 2004), and the Hongtiegou Formation in the
been the subject of considerable recent research and controversy. Chaidam Basin, NW China (Shen et al. 2010). Some authors
However, despite arguments against a primary origin for the have even argued that some of the purported Ediacaran glacial
anomaly (Bristow & Kennedy 2008; Knauth & Kennedy 2009; deposits, such as the Hankalchough (Chumakov 2010) and the
Derry 2010), anomalies of strikingly similar magnitude, structure Hongtiegou (Shen et al. 2010) formations significantly post-date
and approximate age occur in multiple sedimentary basins from the Gaskiers glaciation (Chumakov 2009). Germs et al. (2009)
across the globe (Table 4.1), including Sr-rich limestones in south- argue for a latest Ediacaran (c. 547 Ma) ‘Schwarzrand’ glaciation
central Siberia (Pokrovskii et al. 2006; Melezhik et al. 2009). based on evidence from the upper Nama Group, Namibia.
Thus, the most parsimonious interpretation is that the Shuram Notwithstanding the ambiguity in the age, timing, number and
anomaly reflects a primary oceanographic phenomenon. distribution of Ediacaran-aged glacial deposits, carbon-isotope
Despite the large number of chemostratigraphic data that have data from the basal Ediacaran Nyborg Formation and the overlying
been obtained globally from rocks preserving the Shuram glaciogenic Mortensnes Formation of the Varanger Peninsula,
anomaly, the timing and duration of the anomaly and, indeed, northern Norway, do appear to resolve the relative timing
the number of large Ediacaran negative d13Ccarb anomalies between at least one middle Ediacaran glaciation and the Shuram
remain controversial. For example, whereas d13Ccarb records d13Ccarb anomaly. Dolomite beds within Member E in the upper-
from thick, seemingly continuous carbonate-rich sections in most Nyborg Formation, in the least-truncated section beneath the
Oman (Fike et al. 2006; Le Guerroué et al. 2006) and south-central major erosional unconformity at the base of the glaciogenic Mor-
Siberia (Pokrovskii et al. 2006) clearly show a single, long-lived tensnes Formation (Edwards 1984), have d13Ccarb compositions
negative d13Ccarb anomaly, certain sections from the Doushantuo of –7.6 to –9.9‰, while the matrix of a carbonate-dominated dia-
Formation, South China, show a pair of large-magnitude mictite within the Mortensnes Formation has a d13Ccarb composition
anomalies separated by d13Ccarb values as high as þ6‰ (Jiang of –10.4‰ (see Rice et al. 2011). Because the Shuram anomaly is
et al. 2007; McFadden et al. 2008). Because the Doushantuo the only interval in the Ediacaran Period where d13Ccarb values are
pattern is preserved in highly condensed sections (relative to sec- known to drop below –7‰, these data are interpreted to indicate
tions from other successions) from a basin that was not demonstra- that the Mortensnes glaciation post-dated the nadir of the negative
bly fully open to the global ocean and has not been reproduced anomaly (Halverson et al. 2005). In this regard, an Ediacaran glacia-
globally, the more consistent and higher-resolution patterns tion, although seemingly not global, may have been triggered by a
derived from Oman and Siberia are favoured in the compilation similar mechanism as the Cryogenian glaciations.
in Figure 4.1. However, we acknowledge that the inconsistent pat- The return to positive d13Ccarb values following the Shuram
terns between the Doushantuo Formation and correlative units anomaly occurs late in the Ediacaran, by c. 551 Ma, implying
worldwide pose a problem for carbon-isotope chemostratigraphy. that the Ediacaran fauna first appeared during the anomaly
Separate data sets from the Wonoka Formation, South Australia (Condon et al. 2005). A subsequent, short-lived negative d13C
(Calver 2000), Oman (Fike et al. 2006) and South China (McFad- anomaly coincides with the Precambrian –Cambrian boundary
den et al. 2008) also show a decoupling between the d13Ccarb (Magaritz et al. 1986; Knoll et al. 1995; Saylor et al. 1998;
record spanning the Shuram anomaly and coeval organic carbon Amthor et al. 2003), but there is no evidence to link this
(d13Corg) data (Calver 2000; Fike et al. 2006; McFadden et al. anomaly with a glaciation.
2008), a pattern that lends support to the popular hypothesis that Strontium-isotope ratios from the Mackenzie Mountains
the Ediacaran ocean contained a large, reactive organic carbon indicate little net change in seawater 87Sr/86Sr spanning the end-
pool that strongly modulated d13Ccarb and the oxidation of the Cryogenian glaciation (Fig. 4.2). However, data from strontium-
deep oceans (Rothman et al. 2003; Condon et al. 2005). rich limestones in the lower Maieberg Formation, NW Namibia,
However, the relationship between d13Ccarb and d13Corg is not show a rapid rise in 87Sr/86Sr to values .0.7078 following
the same in these three successions. For example, d13Corg values post-glacial maximum flooding and prior to recovery in d13Ccarb
are much more negative in South China than in Oman or South values to near 0‰ (Halverson et al. 2007; cf. Shields 2007).
Australia, and d13Corg values are more variable through the main This positive spike in 87Sr/86Sr is a predictable result of extremely
part of the anomaly in Oman than elsewhere. Therefore, the signifi- high CO2 levels and corresponding elevated silicate weathering
cance and reliability of the middle Ediacaran d13Corg record rates following deglaciation (Higgins & Schrag 2003). Simple
is unclear. geochemical modelling also implies that the spike to more
Of critical importance in reconstructing the connections radiogenic values should be short-lived, with 87Sr/86Sr approach-
between perturbations to the exogenic carbon cycle, global ing background values after CO2 declined to more typical
climate and biospheric evolution in the Ediacaran Period is the levels within a few million years of deglaciation (Le Hir et al.
connection, if any, between the Shuram anomaly and the short- 2009). This decline may be recorded in data from the Una
lived c. 580 Ma Gaskiers glaciation (Bowring et al. 2003). Is the Group (Irecê basin) on the São Francisco Craton, Brazil (Misi &
Gaskiers glaciation, like the global Cryogenian glaciations, pre- Veizer 1998), where the strontium- and carbonate-isotope
ceded by a large negative d13C anomaly? Unfortunately, no unam- compositions suggest an early Ediacaran age, consistent with
biguous evidence for both the glaciation and the full anomaly recent detrital zircon data from the equivalent Bambui Group
occurs in any single stratigraphic succession. Although Halverson (Sial et al. 2010).
et al. (2005) interpreted major erosional unconformities in succes- A much better established pattern in the Ediacaran strontium-
sions hosting the Shuram anomaly, such as the Wonoka Formation isotope record is the subsequent rise to 87Sr/86Sr  0.7080 that
(South Australia) and the Johnnie Formation (Death Valley), to be characterizes the latter part of the Ediacaran Period (e.g.
related to Gaskiers glacioeustasy, this argument remains speculat- Kaufman et al. 1997; Jacobsen & Kaufman 1999; Shields 2007;
ive. The problem of resolving the relative timing between d13C Melezhik et al. 2009). Sawaki et al. (2010b) suggested additional
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 61

structure in the late Ediacaran record based on data they obtained other major Ediacaran events, either d34S compositions of late
from a drill core through the Doushantuo Formation, including a Neoproterozoic seawater sulphate were highly heterogeneous
drop to ,0.7080 prior to the Shuram anomaly and a peak in (Ries et al. 2009), inferred correlations are inaccurate, or the
87
Sr/86Sr of 0.7090 coinciding with the protracted low d13Ccarb measured sulphur-isotope compositions in some or all of the sedi-
of the anomaly. Regarding the high values, as for Oman, where mentary successions do not preserve reliable signatures of primary
similarly high values have been reported for the late Ediacaran seawater composition.
(Burns et al. 1994), none of these highly enriched values is
preserved in samples retaining high strontium concentrations.
Conversely, strontium-rich limestones in the upper Doushantuo Discussion and conclusions
Formation in the core have 87Sr/86Sr values of c. 0.7085, similar
to strontium-isotope data straddling the Precambrian –Cambrian Precise radiometric ages are the gold standard for drawing precise
boundary in other well-studied successions (Brasier et al. 1996; interbasinal correlations, particularly in the absence of a robust
Kaufman et al. 1997; Saylor et al. 1998). Gaucher et al. (2009) biostratigraphy. Direct ages are also required for calibrating the
report 87Sr/86Sr values of ,0.7080 in the late Ediacaran Arroyo sedimentary record, but they are frustratingly rare in most of the
del Soldado Group on the Rio de la Plata craton, lending support successions that archive important information about the evolution
to the possibility that 87Sr/86Sr may have indeed fluctuated of Earth’s surface environment in the Neoproterozoic. Thus, other
strongly in the late Ediacaran Period, much as it did in the sub- tools and many assumptions are required to merge the fragmentary
sequent Early Cambrian (Halverson et al. 2010). Evidently, records from across the globe into one coherent Neoproterozoic
much work remains in elucidating the fine structure in the late chronology. Isotope chemostratigraphy has long been heralded
Ediacaran 87Sr/86Sr record. as a possible solution (Knoll & Walter 1992). Insofar as the
The relationship between middle Ediacaran 87Sr/86Sr and the popular proxies for ocean chemistry (d13C, d34S, 87Sr/86Sr) are
Gaskiers (or other Ediacaran) glaciation is virtually unconstrained. reliable, geologists can characterize the Neoproterozoic record
However, strontium-isotope data from highly strontium-enriched one fragment at a time. Accordingly, isotope geochemistry has
limestones in southern Siberia show 87Sr/86Sr ¼ 0.7081 at about been widely applied to stratigraphic studies of the Neoproterozoic,
the level of the nadir in the Shuram anomaly, followed by an with a strong focus on the glaciogenic record.
increase towards 0.7086 up-section, but still within the range of An intimate relationship between extreme perturbations to the
very low d13Ccarb (Pokrovskii et al. 2006; Melezhik et al. 2009). global carbon cycle, as recorded in the marine d13Ccarb record,
Anomalously 34S-enriched seawater sulphate compositions and episodes of widespread glaciation (Knoll et al. 1986) has
bookend the Ediacaran Period (Fig. 4.1) in many successions. As long been the starting point for these studies. However, even as
discussed above, variable, but predominantly high d34Ssuphate and this fundamental observation about the behaviour of the Neopro-
low D34S values near 0‰ prevail in the basal Ediacaran Period terozoic Earth has withstood the test of time, diverse studies
(Fig. 4.1). By the early Cambrian, d34Ssulphate, as measured in a from across the world have added important new details to this
variety of minerals, including evaporites, francolite and carbonate, broad pattern. Long-lived positive carbon-isotope trends had
was again high (35 –40‰) in most basins (e.g. Holser & Kaplan been ascribed to pre-glacial oceanographic conditions, while nega-
1966; Strauss 1993; Shields et al. 1999; Kampschulte & Strauss tive carbon-isotope anomalies were attributed to post-glacial ‘cap
2004). However, d34Ssulphate from the late Ediacaran in South carbonates’, providing for elegantly simple models for the inter-
China (McFadden et al. 2008) and southern Namibia (Ries et al. relation between d13Ccarb, carbon burial, atmospheric CO2 and
2009) are generally lower, with concomitant decreases in glaciation (e.g. Kaufman et al. 1997). The recognition that a
D34S. The source of this substantial spatial variability in the late large negative d13Ccarb anomaly actually preceded the end-
Ediacaran sulphur-isotope record, whether oceanographic, Cryogenian glaciation shattered this model of a simple cause-
related to post-depositional influences or analytical artefacts, is and-effect relationship between the two and opened the door for
not resolved. the compelling but controversial Snowball Earth hypothesis
The Ediacaran sulphur-isotope record as compiled in Figure 4.1 (Hoffman et al. 1998), which continues to dominate the debate
comprises a detailed data set of d34Ssulphate (CAS) and d34Spyrite about the Neoproterozoic glaciations.
data from the Nafun Group, Oman (Fike et al. 2006) and pyrite It is now recognized that negative d13C anomalies also precede
data from the Pertatataka Formation, central Australia (Gorjan middle Cryogenian (McCay et al. 2006; Prave et al. 2009; Mac-
et al. 2000). The Oman record, together with other fragmentary donald et al. 2010b) and middle Ediacaran glaciations (Halverson
data (Fig. 4.1), indicates that more typical d34Ssulphate values et al. 2005). Furthermore, in all cases, it appears that the initiation
close to 20‰ are the norm for the Ediacaran Period, although eva- of glaciation followed a return towards positive d13Ccarb values
porite and CAS data appear to be contradictory at times (cf. Fike after the d13Ccarb minima, thus further complicating an already
et al. 2006 with Schröder et al. 2004). At the same time, data complex relationship between the anomalies and global cooling.
from both Oman and central Australia display a spike to positive This motif in the coupled d13C-climate record renders using the
d34Spyrite values in the early Ediacaran Period, followed by a carbon-isotope record alone for global correlations insufficient,
gradual decrease through the middle Ediacaran Period, resulting but fortunately it projects upon a marine strontium-isotope
in a steady increase in d34S, a predictable result of increasing sea- record with a strong unidirectional trend (Fig. 4.1). 87Sr/86Sr
water sulphate concentrations (Halverson & Hurtgen 2007). increases steadily during the Neoproterozoic from 0.7055 around
Average d34Spyrite values in the deepwater Conception Group the Mesoproterozoic– Neoproterozoic boundary to 0.7085 at the
(Avalon Peninsula, SE Newfoundland; Canfield et al. 2007) Neoproterozoic –Palaeozoic boundary, interrupted by only a few
show a decrease from about þ20 to 0‰ across the Gaskiers significant but short-lived declines (Fig. 4.1). Consequently, sea-
glaciation, but the d34Ssulphate record spanning the glaciation is water 87Sr/86Sr ratios are different for each of the large negative
unknown. Iron-speciation data indicate ferruginous (and presum- anomalies associated with the Neoproterozoic glaciations
ably low sulphate) conditions during the Gaskiers glaciation (Table 4.1). The potential of strontium isotopes as a tracer of
(Canfield et al. 2008). Detailed d34Ssulphate (CAS) records have global environmental change remains to be fully realized.
been produced across the onset of the Shuram –Wonoka anomaly. However, abrupt rises, most notably following the Cryogenian
In Oman, a spike in d34Ssulphate to þ29‰ coincides with the glaciations, are consistent with enhanced chemical weathering
anomaly (Fike et al. 2006). Across the presumably equivalent after glaciation. Conversely, changes in seawater 87Sr/86Sr appear
interval in Death Valley (Kaufman et al. 2007) and South China to have been relatively muted during glaciations.
(McFadden et al. 2008), d34Ssulphate decreases to values ,20‰. At this stage, the Neoproterozoic record is now sufficiently well
Thus, although d34S records show tantalizing connections with understood that the combination of carbon- and strontium-isotope
62 G. P. HALVERSON & G. SHIELDS-ZHOU

stratigraphy, particularly if integrated with stratigraphic and sedi- Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to Snowball
mentological data, enables confident correlations and assignment Earth. Nature Geoscience, 1, 817– 825.
of robust, if only relative, ages on carbonate-rich successions Amthor, J. E., Grotzinger, J. P., Schröder, S., Bowring, S. A., Rame-
that otherwise lack firm age control. High-quality data sets span- zani, J., Martin, M. W. & Matter, A. 2003. Extinction of Cloudina
ning variable portions of the Neoproterozoic record and tied to and Namacalathus at the Precambrian – Cambrian boundary in Oman.
an increasing number of precise U –Pb zircon ages have now Geology, 31, 431–434.
been produced from across the world (e.g. Macdonald et al. Bahlburg, H. & Dobrzinski, N. 2011. A review of the Chemical Index
2010b). The result is a convergence in carbon- and strontium- of Alteration (CIA) and its application to the study of Neoproterozoic
glacial deposits and climate transitions. In: Arnaud, E., Halverson,
isotope compilations towards a consistent pattern, even if many
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neo-
of the finer details remain obscure. proterozoic Glaciations. Geological Society, London, Memoirs, 36,
The integrated stratigraphic, d13C and 87Sr/86Sr records serve 81 – 92.
as a chronological template to the Neoproterozoic onto which Bailey, T. R., McArthur, J. M., Prince, H. & Thirlwall, M. F. 2000.
other secular data sets can be added, such as sulphur-isotope Dissolution methods for strontium isotope stratigraphy: whole rock
compositions, redox specific data, biomarkers and biostratigraphy. analysis. Chemical Geology, 167, 313–319.
The sulphur-isotope record is patchy, particularly for the first Banner, J. L. & Hanson, G. N. 1990. Calculation of simultaneous isoto-
half of the Neoproterozoic, and highly variable, with clear pic and trace element variations during water– rock interaction with
instances where data sets spanning equivalent-aged rocks from applications to carbonate diagenesis. Geochimica et Cosmochimica
different parts of the world, and even single basins, are not Acta, 54, 3123– 3137.
consistent with one another. Thus, much work remains to be Bao, H., Lyons, J. R. & Zhou, C. 2008. Triple oxygen isotope evidence
done to establish the reliability of the sulphur-isotope proxy and for elevated CO2 levels after a Neoproterozoic glaciation. Nature,
the causes, either related to oceanographic conditions or diagenetic 453, 504– 506.
process, for the high degree of scatter in the record. Nevertheless, Bartley, J. K., Semikhatov, M. A., Kaufman, A. J., Knoll, A. H.,
several of the salient patterns, mostly notably extremely high Pope, M. C. & Jacobsen, S. B. 2001. Global events across the Meso-
d34Ssulphate and d34Spyrite values, are temporally linked to glaciation, proterozoic – Neoproterozoic boundary: C and Sr isotopic evidence
indicating that the record can be used to help reconstruct the from Siberia. Precambrian Research, 111, 165– 202.
palaeoenvironmental evolution of Neoproterozoic ocean chem- Borg, G., Kärner, K., Buxton, M., Armstrong, R. & van der Merwe,
istry. Specifically, large variations in d34S of sulphate and pyrite S. W. 2003. Geology of the Skorpion supergene zinc deposit,
and the difference between these two values in coeval sediments southern Namibia. Economic Geology, 98, 749–771.
Bowring, S., Myrow, P., Landing, E., Ramezani, J. & Grotzinger, J.
(D34S) is an indirect measure of fluctuations in the redox state of
2003. Geochronological constraints on terminal Proterozoic events
global seawater (Hurtgen et al. 2005). The sustained and very and the rise of the Metazoans. Geophysical Research Abstracts
high d34Spyrite (20 –60‰) values recorded in the post-Sturtian (EGS, Nice), 50, 13219.
Tapley Hill – Aralka formations in Australia (Gorjan et al. 2000) Brand, U. & Veizer, J. 1981. Chemical diagenesis of a multicomponent
and the peak in d34Ssulphate straddling the Precambrian – Cambrian carbonate system – 2: stable isotopes. Journal of Sedimentary Petrol-
boundary stand out as unique chronostratigraphic markers in the ogy, 51, 987– 997.
otherwise noisy sulphur-isotope record (Fig. 4.1). Brasier, M. D. & Lindsay, J. F. 1995. A billion years of environmental
Iron-speciation data has become a popular new tool for recon- stability and the emergence of eukaryotes: New data from northern
structing water column redox conditions (Lyons & Severmann Australia. Geology, 26, 555–558.
2006) that has been successfully applied to the Neoproterozoic Brasier, M. D. & Shields, G. 2000. Neoproterozoic chemostratigraphy
(e.g. Canfield et al. 2007, 2008). Although it remains to be and correlation of the Port Askaig glaciation, Dalradian Supergroup
shown that distinct trends can be correlated beyond a single of Scotland. Journal of the Geological Society, London, 157, 909–914.
basin (Johnston et al. 2010), or are even consistent within a Brasier, M. D., Shields, G. A., Kuleshov, V. N. & Zhegallo, E. A.
single basin, sharp fluctuations in normalizsed highly reactive 1996. Integrated chemo- and biostratigraphic calibration of early
iron content in shales (FeHR/FeT), for example, are closely animal evolution: Neoproterozoic– early Cambrian of southwest
linked to glaciation and other proxy evidence for oscillations in Mongolia. Geological Magazine, 133, 445–485.
the redox chemistry of the oceans (Canfield et al. 2007, 2008). Bristow, T. F. & Kennedy, M. J. 2008. Carbon isotope excursions and
Although not discussed here, many other chemostratigraphic the oxidant budget of the Ediacaran atmosphere and ocean.
proxies show great promise for tracing evolving redox conditions Geology, 36, 863–866.
through the Neoproterozoic and evaluating the highly unusual Burns, S. J. & Matter, A. 1993. Carbon isotopic record of the latest Pro-
terozoic from Oman. Eclogae Geologicae Helvetiae, 86, 595– 607.
environment in the aftermath of global glaciations (Halverson
Burns, S. J., Haudenschild, U. & Matter, A. 1994. The strontium
et al. 2010, and references therein). Furthermore, biostratigraphy isotopic composition of carbonates from the late Precambrian
is increasingly being tightly integrated with chemostratigraphic (,560–540 Ma) Huqf Group of Oman. Chemical Geology, 111,
data in order to evaluate more closely the connections between 269– 282.
biospheric and biogeochemical change (e.g. McFadden et al. Calver, C. 2000. Isotope stratigraphy of the Ediacaran (Neoproterozoic
2008; Nagy et al. 2009; Macdonald et al. 2010a). Thus, there is III) of the Adelaide rift complex, Australia, and the overprint of
little doubt that chemostratigraphy will continue to elucidate the water column stratification. Precambrian Research, 100, 121– 150.
finer details in the number, timing and correlation of Neoprotero- Canfield, D. E. 1989. Reactive iron in marine sediments. Geochimica et
zoic glaciations and the environmental conditions leading up to Cosmochimica Acta, 53, 619– 632.
and following them. Canfield, D. E. & Teske, A. 1996. Late Proterozoic rise in atmospheric
oxygen concentration inferred from phylogenetic and sulfur-isotope
This represents a contribution of the IUGS- and UNESCO-funded IGCP (Inter- studies. Nature, 382, 127–132.
national Geoscience Programme) Project #512. Canfield, D. E., Poulton, S. W., Knoll, A. H., Narbonne, G. M.,
Ross, G., Goldberg, T. & Strauss, H. 2008. Ferruginous conditions
dominated later Neoproterozoic deep-water chemistry. Science, 321,
References 949– 952.
Canfield, D. E., Poulton, S. W. & Narbonne, G. M. 2007. Late-
Alene, M., Jenkin, G. R. T., Leng, M. J. & Darbyshire, D. P. F. 2006. Neoproterozoic deep-ocean oxygenation and the rise of animal life.
The Tambien Group, Ethiopia: An early Cryogenian (ca. 800– 735 Science, 315, 92 – 95.
Ma) Neoproterozoic sequence in the Arabian– Nubian Shield. Pre- Chumakov, N. M. 2009. The Baykonurian Glaciohorizon of the Late
cambrian Research, 147, 79 –99. Vendian. Stratigraphy and Geological Correlation, 17, 373–381.
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 63

Chumakov, N. M. 2010. Neoproterozoic glacial events in Eurasia. In: Frimmel, H. W., Klötzi, U. S. & Siegfried, P. R. 1996. New Pb/Pb
Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, H. E. single zircon age constraints on the timing of Neoproterozoic glacia-
(eds) Neoproterozoic-Cambrian Tectonics, Global Change and tion and continental break-up in Namibia. The Journal of Geology,
Evolution: A Focus on Southwestern Gondwana. Developments in 104, 459–469.
Precambrian Geology, 16. Elsevier, Dordrecht, 389– 403. Gaucher, C., Sial, A. N., Poiré, D., Gómez-Peral, L., Ferreira, V. P.
Condon, D., Zhu, M., Bowring, S., Jin, Y., Wang, W. & Yang, A. & Pimentel, M. M. 2009. Chemostratigraphy. Neoproterozoic –
2005. From the Marinoan glaciation to the oldest bilaterians: Cambrian evolution of the Rı́o de la Plata Palaeocontinent. In:
U –Pb ages from the Doushantou Formation, China. Science, 308, Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, H. E.
95 – 98. (eds) Neoproterozoic –Cambrian Tectonics, Global Change and
Corkeron, M. 2007. ‘Cap carbonates’ and Neoproterozoic glacigenic Evolution: A Focus on Southwestern Gondwana. Developments in
successions from the Kimberly region, north-west Australia. Sedi- Precambrian Geology, 16, 115–122.
mentology, 54, 871– 903. Germs, G. J. B., Miller, R. McG., Frimmel, H. E. & Gaucher, C. 2009.
Corkeron, M. 2011. Neoproterozoic glacial deposits of the Kimberly Syn- to late-orogenic sedimentary basins of southwestern Africa.
Region and northwestern Northern Territory, Australia. In: Neoproterozoic to Early Palaeozoic evolution of Southwestern
Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Africa. In: Gaucher, C., Sial, A. N., Halverson, G. P. &
Geological Record of Neoproterozoic Glaciations. Geological Frimmel, H. E. (eds) Neoproterozoic– Cambrian Tectonics, Global
Society, London, Memoirs, 36, 659–672. Change and Evolution: A Focus on Southwestern Gondwana. Devel-
Corsetti, F. A. & Kaufman, A. J. 2003. Statigraphic investigations of opments in Precambrian Geology, 16, 183– 203.
carbon-isotope anomalies and Neoproterozoic ice ages in Death Gorjan, P., Veevers, J. J. & Walter, M. R. 2000. Neoproterozoic sulfur-
Valley, California. Geological Society of America Bulletin, 115, isotope variation in Australia and global implications. Precambrian
916– 932. Research, 100, 151–179.
Cozzi, A., Allen, P. A. & Grotzinger, J. P. 2004. Understanding Gorjan, P., Walter, M. R. & Swart, R. 2003. Global Neoproterozoic
carbonate ramp dynamics from C profiles: Examples from (Sturtian) post-glacial sulfide-sulfur isotope anomaly recognised in
the Neoproterozoic Buah Formation of Oman. Terra Nova, 16, Namibia. Journal of African Earth Sciences, 36, 89 –98.
62 – 67. Gorokhov, I. M., Semikhatov, M. A., Baskakov, A. V., Kutyavin, E.
Derry, L. A. 2010. A burial diagenesis origin for the Ediacaran Shuram- P., Melnikov, N. N., Sochava, A. V. & Turchenko, T. L. 1995. Sr
Wonoka anomaly. Earth and Planetary Science Letters, 295, isotopic composition in Riphean, Vendian, and Lower Cambrian
152– 162. carbonates from Siberia. Stratigraphy and Geological Correlation,
Derry, L. A., Kaufman, A. J. & Jacobsen, S. B. 1992. Sedimentary 3, 1 – 28.
cycling and environmental change in the Late Proterozoic: evidence Grotzinger, J. P., Bowring, S. A., Saylor, B. Z. & Kaufman, A. J.
from stable and radiogenic isotopes. Geochimica et Cosmochimica 1995. Biostratigraphic and geochronologic constraints on early
Acta, 56, 1317–1329. animal evolution. Science, 270, 598–604.
Detmers, J., Brüchert, V., Habicht, K. S. & Kuever, J. 2001. Diversity Halverson, G. P. 2006. A Neoproterozoic Chronology. In: Xiao, S. &
of sulfur isotope fractionations by sulfate-reducing prokaryotes. Kaufman, A. J. (eds) Neoproterozoic Geobiology and Paleobiology.
Applied and Environmental Microbiology, 67, 888– 894. Springer, 231– 271.
Edmond, J. M. 1992. Himalayan tectonic, weathering processes, and Halverson, G. P. & Hurtgen, M. T. 2007. Ediacaran growth of the
strontium isotope record in marine limestones. Science, 258, marine sulfate reservoir. Earth and Planetary Science Letters, 263,
1594– 1597. 32– 44.
Edwards, M. B. 1984. Sedimentology of the Upper Proterozoic glacial Halverson, G. P., Hoffman, P. F., Schrag, D. P. & Kaufman, A. J.
record, Vestertana Group, Finnmark, North Norway. Norges geolo- 2002. A major perturbation of the carbon cycle before the Ghaub
giske Undersøkelse Bulletin, 394, 1– 76. glaciation (Neoproterozoic) in Namibia: prelude to snowball Earth?
Evans, D. A. D. 2006. Proterozoic low orbital obliquity and axial- Geochemistry, Geophysics, Geosystems, 3, 10.1029/2001GC000244.
dipolar geomagnetic field from evaporite paleolatitudes. Nature, Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
444, 51– 55. glaciation (Neoproterozoic) in Svalbard. Basin Research, 16,
Fairchild, I. J., Marshall, J. D. & Bertrand-Sarfati, J. 1990. Strati- 297– 324.
graphic shifts in carbon isotopes from Proterozoic stromatolitic Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice,
carbonates (Mauritania): influences of primary mineralogy and dia- A. H. 2005. Towards a Neoproterozoic composite carbon-isotope
genesis. American Journal of Science, 290A, 46 –79. record. Geological Society of America Bulletin, 117, 1181–1207.
Fairchild, I. J., Spiro, B., Herrington, P. M. & Song, T. 2000. Controls Halverson, G. P., Dudas, F. O., Maloof, A. C. & Bowring, S. A.
on Sr and C isotope compositions of Neoproterozoic Sr-rich 2007a. Evolution of the 87Sr/86Sr composition of Neoproterozoic
limestones of East Greenland and North China. In: Grotzinger, J. seawater. Palaeogeography, Palaeoclimatology, Palaeoecology,
& James, N. (eds) Carbonate Sedimentation, Diagenesis in an 256, 103–129.
Evolving Precambrian World. SEPM Special Publication, 67, Halverson, G. P., Maloof, A. C., Schrag, D. P., Dudas, F. O. &
297– 313. Hurtgen, M. T. 2007b. Stratigraphy and geochemistry of a ca 800
Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian gla- Ma negative carbon isotope interval in northeastern Svalbard. Chemi-
ciation: data from the Adelaide Geosyncline, South Australia and cal Geology, 237, 5 –27.
Pocatello Formation, Idaho, USA. Selwyn Symposium 2008, Geo- Halverson, G. P., Wade, B. P., Hurtgen, M. T. & Barovich, K. M.
logical Society of Australia, Extended Abstracts, 91, 57– 62. 2010. Neoproterozoic chemostratigraphy. Precambrian Research,
Farquhar, J., Bao, H. & Thiemens, M. 2000. Atmospheric influence of 182, 337–350.
Earth’s earliest sulfur cycle. Science, 289, 756– 758. Hayes, J. M., Kaplan, I. R. & Wedeking, K. W. 1983. Precambrian
Fike, D. A., Grotzinger, J. P., Pratt, L. M. & Summons, R. E. 2006. organic geochemistry, preservation of the record. In: Schopf, J. W.
Oxidation of the Ediacaran ocean. Nature, 444, 744– 747. (ed.) Earth’s Earliest Biosphere. Its Origin and Evolution. Princeton
Fölling, P. G. & Frimmel, H. W. 2002. Chemostratigraphic correlation University Press, 93 – 134.
of carbonate successions in the Gariep and Saldania belts, Namibia Hayes, J. M., Strauss, H. & Kaufman, A. J. 1999. The abundance of 13C
and South Africa. Basin Research, 14, 69 –88. in marine organic matter and isotopic fractionation in the global bio-
Frank, T. D., Kah, L. C. & Lyons, T. W. 2003. Changes in organic geochemical cycle of carbon during the past 800 Ma. Chemical
matter production and accumulation as a mechanism for isotopic Geology, 161, 103– 125.
variation in the Mesoproterozoic ocean. Geological Magazine, 140, Higgins, J. A. & Schrag, A. P. 2003. Aftermath of a snowball Earth.
397– 420. Geochemistry, Geophysics, Geosystems, 4, 1028.
Frimmel, H. 2009. Trace element distribution in Neoproterozoic carbon- Hill, A. C. 2005. Stable isotope stratigraphy, GSWA Lancer 1, Officer
ates as palaeoenvironmental indicator. Chemical Geology, 258, Basin, Western Australia. In: Mory, A. J. & Haines, P. W. (eds)
338– 353. GSWA Lancer 1 Well Completion Report (Interpretive Papers)
64 G. P. HALVERSON & G. SHIELDS-ZHOU

Officer and Gunbarrel Basins, Western Australia. Western Australia Kampschulte, A. & Strauss, H. 2004. The sulfur isotopic evolution of
Geological Survey Record 2005/4, 1– 11. Phanerozoic seawater based on the analysis of structurally substituted
Hill, A. C., Arouri, K., Gorjan, P. & Walter, M. R. 2000. Geochem- sulfate in carbonates. Chemical Geology, 204, 255–286.
istry of marine and non-marine environments of a Neoproterozoic Kasemann, S. A., Hawkesworth, C. J., Prave, A. R., Fallick, A. E. &
cratonic carbonate/evaporite: the Bitter Springs Formation, Central Pearson, P. 2005. Boron and calcium isotope composition in Neo-
Australia. In: Grotzinger, J. P. & James, N. P. (eds) Carbonate proterozoic carbonate rocks from Namibia: evidence for extreme
Sedimentation and Diagenesis in an Evolving Precambrian World. environmental change. Earth and Planetary Science Letters, 231,
SEPM Tulsa Special Publications, 67, 327– 344. 73 – 86.
Hoffman, P. F. & Halverson, G. P. 2011. Neoproterozoic glacial record Kaufman, A. J. & Knoll, A. H. 1995. Neoproterozoic variations in the C-
in the Mackenzie Mountains, northern Canadian Cordillera. In: isotopic composition of seawater. Precambrian Research, 73, 27–49.
Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Kaufman, A. J., Hayes, J. M., Knoll, A. H. & Germs, G. J. B. 1991. Iso-
Geological Record of Neoproterozoic Glaciations. Geological topic composition of carbonates and organic carbon from upper
Society, London, Memoirs, 36, 397–411. Proterozoic successions in Nambia. Precambrian Research, 49,
Hoffman, P. F. & Li, Z.-X. 2009. A palaeogeographic context for Neopro- 301– 327.
terozoic glaciation. Palaeogeography, Palaeoclimatology, Palaeo- Kaufman, A. J., Jacobsen, S. B. & Knoll, A. H. 1993. The Vendian
ecology, 277, 158– 172. record of Sr and C isotopic variations in seawater: Implications for
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: tectonics and paleoclimate. Earth and Planetary Science Letters,
testing the limits of global change. Terra Nova, 14, 129–155. 120, 409– 430.
Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P. Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes, ice
1998. A Neoproterozoic snowball Earth. Science, 281, 1342– 1346. ages, and terminal Proterozoic earth history. Proceedings of the
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., National Academy of Science (USA), 94, 6600–6605.
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635 Kaufman, A. J., Corsetti, F. A. & Varni, M. A. 2007. The effect of
Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary rising atmospheric oxygen on carbon and sulfur isotope anomalies
Science Letters, 258, 114– 131. in the Neoproterozoic Johnnie Formation, Death Valley, USA.
Hoffman, P. F., Macdonald, F. A. & Halverson, G. P. 2011. Chemical Chemical Geology, 237, 47 –63.
sediments associated with Neoproterozoic glaciations: iron for- Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K.-H. &
mation, cap carbonate, barite and phosphorite. In: Arnaud, E., Arthur, M. 1998. Two or four Neoproterozoic glaciations?
Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Geology, 26, 1059–1063.
Record of Neoproterozoic Glaciations. Geological Society, London, Kennedy, M. J., Christie-Blick, N. & Sohl, L. E. 2001. Are Proterozoic
Memoirs, 36, 67– 80. cap carbonates and isotopic excursions a record of gas hydrate
Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L. destabilization following Earth’s coldest intervals? Geology, 29,
2004. A U– Pb zircon date from the Neoproterozoic Ghaub For- 443– 446.
mation, Namibia: constraints on Marinoan glaciation. Geology, 32, Key, R. M., Liyungu, A. K., Njamu, F. M., Somwe, V., Banda, J.,
817– 820. Mosley, P. N. & Armstrong, R. A. 2001. The western arm of the
Holser, W. T. 1997. Geochemical events documented in inorganic carbon Lufilian Arc in NW Zambia and its potential for copper mineraliz-
isotopes. Palaeogeography, Palaeoclimatology, Palaeoecology, 132, ation. Journal of African Earth Sciences, 33, 503– 528.
173– 182. Knauth, L. P. & Kennedy, M. J. 2009. The late Precambrian greening of
Holser, W. T. & Kaplan, I. R. 1966. Isotope geochemistry of sedimen- the Earth. Nature, 460, 728– 732.
tary sulfates. Chemical Geology, 1, 93 –135. Knoll, A. H. 2000. Learning to tell Neoproterozoic time. Precambrian
Hough, M. L., Shields, G. A., Evins, L. Z., Strauss, H., Henderson, R. Research, 100, 3– 20.
A. & Mackenzie, S. 2006. A major sulphur-isotope event at c. 510 Knoll, A. H. & Walter, M. R. 1992. Latest Proterozoic stratigraphy and
Ma: a possible anoxia– extinction –volcanism connection during the Earth history. Nature, 356, 673–677.
Early– Middle Cambrian transition? Terra Nova, 18, 257– 263. Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K. & Lambert, I. B.
Hsu, K. J., Oberhänsli, H., Gao, J. Y., Shu, S., Haihong, C. & Krähen- 1986. Secular variation in carbon isotope ratios from Upper Protero-
bühl, U. 1985. ‘Strangelove ocean’ before the Cambrian explosion. zoic successions of Svalbard and east Greenland. Nature, 321,
Nature, 316, 809–811. 832– 837.
Hurtgen, M. T., Arthur, M. A., Suits, N. & Kaufman, A. J. 2002. The Knoll, A. H., Grotzinger, J. P., Kaufman, A. J. & Kolosov, P. 1995.
sulfur isotopic composition of Neoproterozoic seawater sulfate: Integrated approaches to terminal Proterozoic stratigraphy: An
implications for snowball Earth? Earth and Planetary Science example from the Olenek Uplift, northeastern Siberia. Precambrian
Letters, 203, 413–429. Research, 73, 251– 270.
Hurtgen, M. T., Halverson, G. P., Arthur, M. A. & Hoffman, P. F. Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blick, N.
2006. Sulfur cycling in the aftermath of a Neoproterozoic (Marinoan) 2006. The Ediacaran Period: a new addition to the geologic time
snowball glaciation: Evidence for a syn-glacial sulfidic deep ocean. scale. Lethaia, 39, 13 – 30.
Earth and Planetary Science Letters, 245, 551–570. Kulp, J. L., Turekian, K. & Boyd, D. W. 1952. Strontium content of
Jacobsen, S. B. & Kaufman, A. J. 1999. The Sr, C, and O limestones and fossils. Geological Society of America Bulletin, 63,
isotopic evolution of Neoproterozoic seawater. Chemical Geology, 701– 716.
161, 37 – 57. Le Guerroué, E., Allen, P. A., Cozzi, A., Etienne, J. L. & Fanning,
Jiang, G., Kaufman, A. J., Christie-Blick, N., Zhang, S. & Wu, H. M. 2006. 50 Myr recovery from the largest negative d13C excursion
2007. Carbon isotope variability across the Ediacaran Yangtzee plat- in the Ediacaran ocean. Terra Nova, 18, 147–153.
form in South China: Implications for a large surface-to-deep ocean Le Hir, G., Donnadieu, Y. et al. 2009. The snowball Earth aftermath:
gradient. Earth and Planetary Science Letters, 261, 303– 320. Exploring the limits of continental weathering processes. Earth and
Johnston, D. T., Schmitz, M. D. et al. 2005. Active microbial Planetary Science Letters, 277, 453– 463.
sulfur disproportionation in the Mesoproterozoic. Science, 310, Li, C., Love, G. D., Lyons, T. W., Fike, D. A., Sessions, A. L. & Chu, X.
1477– 1479. 2010. A stratified redox model for the Ediacaran Ocean. Science,
Johnston, D. T., Poulton, S. W., Dehler, C., Porter, S., Husson, J., 328, 80– 83.
Canfield, D. E. & Knoll, A. H. 2010. An emerging picture of Neo- Lyons, T. W. & Severmann, S. 2006. A critical look at iron paleoredox
proterozoic ocean chemistry: Insights from the Chuar Group, Grand proxies: New insights from modern euxinic marine environments.
Canyon, USA. Earth and Planetary Science Letters, 290, 64– 73. Geochimica et Cosmochimica Acta, 70, 5698– 5722.
Jones, D. S., Maloof, A. C., Hurtgen, M. T., Rainbird, R. H. & Macdonald, F. A., Jones, D. S. & Schrag, D. P. 2009a. Stratigraphic
Schrag, D. P. 2010. Regional and global chemostratigraphic corre- and tectonic implications of a newly discovered glacial diamictite-
lation of the early Neoproterozoic Shaler Sueprgroup, Victoria cap carbonate couplet in southwestern Mongolia. Geology, 37,
Island, Northwestern Canada. Precambrian Research, 181, 43 –63. 123– 126.
CHEMOSTRATIGRAPHY AND THE NEOPROTEROZOIC GLACIATIONS 65

Macdonald, F. A., McClelland, W. C., Schrag, D. P. & Macdonald, Pavlov, A. A. & Kasting, J. F. 2002. Mass-independent fractionation of
W. P. 2009b. Neoproterozoic glaciation on a carbonate platform sulfur isotopes in Archean sediments: strong evidence for an anoxic
margin in Arctic Alaska and the origin of the North Slope subterrane. Archean atmosphere. Astrobiology, 2, 27 –41.
Geological Society of America Bulletin, 121, 448–473. Paytan, A., Kastner, M., Campbell, D. & Thiemens, M. H. 1998.
Macdonald, F. A., Cohen, P. A., Dudás, F. O. & Schrag, D. P. 2010a. Sulfur isotopic composition of Cenozoic seawater sulfate. Science,
Early Neoproterozoic scale microfossils in the Lower Tindir Group of 282, 1459–1462.
Alaska and the Yukon Territory. Geology, 38, 143–146. Paytan, A., Kastner, M., Campbell, D. & Thiemens, M. H. 2004.
Macdonald, F. A., Schmitz, M. D. et al. 2010b. Calibrating the Cryo- Seawater sulfur isotopic variations in the Cretaceous. Science, 304,
genian. Science, 327, 1241– 1243. 1663–1665.
Macdonald, F. A., Strauss, J. V., Rose, C., Dudás, F. Ö. & Schrag, Pell, S. D., McKirdy, D. M., Jansyn, J. & Jenkins, R. J. F. 1993, Edia-
D. P. 2011. Stratigraphy of the Port Nolloth Group of Namibia and caran carbon isotope stratigraphy of South Australia — an initial
South Africa, and implications for the age of Neoproterozoic iron for- study. Transactions of the Royal Society of South Australia, 117,
mations. American Journal of Science, 310, 862– 888. 153– 161.
Magaritz, M., Holser, W. T. & Kirschvink, J. L. 1986. Carbon-isotope Pokrovskii, B. G., Melezhik, V. A. & Bujakaite, M. I. 2006. Carbon,
events across the Precambrian/Cambrian boundary on the Siberian oxygen, strontium, and sulfur isotopic compositions in late Precam-
Platform. Nature, 320, 258– 259. brian rocks of the Patom Complex, central Siberia: Communication
Maloof, A. C., Halverson, G. P., Kirschvink, J. L., Schrag, D. P., 1. Results, isotope stratigraphy, and dating problems. Lithology and
Weiss, B. P. & Hoffman, P. F. 2006. Combined paleomagnetic, Mineral Resources, 41, 450– 474.
isotopic, and stratigraphic evidence for true polar wander from the Poulton, S. W., Fralick, P. W. & Canfield, D. E. 2004. The transition
Neoproterozoic Akademikerbreen Group, Svalbard, Norway. Geo- to a sulphidic ocean 1.84 billion years ago. Nature, 431, 173– 177.
logical Society of America Bulletin, 118, 1099– 2014. Prave, A. R. 1999. Two diamictites, two cap carbonates, two d13C
Marenco, P. J., Corsetti, F. A., Kaufman, A. J. & Bottjer, D. J. excursions, two rifts: The Neoproterozoic Kingston Peak Formation,
2008a. Environmental and diagenetic variations in carbonate Death Valley, California. Geology, 27, 339– 342.
associated sulfate: An investigation of CAS in the Lower Triassic Prave, A. R., Fallick, A. E., Thomas, C. W. & Graham, C. M. 2009.
of the western USA. Geochimica et Cosmochimica Acta, 72, A composite C-isotope profile for the Neoproterozoic of Scotland
1570– 1582. and Ireland. Journal of the Geological Society, London, 166,
Marenco, P. J., Corsetti, F. A., Hammond, D. E., Kaufman, A. J. & 845– 857.
Bottjer, D. J. 2008b. Oxidation of pyrite during extraction of car- Rainbird, R. H., Jefferson, C. W. & Young, G. M. 1996. The early Neo-
bonate associated sulfate. Chemical Geology, 247, 124– 132. proterozoic sedimentary Succession B of northwestern Laurentia:
McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial correlations and paleogeographic significance. Geological Society
trinity: Neoproterozoic Earth history within the British –Irish Caledo- of America Bulletin, 108, 454–470.
nides. Geology, 34, 909–912. Raiswell, R., Buckley, F., Berner, R. A. & Anderson, T. F. 1988.
McFadden, K. A., Huang, J. et al. 2008. Pulsed oxidation and biological Degree of pyritisation of iron as a paleoenvironmental indicator of
evolution in Ediacaran Doushantuo Formation. Proceedings of the bottom-water oxygenation. Journal of Sedimentary Petrology, 58,
National Academy of Sciences (USA), 105, 3197– 3202. 812– 819.
McKirdy, D. M., Burgess, J. M. et al. 2001. A chemostratigraphic over- Rice, A. H. N., Edwards, M. B., Hansen, T. A., Arnaud, E. & Halver-
view of the late Cryogenian interglacial sequence in the Adelaide son, G. P. 2011. Glaciogenic rocks of the Neoproterozoic Smalfjord
Fold-Thrust Belt, South Australia. Precambrian Research, 106, and Mortensnes Formations, Vestertana Group, E. Finnmark,
149– 186. Norway. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G.
Melezhik, V. A., Gorokhov, I. M., Kuznetsov, A. B. & Fallick, A. E. (eds) The Geological Record of Neoproterozoic Glaciations. Geo-
2001. Chemostratigraphy of Neoproterozoic carbonates: implications logical Society, London, Memoirs, 36, 593–602.
for ‘blind dating’. Terra Nova, 13, 1 –11. Ries, J. B., Fike, D. A., Pratt, L. M., Lyons, T. W. & Grotzinger, J. P.
Melezhik, V. A., Pokrovsky, B. G., Fallick, A. E., Kuznetsov, A. B. 2009. Superheavy pyrite (d34Spyr . d34SCAS) in the terminal
& Bujakaite, M. I. 2009. Constraints on the 87Sr/86Sr of Late Proterozoic Nama Group, southern Namibia: a consequence of
Ediacaran seawater: insights from high-Sr limestones. Journal of low seawater sulfate at the dawn of animal life. Geology, 37,
the Geological Society, London, 166, 183–191. 743– 746.
Miller, N. R., Stern, R. J., Avigad, D., Beyth, M. & Schilman, B. Rooney, A. D., Selby, D., Houzait, J.-P. & Renne, P. R. 2010. Re– Os
2009. Cryogenian slate-carbonate sequences of the Tambien Group, geochronology of a Mesoproterozoic sedimentary succession, Taou-
Northern Ethiopia (I) ‘Pre-Sturtian’ chemostratigraphy and regional deni basin, Mauritania: Implications for basin-wide correlations and
correlations. Precambrian Research, 170, 129– 156. Re– Os organic-rich sediments systematic. Earth and Planetary
Misi, A. & Veizer, J. 1998. Neoproterozoic carbonate sequences of the Science Letters, 289, 486–496.
Una Group, Irece Basin, Brazil: chemostratigraphy, age and corre- Rose, C. V. & Maloof, A. C. 2010. Testing models for post-glacial “cap
lations. Precambrian Research, 89, 87 –100. dolostone” deposition: Nuccaleena Formation, South Australia. Earth
Misi, A., Kaufman, A. J., Azmy, K., Dardenne, M. A., Sial, A. N. & de and Planetary Science Letters, 296, 165–180.
Oliveira, T. F. 2011. Neoproterozoic successions of the São Fran- Rothman, D. H., Hayes, J. M. & Summons, R. E. 2003. Dynamics of the
cisco Craton, Brazil: The Bambuı́, Una, Vazante and Vaza Barris/ Neoproterozoic carbon cycle. Proceedings of the National Academy
Miaba groups and their glacigenic deposits. In: Arnaud, E., Halver- of Sciences (USA), 100, 124– 129.
son, G. P. & Shields-Zhou, G. (eds) The Geological Record of Sawaki, Y., Kawai, T. et al. 2010a. 87Sr/86Sr chemostratigraphy of Neo-
Neoproterozoic Glaciations. Geological Society, London, Memoirs, proterozoic Dalradian carbonates below the Port Askaig glaciogenic
36, 509– 522. Formation, Scotland. Precambrian Research, 179, 150–164.
Nagy, R. M., Porter, S. M., Dehler, C. M. & Shen, Y. 2009. Biotic Sawaki, Y., Ohno, T. et al. 2010b. The Ediacaran radiogenic Sr isotope
turnover driven by eutrophication before the Sturtian low-latitude excursion in the Doushantuo Formation in the Three Gorges area,
glaciation. Nature Geoscience, 2, 415– 418. South China. Precambrian Research, 176, 46 –64.
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V., Trin- Saylor, B. Z., Kaufman, A. J., Grotzinger, J. P. & Urban, F. 1998. A
dade, R. I. F. & Fairchild, T. R. 2007. Carbon and strontium isotope composite reference section for Terminal Proterozoic strata
fluctuations and paleoceanographic changes in the late Neoprotero- of southern Namibia. Journal of Sedimentary Research, 68,
zoic Araras carbonate platform, southern Amazon craton, Brazil. 1223–1235.
Chemical Geology, 237, 191– 210. Schrag, D. P., Berner, R. A., Hoffman, P. F. & Halverson, G. P. 2002.
Ono, S., Wing, B., Johnston, D., Farquhar, J. & Rumble, D. 2006. On the initiation of a snowball Earth. Geochemistry, Geophysics,
Mass-dependent fractionation of quadruple stable sulfur isotopes as Geosystems, 31, doi: 10.1029/2001GC000219.
a new tracer of sulfur biogeochemical cycles. Geochimica et Cosmo- Schidlowski, M., Eichmann, R. & Junge, C. E. 1975. Precambrian
chimica Acta, 70, 2238–2252. sedimentary carbonates: carbon and oxygen isotope geochemistry
66 G. P. HALVERSON & G. SHIELDS-ZHOU

and implications for the terrestrial oxygen budget. Precambrian Evolution: A focus on Southwestern Gondwana. Developments in
Research, 2, 1 –69. Precambrian Geology, Elsevier, Dordrecht, 16, 31– 69.
Schröder, S., Schreiber, C., Amthor, J. E. & Matter, A. 2004. Strauss, H. 1993. The sulfur isotopic record of Precambrian sulfates: new
Stratigraphy and environmental conditions of the terminal data and a critical evaluation of the existing record. Precambrian
Neoproterozoic – Cambrian Period in Oman: evidence from Research, 63, 225– 246.
sulphur isotopes. Journal of the Geological Society, London, 161, Swanson-Hysell, N. L., Rose, C. V., Calmet, C., Halverson, G. P.,
489– 499. Hurtgen, M. T. & Maloof, A. C. 2010. Cryogenian glaciation
Shen, B., Xiao, S., Zhou, C., Kaufman, A. J. & Yuan, X. 2010. Carbon and onset of carbon-isotope decoupling. Science, 328, 608–611.
and sulfur isotope chemostratigraphy of the Neoproterozoic Quanji Tucker, M. E. 1986. Carbon isotope excursions in Precambrian/
Group of the Chaidam Basin, NW China: Basin stratification in the Cambrian boundary beds. Nature, 319, 48 –50.
aftermath of an Ediacaran glaciation post-dating the Shuram event? Turner, E. C. 2009. Lithostratigraphy and stable isotope values of the
Precambrian Research, 177, 241– 252. early Neoproterozoic Gypsum Formation (Little Dal Group),
Shields, G. A. 1999. Working towards a new stratigraphic calibration Mackenzie Mountains Supergroup, NWT. NWT Open Report,
scheme for the Neoproterozoic –Cambrian. Eclogae Geologicae 2009-002.
Helvetiae, 92, 221– 233. Veizer, J., Compston, W., Clauer, N. & Schidlowski, M. 1983.
87
Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal Sr/86Sr in late Proterozoic carbonates: evidence for a
of existing models and the plumeworld hypothesis. Terra Nova, 17, ‘mantle’ event at 900 Ma. Geochimica et Cosmochimica Acta,
299– 310. 47, 295– 302.
Shields, G. A. 2007. A normalised seawater strontium isotope curve: Walter, M. R., Veevers, J. J., Calver, C. R., Gorjan, P. & Hill, A. C.
possible implications for Neoproterozoic –Cambrian weathering 2000. Dating the 840– 544 Ma Neoproterozoic interval by isotopes of
rates and further oxygenation of the Earth. Earth, 2, 35 – 42. strontium, carbon, and sulfur in seawater and some interpretive
Shields, G. & Veizer, J. 2002. Precambrian marine carbonate isotope models. Precambrian Research, 100, 371–433.
database: Version 1.1. Geochemistry, Geophysics, Geosystems, 3, Williams, G. E. 1979. Sedimentology, stable-isotope geochemistry and
doi: 10.1029/2001GC000266. palaeoenvironment of dolostones capping late Precambrian glacial
Shields, G., Stille, P., Brasier, M. D. & Atudorei, N.-V. 1997. Stra- sequences in Australia. Journal of the Geological Society of Austra-
tified oceans and oxygenation of the late Precambrian environments: lia, 26, 377–386.
a post glacial geochemical record from the Neoproterozoic. Terra Workman, R. K., Grotzinger, J. P. & Hart, S. R. 2002. Constraints on
Nova, 9, 218–222. Neoproterozoic ocean chemistry from C and B analyses of carbonates
Shields, G. A., Strauss, H., Howe, S. S. & Siegmund, H. 1999. Sulphur from the Witvlei and Nama groups, Namibia. In: Goldschmidt Con-
isotope compositions of sedimentary phosphorites from the basal ference Proceedings (Davos, Switzerland).
Cambrian of China — implications for Neoproterozoic – Cambrian Xiao, S., Bao, H. et al. 2004. The Neoproterozoic Quruqtagh Group in
biogeochemical cycling. Journal of the Geological Society, eastern Chinese Tianshan: evidence for a post-Marinoan glaciation.
London, 156, 943 –957. Precambrian Research, 130, 1 –26.
Shields, G., Kimura, H., Yang, J. & Gammon, P. 2004. Sulphur isotopic Xu, B., Xiao, S. et al. 2009. SHRIMP zircon U– Pb age constraints on
evolution of Neoproterozoic – Cambrian seawater: new francolite- Neoproterozoic Quruqtagh diamictites in NW China. Precambrian
bound sulphate S data and critical appraisal of the existing record. Research, 168, 247– 258.
Chemical Geology, 204, 163–182. Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003. Systematic
Sial, A. N., Dardenne, M. A. et al. 2010. The São Francisco Paleocon- variations in C, O, and Sr isotopes and elemental concentrations in
tinent. In: Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, Neoproterozoic carbonates in Namibia: implications for glacial to
H. (eds) Neoproterozoic– Cambrian Tectonics, Global Change and interglacial transition. Precambrian Research, 124, 69 –85.
Chapter 5

Chemical sediments associated with Neoproterozoic glaciation: iron formation,


cap carbonate, barite and phosphorite

PAUL F. HOFFMAN1,2 *, FRANCIS A. MACDONALD1 & GALEN P. HALVERSON3,4


1
Department of Earth and Planetary Sciences, Harvard University, 20 Oxford Street, Cambridge, MA, 02138, USA
2
School of Earth and Ocean Sciences, University of Victoria, Box 1700, Victoria, BC V6W 2Y2, Canada
3
School of Earth and Environmental Sciences, The University of Adelaide, North Terrace, Adelaide, SA 5005, Australia
4
Present address: Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montréal,
PQ H3A 2K6, Canada
*Corresponding author (e-mail: paulfhoffman@yahoo.com)

Abstract: Orthochemical sediments associated with Neoproterozoic glaciation have prominence beyond their volumetric proportions
because of the insights they provide on the nature of glaciation and the records they hold of the environment in which they were preci-
pitated. Synglacial Fe formations are mineralogically simple (haematite jaspilite), and their trace element spectra resemble modern sea-
water, with a weaker hydrothermal signature than Archaean– Palaeoproterozoic Fe formations. Lithofacies associations implicate
subglacial meltwater plumes as the agents of Fe(II) oxidation, and temporal oscillations in the plume flux as the cause of alternating
Fe- and Mn-oxide deposits. Most if not all Neoproterozoic examples belong to the older Cryogenian (Sturtian) glaciation. Older and
younger Cryogenian (Marinoan) cap carbonates are distinct. Only the younger have well-developed transgressive cap dolostones,
which were laid down during the rise in global mean sea level resulting from ice-sheet meltdown. Marinoan cap dolostones have a
suite of unusual sedimentary structures, indicating abnormal palaeoenvironmental conditions during their deposition. Assuming the melt-
down of ice-sheets was rapid, cap dolostones were deposited from surface waters dominated by buoyant glacial meltwater, within and
beneath which microbial activity probably catalysed dolomite nucleation. Former aragonite seafloor cement (crystal fans) found in
deeper water limestone above Marinoan cap dolostones indicates carbonate oversaturation at depth, implying extreme concentrations
of dissolved inorganic carbon. Barite is associated with a number of Marinoan cap dolostones, either as digitate seafloor cement associ-
ated with Fe-dolomite at the top of the cap dolostone, or as early diagenetic void-filling cement associated with tepee or tepee-like brec-
cias. Seafloor barite marks a redoxcline in the water column across which euxinic Ba-rich waters upwelled, causing simultaneous barite
titration and Fe(III) reduction. Phosphatic stromatolites, shrub-like structures and coated grains are associated with a glacioisostatically
induced exposure surface on a cap dolostone in the NE of the West African craton, but this appears to be a singular occurrence of phos-
phorite formed during a Neoproterozoic deglaciation.

The association of chemical sediments with Neoproterozoic Fe and Fe – Mn deposits


glacial deposits has long been known. In fact, Neoproterozoic
glaciation was discovered in a number of regions as a result of Distribution in time and space
the search for economic Fe and Fe–Mn deposits (e.g. SW
Brazil, NW Canada, Namibia). Glacial associated chemical sedi- The palaeogeographic distribution of synglacial Fe and Fe –Mn
ments provide critical evidence concerning the nature of Neopro- sedimentary deposits is shown in Figure 5.1 (Table 5.1). The econ-
terozoic glaciations and their aftermaths, in the form of omic Fe – Mn ores of the Jacadigo Group in the Urucum District of
geochemical records of the waters from which they were precipi- southwestern Brazil and eastern Bolivia (see below) were tenta-
tated, and indirectly the atmosphere with which those waters tively assigned to the terminal Cryogenian (Marinoan) glaciation
interacted. (Hoffman & Li 2009) through correlation with the Puga diamictite
Here, we briefly review the distribution, lithological associa- in the adjacent Paraguay fold belt (Alvarenga & Trompette 1992;
tion, sedimentology and palaeoenvironmental significance of Trompette 1994; Trompette et al. 1998). The Puga diamictite is
orthochemical sediments deposited during Neoproterozoic glacia- capped by a diagnostic Marinoan-type cap-carbonate sequence
tions and deglaciations. They include sedimentary Fe and Fe –Mn (Nogueira et al. 2003, 2007; Trindade et al. 2003; Font et al.
deposits and cap-carbonate sequences, of which ‘cap dolostones’ 2005, 2006; Alvarenga et al. 2008). However, no cap carbonate
sensu stricto form the basal transgressive systems tracts. Barite is preserved in the Jacadigo Group, which occupies a transverse
(BaSO4) and phosphorite mineralization occurs locally within rift-basin on the cratonic foreland of the Paraguay fold belt
Marinoan cap dolostones. More information on geologic setting, (Trompette et al. 1998). The true age of the Jacadigo Group, Stur-
stratigraphic relations, geochemistry and isotopic characteristics tian or Marinoan, is unknown. Similarly, Fe formation in the gla-
is given in the appropriate regional chapters. Palaeogeographies ciogenic Rizu Formation of central Iran was tentatively assigned to
of the deposits (Fig. 5.1) are based on global model maps for the Marinoan glaciation (Hoffman & Li 2009) on the basis of a
715 Ma (Sturtian) and 635 Ma (Marinoan), created by the Tec- reported cap dolostone (Kianian & Khakzad 2008), but details
tonics Special Research Centre in Perth, Western Australia (Li are lacking. The reassignment of the Jacadigo and Rizu Fe for-
et al. 2008; Hoffman & Li 2009). The maps were constructed on mations to the Sturtian glaciation (Fig. 5.1) is therefore permiss-
the basis of palaeomagnetic constraints, the mantle plume record ible, but arbitrary. In either palaeogeographic reconstruction
and palaeocontinental tectonic genealogy, factors that are largely (Fig. 5.1), the Fe- and Fe –Mn deposits formed disproportionately
independent of palaeoclimate. The Sturtian glaciation persisted within 308 of the palaeoequator, and at the margins of ocean basins
until 659 + 6 Ma in its type area (Fanning & Link 2008), by that were both internal (Jiangkou, Sturt, Rapitan, Surprise, Numees
which time the model palaeogeography (Li et al. 2008) was inter- and Chuos) and external (Rizu, Tany) to the fragmented Rodinia
mediate between 715 and 635 Ma (Fig. 5.1). supercontinent.

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 67– 80. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.5
68 P. F. HOFFMAN ET AL.

(a)

Nantuo

Blasskranz 635 Ma Namuskluft
Wildrose
Ghaub Vreeland
Petite Stelfox
Tereeken Khongoryn
Landrigan Ta Pod’em
Olympic Marnya
Blaini Ir Ka Dzemkukan
Au SC TM Si
In
Ma Laur CA
Elatina RP NC
Co
Cottons SF
Fiq Ar Fig. 5.1. Palaeogeographic maps for
Shareef Wilsonbreen 715 Ma and 635 Ma (Hoffman & Li 2009)
Bondo Am Storeelv showing distribution of (a) Sturtian and (b)
Superieure WA
Stralinchy-Reelan Marinoan glaciogenic deposits (open stars),
Kodjari Ba Smalfjord
Bakoye and synglacial Fe or Fe– Mn deposits (black
Jbeliat Palestina Puga stars). Palaeocontinents: Ar, Arabia; Am,
Amazonia; Au, Australia; Ba, Baltica; CA,
Chukotka–Arctic Alaska; Co, Congo; In,
(b) India; Ir; central Iran; Ka, Kalahari; Laur,
Jiangkou 715 Ma Numees
Surprise
Laurentia (including Sc, Scotland and Sv,
Blaubekker Svalbard); Ma, Mawson; NC, North China;
Pocatello SF; São Francisco; RP, Rio de la Plata; SC,
Areyonga Ta Toby South China; Si, Siberia; Ta, Tarim; TM,
Rapitan Tuva-Mongolia; WA, West Africa.
Si Tindir
Palaeocontinents constrained by
In SC
TM Maikhan Ul
Rizu Au Chivida penecontemporaneous palaeomagnetic data
Sturt CA NC Kharlukhtakh have heavy lines. The Sturtian glaciation
Ma persisted to c. 660 Ma in some areas (South
Julius River
Konnarock Ka Laur Australia and western USA), by which time
Sv the model palaeogeography was
Gubrah RP Hula Hula
Tambien Petrovbreen intermediate between 715 and 635 Ma.
Ayn
Ar Co Sc Assignment of Fe formations in the Jacadigo
Ulveso
Grand Port Askaig Group (Am) and Rizu Formation (Ir) to the
Chuos SF Am
Ba Tany Sturtian glaciation is uncertain. Correlation
Akwokwo Baykonur of Fe formation in the Gariep Belt (Ka) with
Inferieure WA
Jequitai Chiquerio Jacadigo the glaciogenic Numees Formation follows
Macdonald et al. (2011).

Lithological associations peri- or subglacial in origin based on the presence of faceted,


striated and/or preferentially oriented clasts. Sizeable Fe and
Impure to pure haematite –jaspilite occurs within intervals of Fe –Mn deposits occur in intervals that are sandwiched between
parallel-laminated argillite, ferruginous argillite and Mn-oxides. composite diamictite horizons, locally of great thickness (e.g.
Those identified in Figure 5.1 host outsize clasts (lonestones) of Braemar, Holowilena, Jakkalsberg, Sayunei, Tany). Smaller
intrabasinal and extrabasinal (crystalline basement) derivation, deposits are intimately interfingered with diamictite (e.g.
generally interpreted as ice-rafted debris. Many occur in contact Braemar, Chuos). Still other deposits lack diamictite but the
with massive to poorly stratified diamictite, inferred to be presence of dropstones of intra- and extrabasinal origin suggest a

Table 5.1. Cryogenian glacigenic Fe and Fe– Mn deposits

Palaeocontinent Location Host strata References

Amazonia Urucum Jacadigo Dorr (1945), Urban et al. (1992), Klein & Ladeira (2004)
Australia Braemar Yudnamutana Whitten (1970), Lottermoser & Ashley (2000)
Baltica Middle Urals Vil’va Chumakov (1992, 2007)
Congo Damara Belt Chuos Martin (1965a), Badenhorst (1988), Clifford (2008)
Kalahari Gariep Belt Numees Macdonald et al. (2011)
Laurentia NWT-Yukon Rapitan Young (1976), Yeo (1981, 1986), Klein & Beukes (1993)
Laurentia* Death Valley Surprise Corsetti & Kaufman (2003)
Lut (central Iran) Kerman Rizu Kianian & Khakzad (2008)
South China Yangtze platform Jiangkou Jiafu et al. (1987)
Tuva-Mongolia Erzin Ongoluk Ilyin (2009)

*In the collisional model for Laramide orogeny (Hildebrand 2009), the Death Valley area as well as most other Cryogenian glacial formations in the western USA and
British Columbia are considered to be exotic with respect to Laurentia before the Late Cretaceous.
CHEMICAL SEDIMENTS ASSOCIATED WITH NEOPROTEROZOIC GLACIATION 69

subaqueous proglacial setting (e.g. Urucum). Parallel-laminated of metres thick. Jaspilite first appears as 2-m-thick septa between
facies hosting haematite –jaspilite are commonly interpreted as individual diamictite bodies in the upper part of the lower
interglacial or interstadial, but they might alternatively represent complex (Mount Berg Formation). In a more distal section
maximum glacial stages, when outlet glaciers were blocked by (Hayhook Lake, SE Mackenzie Mountains), 14 m of ferruginous
thick multi-annual sea ice (Dowdeswell et al. 2000). argillite and haematite –jaspilite conformably overlie 650 m of
maroon-coloured siltstones with graded sandy beds and a sprink-
ling of small dropstones, some of which were redeposited as
Geochemical characteristics debrites. The jaspilite itself contains rounded boulders of quartz-
monzonite and is overlain disconformably by the upper diamictite
Compared with Archaean –Palaeoproterozoic banded Fe for- complex (Shezal Formation). The basal part of the otherwise
mations, Cryogenian Fe deposits are mineralogically simple, con- olive-coloured Shezal diamictite acquired its maroon colour and
sisting of haematite, chert (jasper) and minor carbonate. Fe numerous jaspilite clasts from the underlying Sayunei Formation.
contents range up to 50% Fe2O3. Organic contents are negligible Fe-isotope and Ce anomaly profiles suggest that the haematite –
and carbonates are moderately depleted in 13C, suggesting that jaspilite records subsidence of the basin floor across a redox-
organic matter originally present was respired with Fe(III) cline in the water column (Klein & Beukes 1993; Halverson
serving as electron acceptor. Normalized rare earth element et al. 2011).
profiles are distinct from Archaean–Palaeoproterozoic Fe for-
mations, being more strongly depleted in light rare earth elements, Jakkalsberg Member (Numees Formation, Port Nolloth Group), Gariep
and having much weaker positive Eu anomalies or none at all Belt, Namibia and South Africa. Haematite- and magnetite –jaspilite
(Klein & Beukes 1993; Graf et al. 1994; Klein & Ladeira 2004; with basement-derived lonestones make up the Jakkalsberg
Klein 2005). The Cryogenian profiles are more similar to those Member of the Numees Formation in thrust sheets intersected by
of modern seawater, and hydrothermal contributions are more the Orange River (Frimmel & von Veh 2003). The Numees For-
diluted than in more ancient Fe formations. mation has long been assumed to be the younger of two glacio-
genic horizons in the Gariep Belt, the older being the Kaigas
Formation (Frimmel 2008, 2011). Recently, a mid-Ediacaran age
Notable examples for the Numees glaciation has been advocated from carbonate
Pb/Pb dating (Fölling et al. 2000), 87Sr/86Sr ratios of 0.7082–
Jacadigo Group, Urucum District, Mato Grosso do Sul, Brazil. Eco- 0.7085 (Fölling & Frimmel 2002), correlations with dated South
nomic Fe–Mn oxide ores of the Jacadigo Group occur in strati- American strata (Frimmel 2004, 2008) and micropalaeontological
graphically isolated fault blocks in the Urucum District findings (Gaucher et al. 2005). The microfossils do not appear to
of southwestern Brazil and adjacent Bolivia (Dorr 1945; Walde be diagnostic, however, and the Sr isotope ratios (0.001
et al. 1981; Urban et al. 1992; Trompette et al. 1998). A higher than typical Marinoan values; Halverson et al. 2007) are
lower Urucum Formation, comprising conglomerate, sandstone potentially attributable to diagenesis, as are the Pb/Pb dates. Sub-
and black shale, is overlain by the Santa Cruz Formation, a sequent remapping of the least-deformed and least-allochthonous
sequence of Mn ore horizons, chiefly composed of cryptomelane sections prompted a reassessment of stratigraphic correlations
(K2Mn8O16), interspersed with ferruginous sandstone and up to between the autochthon and thrust sheets within the belt
270 m of haematite –jaspilite. All units of the Santa Cruz For- (Macdonald et al. 2011). The younger autochthonous diamictite,
mation carry dropstones of basement and subordinate carbonate previously correlated with the Numees Formation, is now pro-
lithologies. They are interpreted to be ice-rafted and the Jacadigo posed as an independent glaciogenic horizon, the Namuskluft
Group to have been deposited in a partly ice-covered fjord-like diamictite (Macdonald et al. 2011). It is capped by a pale dolostone
basin (Urban et al. 1992). The alternation of Mn and Fe ores is (Dreigratberg Member, previously conflated with the Bloeddrif
attributed to cycles of glacial advance and retreat. Enhanced Member, of the Holgat Formation) containing hallmark features
fluxes of oxygenated subglacial meltwater during glacial retreats of Marinoan cap dolostones –sheet-crack cements, tubestone
favoured Mn precipitation in ice-covered anoxic parts of the stromatolites and giant wave ripples (Hoffman & Macdonald
basin. Reduced fluxes of subglacial meltwater during glacial 2010). In contrast, the jaspilite-bearing Numees diamictite is
advances favoured Fe deposition (Urban et al. 1992). The Jacadigo overlain by a dark microbial limestone with roll-up structures,
Group lies in the cratonic foreland of southern Amazonia, close to characteristic of Sturtian cap carbonates (see below). Accordingly,
the Paraguay fold belt of Early Cambrian (Brasiliano) age within the Jakkalsberg jaspilite is probably older Cryogenian (Sturtian)
which glaciogenic diamictite of the Puga Formation (Alvarenga in age, and the Kaigas Formation may represent a pre-
& Trompette 1992) is sharply overlain by a typical Marinoan-type Sturtian (c. 0.74 Ga) diamictite of uncertain origin (Macdonald
cap dolostone (Nogueira et al. 2003, 2007; Trindade et al. 2003; et al. 2011).
Font et al. 2005, 2006; Alvarenga et al. 2008). No cap carbonate
is preserved on the Jacadigo Group, however, so the correlation Central Flinders Ranges (Holowilena) and Nackara arc (Braemar),
of the Jacadigo and Puga groups is uncertain, as is, therefore, the South Australia. Like the more ice-proximal (Iron Creek) section
age of the Urucum Fe– Mn deposits. in NW Canada, haematite – jaspilite and associated ferruginous
argillite in South Australia are sandwiched between a conformably
Rapitan Group, northern Canadian Cordillera. Haematite– jaspilite underlying diamictite complex (Pualco ‘Tillite’) and a discon-
and ferruginous argillite occur within glaciogenic diamictite of formably overlying complex (Wilyerpa Formation) dominated
the Rapitan Group and correlatives discontinuously for c. 800 km by diamictite (Whitten 1970; Preiss 1987; Lottermoser & Ashley
along the strike of the Ogilvie and Mackenzie mountains (Young 2000). In the Flinders Ranges, ferruginous argillite hosts sporadic
1976; Yeo 1981, 1986; Eisbacher 1985; Klein & Beukes 1993; lenses of boulders, some of which are grooved and striated,
Macdonald et al. 2011; Hoffman & Halverson 2011; Macdonald ferruginous diamictite, and rare lenses of haematite – jaspilite
& Cohen 2011). The glacial onset coincided with a major flood with rafted dropstones. In the more basinward Nackara arc,
basalt episode across Arctic Laurentia, bimodal representatives haematite –jaspilite is interrupted by thin units of ferruginous
of which underlie and intercalate the basal Rapitan diamictite in diamictite. Sorted sandstones occur within the diamictite com-
the Ogilvie Mountains, constraining its age at 717 Ma (Macdonald plexes in both areas. Correlation of the Holowilena and Braemar
et al. 2010). Haematite –jaspilite is up to 120 m thick in an ice- jaspilites is uncertain, but they and the bounding Pualco and
proximal section (Iron Creek, NW Mackenzie Mtns), where it is Wilyerpa diamictites are referred to the Sturtian glaciation
sandwiched between composite diamictite complexes hundreds (Preiss 1987).
70 P. F. HOFFMAN ET AL.

Cap-carbonate sequences Knoll 1995; Kennedy, 1996; Hoffman et al. 1998, 2007; James
et al. 2001; Kennedy et al. 2001; Higgins & Schrag 2003; Ridgwell
The continuous layers of carbonate that blanket Cryogenian gla- et al. 2003; Trindade et al. 2003; Shields 2005; Allen & Hoffman
ciogenic sequences or their equivalent disconformities are called 2005; Font et al. 2005, 2006; Hurtgen et al. 2006; Jiang et al. 2006;
‘cap’ carbonates. Clearly associated with syndeglacial flooding Bao et al. 2008; Le Hir et al. 2009; Hoffman 2011).
(Kennedy 1996; Bertrand-Sarfati et al. 1997; Hoffman et al.
2007), they occur on virtually every palaeocontinent and even in
siliciclastic-dominated successions (Table 5.2). Lithologically Marinoan-type (basal Ediacaran) cap-carbonate sequences
and isotopically, Sturtian and Marinoan cap carbonates are dis-
tinct, both from each other (Kennedy et al. 1998; Halverson & The extent and uniqueness of Marinoan cap dolostones was first
Shields 2011) and from most other Neoproterozoic carbonates appreciated in Australia (Dunn et al. 1971; Rankama 1973) and
(Hoffman 2011). Hoffman & Schrag (2002) proposed that deposi- the presence of barite in cap dolostones was first recognized in
tional sequences related to Neoproterozoic syndeglacial flooding West Africa (Deynoux & Trompette 1976). The smooth and
be called ‘cap-carbonate sequences’. ‘Cap dolostones’, sensu abrupt, yet conformable, contact between glaciogenic and related
stricto, are the transgressive tracts of cap-carbonate sequences detritus and cap dolostones is so distinctive and widespread that
and feature a suite of idiosyncratic sedimentary features (e.g. size- it was selected as the basis of the Global Stratotype Section and
graded peloids, sheet-crack cements, tubestone stromatolites, Point (GSSP) for the Ediacaran Period (Knoll et al. 2006), the
giant wave ripples, primary and early diagenetic barite). Trans- first Period boundary to be defined strictly on lithologic grounds.
gressive cap dolostones have rarely been documented in Sturtian A tuff at the top of the presumed correlative cap dolostone in
cap-carbonate sequences and are thin (,1.0 m) where present South China yields a 238U – 206Pb (IDTIMS) zircon date of
(Smith et al. 1994). Cap-carbonate sequences have aroused 635.2 + 0.4 Ma (Condon et al. 2005). A statistically indistin-
intense interest because they are unique to Proterozoic glaciations guishable 238U – 206Pb (SHRIMP) zircon date of 636.3 + 4.9 Ma
and record physical, chemical and biological conditions during and was obtained from a tuff near the base of the underlying Nantuo
immediately after global deglaciation (Aitken 1991; Grotzinger & glaciogenic diamictite (Zhang et al. 2008). The presumed

Table 5.2. Average thickness and idiosyncratic features of Marinoan cap dolostones in numerical order of abundance

Palaeocontinent Glaciation Cap dolostone (ref.) Metres LAC PEL GWR SCC TBS TPB DGB SFB
p p p p
Amazonia Puga Mirassol d’Oeste (1)† 24 – – – –
Arabia Gadir Manqil (Fiq) Hadash (2) 4.5 – – – – – – – –
p p p p
Arctic Alaska No deposits Nularvik (3)† 35 – – – –
p p p p p
Australia Olympic Mount Doreen (4) 4 – – –
p p
Australia Elatina Nuccaleena (5) 5 – – – – – –
p p p
Australia Cottons Cumberland Creek (6) 6 – – – – –
p p
Australia Landrigan Lower Stein (7) 8.5 – – – – – –
p p
Baltica Smalfjord Lower Nyborg (8) 5 – – – – – –
p p p p p
Congo Ghaub Keilberg (9)† 38 – – –
Congo Petit Conglomérat Calcaire rosé (10) 10
Congo Upper Tilloid C1 Dolomie rosé (11) 10
India Blaini Upper Blaini (12) 10
p p p p p
Kalahari Namuskluft Dreigratberg (13) 25 – – –
p
Kalahari Blässkranz Tsabesis (14) 21 – – – – – – –
p p p
Kalahari No deposits Bildah (15)† 80 – – – – –
p p p p p p
Laurentia Stelfox (Ice Brook) Ravensthroat (16)† 12 – –
p
Laurentia Storeelv Lower Canyon (17) 10 – – – – – – –
p p p
Laurentia Wilsonbreen Lower Dracoisen (18) 10 – – – – –
p p p
Laurentia Stralinchy-Reelan Cranford (19) 4 – – – – –
p
Laurentia Stelfox (Ice Brook) Hard Luck (20) 4 – – – – – – –
p p p p
Laurentia* Wildrose Noonday (21) 175 – – – –
p
Siberia Ulyakha Lower Ozerki (22) 35 – – – – – – –
p p p p
South China Nantuo Lower Doushantuo (23) 4 – – – –
p
Tarim Tereeken Lower Zhamoketi (24) 6 – – – – – – –
p p p p p p p
Tuva-Mongolia Khongoryn Ol (25)† 15 –
p p
Tuva-Mongolia Khesen Baxha (26) 4 – – – – – –
p p
West Africa Fersiga Oued Djouf (27) 6 – – – – – –
p p p
West Africa Jbéliat Amogjar (28) 5 – – – – –
p p p p p
West Africa Banboli Mid Sud-Banboli (29) 1.5 – – –

*In the collisional model for Laramide orogeny (Hildebrand 2009), Wildrose-Noonday strata and most other Cryogenian glacial formations in the western USA and British
Columbia are considered to be exotic with respect to Laurentia before the Late Cretaceous.

Former-aragonite seafloor cement above cap dolostone. LAC, low-angle cross-laminae; PEL, peloids; GWR, giant wave ripples; SCC, sheet-crack cements; TBS,
tubestone stromatolite; TPB, tepee breccia; DGB, diagenetic barite; SFB, seafloor barite.
References: (1) Nogueira et al. (2003); (2) Allen et al. (2004); (3) Macdonald et al. (2009); (4) Kennedy (1996); (5) Plummer (1978); (6) Calver & Walter (2000);
(7) Corkoron (2007); (8) Edwards (1984); (9) Hoffman et al. (2007); (10) Cahen & Lepersonne (1981); (11) Cahen (1950); (12) Kaufman et al. (2006); (13) Hoffman &
Macdonald (2010); (14) PFH observations; (15) Hegenberger (1993); Prave et al. (2011); (16) James et al. (2001); (17) Hambrey & Spencer (1987);
(18) Halverson et al. (2004); (19) McCay et al. (2006); (20) Macdonald & Cohen (2011); (21) Corsetti & Grotzinger (2005); (22) Sovetov & Komlev (2005);
(23) Jiang et al. (2006); (24) Xiao et al. (2004); (25) Macdonald (2011); (26) Macdonald & Jones (2011); (27) Bertrand-Sarfati et al. (1997); (28) Shields et al. (2007);
(29) Nédélec et al. (2007).
CHEMICAL SEDIMENTS ASSOCIATED WITH NEOPROTEROZOIC GLACIATION 71

correlative Ghaub diamictite in Namibia has an indistinguishable created during Cryogenian glaciations by net erosion, compaction
238
U – 206Pb (IDTIMS) zircon date of 635.6 + 0.5 Ma (Hoffmann and tectonic subsidence. Glacial palaeofjords, for example,
et al. 2004). Marinoan cap dolostones are therefore assumed to contain expanded cap-carbonate sequences compared with adja-
have been deposited at 635 Ma. cent uplands (Fig. 5.3). Similarly, areas undergoing rapid tectonic
subsidence like the Otavi platform, on which the Sturtian and
Distribution and thickness. The median and average thicknesses of Marinoan cap-carbonate sequences are each 250–400 m thick,
29 Marinoan cap dolostones on 15 palaeocontinents (Table 5.2) contrast with areas like the Taoudeni Basin of the West African
are 9 m and 18 m, respectively (Hoffman et al. 2007). As there craton (Deynoux 1985; Shields et al. 2007), where subsidence
are large facies-controlled variations in thickness in some areas, rates were low and cap-carbonate sequences are highly condensed,
these estimates are based on area-weighted average thicknesses, outside of incised palaeovalleys (Fig. 5.3). Condensation may also
where data permit. On the Otavi platform (Congo palaeocontinent) occur in fully marine settings where sedimentation rates were
of Namibia, for example, the Keilberg cap dolostone is 6– 10 m inadequate to fill the existing accommodation space. On the
thick on the distal foreslope, up to 100 m on the upper foreslope, distal foreslope of the Otavi platform, for example, drowning at
up to 75 m on the raised outer platform, and 20– 25 m on the dee- the terminal deglaciations was effectively permanent (Halverson
pened inner platform (Fig. 5.2). In general, cap dolostones are et al. 2005). The cause of condensation of the lower Doushantuo
thickest on palaeotopographic highs and thinnest in lows, converse Formation cap-carbonate sequence of South China (low sedimen-
to underlying glacial deposits. Observed inner-shelf deepening is tation rate or lack of accommodation?) depends on whether tepee-
characteristic of glaciated margins (Anderson 1999). If cap dolo- like breccias and associated early diagenetic barite mineralization
stones averaged 18 m in thickness over 20% of the present conti- in the cap dolostone are submarine (Jiang et al. 2006) or subaerial
nental surface area (including shelves), they would contain (Zhou et al. 2010) in origin, respectively.
c. 2.6  105 Pg (Gt) of C. This dwarfs the 3.8  104 Pg of dis-
solved inorganic C in the present ocean. According to the palaeo- Transgressive cap dolostones in expanded sequences. Cap dolostones
geography in Figure 5.1a, the maximum thickness of cap are typically pale pinkish (yellowish-grey weathering) dolomi-
dolostones increased from 5 m above 408 palaeolatitude, to 10 m crites, with palimpsest micro- to macro-peloidal textures indicat-
above 308, 25 m above 208, 38 m above 108 and 175 m at the ing deposition as silt and sand- and granule-sized aggregates.
palaeoequator (Hoffman & Li 2009). Those in expanded sequences lack the diagenetic overprint experi-
enced by those in condensed sequences, which we treat separately.
Expanded and condensed sequences. The decompacted thickness of The former (e.g. Cumberland Creek, lower Canyon, lower Dracoi-
a cap-carbonate sequence (as distinct from a cap dolostone), minus sen, Dreigratberg, Keilberg, Mirassol d’Oeste, Mount Doreen,
the effect of sediment loading, is a measure of the accommodation Nuccaleena, Ol, lower Pertatataka, Ravensthroat, Sentinel Peak,
created before and during its deposition. The lowering and raising lower Stein, Tsabesis, and lower Zhamoketi; see Table 5.2) are
of global mean sea level attending the growth and decay of ice invariably laminated, with each lamina representing a sedimen-
sheets creates no permanent accommodation if global ice-sheet tation unit composed of well-sorted, graded or reverse-graded
volume was the same before and after a glaciation. Permanent peloids, up to 3 mm in diameter, terminating with a micropeloidal
accommodation (i.e. outlasting isostatic adjustments) was drape (Aitken 1991; Kennedy 1996; Calver & Walter 2000; James

S Distal Slope Upper Slope Outer Platform Inner Platform N


5 km 5 km 100 km
80
LEGEND (e) (f) (g)
Maieberg Formation 20
20
δ C13
-5.0 0.0 60
pink marly limestone rhythmite
100 0
(d)
0 -5.0 δ13C 0.0
with sea-floor crystal fans P4017
40 -5.0 δ13C 0.0
G2008
Keilberg Member cap dolostone
80
marly micropeloidal dololutite turbidites
peloidal dolarenite, giant wave ripples 20
as above, swaley low-angle crossbedding
60 LEGEND
tubestone stromatolitic dolostone
peloidal dolarenite, sheet-crack cements 0 δ 13 C carb sediment
marly micropeloidal dolostone turbidites -5.0 δ13C 0.0
40 P6005 δ 13 C carb sheet-crack cement
10
-5.0 0.0 (c)
-5.0 0.0 10
10 Bethanis Member of the Ghaub Formation
(b) 20
(a) stratified proglacial detrital carbonate with ice-rafted debris
5
5 Ombaatjie Formation cycle b7-b8
5
0 peritial cycles of dolostone ribbonite, stromatolite, grainstone
0 P7016 and microbialaminite)
0 -5.0 δ13C 0.0
0 P7009
δ13C
δ13C P7002
P1607

Fig. 5.2. Representative lithologic columns and d13C profiles of the Keilberg cap dolostone (Marinoan) in contiguous palaeobathymetric zones of the Otavi
platform, northern Namibia (Hoffman et al. 2007). The datum for platform columns (e– g) is the inflection point (arrows) in isotopic profiles. Each column records part of
an overall sigmoidal profile, implying diachronous deposition during progressive marine inundation. The systematic nature of profiles rules out a detrital origin for
dolomite and a diagenetic origin for d13C values. Note that values for isopachous dolomite sheet-crack cement (crosses) in (b) and (c) are indistinguishable from host
dolopelmicrite (dots).
72 P. F. HOFFMAN ET AL.

(m) SW measured sections ( ) NE † tubestone (geoplumb) stromatolites (Cloud et al. 1974; Hegen-
berger 1993; Corsetti & Grotzinger 2005) – metre- to
decametre-scale mounds within which arched microbial
growth laminae are interrupted by tube-like structures, invari-
0
ably oriented palaeovertically, defined by parallel-laminated
dolomicrite with meniscus-like curvature, variably replaced
by late-stage void-filling cement;
† giant wave ripples (Allen & Hoffman 2005; Hoffman & Li
–50 2009) – trochoidal megaripples with high aspect ratios, aggra-
dational development, and crestward-coarsening bidirectional
laminae that interdigitate in the crestal region.
Interpretations of these structures differ. Kennedy et al. (2001)
–100
relate sheet-crack cements to permafrost clathrate destabili-
0 5 10 15 km zation by marine flooding, but d13C values of the isopachous
Teniagouri Group
dolospar cement are not extraordinarily depleted (Fig. 5.2), nor
Atar green and red shale, siltstone is d34S of cap dolostone enriched as predicted (Shields 2005).
West
African
aquamarine chert (silexite), shale Corkeron (2007) attributes sheet-crack cements to pore-fluid over-
craton 6 violet and green limestone pressure associated with differential shale-carbonate compaction.
Jbeliat Group Hoffman & Macdonald (2010) suggest that pore-fluid overpres-
5 dolomitic quartz sandstone, granulestone
4 reddish siltstone grading to fine sandstone CCS
sures signal rapid falls in regional sea level associated with the
3 cap dolostone, tepee breccia, barite crusts disappearance of ice sheets and the loss of their gravitational
2 pebbly regolith, sand-wedge polygons ‘pull’ on adjacent ocean waters. Cloud et al. (1974) interpret the
1 terrestrial lodgement tillite and outwash tubular structure in stromatolites as fluid- or gas-escape channels
Mesoproterozoic Atar, Tifounke, etc. groups on account of their geoplumb orientation. Corsetti & Grotzinger
microbial limestone, dolostone; fine clastics
(2005) interpret them as microbial growth structures. Hoffman
(2011) attempts to estimate the inclination of the upper foreslope
Fig. 5.3. Stratigraphic cross-section of the glaciogenic Jbéliat Group
of the Otavi platform from the mean dip of stromatolitic layering
(Marinoan?) based on sections measured along Atar Cliff, Mauritania, by
after restoring the contained tubes to vertical. Giant wave ripples
present authors PFH and GPH and Adam C. Maloof (Hoffman & Schrag 2002).
The cap-carbonate sequence (CCS) includes cap dolostone (unit 3) and
were first interpreted as supratidal tepee structures (Eisbacher
regressive highstand deposits (units 4 and 5) preserved only in palaeovalleys.
1985), but their crestlines are straight and parallel, not polygonal
The datum is the top of the cap-carbonate sequence. Silicified shale (silexite) like true tepees, and they lack the breccias and void-filling
belongs to the succeeding Téniagouri depositional sequence and need not be cements diagnostic of supratidal tepees (Kendall & Warren
genetically related to glaciation. 1987). Gammon et al. (2005) describe a tepee-like structure in
the Nuccaleena cap dolostone and relate it to a growth fault, but
giant wave ripples in the same formation lack faults and the one
et al. 2001; Halverson et al. 2004; Xiao et al. 2004; Font et al. described by Gammon et al. (2005), which is oriented perpendicu-
2006; Nédélec et al. 2007). Locally, cap dolostones begin with a lar to bedding, could not have had a slip vector that was parallel or
coarsening-upward interval of discrete, parallel-sided, normally subparallel to the outcrop surface (i.e. the plane of their two-
graded, sedimentation units interpreted as turbidites, separated dimensional kinematic analysis), compromising the growth fault
by argillaceous partings (Lithofacies I of Kennedy 1996; interpretation. Allen & Hoffman (2005) interpret them as unu-
Hoffman & Macdonald 2010). The turbidites, where present, sually large, steep and aggradational wave ripples, related to the
grade upward into laminated but less discretely bedded dolostone action of long-period waves in the lower part of the ocean mixed
with ubiquitous low-angle crossbedding, including toplaps, onlaps layer, driven by strong sustained winds (not hurricanes).
and downlaps (Lithofacies II of Kennedy 1996). Kennedy (1996) Hoffman & Li (2009) find that crestal orientations globally had
attributes the low-angle crossbedding to asymmetric synsedimen- meridional (north –south) mean orientations and that zonal
tary dissolution and/or semi-plastic slumping of originally (east – west) orientations were absent, consistent with zonal and
parallel-sided strata deposited below the storm wave base. A not with cyclonic winds.
simpler and more conventional interpretation is that low-angle Sorted peloids, low-angle crossbedding, stromatolites and giant
crossbedding records wave action above the storm wave base. wave ripples signify that cap dolostones were deposited above the
This interpretation is consistent with the sorted macropeloids and storm wave base and at least partly in the euphotic zone. The
implies initial shoaling subsequent to the underlying turbidites occurrence of shallow-water structures over a large range of
(Hoffman & Macdonald 2010). Intertidal– supratidal indicators palaeodepths implies diachronous deposition attending a large
(e.g. fenestral texture or other desiccation features, abraded rise in sea level, consistent with the fragmented d13C records of
intraclasts, beach ‘rosettes’, channels and levees, polygonal cap dolostones from different palaeodepths (Fig. 5.3). The esti-
tepees and related breccias) are absent (Kennedy 1996). The low- mated magnitude of the rise of more than a kilometre implicates
angle cross-bedded lithofacies is dominant and within it occurs global ice-sheet melting (so-called ‘glacioeustasy’) and links the
a triad of highly idiosyncratic structures in a broadly consistent timescales of deglaciation and cap-dolostone sedimentation
vertical sequence, although all three are rarely present in a single (Kennedy 1996; Bertrand-Sarfati et al. 1997; James et al. 2001;
section (Fig. 5.2). The structures are described in more detail Hoffman et al. 2007). Cap dolostones are locally postglacial, but
in Hoffman (2011, and references therein) and are, in order of globally syndeglacial.
appearance:
Seafloor carbonate cement. In expanded sequences, cap dolostones
† sheet-crack cements (Kennedy 1996; Corkeron 2007; Hoffman are overlain conformably by deeper water limestone, limestone
& Macdonald 2010) – bedding-parallel, variably buckled, with dolostone turbidites, marlstone or shale. Seafloor cement, in
fibrous isopachous, void-filling dolospar, confined to a continu- the form of crystal fans of pseudomorphosed prismatic aragonite,
ous metre-thick zone of variable intensity near the base of the built masses up to 100 m thick, localized by sea-bottom topogra-
cap dolostone in distal slope sections, typically just above phy (Hoffman 2011). In magnitude, they dwarf other Neoprotero-
basal turbidites; zoic seafloor cements and differ from volumetrically comparable
CHEMICAL SEDIMENTS ASSOCIATED WITH NEOPROTEROZOIC GLACIATION 73

Archaean –Palaeoproterozoic examples in having formed in the ne Cille, Luoquan, Hankalchough and Croles Hill diamictites).
presence of micritic sedimentation (Sumner 2002). Up to 40 cm of argillaceous limestone overlies glacial marine dia-
mictite in two sections of the short-lived, 582 Ma Gaskiers For-
Highstand deposits of cap-carbonate sequences. Regressive high- mation on the Avalon Peninsula of eastern Newfoundland,
stand deposits in expanded cap-carbonate sequences may be car- Canada (Myrow & Kaufman 1999). This cap carbonate is signifi-
bonate or clastic-dominated. The former begin with limestone cant as the sole carbonate bed in the nearly 15-km-thick host suc-
rhythmite and shoal to dolostone grainstone beneath a well- cession. It lithologically and isotopically resembles the base of
developed subaerial sequence boundary. Grain-sized equivalent many Sturtian cap-carbonate sequences, although the amount of
terrigenous deposits (shales, siltstones and sandstones) conform- carbonate deposited overall is far smaller. Finally, the Egan
ably overlie cap dolostones in many areas (e.g. in the Brachina diamictite represents a short-lived glacial incursion onto a
Formation and the ABC Range Quartzite in South Australia). In mid-Ediacaran carbonate platform in the Kimberley region of
the cratonic Taoudeni Basin of West Africa, highstand deposits Western Australia (Corkeron & George 2001). Diamictite and
are found in palaeovalleys but are missing on palaeotopographic overlying conglomerate are overlain by c.15 m of dolostone with
highs (Fig. 5.3). This has led to misinterpretation. Silicified herringbone and trough crossbedding, and an additional c.15 m
shale (silexite) of the Téniagouri Group (Fig. 5.3) and correlatives of interbedded quartz-arenite and silty dolostone with a distinctive
rest directly on the altered surface of the cap dolostone over stromatolite (Tungussia julia) horizon near the top. The carbonate
vast areas of the craton, leading to the concept of a ‘triad’: sequence following the Egan glaciation resembles no other
tillite –carbonate–silexite. We concur with Bertrand-Sarfati cap-carbonate sequence, including the one on the Landrigan dia-
et al. (1997), in that this concept should be abandoned. mictite in the same succession, stratigraphically well below the
Deposition of the Téniagouri Group (and correlative Azlaf Egan, which has a recognizable Marinoan-type cap dolostone
Group, Deynoux et al. 2006) cannot be related to the glacio- (Corkeron 2007).
eustatic transgression and could be much younger (Bertrand-
Sarfati et al. 1997).
Barite in Marinoan-type (basal Ediacaran) cap dolostones
Transgressive cap dolostones in condensed sequences. Where accom-
modation was lacking, cap dolostones aggraded or prograded to Barite (BaSO4) is a major constituent of certain Marinoan cap
the supratidal zone, where they were subjected to intense meteoric dolostones (Table 5.2). The mineral is easily recognizable in the
and vadose diagenesis (Fig. 5.3). In such cases, primary marine field from its vitreous pearly lustre, bladed crystal habit and high
dolopelmicrite was repeatedly fractured and re-cemented in specific gravity (4.3 – 5.0). Even where pseudomorphosed by
tepee structures and related breccias (Kendall & Warren 1987). calcite, its bladed crystals are distinct from aragonite, which is aci-
Barite is a major void-filling cement in tepee brecciated cap dolos- cular (needlelike) in habit, forming pseudohexagonal prisms due to
tone on the Yangtze platform of South China (Jiang et al. 2006) polysynthetic twinning. The palaeoenvironmental significance of
and the West African craton (Shields et al. 2007). Where cap barite rests with its highly redox-sensitive solubility in S-bearing
dolostones were subaerially exposed (Fig. 5.3), highstand deposits aqueous solutions: the concentration of Ba in modern (oxic) sea-
are typically absent. water is only 1.4 ppb by atom, compared with 51 ppm in crustal
rocks from which Ba in seawater is derived. Consequently,
barite leached by anoxic pore waters from deep-sea carbonate
Sturtian-type (Cryogenian) cap-carbonate sequences ooze reprecipitates cumulatively at the sub-seafloor redox front
as it migrates through the sediment column in response to sediment
The contacts between Sturtian glaciogenic deposits and cap car- accumulation. However, it is doubtful that this process alone could
bonates, like Marinoan ones, are characteristically singular, provide a satisfactory explanation for either barite type in
smooth, abrupt and conformable. Sturtian cap carbonates are litho- Marinoan cap dolostones, given their limited thickness.
logically distinct: where Marinoan cap dolostones are pale in
colour (,0.1 wt% total organic carbon, TOC), arenaceous in Two types of barite in cap dolostones: primary and early diagene-
texture and display wave-generated structures. Sturtian cap- tic. Two types of barite have been described from Marinoan cap
carbonate sequences typically begin with fetid, dark grey, parallel dolostones (Table 5.2): primary barite in the form of seafloor
laminated, micritic limestone (rarely rhodochrosite, MnCO3), cement (Kennedy 1996; Hoffman & Halverson 2011) and early
with or without parallel-sided graded beds (turbidites) and debrites diagenetic barite associated with tepee and tepee-like breccias
(Tojo et al. 2007). In Namibia, the Sturtian cap-carbonate sequence (Jiang et al. 2006; Shields et al. 2007). Seafloor barite forms at a
continues with up to 200 m of continuous sublittoral dolostone laterally continuous horizon in the top few centimetres of cap
microbialaminite, featuring microbial rollup structures, neptunian dolostones. Tiny barite crystals self-assemble into macroscopic
dykes and inclined zones of high primary porosity overlain by crystal fans or digitate aggregates. Interaction between barite
humped microbialaminite (Pruss et al. 2010). The sublittoral growth and particulate sedimentation forms the basis for determin-
microbialaminite extravaganza eventually grades up into cross- ing a seafloor origin, denoting precipitation of barite from the
bedded grainstone, ending at a tepee brecciated sequence boundary. ambient water column. Early diagenetic barite forms void-filling
Compared with Marinoan cap-carbonate sequences, where isopachous crustose cements within breccias associated with
transgressive tracts (i.e. cap dolostones) are well developed, tepee-like structures. The cause of tepee formation and brecciation
shallow-water transgressive tracts are thin (Smith et al. 1994) or is attributed to submarine methane venting (Jiang et al. 2006), or
absent in Sturtian cap-carbonate sequences. Typically, deepwater alternatively to evaporative pumping due to subaerial (supratidal)
deposits directly overlie beveled supraglacial surfaces. The sim- exposure in zones of marine – meteoric groundwater mixing
plest explanation for the difference is a below-critical saturation (Shields et al. 2007). The second alternative is the conventional
state with respect to carbonate in the surface ocean during Sturtian interpretation of tepee structures and associated breccias (Assereto
deglaciation (Hoffman & Schrag 2002). & Kendall 1977; Kendall & Warren 1987).

Seafloor barite in central Australia. In the Amadeus and Ngalia


Mid-Ediacaran cap carbonate sequences basins of central Australia, seafloor barite is associated with ferru-
ginous domal stromatolite at the top of cap dolostones overlying
Most mid-Ediacaran glaciogenic horizons lack cap carbonates Marinoan glaciogenic deposits of the Olympic and Mount
(e.g. Serra Azul, Squantum, Mortensnes, Moelv, Vil’chitsi, Loch Doreen formations, respectively (Kennedy 1996). Barite occurs
74 P. F. HOFFMAN ET AL.

either as isolated bladed crystals (1 –2 mm) or as large upward- Cryogenian (Marinoan) glaciogenic deposits of the Nantuo For-
oriented rosettes of bladed crystal (up to 4 cm high). The crystals mation on the Yangtze platform (Jiang et al. 2006). Void-filling
‘grew at the sediment –water interface, as indicated by the barite fans occur in the lower part (unit C1) of the cap dolostone
upward-oriented growth habit, sediment drape defining crystal ter- on the inner platform and on the slope and basin to the SE. Unit 1
minations, and presence of crystal fragments as detrital material is brecciated everywhere and hosts a variety of sheet-crack
within overlying sediment’ (Kennedy 1996). In the Amadeus cements, cement-filled stromatactis-like cavities, and tepee-like
Basin, the barite-rich horizon is conformably overlain by reddish- structures (Jiang et al. 2006). Importantly, the base of unit C1 is a
grey laminated siltstone with limestone turbidites (Pertatataka sharp, smooth, undulating surface; neither it nor the upper
Formation); in the Ngalia Basin it is overlain by red laminated Nantuo Formation experienced the intense brecciation typically
siltstone (Red Shale Member of the Mount Doreen Formation). observed within unit C1 (Jiang et al. 2006). In addition, possible
In both cases, in our view, the barite horizon marks a conformable seafloor barite is described as ‘layer-parallel barite fans growing
transition from deposition under the influence of storm waves (i.e. out of the dolomicrite substrate’ in the lower part of unit C2 in
size-sorted peloids, low-angle crossbedding) to sedimentation the basin (Jiang et al. 2006). Tepee-like structures occur locally
below the storm wave base. Walter & Bauld (1983) suggested in the lower part of unit C2 on the platform. Unit 3 and the upper
that the barite was secondary after primary anhydrite (CaSO4), con- part of unit C2 are not brecciated, consisting of parallel- and
tingent on their interpretation that the carbonate and sulphate were small-scale cross-laminated peloidal dolopackstone with graded
products of intense evaporation in glacial lakes such as those of the layers. Unit 3 has increased silt and limestone (Jiang et al. 2006).
Antarctic Dry Valleys. This is not a credible scenario in an open
marine setting in the absence of a local hydrothermal source of Ca. Early diagenetic barite on the Dzabkhan Platform, western Mongo-
lia. A characteristic Marinoan cap dolostone (Table 5.2) overlies
Seafloor barite in northwestern Canada. In the Mackenzie Moun- the younger of two Cryogenian glaciogenic horizons (Khongoryn
tains of northwestern Canada, the top 4– 10 cm of the Ravensthroat diamictite) on the Dzabkhan platform of western Mongolia (Mac-
cap dolostone contains digitate barite cement (variably calcitized) donald et al. 2009, 2011). Barite occurs in the mountain pass (Hoh
over a strike length of nearly 200 km (Hoffman & Halverson Davaa) between Bayan-Uul and Jargalan in northern Govi-Altay.
2011). The barite coincides with a change in colour of the peloidal There the cap dolostone includes two units, separated by a sharp
dolostone from pale cream to chocolate brown. Locally, the smooth disconformity veneered by dark brown Fe-dolostone.
barite-rich layer is developed on a train of giant wave ripples. The first unit (11.1 m) is a pale tan peloidal dolostone with low-
The barite first appears as randomly oriented, millimetre-scale, angle crossbedding. Isopachous sheet-crack cements occur near
bladed crystals, which appear to have formed just below the its base, above the basal 0.7 m of marly brown ribbon beds. The
sediment –water interface. Upwards, the crystals self-organize to second unit (8.4 m) consists of variably brecciated medium-grey
form centimetre-scale digitate structures, commonly overstee- dolostone with thick crusts of isopachous void-filling barite
pened toward the SW (seaward). The digitate structures display cement. Toward the top, sheet-crack and stratiform barites form
coral-like growth laminae, defined by films of opaque minerals. domal structures with up to 1.0 m of relief. The barite domes
Where calcitized, they resemble microdigitate stromatolites were ultimately onlapped and buried by a third unit consisting of
except for their diagnostic bladed crystal habit (so-called barite unbrecciated medium-dark grey dolostone ribbons with subaqu-
‘rosettes’). The flanks of the digitate structures are ragged due to eous microbial textured intervals, including rollup structures,
outward growth of barite after each lamina of peloidal Fe-dolomite that lead into the scree-covered maximum flooding horizon of
was deposited between the digits. The interplay between crystal the Ol cap-carbonate sequence. Although onlap relations clearly
growth and burial by peloids demonstrates that the barite precipi- show that the domal barite was exposed to seawater at the top of
tated from the water column, into which the digitate structures the second unit, proof that it formed on the seafloor in the form
projected up to 1.0 cm above the seafloor. The Fe-dolostone of detailed interaction between sedimentation and precipitation
layer with barite cement makes a conformable contact with over- has yet to be observed.
lying micritic limestone (Hayhook Formation) – epigenetically
dolomitized locally – containing acicular crystal fans of pseudo-
morphosed aragonite cement and reworked detrital dolomite Phosphorite in cap-carbonate sequences
basally (James et al. 2001).
Taoudeni Basin, West Africa
Early diagenetic barite in the Taoudeni Basin, West Africa. The
common occurrence of barite in cap dolostones terminating the On the northeastern margin of the Taoudeni Basin, diamictite of
Jbéliat glaciation of the West African craton has been known for the Jbéliat glaciation (Fersiga Formation) is overlain by a cap-
half a century (Deynoux & Trompette 1981). The type area is carbonate sequence correlative with that found in palaeovalleys
the Atar Cliff, Mauritania, on the northern edge of the Taoudeni on the Atar Cliffs (Fig. 5.3), 1250 km to the SW (Bertrand-Sarfati
Basin (Fig. 5.3). There, the cap dolostone directly overlies a peri- et al. 1997; Deynoux et al. 2006). A thin but continuous cap dolos-
glacial permafrost regolith with deep polygonal sand wedges, tone (Oued Djouf Formation) is overlain by a regressive highstand
developed above a terrestrial sub- and proglacial succession of sequence (Grizim Formation) up to 80 m thick, composed of glau-
northerly derivation and vast extent (Deynoux 1982, 1985; conitic green shale, siltstone and sandstone, ultimately aeolian
Deynoux et al. 2006). The cap dolostone is variably brecciated (Bertrand-Sarfati et al. 1997). Apatitic phosphorite is concentrated
and spectacular metre-scale tepee structures (Kendall & Warren at the base of the Grizim Formation, forming stromatolitic domes,
1987) are developed toward the top near Amogjar, east of Atar. microdigitate clusters and shrub-like colonies over moraines of
Isopachous barite forms thick crusts around blocks of brecciated Fersiga diamictite, where the cap dolostone is brecciated due to
dolostone, as well as secondary (remobilized?) vein-fillings subaerial exposure, and crossbedded phospharenite composed of
(Shields et al. 2007). An early diagenetic origin for the cavity- phosphate-coated grains and oncolites in the inter-moraine
filling, crustose barite is indicated by detrital barite clasts in depressions (Bertrand-Sarfati et al. 1997). Authigenic glauconite
coarse-grained quartz sand- and granulestone of the cap-carbonate occurs in both phosphorite facies and void-filling barite is a
sequence highstand tract (Fig. 5.3). minor constituent of the shrub-like phosphate colonies. The phos-
phorites formed rapidly, during the initial stages of the glacio-
Early diagenetic barite on the Yangtze Platform, South China. A eustatic transgression (Bertrand-Sarfati et al. 1997).
relatively thin (2.5 –4.0 m) cap dolostone (lower Doushantuo Phosphorites of early Ediacaran age occur in other areas,
Formation units C1 –3 of Jiang et al. 2006) blankets terminal but their connection to glaciation is tenuous. Khodjari-type
CHEMICAL SEDIMENTS ASSOCIATED WITH NEOPROTEROZOIC GLACIATION 75

phosphorite in the northern Volta basin was tentatively linked to the temperate Quaternary ice sheets, simplified Neoproterozoic
the terminal Cryogenian deglaciation of West Africa (Trompette deglacial simulations, and the apparent absence of polar continents
et al. 1980), but the phosphorite is younger than regionally exten- in Cryogenian palaeogeographic reconstructions (Fig. 5.1) all
sive silicified green argillite (‘silexite’), which itself disconform- suggest that Sturtian and Marinoan ice sheets disappeared
ably overlies the basal Ediacaran (Marinoan) cap-carbonate rapidly. How rapidly? If those middle and low-latitude ice sheets
sequence in the Taoudeni basin (Fig. 5.3; Bertrand-Sarfati et al. disappeared at the rate most temperate mountain glaciers have
1997; Shields et al. 2007). Similarly, economic phosphorite in been lowered in recent decades (1.1 m/a), their demise would
the Khubsugul area of northern Mongolia (Ilyin et al. 1986) was have taken 2 ka assuming their average thickness was compar-
genetically related to the underlying terminal Cryogenian able with the present East Antarctic Ice Sheet (Lythe et al.
Khesen diamictite (Sheldon 1984). However, Ilyin (2009) and 2001). This is close to the pan-deglacial timescale of ,2 ka in a
recent mapping has shown that the phosphorite is separated from climate model (Hyde et al. 2000). Positive feedbacks – ice-albedo,
post-glacial Baxha cap-carbonate sequence by a major hiatus ice-elevation and various greenhouse-gas feedbacks notably water
(Macdonald & Jones 2011). vapour feedback – contribute to rapid deglaciation. The energy
demands for such a deglaciation, amounting to a globally averaged
energy flux of c. 11 W m2 (Wallace & Hobbs 1977, p. 320), are
Genesis and significance of glacial-associated chemical small compared to the Neoproterozoic solar irradiance
sediments of c. 362 W m2 (c. 94% of present), a high surface albedo notwith-
standing. Pending detailed modelling of a global deglaciation, we
Fe and Fe – Mn oxide deposits conservatively take 8 ka, the timescale for Quaternary deglacia-
tions, as an upper limit for the Marinoan deglaciation. Given
Ferrous v. euxinic anoxia. Anoxia due to ice cover has long been median (9 m) and maximum (175 m) thicknesses of Marinoan
invoked as a means of mobilizing reduced Fe and Mn in solution cap dolostones globally (Hoffman et al. 2007), average accumu-
as a source for oxide ores within Cryogenian glaciogenic lation rates were then c. 1 and 22 mm/a for cap dolostones of
sequences (Martin 1965b; Urban et al. 1992; Klein & Beukes typical and extreme thickness, respectively. If deglaciation
1993; Lottermoser & Ashley 2000; Klein & Ladeira 2004). occurred in 2 ka, the respective accumulation rates were four
Fe(II) concentration in the presence of H2S is limited by pyrite times larger. These must be considered minimum rates, as cap
(FeS2) saturation. High degrees of continental ice cover and conse- dolostones at any given location represent only a fraction of the
quent large sea-level falls favour Fe over S in ocean waters because total deglaciation (Fig. 5.2). We consider it likely that unusual
of reduced riverine sulphate supply and higher Fe/S in deep-sea structures in cap dolostones (e.g. sheet-crack cements, tubestone
hydrothermal vent fluids, respectively (Canfield & Raiswell stromatolites and highly aggradational wave ripples) may be fun-
1999; Kump & Seyfried 2005). damentally related to extreme rates of accumulation.
The record of geomagnetic reversals and excursions in the
Subglacial sulphate-rich ferrous waters. The discovery of modern Mirassol d’Oeste and Nuccaleena cap dolostones creates a
Fe- and sulphate-rich (3.45 and 50 mM, respectively) waters dilemma (Trindade et al. 2003; Raub 2008). Conventional esti-
trapped beneath the Taylor Glacier in the McMurdo Dry Valleys mates of their duration and frequency (Johnson et al. 1995;
of East Antarctica gives a new perspective on Cryogenian jaspilites Gubbins 1999; Roberts 2008) imply a timescale on the order of
(Mikucki et al. 2009). Fe is not titrated as pyrite because there is 0.1– 1.0 Ma for those cap dolostones, implying that accumulation
insufficient organic primary production, due to the glacier cover, rates were at least 10 –100 times slower than the ‘conservative’
to fuel bacterial sulphate reduction. The source of Fe in the estimates given in the preceding paragraph. This makes for an
trapped former seawater is subaqueous bedrock ‘weathering’, irreconcilable conflict between conventional interpretations of
enhanced by the low pH (6.2) of the waters. Fe-oxy-hydroxide geomagnetic polarity reversals and excursions on the one hand,
forms instantaneously at Blood Falls, where the anoxic waters dis- and estimated deglaciation rates on the other. If one accepts a
charge at the glacier terminus (Mikucki et al. 2009). This demon- conventional magnetostratigraphic interpretation, cap dolostones
strates that waters rich in both dissolved Fe and sulphate can exist are highly condensed (Fairchild & Kennedy 2007). However,
if ice cover is perennial and sufficiently thick to preclude photo- compared with other depositional sequences in the same
trophic primary production. (or other) successions, Marinoan cap dolostones are relatively
thick and expanded among transgressive tracts (e.g. Hoffman &
Localization of oxidative titration. In anoxic Fe-rich waters, the Halverson 2008). The resolution of geomagnetic field reversals
localization of Fe (and Mn) deposition could relate to the avail- and excursions is limited by the accumulation rates of the sedi-
ability of oxygenated subglacial meltwater at the grounding-lines ments in which the records are kept (Channell & Lehman 1997;
of wet-base ice streams within cold-base ice sheets. This would Roberts & Lewin-Harris 2000; Roberts 2008). If Marinoan cap
be consistent with the observed occurrence of Cryogenian jaspi- dolostones were actually deposited at rates far exceeding the
lites as lenticular bodies in contact with diamictite complexes best Quaternary deep-sea records (Channell & Lehman 1997;
whose lithofacies and internal organization is most consistent Laj et al. 2006), might they not display hitherto unsuspected
with deposition in ice-sheet grounding zones. Alternation of Fe phenomena? This would most likely be the case if the absolute
and Mn ores could relate to variations of the meltwater flux associ- strength of the geomagnetic field was collapsed in 635 Ma, just
ated with glacial retreat and advance, respectively, as proposed by as it does during field reversals and excursions. The slowness
Urban et al. (1992). with which the solid inner core reverses, by diffusion, does not
limit the speed with which the outer core alone reverses in a
weak field (Gubbins 1999). In our view, the issue is unresolved,
Cap-carbonate sequences but neither the sedimentological evidence for high sedimentation
rates nor the physical arguments for rapid deglaciation should be
Accumulation rates for syndeglacial cap dolostones. Marinoan cap dismissed lightly.
dolostones were deposited during marine transgressions resulting
from the meltdown of grounded ice sheets (Preiss et al., 1978;
Kennedy 1996; Bertrand-Sarfati 1997; James et al. 2001; Shields Are syndeglacial cap dolostones non-marine?
2005; Fairchild & Kennedy 2007; Hoffman et al. 2007). The time-
scales for Cryogenian deglaciations have not been specifically tar- If Marinoan ice sheets had an average thickness of c. 2.2 km over
geted in climate modelling. Nevertheless, deglaciation histories of all continents and shelves (Fig. 5.1), their disappearance in 10 ka
76 P. F. HOFFMAN ET AL.

would produce meltwater at a global average rate of source of alkalinity driving carbonate production in deep water?
c. 0.6  106 m3 s21 (2.01  1017 m3/3.15  1010 s), or c. 6.4 Sv One possibility is anaerobic respiration, which operates as an
(Sverdrups, where 1.0 Sv ¼ 106 m3 s21). Deglaciation in 2 ka alkalinity pump in modern super-anoxic fjords (Anderson et al.
would result in a global meltwater flux of c. 32 Sv. These values 1987). Another is that extreme concentrations of dissolved inor-
correspond to approximately 6 and 32 times present total runoff, ganic carbon (DIC), expected for an ocean equilibrated with
respectively. Such an increase in freshwater input, followed by carbonate sediments and a c. 0.1 bar CO2 atmosphere, flattened
strong surface warming after the ice disappeared, must have the negative gradient in carbonate saturation (V) with increasing
impeded ocean mixing by creating a more stable density stratifica- depth, favouring the precipitation of cement on the seafloor,
tion (Shields 2005). Cold saline deepwater that evolved during the irrespective of the oxidant or intensity of organic carbon cycling
Marinoan glaciation would have been capped by a thickening (Higgins et al. 2009).
meltwater-dominated lid. The melting of sea ice, or a sea glacier
in the case of a Snowball Earth (Warren et al. 2002; Goodman
& Pierrehumbert 2003), followed by the melting of ice sheets, Barite in cap dolostones
would have contributed up to 0.7 and c. 1.0 km of freshwater
respectively. Ocean mixing time would have been many times Seafloor barite. Seafloor barites in basal Ediacaran (Marinoan) cap
the present 103 years, ensuring that the surface ocean was brackish dolostones of central Australia and northwestern Canada are
when cap dolostones were deposited (Shields 2005). This strikingly similar in form, dimensions and precise stratigraphic
meltwater-dominated lid has been dubbed ‘plumeworld’ (Shields position within their respective cap-carbonate sequences. We
2005), and we refer to the Marinoan example as ‘Glacial Lake infer that they formed where the seabed intersected the interface
Harland’ (GLH). (W. Brian Harland of Cambridge University between Ba-rich euxinic deeper water and an oxic Fe(III)-
was a pioneer advocate of Neoproterozoic glaciations as chronos- laden mixed layer. Localized upwelling would then have caused
tratigraphic markers.) Shields (2005, p. 305) gives many reasons sulphate production and barite titration. Simultaneously, anaerobic
why GLH was more favourable than seawater for rapid carbonate respiration, where Fe was the electron acceptor, caused Fe(III)
production. reduction, leading to Fe-carbonate (Fe-dolostone) production
coincident with the barite cement horizon. It is well established
that mixed-layer waters from which cap dolostones were deposited
Abiotic or biogenic dolomite? were Fe(III)-laden (Embleton & Williams 1986; Li 2000; Font
et al. 2005), accounting for the pinkish tint of many cap dolostones.
The dolomicrospar of cap dolostones preserves primary structures, Ba in euxinic glacial deep waters would derive from seawater –
makes conformable contact with overlying and (rare) underlying basalt exchange reactions and seafloor weathering of detrital feld-
limestones, and near the base of the former is reworked as intra- spar. We see no need for the involvement of methane venting, for
clastic dolomite grains. These are good reasons to conclude that which there is neither physical nor isotopic evidence at seafloor
the dolomite was either primary or a product of very early diagen- barite horizons in Marinoan cap dolostones (Kennedy 1996;
esis near the sediment – water interface. It has recently been Hoffman & Halverson 2011). Euxinic deep water that evolved
demonstrated that dolomite-forming reactions are catalysed at during the Marinoan glaciation (Hurtgen et al. 2006) may have
low temperatures by microbial activity utilizing different meta- contrasted with Fe(II) sulphate deep water (Mikucki et al. 2009),
bolic pathways: sulphate reduction (van Lith et al. 2003), anaero- which may have developed during the Sturtian glaciation, account-
bic methane oxidation (Moore et al. 2004) and methanogenesis by ing for the deposition of haematite –jaspilite at that time.
Archaea (Kenward et al. 2009). The first of these has been applied
to Marinoan cap dolostones (Shields 2005; Font et al. 2006; Early diagenetic (void-filling) barite. Tepee breccias in cap dolo-
Nédélec et al. 2007). However, methanogenesis by Archaea is stones and associated void-filling isopachous barite cement have
particularly appealing because the formation of ordered dolomite been ascribed to vadose diagenesis in West Africa (Bertrand-
at low temperature (48C) has been demonstrated experimentally Sarfati et al. 1997; Shields et al. 2007). In South China, similar fea-
in a low sulphate, low Mg:Ca (,1), acidic (pH ¼ 6.7), freshwater tures have been attributed to submarine cold seeps (Jiang et al.
environment, not unlike GLH (see above). As the methanogenic 2006) or, alternatively, to vadose diagenesis (Zhou et al. 2010).
consortium (Kenward et al. 2009) is anaerobic, formation of In both interpretations, barite precipitation is thought to have
cap dolostone by this means must have occurred below the resulted from mixing of an oxic sulphate-rich fluid (seawater)
sediment –water interface, assuming deglacial surface waters and a highly reducing, Ba- and possibly methane-rich fluid,
were well oxygenated. derived from the destabilization of permafrost gas hydrate in under-
lying glaciogenic sediments from which the Ba was scavenged
Sources of alkalinity for cap dolostones. Syndeglacial carbonate pro- (Jiang et al. 2006; Shields et al. 2007). It remains to be seen if
duction in surface waters was driven by warming (DT  508C; this attractive scenario is compatible with a basic field observation
Pierrehumbert 2002), inundation of shallow shelves and platforms in both areas, that the contact surface between the glaciogenic dia-
where carbonate burial is favoured (Ridgwell et al. 2003), alka- mictite and the cap dolostone is smooth and unbroken (authors’
linity provided by anaerobic respiration (Kennedy et al. 2001; observations). Is it conceivable that methane and other gases, gen-
but see Shields 2005 for critical discussion), and carbonate weath- erated within the till (permafrost melting), could thoroughly brecci-
ering (Higgins & Schrag 2003; Anderson 2007; see Le Hir et al. ate the cap dolostone on the seafloor, while leaving their mutual
2009 for critical discussion, but note that they consider only sili- contact surface undisturbed? Methane involvement is supported
cate weathering as a source of alkalinity, not carbonate weathering by extreme d13C depletion (below –10‰ Vienna Pee Dee Belem-
which is orders of magnitude more rapid). nite standard) in certain void-filling cement phases in brecciated
cap dolostone on the Yangtze platform (Jiang et al. 2003; Wang
Significance of seafloor aragonite cements. Seafloor cements com- et al. 2008). However, such strong depletions are not observed else-
posed of former aragonite occurring in limestones above Marinoan where, despite thousands of published analyses, and those in South
cap dolostones formed below the storm wave base, at or near the China come mostly from late-stage calcite void-fillings that
maximum flooding horizon (Grotzinger & Knoll 1995; James demonstrably post-date coexisting barite and void-filling dolomite
et al. 2001; Nogueira et al. 2003; Hoffman 2011). They represent cements (Bristow et al. 2011; Zhou et al. 2011). Moreover, d34S
a brief revival of more ancient conditions following a long decline values in cap dolostones are not strongly enriched, as predicted if
in the importance of seafloor cement in the latter half of the Proter- anaerobic methane oxidation (an alkalinity pump) was fueling bac-
ozoic (Grotzinger & James 2000; Sumner 2002). What was the terial sulphate reduction (Shields 2005).
CHEMICAL SEDIMENTS ASSOCIATED WITH NEOPROTEROZOIC GLACIATION 77

Phosphorite in cap dolostones Anderson, L. G., Dyrssen, D. & Skei, J. 1987. Formation of chemogenic
calcite in super-anoxic seawater –Framvaren, Southern-Norway.
It is not possible to generalize about Neoproterozoic glacial-related Marine Chemistry, 20, 361–376.
phosphorite from a single example, no matter how impressive and Assereto, R. L. A. M. & Kendall, C. G. St. C. 1977. Nature, origin and
well documented it is (Bertrand-Sarfati et al. 1997). Its existence classification of peritidal tepee structures and related breccas. Sedi-
supports assumed high productivity during deglaciation (Shields mentology, 24, 153–210.
2005; Font et al. 2006; Nédélec et al. 2007). Badenhorst, F. P. 1988. The lithostratigraphy of the Chuos mixtite in part
of the southern central zone of the Damara Orogen, South West
Africa. Communications of the Geological Survey of South West
Conclusions Africa/Namibia, 4, 103– 110.
Bao, H., Lyons, J. R. & Zhou, C. 2008. Triple oxygen isotope evidence
Synglacial Fe and Fe–Mn deposits are mostly, if not all, associated for elevated CO2 levels after a Neoproterozoic glaciation. Nature,
with the older Cryogenian (Sturtian) glaciation. They indicate 452, 504–506.
anoxic (not euxinic) deep water. Lithofacies associations suggest Bertrand-Sarfati, J., Flicoteaux, R., Moussine-Pouchkine, A. &
that subglacial meltwater plumes provided oxidant. Alternating Aı̈t Kaci Ahmed, A. 1997. Lower Cambrian apatitic stromatolites
Fe and Mn deposits may reflect oscillating plume fluxes associated and phospharenites related to the glacio-eustatic cratonic rebound
with glacial cycles. (Sahara, Algeria). Journal of Sedimentary Research, 67, 957–974.
Sturtian and Marinoan cap carbonates are distinct. Transgres- Bristow, T. F., Boniface, M., Derkowski, A., Eller, J. M. & Grotzin-
ger, J. P. 2011. A hydrothermal origin for isotopically anomalous cap
sive cap dolostones are almost entirely limited to the Marinoan
dolostone cements from south China. Nature, 474, 68 –72.
deglaciation, and were deposited diachronously, mainly above
Cahen, L. & Lepersonne, J. 1981. Proterozoic diamictites of Lower
the storm wave base, during coastal inundation by meltwater- Zaire. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-
dominated surface waters. Biogenic dolomite nucleation in Pleistocene Glacial Record. Cambridge University Press, Cam-
waters of low ionic strength and low Mg:Ca ratio was likely bridge, 153– 157.
microbially mediated. Cahen, L. 1950. Le Calcaire de Sekelolo, le Complexe tillitique et la
Barite occurs in Marinoan cap dolostones both as seafloor Dolomie rose C1 dans l’Anticlinal de Congo dia Kati (Bas-Congo).
cements in terminal Fe-dolomite beds and as early diagenetic void- Annales du Musée du Congo Belge, Sciences Géologiques, 7, 13– 54.
filling crusts within tepee and tepee-like breccias. Seafloor barite Calver, C. R. & Walter, M. R. 2000. The late Neoproterozoic Grassy
cement formed where euxinic Ba-rich deep waters upwelled into Group of King Island, Tasmania: correlation and palaeogeographic
oxic Fe(III)-rich surface waters. Early diagenetic void-filling significance. Precambrian Research, 100, 299– 312.
barite cement in cap dolostones has been attributed to subaerial Canfield, D. E. & Raiswell, R. 1999. The evolution of the sulfur cycle.
and submarine methane seepage, driven by destabilization of American Journal of Science, 299, 697–723.
permafrost hydrates in underlying sediments following marine Channell, J. E. T. & Lehman, B. 1997. The last two geomagnetic
inundation. Although attractive in principle, undisturbed contact polarity reversals recorded in high-deposition-rate sediment drifts.
surfaces between glaciogenic diamictites and internally brecciated Nature, 389, 712– 715.
cap dolostones may be difficult to reconcile with seepage from Chumakov, N. M. 1992. The problems of old glaciations: Pre-Pleistocene
below, unless brecciation was strictly limited to the horizon of glaciogeology in the USSR. Soviet Science Reviews, Section G
methane oxidation. Geology, 1, 1 –208.
Chumakov, N. M. 2007. Climates and climate zonality of the Vendian:
Fieldwork was supported by research grants from the Earth System History geological evidence. In: Vickers-Rich, P. & Komarower, P. (eds)
(ESH), Arctic Natural Science (ANS), and Geobiology & Environmental Geo-
The Rise and Fall of the Ediacaran Biota. Geological Society,
chemistry (GEG) programmes of the US National Science Foundation (NSF).
London, Special Publication, 286, 15– 26.
PFH receives additional support from the Canadian Institute for Advanced Clifford, T. N. 2008. The geology of the Neoproterozoic Swakop– Otavi
Research (CIFAR), Harvard University Center for the Environment (HUCE),
transition zone in the Outjo Distroct, northern Damara Orogen,
US Social Security, and the Canada Pension Plan. FAM and GPH thank the Namibia. South African Journal of Geology, 111, 117– 140.
Yukon Geological Survey for additional support. We gratefully acknowledge
Cloud, P., Wright, L. A., Williams, E. G., Diehl, P. & Walter, M. R.
constructive reviews by G. Jiang and P. K. Link, which prompted several
1974. Giant stromatolites and associated vertical tubes from the upper
improvements. This represents a contribution of the IUGS- and UNESCO- Proterozoic Noonday Dolomite, Death Valley region, eastern Califor-
funded IGCP (International Geoscience Programme) Project #512.
nia. Geological Society of America Bulletin, 85, 1869– 1882.
Condon, D., Zhu, M., Bowring, S. A., Wang, W., Yang, A. & Jin, Y.
2005. U– Pb ages from the Neoproterozoic Doushantuo Formation,
References China. Science, 308, 95 –98.
Corkeron, M. 2007. ‘Cap carbonates’ and Neoproterozoic glacigenic
Aitken, J. D. 1991. The Ice Brook Formation and post-Rapitan, late successions from the Kimberley region, north-west Australia. Sedi-
Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. mentology, 54, 871–903.
Geological Survey of Canada Bulletin, 404, 1 – 43. Corkeron, M. L. & George, A. D. 2001. Glacial incursion on a Neopro-
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves in the terozoic carbonate platform in the Kimberley region, Australia. Geo-
aftermath of a Neoproterozoic glaciation. Nature, 433, 123–127. logical Society of America Bulletin, 113, 1121– 1132.
Allen, P., Leather, J. & Brasier, M. D. 2004. The Neoproterozoic Fiq Corsetti, F. A. & Grotzinger, J. P. 2005. Origin and significance of tube
glaciation and its aftermath, Huqf Supergroup of Oman. Basin structures in Neoproterozoic post-glacial cap carbonates: example
Research, 16, 507– 534. from Noonday Dolomite, Death Valley, United States. Palaios, 20,
Alvarenga, C. J. S. de & Trompette, R. 1992. Glacially influenced sedi- 348– 363.
mentation in the Later Proterozoic of the Paraguay belt (Mato Grosso, Corsetti, F. A. & Kaufman, A. J. 2003. Stratigraphic investigations of
Brazil). Palaeogeography, Palaeoclimatology, Palaeoecology, 92, carbon isotope anomalies and Neoproterozoic ice ages in Death
85 – 105. Valley, California. Geological Society of America Bulletin, 115,
Alvarenga, C. J. S. de, Dardenne, M. A. et al. 2008. Isotope stratigra- 916– 932.
phy of Neoproterozoic cap carbonates in the Araras Group, Brazil. Deynoux, M. 1982. Periglacial polygonal structures and sand wedges in
Gondwana Research, 13, 469–479. the late Precambrian glacial formations of the Taoudeni Basin in
Anderson, J. B. 1999. Antarctic Marine Geology. Cambridge University Adrar of Mauretania (West Africa). Palaeogeography, Palaeoclima-
Press, Cambridge. tology, Palaeoecology, 39, 55 –70.
Anderson, S. P. 2007. Biogeochemistry of glacial landscape systems. Deynoux, M. 1985. Terrestrial or waterlain glacial diamictites? Three
Annual Review of Earth and Planetary Sciences, 35, 375– 399. case studies from the late Proterozoic and late Ordovician glacial
78 P. F. HOFFMAN ET AL.

drifts in West Africa. Palaeogeography, Palaeoclimatology, Frimmel, H. E. 2011. The Kaigas and Numees Formations, Port Nolloth
Palaeoecology, 51, 97– 141. Group, in South Africa and Namibia. In: Arnaud, E., Halverson,
Deynoux, M. & Trompette, R. 1976. Discussion: Late Precambrian G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
mixtites: glacial and/or nonglacial? Dealing especially with the terozoic Glaciations. Geological Society, London, Memoirs, 36,
mixtites of West Africa. American Journal of Science, 276, 223– 231.
1302– 1315. Gammon, P. R., McKirdy, D. M. & Smith, H. D. 2005. The timing and
Deynoux, M. & Trompette, R. 1981. Late Precambrian tillites of the environment of tepee formation in a Marinoan cap carbonate. Sedi-
Taoudeni Basin, West Africa. In: Hambrey, M. J. & Harland, mentary Geology, 177, 195– 208.
W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge Gaucher, C., Frimmel, H. E. & Germs, G. J. B. 2005. Organic-walled
University Press, Cambridge, 123– 131. microfossils and biostratigraphy of the upper Port Nolloth Group
Deynoux, M., Affaton, P., Trompette, R. & Villeneuve, M. (Namibia): implications for latest Neoproterozoic glaciations. Geo-
2006. Pan-African tectonic evolution and glacial events registered logical Magazine, 142, 539–559.
in Neoproterozoic to Cambrian cratonic and foreland basins of Goodman, J. & Pierrehumbert, R. T. 2003. Glacial flow of floating
West Africa. Journal of African Earth Sciences, 46, 397–426. marine ice in ‘Snowball Earth’. Journal of Geophysical Research,
Dorr, J. V. N. II. 1945. Manganese and iron deposits of Morro do Urucum, 108, 3308, doi: 10.1029/2002JC001471.
Mato Grosso, Brazil. United States Geological Survey Bulletin, Graf, J. L., Jr, O’Connor, E. A. & Van Leeuwin, P. 1994. Rare earth
946-A, 1– 47. element evidence of origin and depositional environment of Late
Dowdeswell, J. A., Whittington, J. A., Jennings, A. E., Andrews, Proterozoic ironstone beds and manganese-oxide deposits, SW
J. T., Mackensen, A. & Marienfield, P. 2000. An origin for lami- Brazil and SE Bolivia. Journal of South American Earth Sciences,
nated glacimarine sediments through sea-ice build-up and suppressed 7, 115– 133.
iceberg rafting. Sedimentology, 47, 557– 576. Grotzinger, J. P. & James, N. P. 2000. Precambrian carbonates: evol-
Dunn, P. R., Thompson, B. P. & Rankama, K. 1971. Late Pre-Cambrian ution of understanding. In: Grotzinger, J. P. & James, N. P. (eds)
glaciation in Australia as a stratigraphic boundary. Nature, 231, Carbonate Sedimentation and Diagenesis in the Evolving Precam-
498– 502. brian World. SEPM (Society for Sedimentary Geology) Special
Edwards, M. B. 1984. Sedimentology of the Upper Proterozoic glacial Publication, 67, 3– 20.
record, Vestertana Group, Finnmark, North Norway. Norges Geolo- Grotzinger, J. P. & Knoll, A. H. 1995. Anomalous carbonate precipi-
giske Undersøkelse Bulletin, 394, 76. tates: is the Precambrian the key to the Permian? Palaios, 10,
Eisbacher, G. H. 1985. Late Proterozoic rifting, glacial sedimentation and 578– 596.
sedimentary cycles in the light of Windermere deposition, western Gubbins, D. 1999. The distinction between geomagnetic excursions and
Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 51, reversals. Geophysical Journal International, 137, F1– F3.
231– 254. Halverson, G. P. & Shields, G. 2011. Chemostratigraphy and
Embleton, B. J. J. & Williams, G. E. 1986. Low latitude of deposition for the Neoproterozoic glaciations. In: Arnaud, E., Halverson, G. P.
late Precambrian periglacial varvites in South Australia: implications & Shields, G. (eds) The Geological Record of Neoproterozoic
for palaeoclimatology. Earth and Planetary Science Letters, 79, Glaciations. Geological Society, London, Memoirs, 36, 51 –66.
419– 430. Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
Fairchild, I. J. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the glaciation (Neoproterozoic) in northeast Svalbard. Basin Research,
Earth System. Journal of the Geological Society, London, 164, 16, 297– 324.
895– 921. Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice,
Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian A. H. N. 2005. Toward a Neoproterozoic composite carbon-isotope
glaciation: data from the Adelaide Geosyncline, South Australia record. Geological Society of America Bulletin, 117, 1181– 1207.
and Pocatello Formation, Idaho, USA. Selwyn Symposium Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A.
2008, Geological Society of Australia, Extended Abstracts, 91, 2007. Evolution of the 87Sr/86Sr composition of Neoproterozoic
57 –62. seawater. Palaeogeography, Palaeoclimatology, Palaeoecology,
Fölling, P. G. & Frimmel, H. E. 2002. Chemostratigraphy correlation of 256, 103– 129.
carbonate successions in the Gariep and Saldania Belts, Namibia and Halverson, G. P., Poitrasson, F., Hoffman, P. F., Nédélec, A.,
South Africa. Basin Research, 14, 69 – 88. Montel, J.-M. & Kirby, J. 2011. Fe isotope and trace element geo-
Fölling, P. G., Zartman, R. E. & Frimmel, H. E. 2000. A novel chemistry of the Neoproterozoic syn-glacial Raptian iron formation.
approach to double-spike Pb/Pb dating of carbonate rocks: examples Earth and Planetary Science Letters, 309, 100–112.
from Neoproterozoic sequences in southern Africa. Chemical Hegenberger, W. 1993. Stratigraphy and sedimentology of the Late Pre-
Geology, 171, 97 –122. cambrian Witvlei and Nama Groups, East of Windhoek. Geological
Font, E., Nédélec, A., Trindade, R. I. F., Macouin, M. & Charrière, Survey of Namibia Memoir, 17, 82.
A. 2006. Chemostratigraphy of the Neoproterozoic Mirassol d’Oeste Higgins, J. A. & Schrag, D. P. 2003. Aftermath of a snowball Earth. Geo-
cap dolostones (Mato Grosso, Brazil): an alternative model for physics, Geochemistry, Geosystems, 4, doi:10.1029/2002GC000403.
Marinoan cap dolostone formation. Earth and Planetary Science Higgins, J. A., Fischer, W. W. & Schrag, D. P. 2009. Oxygenation of the
Letters, 250, 89 –103. ocean and sediments: consequences for the seafloor carbonate
Font, E., Trindade, R. I. F. & Nédélec, A. 2005. Detrital remanent mag- factory. Earth and Planetary Science Letters, 284, 25 – 33.
netization in haematite-bearing Neoproterozoic Puga cap dolostone, Hildebrand, R. S. 2009. Did westward subduction cause Cretaceous–
Amazon craton: a rock magnetic and SEM study. Geophysical Tertiary orogeny in the North American Cordillera? Geological
Journal International, 163, 491–500. Society of America, Special Paper, 457, 71.
Frimmel, H. E. 2004. Neoproterozoic sedimentation rates and timing Hoffman, P. F. 2011. Strange bedfellows: glacial diamictite and cap
of glaciations – a southern African perspective. In: Eriksson, carbonate from the Marinoan (635 Ma) glaciation in Namibia. Sedi-
P. G., Altermann, W., Nelson, D. R., Mueller, W. U. & mentology, 58 – 119.
Catuneanu, O. (eds) The Precambrian Earth: Tempos and Hoffman, P. F. & Halverson, G. P. 2008. Otavi group of the western
Events. Elsevier, Amsterdam, 459– 473. northern platform, the eastern Kaoko Zone and the western northern
Frimmel, H. E. 2008. Neoproterozoic Gariep Orogen. In: Miller, Margin Zone. In: Miller, R. McG. (ed.) The Geology of Namibia,
R. McG. (ed.) The Geology of Namibia (In Three Volumes), Vol. 2: Neoproterozoic to Lower Palaeozoic. Handbook of the
Volume 2 (Neoproterozoic to Lower Palaeozoic). Geological Geological Survey of Namibia, Windhoek, 13.69– 13.136.
Survey of Namibia, Windhoek, 14-1–14-39. Hoffman, P. F. & Halverson, G. P. 2011. Neoproterozoic glacial record
Frimmel, H. E. & Von Veh, M. W. 2003. Numees formation (including in the Mackenzie Mountains, northern Canadian Cordillera. In:
the Jakkalsberg Member). In: Johnson, M. R. (ed.) Catalogue of Arnaud, E., Halverson, G. P. & Shields, G. (eds) The Geological
South African Lithostratigraphic Units. South African Committee Record of Neoproterozoic Glaciations. Geological Society, London,
for Stratigraphy, Pretoria, 7-25– 7-28. Memoirs, 36, 397 –411.
CHEMICAL SEDIMENTS ASSOCIATED WITH NEOPROTEROZOIC GLACIATION 79

Hoffman, P. F. & Li, Z. X. 2009. A palaeogeographic context for Klein, C. 2005. Some Precambrian banded iron-formations (BIFs) from
Neoproterozoic glaciation. Palaeogeography, Palaeoclimatology, around the world: their age, geologic setting, mineralogy, meta-
Palaeoecology, 277, 158–172. morphism, geochemistry and origin. American Mineralogist, 90,
Hoffman, P. F. & Macdonald, F. A. 2010. Sheet-crack cements and 1473–1499.
early regression in Marinoan (635 Ma) cap dolostones: regional Klein, C. & Beukes, N. J. 1993. Sedimentology and geochemistry of
benchmarks of vanishing ice-sheets? Earth and Planetary Science the glacigenic Late Proterozoic Rapitan iron-formation in Canada.
Letters, 300, 374– 384. Economic Geology, 88, 542–565.
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: Klein, C. & Ladeira, E. A. 2004. Geochemistry and mineralogy of
testing the limits of global change. Terra Nova, 14, 129– 155. Neoproterozoic banded iron-formations and some selected, siliceous
Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P. manganese formations from the Urucum District, Mato Grosso do
1998. A Neoproterozoic Snowball Earth. Science, 281, 1342– 1346. Sul, Brazil. Economic Geology, 99, 1233– 1244.
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blick, N.
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635 2006. The Ediacaran Period: a new addition to the geologic time
Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary scale. Lethaia, 39, 13– 30.
Science Letters, 258, 114– 131. Kump, L. R. & Seyfried, W. E. Jr. 2005. Hydrothermal Fe fluxes during
Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L. the Precambrian: effect of low oceanic sulfate concentrations and low
2004. U– Pb zircon date from the Neoproterozoic Ghaub Formation, hydrostatic pressure on the composition of black smokers. Earth and
Namibia: constraints on Marinoan glaciation. Geology, 32, 817– 820. Planetary Science Letters, 235, 654– 662.
Hurtgen, M. T., Halverson, G. P., Arthur, M. A. & Hoffman, P. F. Laj, C., Kissel, C. & Roberts, A. P. 2006. Geomagnetic field behavior
2006. Sulfur cycling in the aftermath of a 635-Ma snowball during the Iceland Basin and Laschamp geomagnetic excursions:
glaciation: evidence for a syn-glacial sulfidic deep ocean. Earth a simple transitional field geometry? Geochemistry, Geophysics,
and Planetary Science Letters, 245, 551–570. Geosystems, 7, Q03004, doi:10.1029/2005GC001122.
Hyde, W. T., Crowley, T. J., Baum, S. K. & Peltier, W. R. 2000. Le Hir, G. & Donnadieu, Y. et al. 2009. The snowball Earth aftermath:
Neoproterozoic ‘snowball Earth’ simulations with a coupled exploring the limits of continental weathering processes. Earth and
climate/ice-sheet model. Nature, 405, 425–429. Planetary Science Letters, 277, 453– 463.
Ilyin, A. V. 2009. Neoproterozoic banded iron formations. Lithology and Li, Z. X. 2000. New palaeomagnetic results from the ‘cap dolomite’ of the
Mineral Resources, 44, 78 –86. Neoproterozoic Walsh Tillite, northwestern Australia. Precambrian
Ilyin, A. V., Zaitsev, N. S. & Bjamba, Z. 1986. Khubsugul, Mongolia Research, 100, 359–370.
people’s republic. In: Cook, P. J. & Shergold, J. H. (eds) Phosphate Li, Z. X. & Bogdanova, S. V. et al. 2008. Assembly, configuration and
Deposits of the World, Vol. 1: Proterozoic and Cambrian Phosphor- break-up history of Rodinia: a synthesis. Precambrian Research, 160,
ites. Cambridge University Press, Cambridge. 179– 210.
James, N. P., Narbonne, G. M. & Kyser, T. K. 2001. Late Neoprotero- Lottermoser, B. G. & Ashley, P. M. 2000. Geochemistry, petrology and
zoic cap carbonates: Mackenzie Mountains, northwestern Canada: origin of Neoproterozoic ironstones in the eastern part of the Adelaide
precipitation and global glacial meltdown. Canadian Journal of Geosyncline. Precambrian Research, 101, 49 –67.
Earth Sciences, 38, 1229–1262. Lythe, M. B., Vaughan, D. G. & BEDMAP CONSORTIUM
Jiafu, T., Heqin, F. & Zhioiu, G. 1987. Stratigraphy, type and formation 2001. BEDMAP: a new thickness and subglacial topographic
conditions of the Late Precambrian banded iron ores in South China. model of Antarctica. Journal of Geophysical Research, 106,
Chinese Journal of Geochemistry, 6, 332–351. 11335–11351.
Jiang, G., Kennedy, M. J. & Christie-Blick, N. 2003. Stable isotopic Macdonald, F. A. 2011. The Tsagaan Oloom Formation, southwestern
evidence for methane seeps in Neoproterozoic postglacial cap car- Mongolia. In: Arnaud, E., Halverson, G. P. & Shields, G. (eds)
bonates. Nature, 426, 822– 826. The Geological Record of Neoproterozoic Glaciations. Geological
Jiang, G., Kennedy, M. J., Christie-Blick, N., Wu, H. & Zhang, S. Society, London, Memoirs, 36, 331– 337.
2006. Stratigraphy, sedimentary structures and textures of the late Macdonald, F. A. & Cohen, P. A. 2011. The Tatonduk inlier, Alaska–
Neoproterozoic Doushantuo cap carbonate in South China. Journal Yukon border. In: Arnaud, E., Halverson, G. P. & Shields, G.
of Sedimentary Research, 76, 978– 995. (eds) The Geological Record of Neoproterozoic Glaciations.
Johnson, H. P., Van Patten, D., Tivey, M. & Sager, W. W. 1995. Geo- Geological Society, London, Memoirs, 36, 389–396.
magnetic polarity reversal rate for the Phanerozoic. Geophysical Macdonald, F. A. & Jones, D. S. 2011. The Khubsugul Group, northern
Research Letters, 22, 231– 234. Mongolia. In: Arnaud, E., Halverson, G. P. & Shields, G. (eds)
Kaufman, A. J., Jiang, G., Christie-Blick, N., Banerjee, D. M. & Rai, The Geological Record of Neoproterozoic Glaciations. Geological
V. 2006. Stable isotope record of the terminal Neoproterozoic Krol Society, London, Memoirs, 36, 339– 345.
platform in the Lesser Himalayas of northern India. Precambrian Macdonald, F. A., Mcclelland, W. C., Schrag, D. P. & Macdonald,
Research, 147, 156– 185. W. P. 2009. Neoproterozoic glaciation on a carbonate platform
Kendall, C. G. St. C. & Warren, J. 1987. A review of the origin and margin in Arctic Alaska and the origin of the North Slope subterrane.
setting of tepees and their associated fabrics. Sedimentology, 34, Geological Society of America Bulletin, 121, 448– 473.
1007– 1027. Macdonald, F. A., Schmidtz, M. D. et al. 2010. Calibrating the Cryo-
Kennedy, M. J. 1996. Stratigraphy, sedimentology and isotopic geochem- genian. Science, 327, 1241– 1243.
istry of Australian Neoproterozoic postglacial cap dolostones: degla- Macdonald, F. A., Strauss, J. V., Rose, C., Dudás, F. Ö. & Schrag,
ciation, d13C excursions and carbonate precipitation. Journal of D. P. 2011. Stratigraphy of the Port Nolloth Group of Namibia and
Sedimentary Research, 66, 1050– 1064. South Africa and implications for the age of Neoproterozoic iron
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K.-H. & formations. American Journal of Science, 310, 862–888.
Arthur, M. A. 1998. Two of four Neoproterozoic glaciations? Martin, H. 1965a. The Precambrian Geology of South West Africa and
Geology, 26, 1059– 1063. Namaqualand. University of Cape Town, Precambrian Research
Kennedy, M. J., Christie-Blick, N. & Sohl, L. E. 2001. Are Proterozoic Unit Bulletin, 1, 1– 159.
cap carbonates and isotopic excursions a record of gas hydrate desta- Martin, H. 1965b. Beobachtungen zum Problem der jung-präkam-
bilization following Earth’s coldest intervals? Geology, 29, 443– 446. brischen Glazialen Ablagerungen in Südwestafrika (Observations
Kenward, P. A., Goldstein, R. H., González, L. A. & Roberts, J. A. concerning the problem of the late Precambrian glacial deposits in
2009. Precipitation of low-temperature dolomite from an aerobic South West Africa). Geologische Rundschau, 54, 115– 127.
microbial consortium: the role of methanogenic Archaea. Geobiol- Mikucki, J. A., Pearson, A. et al. 2009. A contemporary microbially
ogy, 7, 1 – 10. maintained subglacial ferrous ‘ocean’. Science, 324, 397–400.
Kianian, M. & Khakzad, A. 2008. Geochemistry of glaciogenic Neopro- Moore, T. S., Murray, R. W., Kurtz, A. C. & Schrag, D. P. 2004.
terozoic banded iron-formations from Kerman District (Iran). 33rd Anaerobic methane oxidation and the formation of dolomite. Earth
International Geological Congress, Oslo, Abstracts, Session CGC-04. and Planetary Science Letters, 229, 141–154.
80 P. F. HOFFMAN ET AL.

Myrow, P. M. & Kaufman, A. J. 1999. A newly discovered cap carbonate Sovetov, Yu. K. & Komlev, D. A. 2005. Tillites at the base of the Oselok
above Varanger-age glacial deposits in Newfoundland, Canada. Group, foothills of the Sayan Mountains, and the Vendian lower
Journal of Sedimentary Research, 69, 784– 793. boundary in the southwestern Siberian Platform. Stratigraphy and
Nédélec, A., Affaton, P., France-Lanord, C., Charrière, A. & Geological Correlations, 13, 337–366.
Alvaro, J. 2007. Sedimentology and chemostratigraphy of the Sumner, D. Y. 2002. Decimetre-thick encrustations of calcite and arago-
Bwipe Neoproterozoic cap dolostones (Ghana, Volta Basin): a nite on the sea-floor and implications for Neoarchaean and Neoproter-
record of microbial activity in a peritidal environment. C. R. ozoic ocean chemistry. Special Publications of the International
Geoscience, 339, 223– 239. Association of Sedimentologists, 33, 107–120.
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V. & Tojo, B., Katsuta, N., Takano, M., Kawakami, S. & Ohno, T. 2007.
Fairchild, T. R. 2003. Soft-sediment deformation at the base of Calcite – dolomite cycles in the Neoproterozoic Cap carbonates,
the Neoproterozoic Puga cap carbonate (southwestern Amazon Otavi Group, Namibia. In: Vickers-Rich, P. & Komarower, P.
craton, Brazil): confirmation of rapid icehouse to greenhouse tran- (eds) The Rise and Fall of the Ediacaran Biota. Geological Society,
sition in snowball Earth. Geology, 31, 613–616. London, Special Publication, 286, 103– 113.
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V., Trin- Trindade, R. I. F., Font, E., D’agrella-Filho, M. S., Nogueira,
dade, R. I. F. & Fairchild, T. R. 2007. Carbon and strontium isotope A. C. R. & Riccimini, C. 2003. Low-latitude and multiple geomag-
fluctuations and paleoceanographic changes in late Neoproterozoic netic reversals in the Neoproterozoic Puga cap carbonate, Amazon
Araras carbonate platform, southern Amazon craton, Brazil. Chemi- craton. Terra Nova, 15, 441–446.
cal Geology, 237, 168–190. Trompette, R. 1994. Geology of Western Gondwana (2000– 500 Ma):
Pierrehumbert, R. T. 2002. The hydrologic cycle in deep-time climate Pan-African –Brasiliano Aggregation of South America and Africa.
problems. Nature, 419, 191–198. Balkema, Amsterdam.
Plummer, P. S. 1978. Note on the palaeoenvironmental significance of the Trompette, R., Affaton, P., Joulia, F. & Marchand, J. 1980. Strati-
Nuccaleena Formation (upper Precambrian), central Flinders Ranges, graphic and structural controls of late Precambrian phosphate depos-
South Australia. Journal of the Geological Society of Australia, 25, its of the Northern Volta Basin in Upper Volta, Niger and Benin, West
395– 402. Africa. Economic Geology, 75, 62– 70.
Prave, A. R., Hoffmann, K.-H., Hegenberger, W. & Fallick, A. E. Trompette, R., de Alvarenga, C. J. A. & Wade, D. 1998. Geological
2011. The Witvlei Group of east-central Namibia. In: Arnaud, E., evolution of the Neoproterozoicv Corumbá graben system (Brazil).
Halverson, G. P. & Shields, G. (eds) The Geological Record Depositional context of the stratified Fe and Mn ores of the Jacadigo
of Neoproterozoic Glaciations. Geological Society, London, Group. Journal of South American Earth Sciences, 11, 587–597.
Memoirs, 36, 211– 216. Urban, H., Stribny, B. & Lippolt, H. J. 1992. Iron and manganese
Preiss, W. V. 1987. The Adelaide Geosyncline: Late Proterozoic stratigra- deposits of the Urucum District, Mato Grosso do Sul, Brazil. Econ-
phy, sedimentation, palaeontology and tectonics. Geological Survey omic Geology, 87, 1375–1892.
of South Australia Bulletin, 53, 438. van Lith, Y., Warthmann, R., Vasconcelos, C. & McKenzie, J. A.
Preiss, W. V., Walter, M. R., Coates, R. P. & Wills, A. T. 1978. Litho- 2003. Sulphate-reducing bacteria induce low-temperature Ca-
logical correlation of the Adelaidean glaciogenic rocks in parts of the dolomite and high Mg-calcite formation. Geobiology, 1, 71 – 79.
Amadeus, Ngalia & Georgina basins. Bureau of Mineral Resources, Walde, D. H. G., Gierth, E. & Leonardos, O. H. 1981. Stratigraphy and
Journal of Australian Geology and Geophysics, 3, 43 –53. mineralogy of the manganese ores of Urucum, Mato Grosso, Brazil.
Pruss, S. B., Bosak, T., Macdonald, F. A., Mclane, M. & Hoffman, Geologische Rundschau, 70, 1077– 1085.
P. F. 2010. Microbial facies in a Sturtian cap carbonate, the Rasthof Wallace, J. M. & Hobbs, P. V. 1977. Atmospheric Science: An Introduc-
Formation, Otavi Group, northern Namibia. Precambrian Research, tory Survey. Academic Press, San Diego.
181, 187– 108. Walter, M. R. & Bauld, J. 1983. The association of sulphate evaporites,
Rankama, K. 1973. The Late Precambrian glaciation, with particular stromatolitic carbonates and glacial sediments: examples from the
reference to the Southern Hemisphere. Journal and Proceedings, Proterozoic of Australia and the Cainozoic of Antarctic. Precambrian
Royal Society of New South Wales, 106, 89– 97. Research, 21, 129– 148.
Raub, T. D. 2008. Prolonged deglaciation of ‘Snowball Earth’. PhD Wang, J., Jiang, G., Xiao, S., Li, Q. & Wei, Q. 2008. Carbon isotope evi-
thesis, Yale University, CT. dence for widespread methane seeps in the ca 635 Ma Doushantuo
Ridgwell, A. J., Kennedy, M. J. & Caldeira, K. 2003. Carbonate depo- cap carbonate in South China. Geology, 36, 347– 350.
sition, climate stability and Neoproterozoic ice ages. Science, 302, Warren, S. G., Brandt, R. E., Grenfell, T. C. & Mckay, C. P. 2002.
859– 862. Snowball Earth: ice thickness on the tropical ocean. Journal of
Roberts, A. P. 2008. Geomagnetic excursions: knowns and unknowns. Geophysical Research, 107, 3167, doi: 10.1029/2001JC001123.
Geophysical Research Letters, 35, L17307, doi: 10.1029/ Whitten, G. F. 1970. The investigation and exploitation of the Razorback
2008GL034719. Ridge iron deposit. Geological Survey of South Australia Reports of
Roberts, A. P. & Lewin-Harris, J. C. 2000. Marine magnetic anomalies: Investigations, 33, 165.
evidence that ‘tiny wiggles’ represent short-period geomagnetic Xiao, S., Bao, H. et al. 2004. The Neoproterozoic Quruqtagh Group in
polarity intervals. Earth and Planetary Science Letters, 183, eastern Chinese Tianshan: evidence for a post-Marinoan glaciation.
375– 388. Precambrian Research, 130, 1 –26.
Sheldon, R. P. 1984. Ice-ring origin of the Earth’s atmosphere and hydro- Yeo, G. M. 1981. The Late Proterozoic Rapitan glaciation in the northern
sphere and Late Proterozoic –Cambrian hypothesis. Geological Cordillera. In: Campbell, F. H. A. (ed.) Proterozoic Basins of
Survey of India Special Publication, 17, 17 –21. Canada. Geological Survey of Canada Paper, 81– 10, 25 –46.
Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal Yeo, G. M. 1986. Iron-formation in the late Proterozoic Rapitan Group,
of existing models and the plumeworld hypothesis. Terra Nova, 17, Yukon and Northwest Territories. In: Morin, J. A. (ed.) Mineral
299– 310. Deposits of Northern Cordillera. Canadian Institute of Mining and
Shields, G. A., Deynoux, M., Strauss, H., Paquet, H. & Nahon, D. Metallurgy Special Volume, 37, 142– 153.
2007. Barite-bearing cap dolostone of the Taoudéni Basin, northwest Young, G. M. 1976. Iron-formation and glaciogenic rocks of the Rapitan
Africa: sedimentary and isotopic evidence for methane seepage Group, Northwest Territories, Canada. Precambrian Research, 3,
after a Neoproterozoic glaciation. Precambrian Research, 154, 137– 158.
209– 235. Zhang, S., Jiang, G. & Han, Y. 2008. The age of the Nantuo Formation
Smith, L. H., Kaufman, A. J., Knoll, A. H. & Link, P. K. 1994. and Nantuo glaciation in South China. Terra Nova, 20, 289–294.
Chemostratigraphy of predominantly siliciclastic Neoproterozoic Zhou, C., Bao, H., Peng, Y. & Yuan, X. 2010. Timing of deposition of
17
successions: a case study of the Pocatello Formation and lower O-depleted barite at the Nantuo glacial meltdown in South China.
Brigham Group, Idaho. Geological Magazine, 131, 301–314. Geology, 38, 903–906.
Chapter 6

A review of the Chemical Index of Alteration (CIA) and its application to the study
of Neoproterozoic glacial deposits and climate transitions

HEINRICH BAHLBURG1* & NICOLE DOBRZINSKI2


1
Institut für Geologie und Paläontologie, Westfälische Wilhelms-Universität, 48149 Münster, Germany
2
RWE Dea AG, Wietze Laboratory, 29323 Wietze, Germany
*Corresponding author (e-mail: hbahlburg@uni-muenster.de)

Abstract: The Chemical Index of Alteration (CIA) is the most accepted of available weathering indices. Past conditions of physical and
chemical weathering can be reliably inferred if application of the CIA is combined with a comprehensive facies analysis. When applied to
the reconstruction of climate conditions during Neoproterozoic times, CIA data provide crucial insights into the changes in the relative
contributions of chemical and physical weathering in the production of sedimentary detritus. CIA data are thus instrumental not only in
documenting changes between icehouse and greenhouse climates, but also in recognizing shorter-term climate oscillations between
glacial and warm– humid conditions. Concerning the Neoproterozoic glacial periods, sedimentological and CIA data sets give strong
evidence of a functioning hydrological cycle, operative sediment routing systems, and variable climate conditions oscillating
between dry–cool and glacial, and warm–humid and interglacial. These findings are incompatible with the hypothesis of a totally
ice-covered Snowball Earth.

Climate exerts the major control on weathering processes affecting Schrag et al. 2002; Baum & Crowley 2003; Fairchild &
the upper continental crust. Direct evidence of past weathering Kennedy 2007; Shields 2008).
conditions and thus climate can be obtained from palaeosols Unfortunately, in many cases, facies analysis of sedimentary
through a combination of, among other things, field observation, rocks does not supply unambiguous information on palaeoclimate
petrography (particularly diagenetic phases), X-ray diffractometry conditions. This conundrum is well illustrated by the mutually
and whole rock and isotope geochemistry. However, palaeosols exclusive interpretations of the Neoproterozoic, that is, Marinoan,
are restricted to continental environments, which commonly Ghaub Formation diamictites (Otavi Group) in northern Namibia.
have a relatively low preservation potential. Even though palaeo- These have been interpreted either as glaciogenic sediments
soils as old as Palaeoproterozoic are reported in the literature (Gevers 1931; Martin 1964; Hoffmann & Prave 1996; Hoffman
(e.g. Gutzmer & Beukes 1998; Nedachi et al. 2005), most infor- & Schrag 2002) or as tectonically triggered mass flow deposits
mation on weathering conditions and thus climate has to be (Schermerhorn 1974; Martin et al. 1985; Eyles & Januszczak
gleaned from reworked siliciclastic material deposited in marine 2007). However, these issues may be moved closer to resolution
or lacustrine environments. In the Precambrian eon, which through a combination of facies analysis with considerations of
represents nearly 90% of Earth history, marine siliciclastic sedi- weathering as a function of climate (e.g. Young & Nesbitt 1999;
mentary rocks are the most widely available source of information Young 2001; Scheffler et al. 2003).
on past climates. In this contribution we will review the most important weather-
In Phanerozoic rocks, information derived from siliciclastic ing proxies applied to siliciclastic sedimentary rocks. We will
sedimentary rocks may be supplemented by ecological indicators concentrate on the CIA (Nesbitt & Young 1982) and its potential
derived from fossils. However, the bearing of palaeofaunal data and limitations for deriving palaeoclimate information from
on the interpretations of palaeoclimate conditions needs to be con- glacial and pre- and post-glacial siliciclastic successions. We
sidered very carefully, because differences between continental will preferentially but not exclusively address Neoproterozoic
regimes of weathering and marine environments of deposition glacial successions based on a combination of published infor-
may be very large. This is illustrated for example by the mation from the Nanhuan –Sinian glacial succession in South
pronounced difference at the coast of northern Chile between con- China (Dobrzinski et al. 2004; Dobrzinski & Bahlburg 2007)
ditions of extreme subtropical aridity in the Atacama desert and the and new data on the Port Askaig Formation (Scotland), the
cold waters of the Humboldt current originating in the peri- Ghaub Formation (northern Namibia) and the Mortensnes and
Antarctic westwind drift. Smalfjord formations (northern Norway).
Extreme climate states are of particular interest to palaeoclimate
research. Among these, those that figure most prominently are the
icehouse climates that happened in at least five cycles of highly Weathering and weathering indices
variable duration and extent during Earth history, that is, the
Palaeoproterozoic and Neoproterozoic, Ordovician, Carbonifer- Exposed rocks are affected to variable degrees by a combination of
ous –Permian and Neogene glaciations (Crowley & North 1991; chemical and physical weathering (Bland & Rolls 1998). Progress-
Crowell 1999). Of these, the Neoproterozoic glaciations are ive chemical weathering of labile minerals like feldspar leads to
some of the most discussed and contentious (e.g. Hoffman et al. the loss of Ca2þ, Kþ and Naþ and the transformation to minerals
1998; Evans 2000; Hoffman & Schrag 2002; Eyles & Januszczak more stable under surface conditions (Fedo et al. 1995). Ulti-
2007; Fairchild & Kennedy 2007; Etienne et al. 2008; Eyles 2008). mately, it results in the formation of shales rich in clay minerals
The timing and causes of the inception and termination of Neopro- such as illite and kaolinite, and Fe-oxyhydrates such as goethite.
terozoic glacial states as well as variations in the severity of the Physical weathering, in turn, leads to the degradation of rocks to
icehouse conditions themselves remain unresolved issues (e.g. smaller grain sizes, ideally without causing geochemical and

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 81– 92. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.6
82 H. BAHLBERG & N. DOBRZINSKI

mineralogical changes. If physical weathering, and the grinding changes of bulk-rock geochemical composition caused by chemi-
action of moving ice-lodged debris in particular, degrades a cal alteration. A very simple proxy is the Ruxton Ratio R (Ruxton
source rock into a clay-sized deposit, it should essentially preserve 1968) given by the SiO2/Al2O3 ratio. This assumes that Al2O3
the mineralogical and geochemical composition of the original remains immobile during weathering, and changes in R therefore
rock (Nesbitt & Young 1982, 1996). Consequently, the character reflect silica loss as a proxy for total element loss. The Ruxton
of the climate framework and the way it governs weathering con- Ratio may be useful when weathering profiles on rocks of felsic
ditions and reactions is reflected by the mineralogical and mobile and intermediate composition are considered, but was found to
element geochemical composition of the resulting deposits, and be poorly correlated to the actual weathering grade of a silicate
the shales and clay-sized materials in particular. rock (Duzgoren-Aydin et al. 2002).
Ultimately, chemical weathering of silicate rocks leads As the transformation of feldspar to clay minerals and the
through hydrolysis to an exchange of the cations Naþ, Kþ, Ca2þ coincident mobility of the main cations is a major process of
and Mg2þ for Hþ, and maybe a loss of Si4þ (Kramer 1968). chemical (i.e. hydrolytical) weathering, Parker (1970) considered
Naþ, Kþ and Ca2þ are commonly supplied by the weathering it more useful to mirror changes in Naþ, Kþ, Ca2þ and Mg2þ
of feldspar and volcanic glass, together accounting for c. 58% of and created a weathering index (WIP, Weathering Index of
the exposed crust. Mg2þ is derived from glasses, sheet silicates Parker) given by
and mafic minerals and resides in chloritic and smectitic clays
(Nesbitt & Young 1984; Pettijohn et al. 1987). In general, hydro-   
lytic weathering causes a progressive transformation of affected Na Mg K Ca
components into clay minerals, ultimately kaolinite. In a qualitat- WIP ¼ :35 þ :9 þ :25 þ :7  100 (1)
0 0 0 0
ive way, mineralogical changes can be detected in shales and
sandstones using X-ray diffractometry. Optical analysis is a
quantitative option in sandstones only. Here, the Mineralogical where the cations* represent the atomic percentage of an element
Index of Alteration (MIA), given by the ratio of quartz to the divided by the atomic weight. Parker (1970) also considered the
sum of quartz þ K-feldspar þ plagioclase, permits an assessment susceptibility of these elements to weathering by including in the
of weathering effects (Johnsson 1993). Quantitative estimations denominator Nicholls’ values of bond strength as a measure of
of weathering in fine-grained as well as coarse-grained rocks are the energy necessary to break the cation-to-oxygen bonds of
relatively easily achieved by using whole-rock geochemical data the respective oxides. These different values are considered to
to calculate geochemical weathering proxies. A comprehensive reflect the probability of an element being mobilized during the
review of a great number of geochemical weathering indices has weathering process. Values of WIP are commonly between
been presented by Duzgoren-Aydin et al. (2002). 100 and 0, with the least weathered rocks having the highest
Recently, von Eynatten et al. (2003) and von Eynatten (2004) values (Fig. 6.1). The WIP implicitly assumes that all Ca2þ in a
presented a quantitative approach, the t-index, to statistically silicate rock is contained in silicate minerals. This simplification
model linear compositional and weathering trends. If corroborated is a source of imprecision of the index, particularly if larger
by further studies the t-index will likely lead to more quantitative amounts of carbonate detritus or cements are present in the rock.
definitions of weathering trends. A related statistical approach has More problematic still is the lack of consideration of a relatively
been taken by Ohta & Arai (2007), whose W index is based on prin- immobile reference phase like Al2O3 in the formula, which
cipal component analysis of eight major oxides. As yet, it has been would help to monitor relative changes of composition of the
applied only to igneous rocks. relevant mineral components.
Geochemical estimations of weathering effects need to be con- The disadvantages of the WIP are overcome in the Chemical
sidered carefully, because the major cations Naþ, Kþ and Ca2þ Index of Alteration (CIA) using whole-rock geochemical data of
may also be mobile under diagenetic conditions (Wintsch & major element oxides (Nesbitt & Young 1982). The index is essen-
Kvale 1994). Geochemical weathering proxies make use of the tially based on the same considerations that led Kramer (1968) to

Fig. 6.1. Relationship between two


weathering proxies, WIP (Parker 1970) and
CIA (Nesbitt & Young 1982). Shown are
data obtained from the matrix of
Neoproterozoic glacial diamictites from
South China (Dobrzinski et al. 2004), the
Port Askaig Formation, Scotland (Panahi &
Young 1997, and this study, Table 6.3), the
Ghaub Formation of northern Namibia
(Table 6.4), and the Smalfjord and
Mortensnes formations, northern Norway
(Table 6.4). Also shown are data from
non-glacial shales and siltstones including
those underlying, intercalated with and
overlying the South China glacial
diamictites, and well-weathered Palaeozoic
shales from the Cantabrian mountain belt of
Hercynian age in northern Spain
(Table 6.1).
A REVIEW OF THE CHEMICAL INDEX OF ALTERATION 83

Table 6.1. Sampled formations and compositional groupings discussed in this contribution

Lithostratigraphic units, Formation Region Rock types Sampled Depositional age Ref.
terminology, this paper lithology

Devonian and Vergaño, Vañes, Northern Spain Quartz Shale, siltstone Late Devonian to Wagner & Wagner-Gentis
Carboniferous Carmen, Potes sandstone, Carboniferous (1963), IGME (1984), ITGE
formations, Group, Murcia sandstone, (1994), Keller et al. (2008)
Cantabria siltstone,
shale
Mortensnes and Mortensnes, Northern Diamictite, Siltstone, shale Ediacaran, Rice & Hoffmann (2001),
Smalfjord Smalfjord Norway shale Neoproterozoic Arnaud & Eyles (2002), Rice
formations et al. (2011)
Port Askaig Port Askaig Scotland Diamictite Diamictite Cryogenian, Panahi & Young (1997),
Formation matrix: shale, Neoproterozoic Arnaud & Eyles (2006),
siltstone Benn & Prave (2006),
Arnaud & Fairchild (2011)
Ghaub diamictites, Ghaub Otavi Diamictite, Silicate shale Cryogenian, Hoffman & Schrag (2002),
Namibia Mountains, siltstone matrix, Neoproterozoic Eyles (2007), Hoffman
northern siltstone (2011)
Namibia
Non-glacial, South Doushantuo, Yangtze Carbonate, Shale Ediacaran, Dobrzinski et al. (2004),
China (post-glacial) Jinjiadong platform, shale Neoproterozoic Dobrzinski & Bahlburg
South China (2007), Zhang et al. (2011)
Glacial, South China, Nantuo, Hongjiang, Yangtze Diamictite Diamictite Cryogenian, Dobrzinski et al. (2004),
(upper diamictites) Leigongwu platform, matrix: shale, Neoproterozoic Dobrzinski & Bahlburg
South China siltstone (2007), Zhang et al. (2011)
Non-glacial, South Datanpo, Xianmeng, Yangtze Carbonate, Shale Cryogenian, Dobrzinski et al. (2004),
China (inter-glacial) Lantian platform, shale Neoproterozoic Dobrzinski & Bahlburg
South China (2007), Zhang et al. (2011)
Glacial, South China, Dongshanfeng, Yangtze Diamictite Diamictite Cryogenian, Dobrzinski et al. (2004),
(lower diamictites) Jiangkou, Tie-si-ao platform, matrix: shale, Neoproterozoic Dobrzinski & Bahlburg
South China siltstone (2007), Zhang et al. (2011)
Non-glacial, South Xieshuihe, Yangtze Sandstone, Shale Tonian and Dobrzinski et al. (2004),
China (pre-glacial) Wuqiangxi, Zitang platform, shale Cryogenian, Dobrzinski & Bahlburg
South China Neoproterozoic (2007), Zhang et al. (2011)

conclude that monitoring the hydrolysis of feldspar and volcanic Ca in phosphates may not be considered in calculation of the
glass and the respective changes in the content of the major CIA, because CIA values will increase by not more than c. 1
cations offers the best quantitative measure of chemical weather- unit if all P2O5 present in common siliciclastic rocks is assigned
ing. It represents a ratio of predominantly immobile Al2O3 to the to apatite.
mobile cations Naþ, Kþ and Ca2þ given as oxides. The CIA is The CIA has been successfully applied in a large number of
defined as studies involving glacial deposits (e.g. Young & Nesbitt 1999;
Young 2001; Condie et al. 2001; Scheffler et al. 2003; Dobrzinski
  et al. 2004; Young et al. 2004; Rieu et al. 2007) and other deposi-
Al2 O3 tional environments (e.g. Gallet et al. 1998; Aristizábal et al. 2005;
CIA ¼  100 (2) Kahmann et al. 2008).
Al2 O3 þ Na2 O þ K2 O þ CaO

where the major element oxides are given in molecular pro- A brief discussion of WIP and CIA
portions. CaO* represents the CaO content of silicate minerals
only (Fedo et al. 1995) and thus eliminates one of the disadvan- Results of a comparison of WIP and CIA are displayed in Fig. 6.1.
tages of the WIP. Kaolinite has a CIA value of 100 and represents It combines data on Neoproterozoic glaciomarine deposits from
the highest degree of weathering. Illite is between 75 and c. 90, South China (Dobrzinski et al. 2004) and the Port Askaig For-
muscovite is at 75, and the feldspars at 50. Fresh basalts have mation (Scotland; Panahi & Young 1997; and our own analyses)
values between 30 and 45, fresh granites and granodiorites with new data on the Ghaub Formation (Namibia) and the
between 45 and 55 (Nesbitt & Young 1982; Fedo et al. 1995; Mortensnes and Smalfjord formations of northern Norway
Fig. 6.1). (Tables 6.1 to 6.4). For comparison we also show hitherto unpub-
Careful attention has to be paid to the potential presence of lished results on a number of late Devonian and early Carbonifer-
clastic carbonate grains or of carbonate cement in the sedimentary ous siliciclastic shales and siltstones from the Palentinian foreland
rock. If undetected but abundant, both would lead to very low and basin (Keller et al. 2007, 2008) of the Hercynian Cantabrian
unrealistic CIA values. The presence of carbonate grains or Mountain belt in northern Spain (Table 6.1). In the Late Palaeo-
cements (calcite or dolomite) and the ratio of calcite to dolomite zoic, the Palentinian Basin was located at near-equatorial latitudes
have to be determined petrographically. Both can be effectively (Weil et al. 2001). We use the Cantabrian data as an example of
accounted for geochemically by determining total MgO and CaO well-weathered detritus developed under non-glacial conditions.
and the total inorganic carbon (TIC) content of a sample, all of The WIP values of the glacial Ghaub, Port Askaig, Mortensnes
which can then be used to calculate relative contributions to and Smalfjord formations fall between 29 and 107, with a majority
CO2, and finally CaO* (Fedo et al. 1995; Tables 6.2, 6.3, 6.4). between 40 and 80 (averages: Ghaub, 76; Mortensnes and
84 H. BAHLBERG & N. DOBRZINSKI

South China consist of a mixture of detritus of glacial and non-


glacial weathering provenance, albeit in different proportions.
The initial incorporation of weathered older detritus into the Neo-
proterozoic glacial deposits was demonstrated by Dobrzinski et al.
(2004) by upward trends to lower CIA values in several sections in
South China (Fig. 6.2). Similar observations were made by Panahi
& Young (1997) in the Port Askaig Formation, Scotland.
K-feldspar and plagioclase are relatively common in the
Chinese glacial and non-glacial deposits. The presence of labile
feldspar indicates that the non-glacial rocks did not form predomi-
nantly under humid and warm –humid conditions. Regarding the
weathering sensitive cations Kþ and Naþ, average values for
Na2O in both the glacial and non-glacial deposits from South
China are 1.3 wt%, whereas the average value of detritus derived
from the upper continental crust is 1.2 wt% (PAAS, Post-Archaean
average Australian shale, Taylor & McLennan 1985). Condie
(1993) gives an estimate of Na2O in the average Proterozoic
shale (APS) derived from the Proterozoic upper crust of 1.1 wt%.
Average K2O values of 3.3 wt% coincide with the Proterozoic
upper crust, upper continental crust and Average Proterozoic
shale values of 3.3, 3.4 and 3.6 wt%, respectively (Condie 1993;
McLennan 2001). Considering the calculations of the respective
indices, an increase in, for example, Na2O of 0.5 wt% causes a
decrease in the CIA of c. 2.5 units, and an increase of WIP of
5 units. A decrease by 0.5 wt% results in similar, but opposing,
changes in magnitude. In the case of K2O the effect is similar;
however, the change in CIA units is c.1.5 per 0.5 wt% change.
The respective changes in WIP are c. + 4 units.
Our data demonstrate that both the WIP and CIA appear to be
equally sensitive to small changes in the concentrations of the
major cations, and thus, by inference, of dominant warm –humid
or dry –cold weathering conditions. Disadvantages of the WIP
include the lack of consideration of mobile relative to immobile
phases, and the fact that it is based on whole-rock CaO and not
on CaO* of the silicate fraction only. Another advantage of the
CIA lies in the way it permits the definition and prediction of
weathering trends of silicate rocks (Nesbitt & Young 1982,
Fig. 6.2. The Yangjiaping section in South China and its Neoproterozoic 1984). Data and trends can then be displayed well in A – CN– K
climate record indicated by CIA values. PG, preglacial units; LD, lower (Al2O3 – CaO* þ Na2O – K2O) ternary diagrams (Fig. 6.3;
diamictite; DF, Datangpo Formation; UD, upper diamictite; DSF, Doushantuo Nesbitt & Young 1984). The combination of these features
Formation (Table 6.1). Note the initially high CIA values in the lower part of the makes the CIA the presently preferred weathering index.
upper diamictite unit followed by a decrease interpreted as the initial and then
waning incorporation of chemically weathered detritus into the glacial deposit.
Source: modified from Dobrzinski et al. (2004).
Provenance and reconstruction of original compositions

Provenance pathways and the distribution of lithologies in source


Smalfjord, 56; Port Askaig, 66; Fig. 6.1). In contrast, the Cantab- regions exert a first-order control on the composition of siliciclastic
rian formations demonstrate their well-weathered character with deposits (Johnsson 1993). Differences in provenance are therefore
values between 0 and 75, with a majority below 40 (average, reflected also in the weathering indices (Fedo et al. 1995). In the
19). Considering these formations, the WIP appears to differentiate ternary A –CN –K diagram of Figure 6.3, the provenance compo-
well between glacial and non-glacial weathering types. sitions and weathering trends can be depicted and predicted.
The CIA values of the glacial units fall within 55 and 77 with a Hydrolytic weathering leads to a loss of Naþ and Ca2þ and
majority below 70 (averages: Ghaub, 63; Mortensnes and Smalf- changes rock compositions towards the A apex and ever higher
jord, 66; Port Askaig, 70; Fig. 6.1). These are typical values of CIA values.
unweathered to slightly weathered detritus and conform well to a Transport sorting exerts another major influence on the compo-
glacial weathering regime. The Cantabrian shales and siltstones, sition of clastic sediments. Along a river and transport path, labile
in turn, reflect the dominance of warm –humid weathering mineral grains will be comminuted preferentially. In the presence
conditions with a CIA between 56 and 100 and an average of 79. of chemical weathering, labile grains like feldspar will be prone to
The glacial and non-glacial deposits from South China seem to progressive decay along cleavages, cracks and other zones of
complicate the picture because they both have relatively low as lattice weakness. Feldspars will eventually be transformed into
well as high WIP values typical of intense and weak chemical clay minerals and a marked compositional difference, in addition
weathering, respectively (Fig. 6.1). The WIP range of the glacial to the comminution of mineral grains, between source and sedi-
deposits is 31–121 with an average of 57. The respective values ment will become evident (Johnsson et al. 1988). Sorting will
of the non-glacial formations are 2 –118 and 52. CIA values are thus lead to a higher proportion of clay minerals downriver and
distributed between 50 and 85 in the glacial deposits (average, consequently higher CIA values (Nesbitt et al. 1997; Fig. 6.3).
68) and between 55 and 85 with the rare sample having a value Thus, when CIA values of different successions are compared,
of close to 100 (average, 71; Fig. 6.1; Dobrzinski et al. 2004). similar grain sizes need to be considered.
The values of both indices seem to demonstrate that both the Only exceptional circumstances permit the transfer of labile
glacial and non-glacial Neoproterozoic sedimentary rocks in mineral grains from source to sink without any mineralogical
A REVIEW OF THE CHEMICAL INDEX OF ALTERATION 85

Fig. 6.3. Major oxides in molecular ratios with compositions of typical magmatic source rock types (Fedo et al. 1995) and average compositions of the UCC
through time (UCC, McLennan 2001; PAAS, Taylor & McLennan 1985; all others, Condie 1993). Note that the lower part of the diagram with A , 40 is not shown. 1,
gabbro; 2, tonalite; 3, granodiorite; 4, granite; 5, A-type granite; 6, charnokite; 7, potassic granite. Ideal weathering trends of UCC-type source lithologies would be
parallel to the predicted weathering trend (Nesbitt & Young 1984). The statistically modelled t-index weathering trend (von Eynatten 2004) is based on data obtained from
the world’s major rivers and erosional products of major denudation areas (McLennan 1993). Trends of K-metasomatism and its graphical correction to pre-metasomatic
values on predicted weathering trends is indicated by grey arrows originating or pointing at the K apex of the diagram (Fedo et al. 1995). Alteration trend of
Neoproterozoic glacial successions estimated from the distribution of samples in Fig. 6.4. The dotted double-headed arrow outlines the effect of grain sorting on the
chemical composition of sediments and consequently on CIA values; c, coarse, sand; f, fine, clay (Nesbitt et al. 1997).

and thus compositional changes, that is, without the influence of effect would disturb the pronounced negative correlation
any chemical weathering. In such an ideal case transport will between the K/Cs ratio and weathering intensity. The distribution
lead only to comminution of mineral grains; the final fine-grained of data in the A –CN –K diagram (Fig. 6.4) indicates that in
deposit will have the same composition as the coarser source particular some of the Chinese sedimentary rocks discussed in
(Nesbitt & Young 1996), and CIA values will be unaffected. this paper show a weak K-metasomatic effect (Dobrzinski et al.
Diagenetic addition of potassium to a deposit during illitization 2004). If projected back to the predicted weathering trend origi-
of kaolinite and the replacement of plagioclase or illite by nating from average granodiorite or the upper continental crust
K-feldspar, the so called K-metasomatic effect of Fedo et al. (UCC) (Fig. 6.3), the originally higher CIA values of these
(1995), causes weathered compositions to plot closer to the K samples can be reconstructed (Tables 6.3 and 6.4), thus permitting
apex (Fig. 6.3), thus resulting in diagenetically lowered CIA the subsequent inference of original provenance compositions
values. This can be checked by comparisons with average upper (Figs 6.3 and 6.4; Fedo et al. 1995).
crustal compositions (see above) and by comparing CIA values As the CIA is founded essentially in considerations of the mobi-
with K/Cs ratios (McLennan et al. 1993); a K-metasomatic lity of the major cations, geochemical estimations of provenance

Fig. 6.4. A–CN– K (Al2O3 –CaO*þ


Na2O– K2O) diagram with data of
Neoproterozoic glacial units and reference
units and their composition (PAAS, Taylor
& McLennan 1985; UCC, McLennan 2001;
all others, Condie 1993). Note that the lower
part of the diagram with A , 40 is not
shown. The left side of the figure shows the
range and averages (black dots) of CIA
values of the Neoproterozoic units corrected
for K-metasomatism according to Fedo
et al. (1995). For comparison, the figure also
shows data obtained from shales of several
late Devonian and Carboniferous
formations from the Palentinian foreland
basin of the Hercynian Cantabrian mountain
belt in northern Spain (Keller et al. 2008),
which conform to the predicted weathering
trend (Tables 6.1, 6.2; Nesbitt & Young
1984).
86 H. BAHLBERG & N. DOBRZINSKI

Fig. 6.5. Zr/Sc v. Th/Sc diagram of


McLennan et al. (1993) demonstrating that
all studied sedimentary rocks have an upper
crustal composition and were affected to
minor degrees by recycling, which is
indicated by high Zr/Sc values. This effect
is most prominently shown in some of the
Late Palaeozoic sedimentary rocks from
Cantabria. It is noteworthy that all the
Neoproterozoic samples cluster around the
composition of the average Proterozoic
shale (APS, Condie 1993). Reference units
and their composition: PAAS, Taylor &
McLennan (1985); UCC, McLennan
(2001); all others, Condie (1993).

should be combined preferably with ratios of immobile elements (sub)parallel to the predicted weathering trend (Nesbitt & Young
including the high field strength elements La, Th, Sc and Zr 1984). Corrections for K-metasomatism and resulting increases
(Bhatia & Crook 1986; McLennan et al. 1993; Bahlburg 1998). in CIA values would be correspondingly smaller. A correction to
All the Neoproterozoic sedimentary rocks considered in this paper the predicted weathering trend thus represents a maximum correc-
have Zr/Sc and Th/Sc ratios similar to or higher than the Upper tion if average granodiorite or UCC is the starting point.
Continental Crust (UCC) and PAAS (Fig. 6.5; Taylor & McLen- The average compositions of post Archaean shale and Protero-
nan 1985; McLennan 2001). The rock compositions all cluster zoic shale (PAAS, Taylor & McLennan 1985; APS, Condie
around the composition of the APS (Fig. 6.5; Condie 1993) as a 1993) plot below the predicted weathering trend towards the K
proxy for material derived from the average Proterozoic upper apex of the A –CN –K diagram. This may be due either to starting
crust. This provides another indication that processes beyond orig- compositions richer in Kþ than UCC and the average Proterozoic
inal weathering and recycling did not significantly alter the compo- upper crust (Fig. 6.3), or a mild K-metasomatic effect including the
sition of the studied Neoproterozoic sedimentary rocks. conversion under increased temperature of illite to secondary
K-feldspar (Nesbitt & Young 1989; Fedo et al. 1995).
The composition of almost all considered Neoproterozoic units
Weathering trends and samples appears to systematically deviate from the predicted
weathering trend, in a way here preliminarily called the observed
Thermodynamic, kinetic, experimental and observational evi- Neoproterozoic alteration trend (Fig. 6.3). In discussing this
dence define weathering trends for silicate rocks that are (sub)par- trend, two things have to be borne in mind: the similarity of the
allel to the CN– A join of the A – CN– K diagram (Nesbitt & Young estimates for the UCC and the average Proterozoic crust in A –
1982, 1984). Weathering of average granodiorite or UCC along the CN –K space, and the fact that glaciogenic deposits commonly rep-
predicted weathering trend (Fig. 6.3) will result in the transform- resent averaged samples of the available upper crust exposed
ation of labile components including the feldspars first to illite, to denudation.
thus causing the sample’s composition to plot ever closer to the It is a reasonable assumption that the original average geochem-
A –K join and the illite composition in A –CN –K space (Nesbitt ical composition of the silicate sources feeding the Neoproterozoic
& Young 1984; Fig. 6.3). Increasing formation of kaolinite glaciogenic deposits was similar to the average upper crustal com-
during progressive weathering will curve the weathering trend positions. Repeated weathering and recycling processes may lead
towards the A apex when approaching the A – K join (Nesbitt & to a stepwise increase of Kþ in the resulting deposits, as indicated
Young 1984) in a way similar to the calculated weathering trend by the offsets of the average shale compositions from the predicted
of von Eynatten et al. (2004; Fig. 6.3). The validity of this weathering trend (Fig. 6.3). The observed Neoproterozoic altera-
general weathering behaviour and trend is substantiated by the tion trend may consequently reflect the degree to which unweath-
fact that the compositions of shale composites like PAAS and ered and weathered crust and recycled sedimentary rocks have
APS as averaged samples of the weathered and recycled upper been mixed and incorporated into the glaciogenic deposits. The
crust plot very near the predicted weathering trend (Fig. 6.3). correction of CIA values for a K-metasomatic effect according
Also, the Late Palaeozoic reference samples from Cantabria, to Fedo et al. (1995) will thus reconstitute CIA values to the pre-
Spain, which were well weathered under non-glacial conditions, dicted weathering trend originating in fresh upper crust, which
plot along this trend as predicted by CIA systematics (Figs 6.3 & may, however, be unrepresentative of the actually eroded conti-
6.4; Nesbitt & Young 1982, 1984). nental surface. If CIA values were referred back only to unweath-
In many cases, sedimentary rocks do not plot on the predicted ered compositions, irrespective of the regional context and the
weathering trend but below it, thus reflecting an increase in Kþ recycling history, estimates of weathering effects during the last
in the samples (e.g. Panahi & Young 1997; Rieu et al. 2007). A sedimentary cycle will be exaggerated and may consequently
correction of this K-metasomatic effect according to Fedo et al. be misleading.
(1995) of those CIA data points below the predicted weathering
trend line results in an increase of CIA values (Figs 6.3, 6.4;
Tables 6.3, 6.4). This increase depends on the compositional The evidence of glaciomarine deposits
difference between the sample and the predicted original value
and reflects the degree of K-metasomatism having affected a Successions of modern and ancient glacial deposits and of glacially
sample’s detritus if the original composition was granodioritic derived detritus are ideally characterized by low CIA values
and close to that of the UCC (Fig. 6.3; Fedo et al. 1995). If the orig- reflecting the dominance of physical over chemical weathering
inal composition was richer in Kþ than granodiorite or UCC, the processes (Nesbitt & Young 1982; Young 2001). A typical
predicted weathering trend would shift to the right but remain example is the Palaeoproterozoic Gowganda Formation (Young
A REVIEW OF THE CHEMICAL INDEX OF ALTERATION 87

Table 6.2. Major element whole-rock compositions and weathering indices of shales

Sample* SiO2 Al2O3 Fe2O3 MgO CaO‡ Na2O K2O TiO2 P2O5 MnO LOI Total TC TIC WIP CIA†

BT Nr.64 96.83 1.85 0.60 0.04 – – – 0.23 0.08 0.01 0.42 100.06 0 100
MB 87-K 95.33 2.40 0.61 0.02 – – 0.04 0.23 0.04 0.01 0.72 99.41 0 98
MB P-05 87.40 3.54 0.83 0.12 2.49 0.07 0.41 0.20 0.07 0.05 3.02 98.21 0.72 0.54 11 86
MB P-25 83.00 3.47 2.19 0.29 0.61 0.03 0.73 0.09 0.03 0.03 2.20 93.28 9 64
MB P-16 60.87 19.13 5.87 0.60 3.14 0.60 2.68 1.02 0.09 0.07 6.79 100.87 0.94 0.45 38 83
BT PO 28 84.47 6.02 2.79 0.66 1.40 0.15 1.08 0.62 0.08 0.03 1.57 100.28 16 60
BT PO 29 92.10 3.78 0.82 0.14 1.02 0.21 0.76 0.11 0.02 0.03 1.57 101.58 11 56
PO 1 75.34 8.04 7.37 1.04 0.55 0.03 1.06 0.52 0.07 0.18 4.30 98.50 0.74 0.58 14 87
PO 2-1 79.79 9.72 3.36 0.62 0.05 0.10 1.45 0.56 0.04 0.01 2.55 98.30 15 84
PO 3 88.09 7.16 1.74 0.15 – – 0.75 0.53 0.02 0.00 1.62 100.06 7 90
PO 4-2 74.77 9.96 6.38 0.42 0.92 0.40 1.23 0.58 0.09 0.15 3.67 98.56 0.30 0.16 18 83
PO 4-4 67.92 6.79 9.36 0.87 4.40 0.34 1.43 0.40 0.07 0.28 6.74 98.59 1.31 1.21 29 76
PO 5 37.88 3.92 4.89 0.45 27.10 0.03 0.49 0.19 0.07 0.09 23.38 98.49 6.04 5.92 75 87
PO 6 45.19 2.17 1.41 0.34 26.81 0.02 0.16 0.11 0.14 0.12 21.91 98.38 5.86 5.76 71 91
PO 9 54.00 3.36 2.81 0.39 19.65 0.12 0.34 0.16 0.16 0.17 16.83 97.99 4.42 4.33 55 86
PO 11 84.53 6.15 3.20 0.18 0.34 0.05 0.93 0.46 0.09 0.03 2.04 98.35 10 78
BT PO 4 71.88 6.79 4.90 1.22 4.29 0.25 1.19 0.37 0.08 0.08 6.91 97.97 1.80 1.59 27 80
BT PO 5 68.46 6.44 5.28 1.15 6.87 0.08 0.94 0.35 0.08 0.08 9.26 98.99 2.40 2.24 29 85
BT PO 16 72.33 7.15 8.37 1.27 2.11 0.21 1.04 0.46 0.20 0.13 7.16 100.44 1.83 1.22 20 83
BT PO 30/2 68.44 12.29 5.23 1.17 2.68 0.31 1.72 0.70 0.09 0.06 5.60 98.30 1.21 0.75 28 84
FC 1 73.68 2.87 7.31 0.50 6.40 0.15 0.55 0.21 0.14 0.24 7.90 99.98 1.92 1.81 24 77
FC 2 77.89 3.83 10.22 0.80 0.40 0.15 0.53 0.24 0.09 0.21 5.60 99.97 9 71
FC 3 87.51 4.35 4.40 0.43 0.07 0.20 0.54 0.33 0.07 0.06 2.20 100.18 8 81
FC 5 74.16 6.86 11.27 0.57 0.13 0.40 1.20 0.50 0.14 0.23 4.60 100.10 16 76
CRM 1A 85.36 6.00 3.72 0.28 0.38 0.04 1.21 0.41 0.10 0.05 2.10 99.67 12 74
CRM 2 84.48 4.90 3.58 0.30 1.78 0.05 1.03 0.34 0.08 0.06 3.10 99.72 15 52
CRM 3 84.32 6.80 3.88 0.17 0.11 0.40 1.05 0.41 0.11 0.06 2.40 99.74 13 77
CM 1A 79.37 3.11 9.94 0.86 0.64 0.11 0.36 0.21 0.06 0.20 5.00 99.87 8 64
CM 3A 84.89 3.25 6.37 0.60 0.22 0.12 0.36 0.25 0.04 0.16 3.60 99.87 6 77
CM 3B 90.78 3.82 2.13 0.32 0.04 0.22 0.47 0.25 0.03 0.01 2.00 100.08 7 80
CM 4A 89.18 4.10 3.27 0.44 0.04 0.23 0.51 0.30 0.05 0.01 1.40 99.55 8 80
CM 4A_2 89.09 4.16 3.43 0.45 0.05 0.23 0.51 0.31 0.06 0.01 1.40 99.71 8 80
CM 5A 76.72 5.54 9.19 1.02 0.34 0.36 0.81 0.61 0.08 0.33 4.80 99.82 14 73
VN 1 79.31 10.69 3.19 0.82 0.06 0.51 1.65 0.65 0.08 0.02 2.50 99.52 21 80
VN 2 79.70 10.79 2.82 0.81 0.06 0.52 1.76 0.66 0.10 0.01 2.50 99.77 22 79
VN 3A 81.63 9.43 2.84 0.79 0.07 0.55 1.50 0.59 0.09 0.02 2.10 99.64 20 78
VN 3B 82.34 8.93 2.78 0.75 0.06 0.52 1.50 0.55 0.10 0.01 2.30 99.87 20 78
VN 4 79.00 11.45 2.83 0.82 0.04 0.48 2.00 0.66 0.07 0.01 2.60 100.00 24 79
VN 5 81.63 9.40 2.62 0.75 0.11 0.42 1.44 0.81 0.14 0.01 2.40 99.77 18 79

*Formations, late Devonian– Carboniferous, of the Cantabrian Mountains, northern Spain: Vergaño, VN; Vañes, CM; Carmen, FC, CRM; Potes Group, PO, BT PO; Murcia,
BT, MB. Major elements were determined by ICP-ES after a LiBO2 fusion at ACME Analytical Labs, Vancouver, Canada.

CIA has been calculated according to Nesbitt and Young (1982) and Fedo et al. (1995). Ca in phosphates has not been considered in the calculation of the CIA, because CIA
values increase by only c.1 unit if all P2O5 is assigned to apatite.

Calculation of CaO* is based on values of TC (total carbon) and TIC (total inorganic carbon) obtained by analysis using CS-Mat 5500, in the Department of Geology and
Paleontology, Münster University. Only samples with increased values of CaO were selected. Those with low values did not show a reaction with HCl (10%); dolomite
was not observed in thin section. CaO was consequently considered to be equal to CaO*.
–, below detection limit; blank spaces, not measured.

& Nesbitt 1999). Within this unit, rocks of diamictite facies com- pre-weathered rocks and soils generally have lower CIA values
monly have the lowest CIA values (between 50 and 70). Diamic- and plot lower in the A –CN –K triangle than associated muds,
tites in general represent either lodgement tills or moraine because feldspars and other labile minerals tend to be concentrated
material that was rapidly redeposited by mass wasting in glacio- in the coarser-grained fraction (Nesbitt et al. 1996; Fig. 6.3).
marine environments (e.g. Scheffler et al. 2003; Dobrzinski et al. Furthermore, sorting or mixing of known and cryptic sources
2004; Young et al. 2004; Rieu et al. 2007; Fig. 6.4). Such deposits may result in an unexpected change of the amount of feldspar in
with low CIA values are frequently associated with finer-grained a deposit.
and laminated glaciomarine sandstones, siltstones and shales, A sedimentary enrichment of K-feldspar relative to the assumed
which may include ice-rafted debris, with higher CIA values source rock causes the distribution of data in the A –CN –K
between 70 and 85 (Fig. 6.4; Panahi & Young 1997; Dobrzinski diagram to mimic a K-metasomatic effect. In such a case the pre-
et al. 2004; Young et al. 2004; Rieu et al. 2007). dicted weathering trend would shift to the right and point more
Lower CIA values in glaciomarine deposits can be caused by closely to a muscovite rather than an illite composition on the
several factors acting alone or in concert. K-metasomatism in con- A –K join of the diagram, or a CIA of 80 instead of 90 (Fig. 6.4;
nection with the conversion of illite to K-feldspar, or the illitization Nesbitt et al. 1997). A correction of the CIA data for an alleged
of kaolinite, is the most commonly cited cause (Nesbitt & Young K-metasomatism according to Fedo et al. (1995; Figs 6.3 and
1989; Fedo et al. 1995). However, the grain-size of the analysed 6.4) would in this case lead to an overadjustment. A cause of
rock is also an important factor. Sands formed from chemically increased or high CIA values in glaciomarine rocks may also be
88 H. BAHLBERG & N. DOBRZINSKI

Table 6.3. CIA and WIP values of the Neoproterozoic Port Askaig Formation, the incorporation of older, strongly weathered material into the
Scotland, based on the data of Panahi & Young (1997) and our own analyses glacial deposits (Nesbitt & Young 1997; Dobrzinski et al. 2004;
Fig. 6.2).
Sample TC TIC WIP CIA* CIAcorr† Many Neoproterozoic glaciomarine successions are character-
ized by a deviation of the observed weathering trend from the pre-
94-102 0.58 0.45 33 68 84 dicted one and towards the K apex (Figs 6.3 and 6.4), including the
94-106 0.68 0.44 56 65 73 lower and upper diamictites of the Yangtze platform in South
94-108 0.90 0.62 61 64 72 China (Table 6.1; Dobrzinski et al. 2004), the Port Askaig For-
94-78 4.14 3.39 75 74 83 mation (Nesbitt & Young 1997; Tables 6.1 and 6.3), the Ghaub
94-79 4.52 4.17 72 74 86 Formation of northern Namibia, the Norwegian Smalfjord and
94-87 1.95 1.23 57 63 72 Mortensnes formations (Tables 6.1 and 6.4), and the Fiq Formation
94-97 0.56 0.45 36 64 71
in Oman (Rieu et al. 2007). The same effect is also evident in the
95-62 3.39 2.85 69 77 87
glacial Ordovician Table Mountain Group in South Africa (Young
95-66 5.80 5.00 84 74 87
et al. 2004).
95-67 5.11 4.75 75 72 87
95-69 3.37 2.91 75 70 86
Regarding the Port Askaig and Fiq formations, the Chinese
95-70 3.92 3.37 70 68 78 diamictites and the Table Mountain Group, K-metasomatism has
95-71 6.19 4.90 85 71 87 in fact been cited as the cause of the deviation from the commonly
95-73 4.93 4.04 79 71 84 used predicted weathering trend (Nesbitt & Young 1997;
95-76 5.21 4.25 78 75 83 Dobrzinski et al. 2004; Young et al. 2004; Rieu et al. 2007).
95-79 3.57 2.93 64 73 87 However, almost all the mentioned cases include analyses of
95-80 5.03 4.66 80 70 81 matrix material of glaciomarine diamictites and associated coarse-
95-83 1.69 1.32 53 66 74 grained strata. Some of the Neoproterozoic diamictites, including
95-84 1.56 0.51 56 61 66 the ones from South China (Dobrzinski & Bahlburg 2007), have an
impure shaley matrix rich in silt or fine sand. The presence of this
*Ca in phosphates has not been considered in the calculation of the CIA, because coarser material in the matrix of the diamictites may have contrib-
CIA values increase by only c.1 unit if all P2O5 is assigned to apatite. CIA has uted consequently to a deviation of the observed weathering trends
been calculated according to Nesbitt & Young (1982) and Fedo et al. (1995). from the predicted one.

CIAcorr: CIA value corrected to the predicted weathering trend according to Fedo A combination of grain-size effects and the variable incorpor-
et al. (1995). ation of older weathered material into the Neoproterozoic glacio-
Calculation of CaO* is based on values of TC (total carbon) and TIC (total marine deposits may account for the measured CIA values
inorganic carbon) obtained by analysis using CS-Mat 5500, in the (Fig. 6.4; Tables 6.2 to 6.4). In view of the corrected CIA values
Department of Geology and Paleontology, Münster University.

Table 6.4. Major element whole-rock compositions and weathering indices

Sample*† SiO2 Al2O3 Fe2O3 MgO CaO} Na2O K2O TiO2 P2O5 MnO LOI Total TC TIC WIP CIA§ CIAcorrk

GF 1 70.77 6.66 15.52 0.4 0.05 0.05 3.17 1.24 0.12 0.01 1.50 99.50 0.03 –‡ 29 65 87
GF 2 81.78 8.45 2.31 0.42 0.01 0.1 5.22 0.43 0.06 0.01 1.20 99.98 0.01 – 46 59 87
GF 3 72.25 7.50 10.25 0.95 0.47 0.09 4.79 0.34 0.35 0.04 2.70 99.73 0.06 0.01 45 55 79
GF 4 28.42 6.46 2.82 10.93 19.89 0.86 2.14 0.40 0.12 0.06 27.80 99.91 7.48 7.26 107 63 72
GF 5 38.51 7.33 3.44 9.29 15.52 0.83 2.55 0.43 0.14 0.06 21.70 99.80 5.66 5.34 94 64 74
GF 7 27.84 4.28 3.20 12.98 19.08 0.22 1.79 0.24 0.10 0.08 30.00 99.82 7.91 7.65 102 65 81
GF 8 59.81 16.53 7.32 3.15 0.70 0.74 5.36 0.87 0.18 0.11 5.00 99.80 0.23 0.09 63 69 80
GF 9 33.92 7.96 4.97 10.47 15.03 0.91 2.82 0.47 0.14 0.14 23.00 99.83 6.09 5.62 99 64 74
GF 11 71.33 13.72 3.68 1.11 0.56 2.20 4.93 0.44 0.14 0.03 1.70 99.84 0.07 0.06 67 59 68
GF 12 73.82 12.63 3.39 0.96 0.31 2.06 4.49 0.42 0.12 0.05 1.60 99.86 0.03 0.02 61 59 68
GF 13 7.93 0.84 0.48 19.65 27.73 0.02 0.29 0.06 0.07 0.60 42.30 99.98 11.13 10.92 127 71 85
S1 73.68 10.67 3.59 1.89 1.42 1.45 2.82 0.5 0.12 0.03 3.20 99.38 0.36 0.27 46 65 70
S2 73.22 11.95 4.10 1.72 0.41 1.03 3.66 0.65 0.11 0.02 2.50 99.38 0.08 – 46 65 73
S3 67.36 13.36 5.94 2.94 0.27 1.41 3.91 0.83 0.17 0.10 3.10 99.40 0.09 0.04 55 67 74
S4 59.38 11.26 4.65 4.85 4.74 1.37 3.48 0.79 0.16 0.06 8.80 99.55 1.88 1.56 68 65 73
S5 69.49 12.75 3.91 2.33 1.48 1.91 3.39 0.61 0.11 0.03 3.40 99.42 0.46 0.32 57 65 70
S6 65.02 14.83 5.73 2.6 1.05 1.86 4.00 0.70 0.13 0.04 3.50 99.48 0.24 0.17 61 65 69
M1 64.35 15.52 6.23 3.13 0.33 2.05 3.58 0.78 0.14 0.05 3.30 99.48 0.06 0.06 59 68 72
M2 64.14 15.93 5.49 2.67 0.89 2.36 3.76 0.82 0.15 0.04 3.20 99.46 0.16 0.14 63 65 69
M3 69.19 10.5 3.88 3.05 2.98 1.45 2.45 0.51 0.11 0.05 5.20 99.38 0.95 0.73 50 68 73
M4 67.96 11.91 4.44 2.88 2.35 1.96 2.42 0.58 0.12 0.06 4.70 99.40 0.78 0.56 53 67 69
M5 65.56 11.05 5.19 3.01 3.43 1.85 2.13 0.54 0.10 0.10 6.20 99.17 1.11 0.79 52 67 69
M6 63.64 14.2 5.79 3.39 2.00 2.45 2.61 0.69 0.13 0.10 4.40 99.43 0.46 0.29 59 64 64

*Neoproterozoic Ghaub Formation (GF), northern Namibia, and the Smalfjord (S) and Mortensnes (M) formations, northern Norway. Samples of the Ghaub Formation
diamictites were taken on the Ghaub and Jakkalumuramba farms NE of the town of Otavi. Analyses represent matrix compositions.

Major elements were determined by ICP-ES after a LiBO2 fusion at ACME Analytical Labs, Vancouver, Canada.
–below detection limit.
§
CIA has been calculated according to Nesbitt & Young (1982) and Fedo et al. (1995). Ca in phosphates has not been considered in the calculation of the CIA, because CIA
values increase by only c.1 unit if all P2O5 is assigned to apatite.
k
CIAcorr: CIA value corrected to the predicted weathering trend according to Fedo et al. (1995).
}
Calculation of CaO* is based on values of TC (total carbon) and TIC (total inorganic carbon) obtained by analysis using CS-Mat 5500, in the Department of Geology and
Paleontology, Münster University.
A REVIEW OF THE CHEMICAL INDEX OF ALTERATION 89

almost meaningless as climate indicators when not considered in


the context of the stratigraphic framework and facies of the ana-
lysed sedimentary rock. This is demonstrated very well in a
study of the Fiq Formation in Oman (Rieu et al. 2007; Fig. 6.6),
an association of alternating glacial diamictites, debris-flow depos-
its, turbidite sandstones, hemipelagic shales, and wave-rippled
shoreface deposits. The facies changes are connected to coincident
changes between low CIA values of 62 –70 for the glacial diamic-
tite matrix, and high ones between 80 and 85 (corrected after Fedo
et al. 1995) for shales, siltstones and fine-grained sandstones.
Facies changes and CIA data give evidence of variations in the
intensity of chemical weathering as a function of climate oscil-
lations between glacial and warm –humid interglacial conditions
(Rieu et al. 2007). Scheffler et al. (2003) presented similar data
in a study of the waning phase of the Late Palaeozoic glaciation
in South Africa recorded in the Dwyka Group. In the Neoprotero-
zoic as well as in the Late Palaeozoic, these climate oscillations
together with the facies of the respective deposits give evidence
of a functioning hydrological cycle and sediment dispersal
systems. In the Neoproterozoic case these data are in direct oppo-
sition to assumptions of a totally ice-covered Snowball Earth as
proposed by Hoffman & Schrag (2002).
These CIA values are compatible with a prominent influence of
physical weathering on the production of the diamictite detrital
silicate matrix. Together with sedimentological features including
the presence of dropstones in laminated facies, these values
strengthen interpretations of a glacially connected origin of the
Ghaub diamictites.
The interpretation of the Neoproterozoic Smalfjord and Mor-
tensnes formations of northern Norway is less controversial and
Fig. 6.6. Variations in the CIA with stratigraphic height in the Fiq Formation,
centres around the question whether the strata represent redepos-
Oman. MB Fm., Masirah Bay Formation. Modified from Rieu et al. (2007). ited detritus originally produced in a glacial environment
(Arnaud & Eyles 2002) or deposits of direct glacial action
(Edwards 2004; Rice 2004). Application of the CIA cannot solve
these differences in a fundamentally sedimentological contro-
shown in Figure 6.4 and the discussion in the preceding para- versy. The uncorrected CIA values of the Smalfjord and Mor-
graphs, we conclude that in particular the high values are overcor- tensnes formations (Table 6.4) are uniformly low between 65
rected maximum values very likely exaggerating the magnitude of and 68, and between 64 and 74 (averages of 72 and 69, respect-
changes due to chemical weathering. ively) if corrected according to Fedo et al. (1995; Fig. 6.4). It is
thus legitimate to infer at least a glacial origin for the detritus.
The Neoproterozoic lower and upper glacial diamictites of the
Implications for the study of Neoproterozoic climate change Yangtze platform represent associations of lodgement till, subgla-
cial melt-out deposits of grounded glaciers, sediments of concen-
The origin of Neoproterozoic diamictite successions was debated trated density flows and glacial outwash, and turbidites and
long before the ‘Snowball Earth’ hypothesis was proposed laminated deposits containing ice-rafted debris (Dobrzinski &
(Kirschvink 1992; Hoffman et al. 1998; Hoffman & Schrag Bahlburg, 2007). CIA values of both diamictite units are
2002). Classic examples like the Ghaub diamictites in northern between 50 and 90 (Fig. 6.4; Table 6.1; Dobrzinski et al. 2004).
Namibia were repeatedly interpreted either as tillites and glacio- Correction of values following Fedo et al. (1995) does not
marine sediments (e.g. Gevers 1931; Martin 1964; Hoffmann & change their range but leads to a shuffling of values within this
Prave 1996; Hoffman & Schrag 2002) or as tectonically triggered range. The corrected average value is 72 (Fig. 6.4). The combi-
mass flow deposits (e.g. Schermerhorn 1974; Martin et al. 1985; nation of low values in the matrix of glacial diamictites and
Eyles & Januszczak 2007). Furthermore, there are also cases higher values in associated glacial facies is interpreted as a result
where alleged glacial deposits were proven to be in fact of deposi- of reworking of older, weathered material into the glacial deposits
tional origin unrelated to ice action and cold climates. Examples (Dobrzinski et al. 2004). This is evident in some sections showing
include purported tillites of Tremadoc age in the Andes of north- a decrease in CIA values upsection (Fig. 6.2).
western Argentina (Keidel 1943), which were demonstrated to The non-glacial units of the China platform consisting of
be coastal conglomerates (Bahlburg 1990), and conglomeratic preglacial siliciclastics, the shales and carbonates of the Datangpo
deposits in northern Chile allegedly linked to the Late Palaeozoic Formation intercalated between the glacial units, and the post-
Gondwana glaciation (Cecioni 1979, 1981), which were shown to glacial carbonates and shales of the Doushantuo Formation
be of mass-flow origin in turbidite complexes (Charrier 1986; (Table 6.1) have CIA values ranging between 59 and 97 with
Bahlburg & Breitkreuz 1993). abundant higher values. The corrected ranges are rather similar,
It is usually necessary to support one line of evidence with at with a corrected average of 76 (Fig. 6.4). The higher values of
least a second, independent one in order to have confidence that the non-glacial deposits relative to the diamictite successions are
an interpretation is correct. This applies not only to controversial taken to reflect more humid and warmer weathering and climate
sedimentological cases. The analysis of the CIA is a powerful conditions. Particularly noteworthy is the presence of K-feldspar
tool in studies of the palaeoclimatic record preserved in siliciclas- and plagioclase in the silicate fraction contained in the Marinoan
tic sedimentary successions. However, as discussed previously, cap carbonates in South China indicating the minor influence
numerous factors can influence the major element geochemical of chemical weathering on the silicate detritus in these strata
composition of a siliciclastic rock. Thus, CIA values alone are (Dobrzinski 2005).
90 H. BAHLBERG & N. DOBRZINSKI

Conclusions Charrier, R. 1986. The Gondwana glaciation in Chile: description of


alleged glacial deposits and paleogeographic conditions bearing on
The CIA (Nesbitt & Young 1982) is the most widely applied and the extension of the ice cover in Southern South America. Palaeogeo-
most indicative of the available weathering indices. When the graphy, Palaeoclimatology, Palaeoecology, 56, 151–175.
intricacies of the weathering systems and of applying the index Condie, K. C. 1993. Chemical composition and evolution of the upper
are appropriately considered (Fedo et al. 1995; Nesbitt & Young continental crust: contrasting results from surface samples and
1996; Nesbitt et al. 1997; Young 2001; and discussion of shales. Chemical Geology, 104, 1– 37.
Condie, K. C., Des Marais, D. J. & Abbott, D. 2001. Precambrian super-
Fig. 6.4), the index is a very valuable tool in the assessment of
plumes and supercontinents: a record in black shales, carbon isotopes,
past climate change as recorded by siliciclastic sedimentary
and paleoclimates? Precambrian Research 106, 239–260.
rocks, with one critical caveat: it needs to be applied in conjunction Crowell, J. C. 1999. Pre-Mesozoic ice ages: their bearing on understand-
with a comprehensive facies analysis. Concerning the Neoproter- ing the climate system. Geological Society of America Memoir, 192,
ozoic glacial periods, the combination of both data sets gives 1 –106.
strong evidence of (i) a functioning hydrological cycle, (ii) operat- Crowley, T. J. & North, G. R. 1991. Paleoclimatology. Oxford Mono-
ive sediment routing systems, and (iii) variable climate conditions graphs on Geology and Geophysics, 18, 349.
oscillating between dry –cool and glacial, and warm –humid and Dobrzinski, N. 2005. Das Paradoxon äquatornah abgelagerter
interglacial. These findings are incompatible with the hypothesis glazialer Sedimentfolgen: Sedimentologische und geochemische
of a totally ice-covered Snowball Earth. Klimaindizien von der neoproterozoischen Yangtze Plattform
(Südchina). Dissertation, Westfälische Wilhelms-Universität,
H. Rice (Vienna, Austria) kindly supplied samples of the Smalfjord and Mor-
Münster, Germany.
tensnes formations of northern Norway. We thank G. Young (London, Ontario,
Dobrzinski, N. & Bahlburg, H. 2007. Sedimentology and environ-
Canada) for sending us sample powders of the Panahi & Young (1997) samples
mental significance of the Cryogenian successions of the Yangtze
of the Port Askaig Formation. This study was supported by the German Research
platform, South China block. Palaeogeography, Palaeoclimatology,
Foundation DFG (grants Ba 1011/23-1,2,3). This paper is a contribution to
Palaeoecology, 254, 100–122.
IGCP project 512 ‘Neoproterozoic Ice Ages’. We thank A.R. Prave
Dobrzinski, N., Bahlburg, H., Strauss, H. & Zhang, Q. R. 2004.
(St. Andrews, Scotland), C. Augustsson and C. Reimann (Münster, Germany),
Geochemical climate proxies applied to the Neoproterozoic glacial
for commenting on earlier versions of the manuscript. We appreciate the very
succession on the Yangtze Platform, South China. In: Jenkins, G.,
constructive reviews by Hilmar v. Eynatten (Göttingen, Germany) and an anon-
McMenamin, M., McKay, C. P. & Sohl, L. (eds) The Extreme Pro-
ymous reviewer. B. Fister (Münster) kindly re-drafted Figure 6.6.
terozoic: Geology, Geochemistry and Climate. American Geophysi-
cal Union Monograph Series, 146, 13– 32.
References Duzgoren-Aydin, N. S., Aydin, A. & Malpas, J. 2002. Re-assessment of
chemical weathering indices: case study of pyroclastic rocks of Hong
Aristizábal, E., Roser, B. & Yokota, S. 2005. Tropical chemical weath- Kong. Engineering Geology, 63, 99 –119.
ering of hillslope deposits and bedrock source in the Aburrá Valley, Edwards, M. B. 2004. Glacial influence on Neoproterozoic sedimen-
northern Colombian Andes. Engineering Geology, 81, 389– 406. tation: the Smalfjord Formation, northern Norway — discussion.
Arnaud, E. & Eyles, C. 2002. Glacial influence on Neoproterozoic sedi- Sedimentology, 51, 1409– 1417.
mentation: the Smalfjord Formation, northern Norway. Sedimentol- Etienne, J. L., Allen, P. A., Rieu, R. & Le Gerroué, E. 2008. Neopro-
ogy, 49, 765– 788. terozoic glaciated basins: a critical review of the Snowball Earth
Arnaud, E. & Eyles, C. 2006. Neoproterozoic environmental change hypothesis by comparison with Phanerozoic glaciations. Special
recorded in the Port Askaig Formation, Scotland: climatic vs tectonic Publication of the International Association of Sedimentologists
controls. Sedimentary Geology, 183, 99 –124. Special Publication 39.
Arnaud, E. & Fairchild, I. J. 2011. The Port Askaig Formation, Dalra- Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic
dian Supergroup, Scotland. In: Arnaud, E., Halverson, G. P. & constraints upon the Neoproterozoic climatic paradox. American
Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Journal of Science, 300, 347– 433.
Glaciations. Geological Society, London, Memoirs, 36, 635–642. Eyles, N. 2008. Glacio-epochs and the supercontinent cycle after 3.0
Bahlburg, H. 1990. The Ordovician basin in the Puna of NW Argentina Ga: Tectonic boundary conditions for glaciation. Palaeogeography,
and N Chile: geodynamic evolution from back-arc to foreland basin. Palaeoclimatology, Palaeoecology, 258, 89 – 125.
Geotektonische Forschungen, 75, 1 –107. Eyles, N. & Januszczak, N. 2007. Syntectonic subaqueousmass flows of
Bahlburg, H. 1998. The geochemistry and provenance of Ordovician tur- the Neoproterozoic Otavi Group, Namibia: where is the evidence of
bidites in the Argentinian Puna. In: Pankhurst, R. J. & Rapela, global glaciation? Basin Research, 19, 179– 198.
C. W. (eds) The Proto-Andean Margin of Gondwana. Geological Fairchild, I. J. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the
Society, London, Special Publication, 142, 127– 142. Earth system. Journal of the Geological Society, London, 164,
Bahlburg, H. & Breitkreuz, C. 1993. Differential response of a Devo- 895– 921.
nian– Carboniferous platform-deeper basin system to sea-level Fedo, C. M., Nesbitt, H. W. & Young, G. M. 1995. Unraveling the
change and tectonics, N. Chilean Andes. Basin Research, 5, 21– 40. effects of potassium metasomatism in sedimentary rocks and paleo-
Baum, S. K. & Crowley, T. J. 2003. The snow/ice instability as a mech- sols, with implications for paleoweathering conditions and prove-
anism for rapid climate change: a Neoproterozoic snowball Earth nance. Geology, 23, 921– 924.
model example. Geophysical Research Letters, 30, 2030, Gallet, S., Jahn, B., Van Vliet Lano, B., Dia, A. & Rosello, E. 1998.
doi: 10.1029/2003GL017333. Loess geochemistry and its implications for particle origin and com-
Benn, D. I. & Prave, A. R. 2006. Subglacial and proglacial glacitectonic position of the upper continental crust. Earth and Planetary Science
deformation in the Neoproterozoic Port Askaig Formation, Scotland. Letters, 156, 157– 177.
Geomorphology, 75, 266– 280. Gevers, T. W. 1931. An ancient tillite in South West Africa. Transactions
Bhatia, M. R. & Crook, K. A. W. 1986. Trace element characteristics of of the Geological Society of South Africa, 34, 1 –17.
graywackes and tectonic setting discrimination of sedimentary Gutzmer, J. & Beukes, N. J. 1998. Earliest laterites and possible evi-
basins. Contributions to Mineralogy and Petrology, 92, 181–193. dence for terrestrial vegetation in the Early Proterozoic. Geology,
Bland, W. & Rolls, D. 1998. Weathering. An Introduction to the Scien- 26, 263– 266.
tific Principles. Arnold Publishers, London. Hoffman, P. F. 2011. Glaciogenic and associated strata of the Otavi car-
Cecioni, G. 1979. Grupo El Toco, desierto de Atacama, Chile. Revista de bonate platform and foreslope, northern Namibia: evidence for large
la Asociación Geológica Argentina, 34, 211–223. base-level and glacioeustatic changes. In: Arnaud, E., Halverson,
Cecioni, G. , 1981. Triassic El Toco Group, Atacama Desert, Chile. In: G. P. & Shields-Zhou, G. (eds) The Geological Record of the Neo-
Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene proterozoic Glaciations. Geological Society, London, Memoirs, 36,
Glacial Record. Cambridge University Press, Cambridge. 195– 209.
A REVIEW OF THE CHEMICAL INDEX OF ALTERATION 91

Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: sorting on bulk composition and mineralogy. Sedimentology, 42,
testing the limits of global change. Terra Nova, 14, 129– 155. 341– 358.
Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P. Nesbitt, H. W., Fedo, C. M. & Young, G. M. 1997. Quartz and
1998. A Neoproterozoic snowball Earth. Science, 281, 1342–1346. feldspar stability, steady and non-steady-state weathering, and petro-
Hoffmann, K. H. & Prave, A. R. 1996. A preliminary note on a revised genesis of siliciclastic sands and muds. Journal of Geology, 105,
subdivision and regional correlation of the Otavi group based on gla- 173– 191.
cigenic diamictites and associated cap dolostones. Communications Nesbitt, H. W., Young, G. M., McLennan, S. M. & Keays, R. R. 1996.
of the Geological Survey of Namibia, 11, 81 – 86. Effects of chemical weathering and sorting on the petrogenesis of sili-
IGME 1984. Mapa Geológica de España a escala 1:50.000, Hoja no. 107 ciclastic sediments, with implications for provenance studies. Journal
(Barruelo de Santullan), 113. of Geology, 104, 525–542.
ITGE 1994. Mapa Geológica de España a escala 1:50.000, Hoja no. 81 Ohta, T. & Arai, H. 2007. Statistical empirical index of chemical weath-
(Potes), 128. ering in igneous rocks: a new tool for evaluating the degree of weath-
Johnsson, M. J. 1993. The system controlling the composition of clastic ering. Chemical Geology, 240, 280– 297.
sediments. In: Johnsson, M. J. & Basu, A. (eds) Processes Control- Panahi, A. & Young, G. M. 1997. A geochemical investigation into the
ling the Composition of Clastic Sediments. Geological Society of provenance of the Neoproterozoic Port Askaig Tillite, Dalradian
America Special Paper, 285, 1 – 19. Supergroup, western Scotland. Precambrian Research, 85, 81 –96.
Johnsson, M. J., Stallard, R. F. & Meade, R. H. 1988. First-cycle Parker, A. 1970. An index of weathering for silicate rocks. Geological
quartz arenites in the Orinoco River Basin, Venezuela and Colombia. Magazine, 107, 501– 504.
Journal of Geology, 96, 263– 277. Pettijohn, F. J., Potter, P. E. & Siever, R. 1987. Sand and Sandstone.
Kahmann, J. A., Seaman, J. III & Driese, S. G. 2008. Evaluating trace Springer, New York.
elements as paleoclimate indicators: multivariate statistical analysis Rice, A. H. N. 2004. Glacial influence on Neoproterozoic sedimentation:
of Late Mississippian Pennington Formation paleosols, Kentucky, the Smalfjord Formation, northern Norway – discussion. Sedimentol-
U.S.A. Journal of Geology, 116, 254– 268. ogy, 51, 1419–1422.
Keidel, J. 1943. El Ordovı́cico inferior en los Andes del norte Argentino y Rice, A. H. N. & Hofmann, C. C. 2001. The transition from
sus depósitos marino glaciales. Boletin de la Academı́a Nacional de Neoproterozoic glacial to interglacial sedimentation near Ham-
Ciencias de Córdoba, 36, 140– 229. marnes, East Finnmark, North Norway. Norsk Geologisk Tidsskrift,
Keller, M., Bahlburg, H. & Reuther, C.-D. 2008. The transition from 81, 257– 262.
passive to active margin sedimentation in the Cantabrian Mountains, Rice, A. H. N., Edwards, M. B., Hansen, T. A., Arnaud, E. & Halver-
Northern Spain: Devonian or Carboniferous? Tectonophysics, doi: son, G. P. 2011. Glaciogenic rocks of the Neoproterozoic Smalfjord
10.1016/j.tecto.2008.06.022 and Mortensnes Formations, Vestertana Group, E. Finnmark, Norway.
Keller, M., Bahlburg, H., Reuther, C.-D. & Weh, A. 2007. Flexural to In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds)
broken foreland basin evolution as a result of Variscan collisional The Geological Record of Neoproterozoic Glaciations. Geological
events in northwestern Spain. In: Hatcher, R. D. Jr, Carlson, M. Society, London, Memoirs, 36, 593– 602.
P., McBride, J. H. & Martı́nez Catalán, J. R. (eds) The 4D Frame- Rieu, R., Allen, P. A., Plotze, M. & Pettke, T. 2007. Compositional
work of Continental Crust. Geological Society of America Memoir, and mineralogical variations in a Neoproterozoic glacially influenced
200, 489– 510. succession, Mirbat area, south Oman: Implications for paleoweather-
Kirschvink, J. L. 1992. Late Proterozoic low-latitude global glacia- ing conditions. Precambrian Research, 154, 248– 265.
tion: the snowball earth. In: Schopf, J. W & Klein, C. (eds) The Ruxton, B. P. 1968. Measures of the degree of chemical weathering of
Proterozoic Biosphere. Cambridge University Press, Cambridge, rocks. Journal of Geology 76, 518–527.
51 – 52. Scheffler, K., Hoernes, S. & Schwark, L. 2003. Global changes during
Kramer, J. R. 1968. Mineral-water equilibria in silicate weathering. Inter- Carboniferous–Permian glaciation of Gondwana: Linking polar and
national Geological Congress, 23rd session, Section 6, 149–160. equatorial climate evolution by geochemical proxies. Geology, 31,
Martin, H. 1964. Beobachtungen zum Problem der jung-präkambrischen 605– 608.
glazialen Ablagerungen in Südwestafrika. Geologische Rundschau, Schermerhorn, L. J. G. 1974. Late Precambrian mixtites: glacial and/or
54, 115– 127. non-glacial? American Journal of Science, 274, 673– 824.
Martin, H., Porada, H. & Walliser, O. H. 1985. Mixtite deposits of the Schrag, D. P., Berner, R. A., Hoffman, P. F. & Halverson, G. P. 2002.
Damara sequence, Namibia: problem of interpretation. Palaeogeo- On the initiation of Snowball Earth. Geochemistry, Geophysics, Geo-
graphy, Palaeoclimatology, Palaeoecology, 51, 159– 196. systems, 3, doi:10.1029/2001GC000219.
McLennan, S. M. 1993. Weathering and global denudation. Journal of Shields, G. A. 2008. Palaeoclimate – Marinoan meltdown. Nature
Geology, 101, 295– 303. Geoscience, 1, 351– 353.
McLennan, S. M. 2001. Relationships between the trace element compo- Taylor, S. R. & McLennan, S. M. 1985. The Continental Crust: Its
sition of sedimentary rocks and upper continental crust. Geochemis- Composition and Evolution. Blackwell, Oxford.
try, Geophysics, Geosystems (G3), 2, doi:10.1029/2000GC000109. von Eynatten, H. 2004. Statistical modelling of compositional trends in
McLennan, S. M., Hemming, S., McDaniel, D. K. & Hanson, G. N. sediments. Sedimentary Geology, 171, 79– 89.
1993. Geochemical approaches to sedimentation, provenance and von Eynatten, H., Barceló-Vidal, C. & Pawlowsky-Glahn, V.
tectonics. In: Johnsson, M. J. & Basu, A. (eds) Processes Control- 2003. Modelling compositional change: the example of chemi-
ling the Composition of Clastic Sediments. Geological Society of cal weathering of granitoid rocks. Mathematical Geology, 35,
America Special Paper, 285, 21 – 40. 231– 251.
Nedachi, Y., Nedachi, M., Bennett, G. & Ohmoto, H. 2005. Geochem- Wagner, R. H. & Wagner-Gentis, C. H. T. 1963. Summary of the stra-
istry and mineralogy of the 2.45 Ga Pronto paleosols, Ontario, tigraphy of Upper Paleozoic rocks in NE. Palencia, Spain. Proceed-
Canada. Chemical Geology, 214, 21 – 44. ings Kongelige Nederlandse Akademie Wetenschappen (B) LXVI, 3,
Nesbitt, H. W. & Young, G. M. 1982. Early Proterozoic climates and 149– 163.
plate motions inferred from major element chemistry of lutites. Weil, A. B., Van der Voo, R. & Van der Pluijm, B. A. 2001. Oroclinal
Nature, 199, 715– 717. bending and evidence against the Pangea megashear; the Cantabria –
Nesbitt, H. W. & Young, G. M. 1984. Prediction of some weathering Asturias Arc (northern Spain). Geology, 29, 991– 994.
trends of plutonic and volcanic rocks based on thermodynamic and Wintsch, R. P. & Kvale, C. M. 1994. Differential mobility of elements in
kinetic considerations. Geochimica et Cosmochimica Acta, 48, burial diagenesis of siliciclastic rocks. Journal of Sedimentary
1523– 1534. Research, A64, 349– 361.
Nesbitt, H. W. & Young, G. M. 1989. Formation and diagenesis of Young, G. M. 2001. Comparative geochemistry of Pleistocene and Paleo-
weathering profiles. Journal of Geology, 97, 129–147. proterozoic (Huronian) glaciogenic laminated deposits: relevance to
Nesbitt, H. W. & Young, G. M. 1996. Petrogenesis of sediments crustal and atmospheric composition in the last 2.3 Ga. Journal of
in the absence of chemical weathering: effects of abrasion and Geology, 109, 463– 477.
92 H. BAHLBERG & N. DOBRZINSKI

Young, G. M. & Nesbitt, H. W. 1999. Paleoclimatology and provenance rocks in the Table Mountain Group, South Africa. Palaeogeography,
of the glaciogenic Gowganda Formation (Paleoproterozoic), Ontario, Palaeoclimatology, Palaeoecology, 214, 323– 345.
Canada: a chemostratigraphic approach. Geological Society of Zhang, Q.-R., Chu, X.-L. & Feng, L.-J. 2011. Neoproterozoic glacial
America Bulletin, 111, 264–274. records in the Yangtze Region, China. In: Arnaud, E., Halverson,
Young, G. M., Minter, W. E. L. & Theron, J. N. 2004. Geochemistry G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproter-
and palaeogeography of upper Ordovician glaciogenic sedimentary ozoic Glaciations. Geological Society, London, Memoirs, 36, 357–366.
Chapter 7

Neoproterozoic glacial palaeolatitudes: a global update

D. A. D. EVANS1* & T. D. RAUB2


1
Department of Geology & Geophysics, Yale University, 210 Whitney Avenue, New Haven, CT 06520-8109, USA
2
Division of Geological and Planetary Sciences, 100-23 Caltech, Pasadena CA 91125, USA
*Corresponding author (e-mail: dai.evans@yale.edu)

Abstract: New stratigraphic, geochronological and palaeomagnetic constraints allow updates to be made to a synthesis of Neoprotero-
zoic glacial palaeolatitudes, including modifications to some reliability estimates. The overall pattern of a Neoproterozoic climatic
paradox persists: there is an abundance of tropical palaeolatitudes and near to complete absence of glaciogenic deposits demonstrably
laid down between latitudes of 608 and 908. In addition to 12 units with palaeolatitude estimates that are somewhat reliable, estimates
with moderate to high reliability now include Konnarock (less than 108 from the palaeo-equator), Elatina, Rapitan, Mechum River, Grand
Conglomerat (10– 208), Upper Tindir, Puga (20– 308), Nantuo, Gaskiers (30–408) and Walsh (40– 508). Among these, Elatina, Upper
Tindir and Nantuo are considered to have the highest reliability, all with estimates of low to moderate palaeolatitude. The Elatina
result stems from sedimentary rocks with quantitative correction of inclination-shallowing effects, and the Upper Tindir result stems
from data collected from igneous rocks that are precisely coeval with the glacial deposits. Despite continuing debate on the global char-
acter of Neoproterozoic ice ages, their pan-glacial extent (ice extending to low latitude in a low-obliquity world) is well demonstrated.

Palaeomagnetism of glaciogenic deposits has provided a quantita- as well as those of Hambrey & Harland (1981), Eyles &
tive basis for hypotheses of extreme climatic shifts in the late Januszczak (2004, table 1), Trindade & Macouin (2007) and
Neoproterozoic Era. The most recent global compilations of Hoffman & Li (2009). Deposits are numbered by present geo-
palaeomagnetic depositional latitudes for Proterozoic ice ages graphical location, differentiating units that are separated from
indicate a dominant mode near the palaeo-equator (Evans 2000, each other on the scale of 100 km or more, or in some cases dis-
2003), with no robust palaeo-polar deposits yet discovered. Such tinguishing units of uncertain relative correlation that are presently
results could therefore support either the Snowball Earth (Kirsch- adjacent via tectonic stacking in orogenic belts. Within each
vink 1992) or the high-obliquity (Williams 1993) hypotheses for numbered location, the units are denoted (a, b, c, etc.) in ascending
Precambrian ice ages, but would appear to reject the uniformitarian stratigraphic order. Many global chronostratigraphic schemes
comparison to polar/temperate-restricted Phanerozoic glaciogenic have been proposed, that assign ages to undated deposits via cor-
deposits (Evans 2000). Hoffman (2009) has suggested that relation, but herein, we adopt the approach used by Evans (2000)
Neoproterozoic ice ages represent a globally all-encompassing to consider each unit’s age constraints in isolation. This splitting,
‘pan-glacial’ state of the Earth’s climate system, fundamentally rather than lumping, approach serves to illustrate how few of the
distinct from either partially ice-covered ‘glacial –interglacial’ deposits are precisely dated or studied palaeomagnetically. Many
or ice-free ‘nonglacial’ palaeoclimates experienced during the of the deposits have been classified using the loose terms
Phanerozoic Eon. ‘Sturtian’, ‘Marinoan’, or ‘Ediacaran’ based on the lithographic
Several reviews of stratigraphic (Halverson et al. 2005, 2007), character of either diamictites or their overlying cap carbonates
sedimentological (Hoffman & Schrag 2002; Eyles & Januszczak (e.g. Hoffman & Li 2009). Recent U – Pb geochronology has
2004; Fairchild & Kennedy 2007; Hoffman et al. 2007; Allen & thus far permitted many so-called Marinoan deposits to record
Etienne 2008; Hoffman 2009) and palaeomagnetic (Trindade & the synchronous end of a widespread ice age of unknown duration,
Macouin 2007; Hoffman & Li 2009) data sets pertaining to which ended at 635 Ma (Hoffmann et al. 2004; Condon et al. 2005;
Neoproterozoic ice ages have appeared recently. These discussions Zhang et al. 2008), and a mid-Ediacaran ice age at c. 580 Ma
of glacial deposits owe much to the pioneering synthesis of (Bowring et al. 2002, abstract only). U –Pb ages of so-called
Hambrey & Harland (1981), but the recent reviews arrive at differ- Sturtian deposits range from c. 765 Ma (Key et al. 2001) to poss-
ing conclusions regarding the extent and severity of Neoprotero- ibly as young as c. 660 Ma (Fanning & Link 2008; from the type
zoic ice ages. In particular, the study by Eyles & Januszczak area in South Australia). Use of the lithological characteristics of
(2004) is commonly cited as a palaeogeographic alternative that cap carbonates as global correlation tools appears successful in
avoids the need for nonuniformitarian processes to account for some instances, but may be problematic in others (Corsetti &
the advance of widespread continental ice sheets into tropical Lorentz 2006; Kendall et al. 2009).
palaeolatitudes. It becomes useful, then, to review the global evi- Finally, several of the units discussed by Evans (2000) are now
dence for or against low-latitude glaciation. Herein, we reassess considered irrelevant to discussions on Neoproterozoic ice ages, as
Neoproterozoic glacial palaeolatitudes in light of new strati- they are likely nonglacial, or demonstrably Cambrian or younger
graphic, geochronological and palaeomagnetic data obtained in age. Those units are included in Table 7.1, but are now stripped
within the last decade, providing the first comprehensive update of their numerical codes. In a few cases, questions remain about
and revision of the palaeographic analyses of Evans (2000). these issues, and the deposits retain their numerical status until
documented otherwise.
Table 7.1 also lists our preferred interpretation on the reliability
Methods of palaeolatitude determination for the deposits, using the three-
level qualitative scale after Evans (2000). Although quantitative
Table 7.1 lists the known or alleged Neoproterozoic glaciogenic measures of palaeomagnetic pole reliability exist (e.g. Van der
deposits, revising the unit-identifying numbering scheme intro- Voo 1990), they have generally been tailored towards plate recon-
duced by Evans (2000), but with cross-references to that study, structions rather than palaeoclimatic problems. Herein, we regard

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 93– 112. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.7
94
Table 7.1 Published assessments of glacial influence and depositional palaeolatitude for Neoproterozoic strata

# Deposit name Lat (88 N) Long (88 E) HH81 E00 EJ04 TM07 HL09 This study

Laurentia and environs


1a Hula Hula 69.5 215.5 – – – – Hu(S) 0 –15 No pmag data
1b Katakturuk unit 2 69.5 214.5 – – – – – No pmag data
2a Upper Tindir: unit 2 65 222 Yes (F10) No pmag data (#3) – – Ti(S) 0 –15 *** 21 +3
2b Upper Tindir: unit 3b 65 219 – – – – – No pmag data
3a Rapitan 64 230 Yes (F11) ** 06 þ8/27 (#1) Yes Q ¼ 4; 06 +4 Ra(S) 0 –15 ** 18 +3
3b Ice Brook (Stelfox) 63.5 231.5 – Unreliable pmag (#2) ?? – IB(M) 0 –15 Unreliable pmag
4 Mt Lloyd George 58 235 Yes (F12) No pmag data (#4) – – – No pmag data
5 Deserters Range 57 235.5 – No pmag data (#4) – – – No pmag data
6 Mount Vreeland 54.5 239 Uncertain (F12) No pmag data (#5) – – Vr(M) 0–15 No pmag data
7 Toby 49.5 243 Yes (F13) * 08 +4 (#6) – – To(S) 0 –15 No pmag data
8 Edwardsburg 45 244.5 – No pmag data (#7) – No pmag data – No pmag data

D. A. D. EVANS & T. D. RAUB


9 Pocatello: Scout Mtn 42.5 248 Yes (F14) No pmag data (#7) – Q ¼ 6; 08 +3 Po(S) 0– 15 No pmag data
10 Mineral Fork/Dutch Pk 40.5 247.5 Yes (F15) No pmag data (#7) – – – No pmag data
11a Kingston Peak: Surprise 36 243.5 Yes (F16) No pmag data (#8) No – Su(S) 0– 15 No pmag data
11b Kingston Peak: Wildrose 36 243.5 – No pmag data (#8) No – Wr(M) 0– 15 No pmag data
11c Ibex 36 243.5 – – – – – No pmag data
11d Johnnie Rainstorm 36 244 – * 01 +4 (#9) – – – * 01 +4
12 Cerro Las Bolas 29 250.5 – – – – – No pmag data
– Florida Mtns 32 252.5 Yes (F17) * 03 +8 (#10) – – – Too young
13 Konnarock/Grandf. Mtn 36.5 278.5 Yes (F18) See Mechum R (#11) No – Kn(S) 15–30 ** 06 +5
14a Mechum River 38.5 281.5 – ** 20–21 +4 (#11) – – – ** 16 +3
14b Fauquier 39 282.5 – – – – – * 78 +12
15a Port Askaig 55.5 353.5 Yes (E18 –E20) Unreliable pmag (#12) Yes No pmag data Pt(S) 30– 45 Unreliable pmag
15b Stralinchy (Reelan) 55.5 353.5 – – – – Re(M) 30– 45 No pmag data
15c Loch na Cille 55.5 353.5 – No pmag data (#12) – – Lo(E) 75–90 Unreliable pmag
16 Gåseland/Charcot Land 71 330.5 Yes (F23,F24) No pmag data (#14) – – – No pmag data
17a Tillite Gp: Ulvesø 73 336 Yes (F25) Unreliable pmag (#15) Yes – Ul(S) 15– 30 Unreliable pmag
17b Tillite Gp: Storeelv 73 336 Yes (F25) Unreliable pmag (#15) Yes – St(M) 30– 45 Unreliable pmag
18a Elbobreen (Petrovbreen) 79.5 18 Yes (E10) No pmag data (#16) – – Pb(S) 15– 30 No pmag data
18b Wilsonbreen 79.5 18 Yes (E10) No pmag data (#16) – – Wb(M) 30–45 No pmag data
19 West Spitsbergen 78 13 Yes (E10) No pmag data (#17) – – – No pmag data
20 Moraenesø 82 326 Yes (F26) No pmag data (#18) – – – No pmag data
21 Pearya 82.5 281.5 – No pmag data (#19) – – – No pmag data
Baltica
22a Vestertana: Smalfjord 70 28 Yes (E12) ** 33 þ14/212 (#20) Yes (a or b) Q ¼ 3; 33 +9 Sm(M) 45–60 * 33 þ14/212
22b Vestertana: Mortensnes 70 28 Yes (E12) ** 33 þ14/212 (#20) Yes (a or b) Q ¼ 3; 33 +9 Mt(E) 45– 60 No pmag data
– Sito/Vakkejokk 68 19 Uncertain (E13) Too young (#21) – – – Too young
23 Långmarkberg/Lillfjället 64 14.5 Yes (E14/E15) No pmag data (#22) – – – No pmag data
24 Moelv 61.5 11 Yes (E16) Unreliable pmag (#23) Yes No pmag data Mo(E) 45– 60 Unreliable pmag
25 Vilchitsy/Blon 53 32 Yes (E24 –E28) Unreliable pmag (#24) – – Vi(E) 45–60 Unreliable pmag
26 N Urals: Churochnaya 60 57.5 Yes (E30) No pmag data (#25) – – Cn(E) 30– 45 No pmag data
27a C Urals: Tany 58.5 59 Yes (E31) No pmag data (#26) – – Ty(S) 45– 60 No pmag data
27b C Urals: Koyva 58.5 59 Yes (E31) No pmag data (#26) – – – No pmag data
27c C Urals: Staryye Pechi 58.5 59 Yes (E31) No pmag data (#26) – – – No pmag data
28 S Urals: Kurgashlya 53.5 57.5 Yes (E32) Unreliable pmag (#27) – – – Unreliable pmag
Altaids and Siberia
29a West Altaids (Satan/Dzhetym) 44 72 Yes (C27) No pmag data (#31) – – – No pmag data
29b West Altaids (Baykonur) 44 72 Variable (C22– C28) No pmag data (#31) – – Br(S) 45– 60 No pmag data
30a Tsagaan Oloom (Maikhan Ul) 47 96 – No pmag data (#32) Blank – Mk(S) 0–15 Unreliable pmag
30b Tsagaan Oloom (Khongoryn) 47 96 – – – – Kg(M) 0– 15 * 03 + 11
31 East Sayan/Khubsugul: Zabit 51 100 – No pmag data (#30) – – – Unreliable pmag
32a Chivida 61 92 Nonglacial (C30) Unreliable pmag (#28) – – Cv(S) 15–30 Unreliable pmag
32b Pod’em 61 92 – – – – Pd(M) 0–15 No pmag data
33 Marnya 54.5 99 – – – Ma(M) 0 –15 No pmag data
34a Patom: Kharluktakh 59 115 – – – – Kh(S) 0 –15 No pmag data
34b Patom: Dzhemkukan 59 115 Uncertain (C31) No pmag data (#29) – – Dz(M) 0– 15 No pmag data
China cratons
35a Aksu: Qiaoenbrak 41 79.5 Yes (C33) * 08 +8 (#33) – – – No pmag data

GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES


35b Aksu: Yuermeinbrak 41 79.5 Yes (C33) * 08 +8 (#33) – – – * 27 + 9
36a Quruqtagh: Bayisi 41.5 87.5 Yes (C33) * 08 +8 (#33) – Q ¼ 4; 01 +3 By(S*) 30–45 * 01 þ4/22
36b Quruqtagh: Tereeken 41.5 87.5 Yes (C33) * 08 +8 (#33) – – Te(M) 15– 30 No pmag data
36c Quruqtagh: Hankalchough 41.5 87.5 Yes (C33) * 08 +8 (#33) – – Ha(E) 15– 30 No pmag data
37 Qaidam: Hongtiegou 37.5 96 Yes (C33) – – – – No pmag data
38 Luoquan 34 115 Yes (C33,C34) Unreliable pmag (#34) Yes – Lq(E) 0 –15 Unreliable pmag
39a Chang’an/Tiesiao/Jiangkou 27 111 Yes (C33,C35) *** 30–40 +12 (#35) Yes (a or b) – Ji(S) 15–30 No pmag data
39b Nantuo 27 111 Yes (C33,C35) *** 30–40 +12 (#35) Yes (a or b) Q ¼ 6; 04 +4 Na(M) 15– 30 *** 37 +9
India to Nubia
40 Blaini 30 78.5 Yes (C14) Unreliable pmag (#37) – – Bl(M) 0– 15 Unreliable pmag
– Penganga 19.5 75 – No pmag data (#38) – – – Nonglacial
41 Rizu 31 56 – – – – Ri(M) 0– 15 Too young?
42a Ghubrah 23 58 – Unreliable pmag (#36) – No pmag data Gu(S) 15–30 No pmag data
42b Fiq 23 58 – No pmag data (#36) Yes (’Shuram’) Q ¼ 6; 15 +4 Fi(M) 30– 45 * 13 +7
43a Ayn 17.5 55 – Unreliable pmag (#36) – – Ay(S) 15–30 Unreliable pmag
43b Shareef 17.5 55 – No pmag data (#36) – – Sh(M) 30– 45 * 18 +7
44 Tambien 14 39 – – – – Ta(S) 15–30 No pmag data
45 Atud 26 35 – – – – – No pmag data
Australia and Mawsonland
46 Walsh 217 126 Yes (D16) ** 45 þ14/212 (#42) – Q ¼ 5; 45 +5 – ** 45 þ14/212
47a Landrigan 218 126.5 Yes (D16) No pmag data (#44) – – La(M) 0– 15 No pmag data
47b Egan 218.5 126.5 Yes (D16) * 21 +8 (#43) – – Eg(E) 15–30 * 21 +8
48a Fargoo 217 129 Yes (D16) No pmag data (#44) – – – No pmag data
48b Moonlight Valley 217 129 Yes (D16) No pmag data (#44) Yes – – No pmag data
49 Little Burke 221.5 140 Yes (D19) Unreliable pmag (#40) – – – Unreliable pmag
50a Areyonga/Naburula/Yardida 223 134 Yes (D17,D18) No pmag data (#39) – – Ar(S) 15– 30 No pmag data
50b Olympic/Mount Doreen 223 135 Yes (D17) No pmag data (#40) – – Ol(M) 0 –15 No pmag data

(Continued)

95
96
Table 7.1 Continued

# Deposit name Lat (88 N) Long (88 E) HH81 E00 EJ04 TM07 HL09 This study

51 Boondawari/Wahlgu/Turkey Hill 225 124 – No pmag data (#40) – – – Unreliable pmag


52 Chambers Bluff 227 134 – No pmag data (#39) – – – No pmag data
53a Sturtian (type area) 232 139 Yes (D21) Unreliable pmag Yes No pmag data St(S) 0– 15 Unreliable pmag
53b Elatina 232 139 Yes (D21) *** 03–09 +4 (#40) Blank Q ¼ 4,6; 05 +6 El(M) 0–15 *** 10– 14
54a Yancowinna 231.5 141.5 Yes (D20) No pmag data (#39) – – – No pmag data
54b Teamsters Creek 231.5 141.5 Yes (D20) No pmag data (#40) – – – No pmag data
55 Cottons 240 144 Uncertain (D22) Unreliable pmag (#41) – No pmag data Co(M) 0– 15 Unreliable pmag
56a Julius River 241 145 Uncertain (D22) No pmag data (#41) – No pmag data Ju(S) 0 –15 No pmag data
56b Croles Hill 240.5 145 – – – No pmag data Cr(E) 0 –15 No pmag data

D. A. D. EVANS & T. D. RAUB


57 Goldie 283 160 – No pmag data (#45) – – – No pmag data
Kalahari and environs
58a Blaubeker/Court 223.5 17.5 Yes (A29) Unreliable pmag (#46) – – Bb(S) 0– 15 Unreliable pmag
58b Blasskrans/Naos 223.5 17.5 Map only (A29) No pmag data (#47) – – Bk(M) 0– 15 No pmag data
59a Kaigas 228.5 16.5 Nonglacial (A29) No pmag data (#48) – No pmag data Ka(S*) 0 –15 No pmag data
59b Numees 228.5 16.5 Yes (A29) No pmag data (#49) – – Nu(M) 0– 15 No pmag data
59c Namaskluft 228.5 16.5 Yes (A29) No pmag data (#49) – – 0–15 No pmag data
59d Schwarzrand 227 18 Yes (A29) * 38 +3 (#51) – – – * 38 +3
60a Karoetjes Kop 231.5 18.5 – – – – – No pmag data
60b Aties 232 18.5 Nonglacial (A29) No pmag data (#50) – – – No pmag data
61 Dernburg 228 15.5 – – – – – No pmag data
Congo-São Francisco
62a Grand Conglomérat 211 27 Yes (A28) No pmag data (#52) – Q ¼ 5; 10 +6 Gr(S*) 15–30 ** 10 +5
62b Petit Conglomérat 211 27 Uncertain (A28) No pmag data (#53) – – Pe(M) 0 –15 No pmag data
63 Geci 213 35 – – – – – No pmag data
64 Tshibangu 22 29 – No pmag data (#54) – – – No pmag data
65 Bunyoro 1.5 31.5 Yes (A25) No pmag data (#55) – – – No pmag data
66a Akwokwo/Bandja 3 24 – No pmag data (#55) – – Ak(S) 30–45 No pmag data
66b Bondo 5 20 – No pmag data (#56) – – Bo(M) 15– 30 No pmag data
67 Mintom 3 13 – – – – – No pmag data
68a Sergipe: Juetê/Ribeirópolis 211 322 – No pmag data (#56) – – – No pmag data
68b Sergipe: Palestina 211 322 – No pmag data (#56) – – Pa(M) 30–45 No pmag data
69 Bebedouro 211 319 Uncertain (G28) No pmag data (#58) – – – Unreliable pmag
70 Rio Preto: Canabravinha 211 314 – – – – – No pmag data
71 Brasilia int: Ibiá/Cristalina 218 313 Uncertain (G27) No pmag data (#59) – – – No pmag data
72a Vazante: St. Antônio do Bonito 217.5 313 – – – – – No pmag data
72b Vazante: Lapa 217.5 313 – – – – – No pmag data
73 Carandaı́ 221 316 Uncertain (G30) No pmag data (#60) – – – No pmag data
74a Bambui: Jequitaı́/Macaúbas 217 316 Yes (G31) Unreliable pmag (#57) No – Je(S) 30–45 Unreliable pmag
74b Bambui: Inhaúma 219.5 315.5 – – – – – No pmag data
74c Bambui: Lagoa Formosa 218.5 313.5 – – – – – No pmag data
75 Rio Pardo: Salobro 215.5 320.5 Uncertain (G29) Uncertain glacial (#57) – – – Nonglacial
76a West Congo Inférieure 25 14 Nonglacial (A26,A27) No pmag data (#61) – – In(S) 30–45 No pmag data
76b West Congo Supérieure 25 14 Nonglacial (A26,A27) No pmag data (#62) – – Sp(M) 15– 30 No pmag data
77a Chuos/Varianto 221 16 Variable (A29) * 10 +5 (#63) No – Ch(S) 15–30 * 09 +5
77b Ghaub 221 16 Uncertain (A29) Unreliable pmag (#64) No No pmag data Gh(M) 0– 15 Unreliable pmag
West Africa and Hoggar
78a Basal Atar Gp 22 352 – Unreliable pmag (#66) – – – Nonglacial?
78b Jbéliat/Bthaat Ergil (triad) 22 352 Yes (A19) * 30– 70 (#65) Yes – Jb(M) 45–60 No pmag data
79 Mali Gp/Bakoye 15 351 Yes (A19) Same as #65 (#67) – – Ba(M) 30– 45 No pmag data
80 Rokel River 8.5 348 Yes (A20) No pmag data (#68) – – – No pmag data
81a Kodjari 10 0 Yes (A21) No pmag data (#69) – – Ko(M) 30–45 No pmag data
81b Tamale/Obosum 8 0 Yes (A21) No pmag data (#70) – – – No pmag data
82a Série Verte (Tafeliant) 21 2 Yes (A23) Unreliable pmag (#71) – – – Unreliable pmag
82b Série Pourprée 24 2 Yes (A22) Same as #65 (#72) – – – No pmag data
83 Série Tirririne 21 8.5 Yes (A23) Unreliable pmag (#71) Unreliable pmag

GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES


84a Sarhro 31 352.5 – No pmag data (#73) – – – No pmag data
84b Tiddiline 30 352 Nonglacial (A18) No pmag data (#73) – – – Unreliable pmag
Avalonia and Cadomia
85 Squantum 42.5 289 Uncertain (F19) * 55 þ8/27 (#74) No Q ¼ 3; 55 +3 Sq(E) 75– 90 * 55 þ8/27
86 Gaskiers 47.5 307 Yes (F20) * 31 þ10/28 (#75) Yes No pmag data Ga(E) 75– 90 ** 34 þ8/27
87 Gwna/Mona 53 355.5 – – – – – No pmag data
88 Brioverian (Granville) 48.5 358 Nonglacial (E21) No pmag data (#76) No – Gr(E) 45– 60 No pmag data
89 Clanzschwitz/Weesenstein 51 13.5 Correlative to (E22)? – – – – No pmag data
Amazonia and environs
90a Puga 215.5 303 Yes (G26) No pmag data (#77) – Q ¼ 5; 22 +6 Pu(M) 30– 45 ** 22 þ6/25
90b Serra Azul 215.5 303 – – – – Az(E) 60– 75 No pmag data
91 Chiquerı́o 215.5 285 – No pmag data (13) – – Cq(S) 15–30 No pmag data
Rio Plata and environs
92a Zanja del Tigre 234.5 305 – – – – – Nonglacial?
92b Las Ventanas/Playa Hermosa 234.5 305 – – – – – Unreliable pmag
93 Picada das Graças 231 305.5 – Unreliable pmag (#78) – – – Unreliable pmag
94 Iporanga 224.5 311.5 – – – – – No pmag data
95 Sierra del Volcán 238 302 – – – Q ¼ 5; 48 +7 – Too young?

Deposits are numbered by geographical area; within each region, they are lettered by ascending stratigraphic order. Glaciogenic influence assessments: HH81, Hambrey & Harland (1981); EJ04, Eyles & Januszczak (2004, table 1).
The terms ‘yes’, ‘no’ and ‘uncertain’ refer to the question of whether a glaciogenic influence has been demonstrated according to each compilation. ‘Blank’ refers to blank entries in Eyles & Januszczak (2004, table 1).
Palaeomagnetic constraint assessments: E00, Evans (2000); TM07, Trindade & Macouin (2007); HL09, Hoffman & Li (2009). Correlations by Hoffman & Li (2009) presented in parentheses, follow the deposit abbreviations
used in their work: E, Ediacaran; M, Marinoan; S, Sturtian; S*, nominally Sturtian but with possibly older ages than the deposits labelled ‘S’. Asterisks in the columns for E00 and this study indicate the relative reliability
(*, low; **, moderate; ***, high) of glacial palaeolatitudes assessed by Evans (2000) or herein. Q values use the reliability scheme of Van der Voo (1990). Throughout the table, dashed entries indicate that the deposit was not
mentioned by the particular compilation.

97
98 D. A. D. EVANS & T. D. RAUB

reliability of glacial palaeolatitudes to depend mainly on the


(a)
quality of palaeomagnetic data, but also on the confidence in Shareef Somewhat reliable
chrono/lithostratigraphic correlations, if necessary for palaeolati-
tude estimation, and to a lesser extent the amount of consensus on Moderately reliable
Chuos Fiq
the glaciogenicity of the deposits. In addition, some of our assess-
ments in Table 7.1 are labelled ‘unreliable pmag [palaeomagnetic Very reliable
data]’, which merely indicates that the data are not applicable for Bayisi Grand Egan Schwarz-
Uniform distribution (n=22)
Congl. rand
reliable estimates of glacial palaeolatitude, although in some cases
they could be useful for other purposes such as tectonic reconstruc- Khon- Yuer- Smal-
tions. We do not provide references to all the stratigraphic, geo- goryn Mechum meinbrak
fjord
chronological, and palaeomagnetic constraints on the deposits,
but rather only discuss those instances in which our current assess- Johnnie Rapitan Puga Gaskiers
ment differs from previously published global compilations. Our
global palaeolatitude estimates for Neoproterozoic ice ages,
Konna- U.Tindir Squan-
using this quality filtering method, are summarized graphically rock Elatina unit 2
Nantuo Walsh tum Fauquier
in Fig. 7.1.
We begin with the assumption that Earth’s time-averaged (.103 0 10 20 30 40 50 60 70 80 90

years) magnetic field has the geometry of a geocentric-axial dipole Palaeolatitude (°)
(GAD). This assumption is justified to first order by the compari-
Shareef (b)
son of evaporite palaeolatitudes in the palaeomagnetic reference Chuos Fiq
Area-normalized distribution
Schwarz.
frame, relative to their expected palaeolatitudes according to inter- Bayisi Grand C. Egan
Smalfj.
tropical convergence of zonal atmospheric circulation. Agreement Khong. Mechum Yuerm.
Fauquier
Johnnie Rapitan Puga Gaskiers
between these two reference frames for most Proterozoic evaporite Konnar. Elatina U.Tindir Nantuo Walsh Squant.
deposits implies that the GAD hypothesis is tenable for most of the 0 10 20 30 40 50 60 70 80 90
last two billion years (Evans 2006). Departures from this model are Fig. 7.1. Reliability of palaeomagnetic depostional latitudes for Neoproterzoic
possible, due to data gaps, over intervals as long as 100– glacial deposits. (a) Unit-weight given to each estimate, following Evans
200 million years; notably, the Neoproterozoic glacial interval (2000). Thick lines indicate the normalized probability at each latitude band, for
does indeed lack voluminous evaporite deposits. Ephemeral, dom- a uniform (random) distribution over the surface of the Earth. (b) Latitude
inantly non-GAD fields (on the order of 10 million years duration band-area normalization of the palaeolatitude estimates.
or shorter) are also permissible, as will be discussed further below,
but in light of the broader record of a uniformitarian geodynamo
(see also Swanson-Hysell et al. 2009), a burden of proof must lie which was determined by direct measurements of the Rapitan suc-
with these alternatives for any specific age. cession that apparently were substantially affected by sedimentary
inclination shallowing (implied flattening factor, f  0.3 calcu-
lated here; see Tauxe et al. 2008). Nonetheless, general concor-
dance of 750– 700 Ma palaeomagnetic results from Laurentia
Palaeomagnetic constraints on Neoproterozoic glacial (Fig. 7.2) indicates that the Earth’s geomagnetic field was stable
palaeolatitudes at that time. Distinctly higher in the Tindir stratigraphy, a second
diamictite level (unit 3b) and overlying cap carbonate bear striking
Laurentia and environs resemblance to the Stelfox/Ravensthroat diamictite/cap carbon-
ate couplet in the nearby Mackenzie Mountains (Macdonald et al.
Macdonald et al. (2010a) have produced direct, high-precision, 2010b). Neither of the younger diamictite/cap carbonate pairs at
U –Pb age constraints on the Upper Mount Harper Group the two localities have palaeomagnetic constraints on depositional
diamictite-bearing succession, at 716.5 Ma, which is precisely palaeolatitude.
coeval with the immediately underlying Mount Harper Volcanic Inliers on Alaska’s North Slope belong to an Arctic Alaska –
Complex and the transcontinental Franklin large igneous province. Chukotka microplate of questionable Laurentian or Siberian affi-
The latter has yielded a high-quality palaeomagnetic pole repre- nity. Regionally, diamictite (informally named Hula Hula) lies
senting data measured in several laboratories and spanning the above post-760 Ma volcanic rocks, and locally, it is interbedded
width of Arctic Canada plus northwestern Greenland (Denyszyn with basaltic flows (Macdonald et al. 2009a). The overlying
et al. 2009). The Mount Harper and underyling Fifteenmile Katakturuk dolomite contains a diagnostic unit (Katakturuk 2)
groups are well correlated, lithostratigraphically, to the Tindir considered correlative to classic Marinoan cap carbonates else-
Group across the length of the Oglivie Mountains, and the Franklin where in the world. Although the Hula Hula-associated flows
palaeomagnetic pole implies a depositional palaeolatitude of might correlate to the Franklin event and merit application of
21+38 for the Tindir/Mount Harper glaciogenic succession. We that palaeomagnetic pole, the allochthonous nature of the terrane
assign this result to the highest reliability in the present analysis, and uncertain timing and style of its accretion to Alaska render
and it is noteworthy by the fact that it derives from igneous these deposits palaeomagnetically unconstrained.
rocks that are immune from any possible effects of systematic New U –Pb ages on zircon from Neoproterozoic successions in
inclination shallowing. We assign only slightly less reliability to Idaho support several distinct pulses of glaciation in that region. In
the application of the Franklin mean palaeomagnetic pole to the the Pocatello region, Fanning & Link (2004) produced U –Pb
Rapitan glaciogenic deposits in the adjacent Mackenzie Mountains SHRIMP ages constraining the Scout Mountain Member (mainly
(implied depositional palaeolatitude of 18 + 38) because that suc- diamictite) of the Pocatello Formation, to slightly younger than
cession lacks the volcanic manifestation of the Franklin event. 709 + 5 Ma and substantially older than 667 + 5 Ma. The
Correlation among the Upper Tindir, Upper Mount Harper and younger age was determined using a zircon subpopulation from
Rapitan diamictite successions is moderately strong, because of a channelized tuff horizon that lies closely beneath a crystal fan-
their close associations with Fe formation, similarity of overlying bearing marble. The fans resemble those in post-glacial cap
shale units, and reasonable matches of chemostratigraphic profiles carbonates. It was suggested that the Scout Mountain Member rep-
from underlying units (Macdonald et al. 2010a). The palaeolati- resents a distinct glacial event from the Edwardsburg metadiamic-
tudes determined herein for these deposits are slightly higher tites dated by Lund et al. (2003) at c. 685 Ma. More recently,
(c. 208) than the previous estimate (06 þ8/–78; Evans 2000), however, the older age has been ‘corrected’ to 686 + 4 Ma by
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 99

in his summary figure on glacial palaeolatitudes, that unit is


demonstrably early Palaeozoic in age and is therefore excluded
17a/ from the present analysis.
18a In the central Appalachians of eastern Laurentia, pebbly mud-
1a stone diamictites of the Grandfather Mountain Formation lie
immediately above rhyolitic flows dated at 742 + 2 Ma; this suc-
2a cession has been correlated with the Konnarock Formation diamic-
tites and rhythmites with dropstones. In the Blue Ridge Mountains
15a of Virginia, the c. 730–700 Ma Mechum River succession also
3a contains glaciogenic features (Bailey et al. 2007). Evans (2000)
correlated all three successions for the sake of simplicity, and esti-
mated depositional palaeolatitudes based on a palaeomagnetic pole
from the Franklin LIP at c. 720 Ma. In the present compilation,
7 however, we distinguish the Grandfather Mountain and Konnarock
formations from the Mechum River succession, based on slightly
different ages as reviewed above. For the former units, we apply
9/10 a new palaeomagnetic pole from the Kwagunt Formation in the
Grand Canyon (Weil et al. 2004), which is precisely coeval at
14a 742 + 6 Ma (Karlstrom et al. 2000). Assuming negligible rotation
11a of the Colorado Plateau relative to cratonic North America, as con-
13 sidered by Weil et al. (2004), this pole implies a palaeolatitude of
06 + 58 for the Konnarock/Grandfather Mountain region. Alter-
natively, a near-maximum estimate of Colorado Plateau rotation
of 128 anticlockwise would imply a palaeolatitude for Konnar-
ock/Grandfather Mountain at 04 + 58, within the error of the
no-rotation model. This estimate is rated as only moderately
reliable, because the dated Grandfather Mountain diamictites are
not yet conclusively demonstrated as glacially influenced, and
the presumed correlative Konnarock glaciogenic strata have
Fig. 7.2. Palaeoequators plotted across Laurentia, according to the named more lax age constraints.
palaeomagnetic poles from 750 to 700 Ma, as discussed in the text and in Evans The Mechum River succession remains within the age range of
(2000). Numbered stars correspond to glaciogenic deposits keyed to Table 7.1. the c. 720 Ma Franklin LIP, and the new grand-mean pole for that
Star colours are keyed to the reliability scheme of Figure 7.1, or if uncoloured, event (Denyszyn et al. 2009), implies a depositional palaeolatitude
stars represent deposits typically correlated with the Rapitan glaciation. of 16 + 38 (assuming an internally rigid Laurentian plate and a
References to palaeoequators given in Evans (2000) except for the recently geocentric axial dipole geomagnetic field). Stratigraphically
updated Franklin LIP mean (Denyszyn et al. 2009). higher in the succession, the Fauquier Formation contains a
basal diamictite member, followed by a thick sandstone interval
that is capped by a thin carbonate unit with a negative d13C excur-
the same authors (Fanning & Link 2008). These dates strengthen sion; in turn, this unit is followed conformably by basalts of the
the evidence for glaciogenic deposits in the Cordilleran rift c. 570 Ma Catoctin Formation (Hebert et al. 2010). Palaeomag-
basins that are younger than 700 Ma. Correlations across the Cor- netic results from the latter volcanic succession include two rema-
dilleran margin of Laurentia have emphasized the close association nence components of greatly differing inclination, hence, implied
of allegedly glaciogenic diamictites and volcanic rocks in that time palaeolatitude (Meert et al. 1994). The coexistence of both high-
frame. Evans (2000) suggested that all of these deposits could be and low-inclination directions in these rocks is symptomatic of
correlative, with an age of c. 740–720 Ma. In that case, palaeo- many broadly coeval igneous formations in eastern North America,
magnetic poles from the c. 720 Ma Franklin event could be used leading to a fundamental debate on their general interpretation
to infer palaeolatitudes of the Cordillera (applying the GAD field (e.g. Meert & Van der Voo 2001; Pisarevsky et al. 2001). The high-
assumption and a rigid Laurentian plate). However, new data palaeolatitude Catoctin component is corroborated by other data
suggest some of the deposits of this diamictite –volcanic associ- of generally higher quality than the low-palaeolatitude data from
ation are demonstrably younger, and therefore applications of the Iapetus margin of Laurentia (McCausland et al. 2007), so we
the Franklin palaeomagnetic pole to undated Cordilleran deposits prefer its implied palaeolatitude of 78 + 128 for the conformably
are tenuous. For these reasons, the Toby Formation’s deep-tropical underlying Fauquier Formation. This is the first near-polar esti-
palaeolatitude estimate (Evans 2000) is no longer considered to mate for a Neoproterozoic glacial deposit among recent palaeo-
be valid. geographic syntheses, but it is only somewhat reliable for the
Elsewhere in western Laurentia, the Johnnie Rainstorm following reasons: (i) the Catoctin palaeomagnetic data remain
Member’s incised canyons and features reminiscent of postglacial ambiguous with respect to the two remanence components, (ii)
cap carbonates (Corsetti & Kaufman 2003) still retain the near- the corroborating data from elsewhere in eastern Laurentia are
equatorial palaeolatitude of 01 + 48 as assessed by Evans also not of the highest quality, (iii) although the contact between
(2000). Stratigraphically below that unit, but above the Kingston the Fauquier Formation and the overlying basalts appears to be
Peak glaciogenic unit, the Ibex Formation contains a polymictic conformable, the basal Fauquier diamictites lie as much as hun-
conglomerate and overlying pink dolostone with negative carbon dreds of metres lower in the stratigraphy, and (iv) the 580–
isotope values similar to many postglacial cap carbonates (Corsetti 570 Ma interval appears to harbour very rapid apparent polar
& Kaufman 2005). However, it is impossible to apply palaeomag- wander for Laurentia (McCausland et al. 2007), so absolute age
netic constraints from better dated units elsewhere in Laurentia uncertainties of only a few million years can permit a wide
without precise age constraints. A newly recognized glacial/cap- range of possible palaeolatitudes (tens of degrees) for sedimentary
carbonate succession has been described in Sonora, Mexico (Cor- basins on that continent.
setti et al. 2007), lacking both a precise age and a palaeomagnetic In Scotland and stratigraphically correlative areas of northern
estimate of depositional palaeolatitude. Finally, although Evans Ireland, the Port Askaig glaciogenic level is now generally corre-
(2000) included the Florida Mountains (New Mexico) diamictite lated to the Sturtian ice ages (Prave 1999; Condon & Prave 2000;
100 D. A. D. EVANS & T. D. RAUB

McCay et al. 2006), but with the uncertainty surrounding the Norwegian sparagmite basins: younger than 620 + 15 Ma detrital
numerical age of Sturtian deposits, it is not possible to apply extra- zircons in a distally underlying quartzite (Bingen et al. 2005) or
basinal palaeomagnetic data to address the palaeolatitude of Port perhaps even younger than 561 + 4 Ma based on Re – Os dating
Askaig deposition. Older palaeomagnetic data obtained directly of the immediately underlying Biri black shale (Hannah et al.
from the Port Askaig succession were summarized by Evans 2007, abstract only). Nonetheless, the mid-Ediacaran apparent
(2000) as unreliable, a conclusion followed here. polar wander path for Baltica is highly uncertain (Iglesia-Llanos
McCay et al. (2006) also identified a so-called Marinoan equiv- et al. 2005; Meert et al. 2007). Where Vilchitsy and Blon diamictite
alent in northern Ireland, comprising the Stralinchy and Reelan units are in close stratigraphic proximity to c. 551 Ma Volhyn volca-
Formations, based on the presence of diamictite, exotic clasts nic rocks of the southwestern Baltic craton, palaeomagnetic results
interpreted as dropstones, and a cap carbonate with a negative from the latter units imply a depositional palaeolatitude of 50 + 38
d13C anomaly. As with other putative Laurentian glaciogenic (A component of Nawrocki et al. 2004); however, nearly coeval
deposits of this age (assumed to be c. 635 Ma by correlation; see sedimentary rocks of the White Sea region (555.3 + 0.3 Ma;
above), there are no available palaeomagnetic data from Laurentia Martin et al. 2000) yield a more convincingly primary palaeomag-
that can be applied either directly from the relevant sedimentary netic pole (Popov et al. 2002, 2005; Iglesia-Llanos et al. 2005) that
basin or from elsewhere across the palaeocontinent. implies a palaeolatitude of 28 + 38 for the Vilchitsy/Blon region.
The youngest purported Neoproterozoic glaciogenic unit in the Any of these younger poles could also apply to the Moelv diamic-
Scottish and northern Irish Dalradian stratigraphic succession is tite and correlative units in the Caledonides, now dated as late
the Loch na Cille boulder bed (and its correlative units). It Ediacaran as noted above.
closely overlies the Tayvallich Volcanics that have been recently Ediacaran palaeogeography of Baltica is usually paired with
re-dated at 601 + 4 Ma (Dempster et al. 2002). The two relevant eastern Laurentia, but both cratons’ palaeomagnetic data appear
palaeomagnetic poles from Laurentia that approximate this age to indicate rapid motions that challenge uniformitarian plate-
are from the Long Range dykes of Newfoundland and Labrador, tectonic interpretations (Abrajevitch & Van der Voo 2010;
dated by U – Pb at 615+2 and 614 þ 6/–4 Ma (Kamo et al. McCausland et al. 2011). Amid these uncertainties, the mid –late
1989; Kamo & Gower 1994) and the Grenville Dykes of southeast- Ediacaran glacial palaeolatitudes of Baltica must be considered
ern Ontario, dated at 590 þ 2/ –1 Ma (Kamo et al. 1995). Both unresolved at present.
data sets contain an unusually large spread of inclinations and
hence implied palaeolatitudes of emplacement, confounding sim-
ple interpretations (reviewed by Hodych & Cox 2007; McCaus- Altaids and Siberia
land et al. 2007). Low to moderate palaeolatitudes for the Long
Range dykes (Murthy et al. 1992) are supported by subsequent Chumakov (2009) has summarized the terminal Proterozoic to
work (McCausland et al. 2009, abstract only). For the younger, Cambrian stratigraphy across Kazakhstan and Kyrgyzstan, dis-
Grenville Dykes, a positive baked-contact test demonstrates that tinguishing between two prominent glacial levels (Satan, Rang,
the steep palaeomagnetic B remanence found in some of the or Dzhetym Formations below, and Baykonur above). Both
dykes is primary (Murthy 1971; Hyodo & Dunlop 1993). If these levels are younger than the c. 700 Ma volcanic complexes lying
results withstand further scrutiny, then the Loch na Cille boulder unconformably below. He also has described the Zabit Formation,
bed would have been deposited during a migration of Laurentia in the East Sayan region, as glaciogenic and correlative with the
from low to high latitudes. Baykonur level, which is in turn correlated with the uppermost
For the East Greenland Neoproterozoic succession, Mac glacial level in Tarim (Hankalchough Formation; see below). A
Niocaill et al. (2008) report in abstract some preliminary palaeo- complex palaeomagnetic data set was obtained from the succes-
magnetic data that would appear to indicate rapid motion of sion c. 1000 m or more above the Zabit diamictites into Cambrian
Laurentia across palaeolatitudes. The upper Tillite Group glacio- strata (Kravchinsky et al. 2010), which may be useful for tectonic
genic unit bears a high-palaeolatitude magnetic remanence direc- reconstructions but shed no direct light on glacial palaeolatitudes.
tion, but further sample analysis is needed to confirm that result. Two glaciogenic levels are also now recognized in the Dzhab-
khan region of western Mongolia (Macdonald et al. 2009b): the
basal, Maikhan Ul Member of the Tsagaan Oloom Formation,
Baltica and the Khongoryn Member in the middle of the same formation.
Chemostratigraphy of carbon isotopes in this succession supports
Several new geochronological and palaeomagnetic constraints have correlation with the mid-Cryogenian (so-called Sturtian) and late
forced revision of Evans’s (2000) estimate of depositional palaeola- Cryogenian (Marinoan) ice ages worldwide. Kravchinsky et al.
titude of Neoproterozoic glaciogenic deposits on Baltica. First, (2001) found several components of magnetization through
the Vestertana Group, in which the palaeomagnetically studied mainly the middle and upper parts of the Tsagaan Oloom
Nyborg Formation lies between diamictite units of the underlying Formation, and preferred a high-unblocking-temperature, low-
Smalfjord Formation and overlying Mortensnes Formation, has an palaeolatitude component as tentatively primary. That component,
imprecise palaeolatitude of 33 þ 14/– 128 (Torsvik et al. 1995). although of dual polarity, is not supported by field stability or
The reliability of this estimate has been downgraded because recent reversals tests on the magnetization age, so is only somewhat
syntheses of Baltica’s apparent polar wander path (Iglesia-Llanos reliable. Levashova et al. (2010) found a high-unblocking-
et al. 2005) show the Nyborg pole to coincide with a c. 510 Ma temperature component of magnetization in the underlying
overprint pole from elsewhere in Baltica. Because Middle Cam- Dzabkhan volcanic rocks, with U –Pb laser ablation ages of
brian Finnmarkian deformation is prevalent in the northernmost c. 800– 770 Ma, with moderate palaeolatitudes of igneous
Caledonides (Roberts 2003), there is a distinct possibility that emplacement. This estimate, however, would precede Maikhan
the Nyborg palaeomagnetic remanence is an overprint acquired Ul glaciogenic deposition by nearly 100 million years if the
shortly before Finnmarkian folding, and thus not indicative of latter is correlated to the Rapitan ice age as discussed above
depositional palaeolatitudes. In addition, given the possibility of (Macdonald et al. 2010a).
substantial age difference between the Smalfjord and Mortensnes New recognition of late Neoproterozoic glaciation in southern
Formations (e.g. correlations by Hoffman & Li 2009), the Siberia has focused on the Yenisei ridge and the pre-Sayan
Nyborg pole is now considered as only applicable (if primary) to region NW of Irkutsk. Sovetov (2002) and Sovetov & Komlev
the conformably underlying Smalfjord Formation. (2005) describe regional stratigraphic correlations around two dis-
Elsewhere in Baltica, depositional palaeolatitudes on Ediacaran tinct levels of diamictites with various glaciogenic features. The
glacial deposits are similarly or even more poorly constrained. older Chivida unit of the Chingasan Group is more limited in geo-
New ages exist for the Moelv diamictites of the southern graphic distribution in the Teya-Chapa trough (Yenisei region)
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 101

relative to the younger Pod’em Formation that is correlated to the palaeomagnetic studies from Tarim demand further work to
Marnya Formation best exposed along the Biryusa and Uda Rivers verify their applicability to Neoproterozoic glacial palaeolatitudes.
(pre-Sayan). The Chivida diamictites are closely associated with In the Qaidam terrane near the boundary between the Tarim and
volcanic rocks and are traditionally associated with an age of North China cratons, the Hongtiegou diamictite and overlying cap
c. 700– 750 Ma (Khomentovsky 1986), but they are not directly carbonate have been correlated, via stable isotope stratigraphy, to
dated with modern methods. The Marnya Formation contains the uppermost Ediacaran (Sinian) successions in South China
clasts that are speculated to derive from the Nersa Complex, a (Shen et al. 2010). Palaeomagnetic data are not available from
nearby mafic intrusive suite, according to Sovetov & Komlev the Neoproterozoic of Qaidam.
(2005). Those authors quoted an Ar/Ar age of 611 + 3 Ma for Chronology of the Sinian sedimentary basins in South China has
the Nersa Complex, but that age (now revised to 612 + 6 Ma), improved vastly during the past few years. Two glacial levels are
in fact, derives from other mafic intrusive suites in southern described with various local names, but generally referred to as the
Siberia (Gladkochub et al. 2006); thus the Marnya Formation dia- older Chang’an (or Tiesiao, or Jiangkou) episode and the younger
mictites lack any precise age constraints. Nantuo ice age. The Chang’an level is now bracketed in age
In the Patom region NE of Lake Baikal, at least two widespread between 725 + 10 Ma (youngest U – Pb SHRIMP population
glaciogenic levels have been reported: the older Kharlaktakh unit from detrital zircons in the underlying Danzhou Group; Zhang
of the Ballaganakh Group and the younger Dzhemkukan or et al. 2008) and 663 + 4 Ma (Zhou et al. 2004). The Nantuo
Bol’shoy Patom Formation of the Dal’nyaya Taiga Group (Pok- diamictite is younger than 654.5 + 3.8 Ma and contains a basal
rovskii et al. 2006; Sovetov 2008). No results from these succes- tuff layer dated at 636.3 + 4.9 Ma (Zhang et al. 2008), with a
sions are available in the global palaeomagnetic database. cap carbonate dated at 635.2 + 0.6 Ma (Condon et al. 2005).
The late Neoproterozoic to Early Cambrian palaeomagnetic Prior to the acquisition of these ages, Evans (2000) assigned a
apparent polar wander path for Siberia is highly contentious moderate depositional palaeolatitude to the Nantuo Formation
with a wide scatter of poles that universally imply low palaeolati- based on results from the immediately underlying Liantuo
tude for the craton; and if one includes the full range of data redbeds (Evans et al. 2000). However, given that the Liantuo For-
(Kirschvink & Rozanov 1984; Kravchinsky et al. 2001; Gallet mation contains an ash bed dated at 748 + 12 Ma (Evans et al.
et al. 2003; Metelkin et al. 2005) rather than a simplified selection 2000), the new dates from South China now preclude use of the
(Cocks & Torsvik 2007), the problem is compounded further. Liantuo pole to constrain the significantly younger ice ages
(Zhang et al. 2009).
Macouin et al. (2004) conducted a palaeomagnetic study on the
China cratons Doushantuo Formation, which conformably overlies the Nantuo
diamictites across most of South China. They obtained consistent
Recent work has helped refine the chronology and palaeomagnet- results throughout the entire formation, implying near-equatorial
ism of Neoproterozoic volcanic-sedimentary successions on the palaeolatitudes, except for the postglacial cap carbonate at its
margin of the Tarim block in NW China. Two principal regions base, which was palaeomagnetically unstable. Given that the
contain diamictites interpreted as glaciogenic deposits: Aksu in Doushantuo Formation is now constrained to span more than 80
the west, and Quruqtagh in the east. The Aksu succession contains million years of time (Condon et al. 2005), and that the direction
two diamictite levels, and a coarse magnetostratigraphic study by obtained by Macouin et al. (2004) is of single geomagnetic
Li et al. (1991) yielded near-equatorial palaeolatitudes throughout polarity, we suspect that it is a secondary overprint. The positive
the succession. More recently, Zhan et al. (2007) conducted a more fold test only constrains the age of magnetization to pre-Mesozoic,
detailed palaeomagnetic investigation of only the Sugetbrak For- and the direction is similar to both Early Cambrian and Silurian
mation, which rests with disconformity upon the upper diamictite directions for South China. Zhang et al. (2009) correct numerous
level (Yuermeinbrak Formation). Although Zhan et al. (2007) misconceptions about the stratigraphy and palaeomagnetic data-
suggested that their high-stability, dual-polarity direction for the base for the Sinian of South China, and the best estimate for
upper Aksu succession, which was acquired prior to folding of depositional palaeolatitude of the Nantuo diamictite reverts to
unspecified age, implies a palaeolatitude of c. 278 for the the value obtained by Zhang & Piper (1997) from Yunnan Pro-
Yuermeinbrak ice age, there is lingering concern over why the vince, at 37 + 98. An identical result has been obtained by
palaeomagnetic results by Li et al. (1991) and Zhan et al. Zhang et al. (2006, abstract only) from Guizhou Province, support-
(2007), in part from the same sampling region on the same strati- ing this conclusion. The soft-sediment fold test reported in the
graphic units, are so different. We tentatively accept the new con- earlier study (Zhang & Piper 1997) imparts a high level of confi-
straint as more reliable, because the pole by Li et al. (1991) is dence to the moderate-palaeolatitude determination.
similar to Cretaceous poles from Tarim and could represent an over-
print of that age.
Farther east in the Tarim block, the Quruqtagh region contains India to Nubia
three Neoproterozoic diamictite levels: Bayisi, Altungol-Tereeken
and Hankalchough. The oldest level lies immediately above Although Evans (2000) speculated that conglomerates within
bimodal volcanic rocks dated by U –Pb SHRIMP at 740 + 7 Ma the Penganga Group might be glaciogenic, more thorough study
(previously 755 + 15 Ma from the same sample; Xu et al. of that region suggests no demonstrable glacial influence, as is
2005), and below volcanic rocks dated at 725 + 10 Ma (Xu evident, for example, in the discussions by Mukhopadhyay &
et al. 2009). The proximity of ages between volcanism and diamic- Chaudhuri (2003) and Chakraborty et al. (2010). Other Neoproter-
tite deposition allows the use of palaeomagnetic data from the ozoic glaciogenic deposits in southern Asia are listed in Table 7.1.
former rocks (Huang et al. 2005) to constrain the latitude of glacia- In Iran, the so-called Infracambrian stratigraphic succession,
tion, assuming the Bayisi diamictites are glaciogenic. That palaeo- between Neoproterozoic metamorphic basement and overlying
magnetic work found a characteristic remanence component of nonmetamorphosed Cambrian sedimentary rocks, is variable in
high consistency but no field tests were performed on its stability; lithology but contains diamictite and lonestone-bearing units in
it is therefore unclear whether its near-equatorial palaeolatitude its lower part, named the Rizu Formation (Huckriede et al. 1962;
(01 þ 4/ –28) is primary or secondary. It is rated here as only Hamdi 1992). The upper part of the succession is equally litho-
somewhat reliable. The Altungol-Tereeken diamictites, overlain logically variable, as currently correlated across the country
by a notable pink cap carbonate, lie above the 725 Ma volcanic (Alsharhan & Nairn 1997), but contains rhyolites thought to be
rocks and below another volcanic horizon dated at 615 + 6 Ma consanguineous with c. 530 Ma intrusions (Ramezani & Tucker
(Xu et al. 2009). The Altungol-Tereeken interval has not been 2003). No reliable palaeomagnetic data have yet been published
studied palaeomagnetically. In summary, both of the two recent on the Rizu strata or correlative rocks.
102 D. A. D. EVANS & T. D. RAUB

Despite a new, precise U–Pb TIMS age of 711.5+ 1.1 Ma for in the range of c. 12 –218 (McWilliams & Schmidt 2003) that
the Ghubrah diamictite in the Jebel Akhdar region of northeastern could be representative of Croles Hill deposition. As these data
Oman (Bowring et al. 2007; all sources of error included), its palaeo- are reported in abstract only, they are omitted from our final
latitude remains poorly constrained. However, unmetamorphosed compilation.
deposits of the possibly Ghubrah-correlative, lower Mirbat sand- As for the Elatina glacial deposits and their continent-wide cor-
stone/diamictite assemblage, now renamed as the Ayn Formation relative units, new palaeomagnetic results add to an already long
(Rieu et al. 2006, 2007), yield a low-reliability palaeomagnetic history of palaeomagnetic studies of this unit (reviewed by
direction (Kempf et al. 2000). This under-defined magnetization Evans 2000; Williams et al. 2008). Schmidt et al. (2009) presented
(n ¼ 10 samples from two sites) would imply a palaeolatitude of a detailed analysis of palaeomagnetic data from both the Elatina
09 + 48 for the Ayn Formation, but we cannot include it in our and Nuccaleena formations in the Flinders Ranges, combining
final compilation due to the severely limited sample size. new results with previously published work. Based on moderate
On the younger and less precisely dated Fiq diamictite and its differences in remanent inclination between the two formations,
overlying Hadash cap carbonate, a two-polarity, pre-fold palaeo- they concluded that deglaciation coincided with rapid drift of the
magnetic remanence implies a palaeolatitude of 13 + 78 for the proto-Australian continent during a hiatus in deposition between
Jebel Akhdar region, and 18 + 78 for the Mirbat region (Kilner the two units. The observed mean remanent inclination of 12.98
et al. 2005), the latter area hosting the erosional remnants of an from the Elatina Formation would correspond to a depositional
upper diamictite unit (Shareef Formation) that is correlated with palaeolatitude of 6.58 if the geomagnetic field was dipolar and
the Fiq (Rieu et al. 2007). The folding is probably Palaeozoic in sedimentary inclination shallowing was negligible. For the Nucca-
age, suggesting that the remanence could be primary. However, leena Formation, the mean inclination of 34.98 would correspond
Rowan & Tait (2010, abstract only) have undertaken a more to a palaeolatitude of 19.28. Schmidt et al. (2009) also reported
detailed magnetostratigraphic study of the Hadash cap carbonate results from four measurements designed to quantify the effects
in two sections with greatly differing structural attitudes, and of remanence shallowing in the sedimentary rocks. In the first
have found the same remanence direction to be definitively post- measurement, a correction for their measured values of remanence
folding, and possibly associated with regional remagnetization anisotropy would increase the Elatina inclination to 14.08;
during emplacement of the Semail ophiolite. Given that only however, Tauxe et al. (2008) showed that this method commonly
about 20% of the samples in the study by Kilner et al. (2005) underestimates the total amount of inclination shallowing. Indeed,
yielded the preferred remanence direction, and only a few of two other measurements by Schmidt et al. (2009) appear to justify
their sites contributed significantly to their fold test, the low- such an assessment. Their second test, elongation/inclination
palaeolatitude estimates for Fiq and Shareef diamictites can only analysis on the distribution of data around the mean direction
be considered as somewhat reliable. Allen (2007) points out (using the assumed TK03 model of geomagnetic secular vari-
additional problems with the published data from Oman, in ation), would correct the Elatina inclination to 198 (bootstrap-
terms of their stratigraphic correlations and tectonic implications. generated uncertainties were not shown but are typically large
In the northern East African Orogen of Ethiopia, the Tambien for such tests). The third analysis, a dip test on crossbedded sand-
Group contains a glaciogenic/cap-carbonate succession that stones in the succession, would correct the Elatina inclination to
has only recently been described in detail (Beyth et al. 2003; between 18.4 and 20.28. Because sandstones should be the least
Miller et al. 2003). The Tambien Group is tightly folded into struc- compacted of all the Elatina siliciclastic facies, this correction of
tures that are crosscut by the Mereb granite suite, dated by zircon the aggregate palaeomagnetic results include data from both sand-
Pb-evaporation on two plutons at 606.0 + 0.9 and 613.4 + 0.9 stones and siltstones. The fourth test considered differences in
Ma (Miller et al. 2003), and by zircon U –Pb SHRIMP at inclination between carbonates and mudstones in the Nuccaleena
612.3 + 7.5 Ma (Avigad et al. 2007). Detrital zircons in the Formation, the latter of which are expected to be more compacted
Negash diamictite have yielded near-concordant U – Pb SHRIMP than the former. The carbonates’ mean inclination was 33.68
analyses as young as c. 750 Ma (Avigad et al. 2007; all younger (a95 ¼ 7.78), substantially greater than the mudstones’ mean incli-
diamictite detrital ages quoted in that work are more than 10% nation of 21.18 (a95 ¼ 7.88). Schmidt et al. (2009) claimed that
discordant). No palaeomagnetic data are available for the these two means (taking into account both declination and incli-
Tambien Group. Farther to the north in the Arabian –Nubian nation) were indistinguishable, using a test from McFadden &
Shield, widespread banded Fe-formation is associated with dia- Lowes (1981). However, the test they used from that paper
mictites of the Atud and Nuwaybah Formations (Stern et al. (equation 25) assumes an equivalency in concentration parameters
2006; Ali et al. 2009). between the two sample sets, which for the Nuccaleena data can be
rejected with .99% confidence (Fisher et al. 1987, p. 219). If the
correct test is used (the more general equation 23 from the same
Australia and Mawsonland paper), the Nuccaleena data-derived statistic increases to 4.40,
exceeding the p ¼ 0.05 F[2,76] value of 3.12; thus, the carbonates’
The palaeolatitude assessments by Evans (2000) remain valid for and mudstones’ inclinations are significantly distinct at more than
most of the many glaciogenic deposits scattered across the late 95% confidence, and the Nuccaleena data therefore do support
Neoproterozoic glacial record of Australia. Ediacaran stratigraphic inclination shallowing having affected any fine-grained sedimen-
correlations in Australia are now under intense debate, following tary rocks contributing to the mean directions in the sampling
the recognition of the possibly glaciogenic Croles Hill Diamictite area (contra Schmidt et al. 2009).
in Tasmania coincident with 582 + 4 Ma rhyolite volcanism In summary, the various inclination corrections to the Elatina
(Calver et al. 2004). Whether those diamictites correlate to the palaeomagnetic data, as reported by Schmidt et al. (2009), imply
Elatina Formation in South Australia (by way of the Cottons a ‘true’ inclination value around 208, with an implied palaeolati-
Breccia on King Island), or whether they are substantially tude of approximately 108 for the sampled areas in the Flinders
younger than Elatina, carries implications as profound as a factor- Ranges. Nuccaleena inclinations are steeper, due to the less-
of-two uncertainty in the duration of the Ediacaran Period that compacted carbonate lithology and perhaps also partly due to Aus-
begins, as formally defined, with deposition of the post-Elatina, tralian plate motion between the ice age and the deglacial interval.
Nuccaleena cap carbonate. No published palaeomagnetic data Raub (2008) found a mean inclination of 278 (a95 ¼ 3.58) for the
are available for the Tasmanian diamictites, but data are reported Nuccaleena Formation, with several correlatable magnetozones
in abstract for the Cottons Breccia, with an implied palaeolatitude spanning geodynamo reversals during deposition. That inclination
of 21 þ 9/ –88; and the subsequent igneous activity dated at value indicates a primary palaeolatitude of 14 + 28, between the
c. 580 Ma (Meffre et al. 2004), with implied palaeolatitudes estimates by Schmidt et al. (2009) for their compaction-corrected
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 103

Elatina and Nuccaleena data. Rare stratigraphic sections preserve Figure 7.3 shows two possible palaeogeographies of Australia
sedimentological conformity (mixed transition) between siliciclas- during the time of terminal Cryogenian glaciation. The first rep-
tic rocks indistinguishable from Elatina Formation sandstones and resentation (Fig. 7.3a) retains coherence of cratonic Australia in
Nuccaleena Formation cap-carbonate beds (Raub et al. 2007a), present-day coordinates, with palaeoequators according to the
and the onlap relationships between the two formations described various palaeomagnetic studies of the Elatina or Nuccaleena for-
by Schmidt et al. (2009) can be accounted for by sediment draping mations. The second (Fig. 7.3b) incorporates a new model of Pro-
of an irregular post-glacial topography, perhaps with diachronous terozoic Australia, with the northern part of the continent displaced
onset of cap-carbonate deposition, without a substantial lacuna from the western and southern parts by a local Euler rotation angle
between the formations (Rose & Maloof 2010). For this reason, of 408 (Li & Evans 2011). In this model, the palaeoequator deter-
we propose that 10–148 of palaeolatitude best characterizes mined from the Walsh Tillite cap carbonate (Li 2000) returns to a
the Elatina/Nuccaleena glacial/postglacial palaeogeography. position that is much closer to the palaeoequator attributed to the
The 10–208 palaeolatitude range (as indicated in Fig. 7.1) also Nuccaleena cap carbonate in South Australia.
well represents the distribution of likely Elatina-correlative depos- Southern Australia may also have been glaciated in the mid-
its across the Australian –Mawsonland palaeocontinent (Fig. 7.3). Ediacaran interval. Gostin et al. (2011) describe possible ice-
The compaction-corrected Elatina/Nuccaleena result is con- rafting features in the Bunyeroo Formation and equivalent units,
sidered here to be one of the two most reliable palaeolatitude deter- which are demonstrated to have a pre-folding palaeomagnetic
minations among all Neoproterozoic glaciogenic deposits. remanence that was acquired about 158 from the equator. In
Grey et al. (2005) produced a landmark correlation of Neopro- addition, Jenkins (2011) proposes a glacial influence on strata of
terozoic strata among deep boreholes throughout the Officer Basin the Billy Springs Formation higher in the stratigraphy.
of Western Australia, using diamictites and other indicators. New Most tectonic syntheses of the region formerly known as East
palaeomagnetic data from the Wahlgu diamictite level (correlated Gondwanaland now subdivide the cratonic areas of Antarctica
with the Elatina Formation) in the Lancer-1 borehole (Pisarevsky into (at least) three fragments that collided during Cambrian
et al. 2007) generate a palaeolatitude of 7 + 178, but the data time (e.g. Fitzsimons 2000). The Mawson Continent, or Mawson-
are highly scattered and merely consistent with the range of pre- land, comprises South Australia and adjacent regions of Antarctica
vious results from the Elatina Formation. Haines et al. (2008) in a pre-breakup Gondwanaland fit; those regions extend much of
revise the correlation of one diamictite and dropstone-bearing the length of the Transantarctic Mountains, from Victoria Land
unit (therein named the Pirrilyungka Formation) to an age sub- toward the South Pole. The Goldie Formation (redefined by
stantially older than the Wahlgu and Turkey Hill Formations, Myrow et al. 2002) contains diamictites of uncertain depositional
perhaps correlating with the Chambers Bluff diamictite in South setting. Although no estimates of palaeolatitude are available from
Australia. the Goldie Formation in the Transantarctic Mountains, its age is
In the Kimberley region of northern Australia, the cap carbonate now tightly constrained by U –Pb on zircon, 668 + 1 Ma, from
of the Walsh Tillite (western Kimberley) has yielded a moderate- interleaved mafic (meta)volcanic rocks (Goodge et al. 2002).
palaeolatitude magnetic remanence supported by a tectonic fold
test and two polarities of magnetization (Li 2000). Although the
Walsh Tillite is variously correlated with either the Sturtian or Kalahari and environs
Marinoan glacial level across Australia (see Corkeron 2007 for a
summary), that issue is irrelevant for the reliability of the palaeo- Neoproterozoic glaciogenic deposits on the Kalahari craton are
magnetic pole and its implications for deglaciation of the Walsh restricted to its present western margin, where as many as two
ice age. We accept the correlation by Corkeron & George (2001) diamictite levels are preserved in any single succession. In
of the Egan glaciogenic deposit (central Kimberley) with the addition, several levels of canyon incision with diamictic infill
Boonall dolomite in eastern Kimberley, and by further correlation are present within the Schwarzrand Subgroup of the Nama
using stromatolites, the mid-Ediacaran Julie Formation in central Group, spanning the Ediacaran –Cambrian boundary (reviewed
Australia. A palaeomagnetic pole from the latter region (Kir- by Evans 2000). No new palaeomagnetic constraints are available
schvink 1978) implies a depositional palaeolatitude of 21 + 88 for any of the pre-Schwarzrand glacial levels, and the Schwarzrand
for the Egan, assuming that the pole is primary and the correlations palaeolatitude of 38 + 38 (implied by the pole from the
are correct. 547 + 4 Ma Sinyai metadolerite on the Congo craton; Meert &

Fig. 7.3. (a) Palaeoequators plotted across


Australia and Mawsonland, according to
various palaeomagnetic studies of the
Elatina Formation, or, as indicated, of the
Nuccaleena Formation (Nucc). (b)
Restoration of Ediacaran tectonic rotation
between the northern and western/southern
portions of cratonic Australia, following Li
& Evans (2011). For star legend, see
caption to Figure 7.2 (uncoloured stars
representing deposits typically correlated
with the Elatina glaciation. Palaeoequators
are referenced as follows: EW86, Embleton
& Williams (1986); SWE91, Schmidt et al.
(1991); SW95, Schmidt & Williams (1995);
SCBK, Sohl et al. (1999); R08, Raub
(2008); SWM09, Schmidt et al. (2009);
Walsh, Li (2000).
104 D. A. D. EVANS & T. D. RAUB

Van der Voo 1996) is assigned only low reliability because, despite Several recent papers provide additional insight on correlations
excellent age constraints and the consistency of earliest Cambrian of diamictite units across the present eastern and northern edges
Gondwanaland palaeomagnetic poles from a variety of areas of the Congo – São Francisco craton. Melezhik et al. (2006)
within the megacontinent (Mitchell et al. 2010), it remains provide stable isotopic data bearing on regional and global strati-
unclear whether the Schwarzrand succession is truly glaciogenic. graphic correlations of the Geci metacarbonates in northern
Among older glaciogenic strata of the Gariep belt, Macdonald Mozambique, which are associated with metadiamictites that may
et al. (2010c) have revised correlations to distinguish two separate be glaciogenic (Pinna et al. 1993). To the NW, Poidevin (2007)
stratigraphic levels among outcrops previously lumped together reviews stratigraphic data from the Oubanguide belt in central
as the Numees Formation. In this new correlation, the name Africa, whereas Caron et al. (2010) report the discovery of a
‘Numees’ is retained for an older level associated with an organic- new diamictite cap-carbonate succession in southern Cameroon.
rich cap carbonate typical of ’Sturtian’ strata worldwide, whereas a Lack of precise ages on all of these units, as well as a general
new name, ‘Namaskluft’, is introduced for a younger level asso- dearth of late Neoproterozoic palaeomagnetic data from the
ciated with a light-coloured peloidal cap carbonate typical of Congo craton, prohibit assignment of depositional palaeolatitudes.
‘Marinoan’ strata in other regions. Kaufman et al. (2009) review Continuing around the São Francisco side of the craton,
these successions and present an unconventional set of correlations in Brazil, the various lithostratigraphic names for late Neoprotero-
with the global stratigraphic record, but it is not clear how such zoic diamictites are well summarized by Misi et al. (2007) and Sial
correlations might change according to the proposed distinction et al. (2009). As many as two or perhaps three glaciogenic levels
between Numees and Namuskluft glacial horizons. Farther south are known from each area, with few precise age constraints and
in the parautochthonous Gariep belt stratigraphy, the Aties For- thus, many possible correlations across the region (see also
mation as listed by Evans (2000) has been subdivided into Kaufman et al. 2009). A comprehensive summary of all these poss-
several new units, which include two glaciogenic levels recog- ible relationships is beyond the scope of the present paper, but
nized by Frimmel (2011): the lower Karoetjes Kop Formation, Table 7.1 splits the deposits into many distinct geographical and
and the upper Bloupoort Formation. The former is correlated tectonostratigraphic subdivisions, which later can be combined
with the Kaigas diamictites in the Gariep belt to the north, as correlations gain reliability. In the cratonic autochthon, the
whereas the latter is correlated to the Numees diamictites (s.l.), basal Bambuı́ cap carbonate (Sete Lagoas Formation) has
although further work is required to discern whether Bloupoort yielded a Pb/Pb isochron age of 740 + 22 Ma (Babinski et al.
would correspond to Numees or Namuskluft, as distinguished by 2007), and the authors of that study infer a depositional palaeola-
Macdonald et al. (2010c). titude according to c. 750 Ma palaeomagnetic data from Africa;
In the allochthonous Marmora terrane of the Gariep Belt, however, we consider the Sete Lagoas age constraints lax
Namibia, diamictites and dropstone-bearing stratigraphic units enough to render its palaeolatitude assignment too tentative.
are identified within the Dernburg Formation. These deposits are Another age constraint of particular note is an astounding Re –
interpreted as having been deposited on a seamount in the Adamas- Os black shale isochron in the 1100–1000 Ma range (Azmy
tor Ocean that lay between the Kalahari and Rio de la Plata cratons et al. 2008) for the Lapa Formation (uppermost formation in the
prior to their collision (Frimmel & Jiang 2001; Frimmel et al. diamictite-bearing Vazante Group) that, if correct, would represent
2002). No palaeomagnetic data are available for the Marmora the first discovery of a Mesoproterozoic ice age, anywhere in the
terrane, although extensive carbonates and evaporites in the suc- world. However, Rodrigues et al. (2008, extended abstract) find
cession suggest low to moderate palaeolatitudes of deposition. ,1000 Ma detrital zircons in all Vazante units, in direct contrast
Gaucher et al. (2009) included the Marmora terrane within a to the Re –Os ages. As for palaeomagnetic constraints, the
crustal block they named ‘Arachania’, inferred to have transferred Salitre cap carbonate of the Bebedouro Formation has only
from the Rio de la Plata craton to the Kalahari craton between yielded a Cambrian remagnetization (Trindade et al. 2004). No
c. 630 and 550 Ma. reliable palaeomagnetic data are available for the other diamictites
of the São Francisco craton.

Congo – São Francisco


West Africa and Hoggar
The Katangan succession in east-central Africa contains two pro-
minent diamictite horizons with overlying cap carbonates, Although Evans (2000), following earlier syntheses across the
named the Grand Conglomerat and Petit Conglomerat, in strati- Taoudeni basin, considered the possibility that the glaciogenic
graphic order. U– Pb SHRIMP ages of 765–735 Ma from volca- ‘triad’ succession is late Ediacaran to earliest Cambrian in age, a
nic strata, interbedded with diamictites that are correlated with more detailed study of that succession in the Volta basin (Porter
the Grand Conglomerat (Key et al. 2001), overlap with a et al. 2004) provides compelling chemostratigraphic comparisons
748 + 6 Ma date on the syenitic phase of the Mbozi complex with basal Ediacaran cap carbonates elsewhere in the world that
from the same craton (Mbede et al. 2004, abstract only). This are dated at 635 Ma (South China and Namibia, see above). Fur-
allows the Grand Conglomerat’s depositional latitude to be esti- ther, the assigned Early Cambrian age of the glaciogenic Hassanah
mated indirectly at 10 + 58, using the near-coeval Mbozi palaeo- Diallo Formation in Senegal, based on a small shelly fossil assem-
magnetic pole (Meert et al. 1995). This estimate is considered blage in dolomite resembling cap carbonate, is now retracted with
more reliable than the same pole applied for the Chuos Formation the interpretation that the fossiliferous sample was collected from a
in northern Namibia, which might be substantially younger than loose boulder fallen from overlying strata (Shields et al. 2007). For
the c. 750 Ma Mbozi pole. Both of these units are included separ- these reasons, the indirect estimation of triad palaeolatitudes from
ately in our new compilation because of the c. 1000 km lateral dis- an Early Cambrian aggregate Gondwana Land apparent polar
tance between their exposures, as well as the possibility that they wander path (Evans, 2000) now appears inappropriate. As dis-
are not precisely coeval. Our new estimate of Chuos depositional cussed in Evans (2000), previous palaeomagnetic data from the
palaeolatitude differs slightly from that of Evans (2000) due to a Taoudeni sedimentary succession are of questionable reliability.
refined location for the centre of Chuos-correlated strata in the However, a new result from one site of the ‘triad’ cap carbonate
northern and central Damara belt. As for the younger portion of in the Gourma region indicates a magnetic remanence that was
these successions, there is still no reliable palaeomagnetic con- acquired prior to Pan-African folding in the area (Boudzoumou
straint on that unit, despite the recent U –Pb TIMS age constraint et al. 2011). If primary, the result would imply a depositional
of 635.5 + 1.2 Ma on a likely correlative of the Ghaub Formation palaeolatitude of 08 + 78, but because of the small sample size,
in the internal sector of the Damara orogen (Hoffmann et al. 2004; the result warrants further study before it can be considered as
but see also Kaufman et al. 2009). reliable as the other determinations shown in Figure 7.1.
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 105

Alleged periglacial features from the base of the Atar Group Gaskiers-correlative age in an open ocean setting between 595
(Trompette 1994), which were discussed by Evans (2000) with and 550 Ma (Kawai et al. 2008). The authors of the study inferred
additional references, must now be considered as Mesoproterozoic, a moderate depositional palaeolatitude from the presumably
given multi-sample, black shale Re –Os isochron ages of adjacent Avalonian arc: the 580– 575 Ma Marystown Volcanics
c. 1105 Ma (Rooney et al. 2010) from somewhat higher levels (see above) and the c. 603 Ma Caldecote Volcanics plus related
in the succession. The ages are reminiscent of those from the intrusions (Vizan et al. 2003). We suspect that this estimate is
Vazante Group and its diamictites, marginal to the São Francisco unreliable, given the combined uncertainties in glacial influence
craton (see above), but the Atar Re –Os isochrons have lower of the Gwna beds, palaeogeographic affinity of the Mona
mean square of weighted deviates values and thus, are more complex relative to the Avalonian arc, and age of the succession.
reliable. The alleged periglacial influence on the basal Atar sedi- In the Cadomian terrane, Bohemian massif, marine diamictite
ments should be tested by further study. lenses associated with inferred, glacially influenced, base level
Despite new U –Pb ages from the Anti-Atlas Mountains of changes are found in the Clanzschwitz and Weesenstein Groups.
Morocco, late Neoproterozoic glaciogenic deposits there remain These lenses contain detrital zircons as young as c. 570 Ma, and
unconstrained in depositional palaeolatitude. The older Sarhro dia- therefore, their deposition is also approximately the same age (Lin-
mictites in the Sirwa region were deformed and metamorphosed nemann et al. 2007). These formations, in the Saxo-Thuringian
c. 660 Ma (Thomas et al. 2002, 2004). Although Thomas et al. Zone, were likely proximal to West Africa during Ediacaran to
(2004) advocate the traditional correlation of the Sarhro Group early Palaeozoic time; however, lack of reliable Ediacaran palaeo-
with the Tiddiline succession in the Bou Azzer –El Graara inlier, magnetic data from either the Cadomian Belt itself or the West
the latter diamictite-bearing succession contains boulder-sized African craton prohibit assignment of depositional palaeolatitudes
clasts of the Bleı̈da granodiorite now dated at 579.4 + 1.2 Ma for the Clanzschwitz and Weesenstein units.
(Inglis et al. 2004). Therefore, the new stratigraphic constraints
appear to distinguish the two successions in age by c. 100
million years. No reliable palaeomagnetic data exist for the older Amazonia and environs
age interval, but the Tiddiline succession depositional palaeolati-
tude could be estimated by the immediately overlying, and pene- In the Paraguay belt at the southeastern margin of the Amazon
contemporaneous (c. 575 –560 Ma; Thomas et al. 2002; Maloof craton, the Puga diamictite is overlain by the Mirassol d’Oeste
et al. 2005), Ouarzazate volcanic group. Unfortunately, the exist- cap carbonate. A characteristic component of magnetization has
ing palaeomagnetic results from those rocks are of near-zero been found in the cap dolostone (Trindade et al. 2003), carried
reliability, as summarized by Tohver et al. (2006). by both magnetite and haematite with high rock-magnetic stability
(Font et al. 2005) and multiple reversals were observed through
the stratigraphic section (Font et al. 2010) exposed in a single
Avalonia and Cadomia rock quarry. If primary, that magnetic remanence would indicate
a depositional palaeolatitude of 22 þ 6/– 58 for Puga deglaciation.
The Gaskiers glaciation is now dated by U –Pb TIMS to c. 580 Ma The direction, however, is coincident with the Mesozoic to recent
(Bowring et al. 2002, abstract only), and serendipitously, the age of expected directions at the sampling sites. Although the presence of
the Marystown volcanic rocks (used by Evans 2000 to estimate two magnetic polarities is encouraging, remagnetizations of that
Gaskiers depositional palaeolatitudes) has also been revised to nature are well known in the palaeomagnetic literature (e.g. Kent
c. 585–575 Ma (McNamara et al. 2001). The concordance of & Dupuis 2003), and the Mirassol d’Oeste characteristic rema-
these ages increases the likelihood that the Marystown palaeolati- nence is so far observed at only a single sampling locality with
tude of 34 þ 8/– 78 applies to the Gaskiers deposit. It is considered no field stability tests on its age. For this reason, it is considered
here as only moderately reliable as an indicator for global palaeo- here to be only moderately reliable.
climate models, because the Gaskiers unit is geographically Higher in the stratigraphy, the Serra Azul Formation is now
limited and of short-lived duration (Bowring et al. 2002), and recognized as a second glaciogenic level in the Paraguay belt
thus, it may represent only local, alpine glaciation in an active- (summarized by Alvarenga et al. 2009). Age constraints are lax,
margin tectonic setting. and its depositional palaeolatitude is unknown.
In the Boston basin, the Roxbury Conglomerate has yielded a Along coastal Peru, the Arequipa massif contains a diamictite
magnetization in nonglacial red slates bracketing the glaciogenic unit, the Chiquerı́o Formation, which has been correlated to glacio-
Squantum Member, which is consistent with that of unconform- genic deposits in eastern Laurentia (see references in Evans 2000).
ably underlying volcanic rocks, and partially supported by a posi- Recent U –Pb dating of its host terrane, the Arequipa-Antofalla
tive conglomerate test on those volcanic rocks (Fang et al. 1986). block (AAB), however, has identified more compelling tectonic
Further palaeomagnetism and geochronology of this same succes- affinities to the Amazon craton (Chew et al. 2007; Casquet et al.
sion (Thompson et al. 2007) has found that the c. 600– 595 Ma, 2010). In either case, without a precise age, the Chiquerı́o diamic-
pre-Squantum, Mattapan volcanic rocks yielded a distinct magne- tite’s depositional palaeolatitude remains unconstrained.
tization that had not been recovered previously, and the earlier
characteristic magnetization of Fang et al. (1986) was not repro-
duced. Thompson et al. (2007) reported a possibly primary com- Rio de la Plata craton and environs
ponent from only one site of the Squantum unit, and did not
discuss the discrepancy between their results and those of the The Las Ventanas and Playa Hermosa Formations, near the eastern
earlier study. Because the Squantum Member is younger than edge of the Rio de la Plata craton in southern Uruguay, contain the
593 Ma (U– Pb on detrital zircons; Thompson & Bowring 2000) most convincingly glaciogenic features in the entire region (Pazos
but otherwise poorly constrained in age, it is possible that the et al. 2003, 2008; Pecoits et al. 2008). Their age constraints are not
difference is due to plate motions and rotations of the Avalonian well defined within the Ediacaran Period. Palaeomagnetic results
arc, between deposition of the volcanic rocks and the overlying obtained directly from the Playa Hermosa Formation suggest a tro-
glaciogenic succession. We continue to use the directly obtained pical palaeolatitude of 13 þ 9/ –88 (Sánchez-Bettucci & Rapalini
result of Fang et al. (1986) in our analysis, implying a depositional 2002), but we exclude this result from our final analysis because of
palaeolatitude of 55 þ 8/– 78. very limited sampling in that preliminary study.
In the Anglesey region of Wales, an East Avalonian accre- Lower in the stratigraphy, within the same tectonic unit (Nico
tionary ‘Mona’ complex includes heterolithic, cobble-supporting, Pérez terrane and adjacent portion of the Dom Feliciano belt),
volcaniclastic mudstone (Gwna Group) of interpreted ophio- the Zanja del Tigre Formation is reported to contain a thin meta-
litic affinity, plausibly representing ice-rafted debris of diamictite horizon as part of a succession (Lavalleja Group) that
106 D. A. D. EVANS & T. D. RAUB

metamorphosed at c. 630 Ma (Pazos et al. 2008, fig. 4). A glacio- verify the primary age of magnetization; reproducibility by four
genic influence is disputed for these metamorphic rocks (Gaucher different palaeomagnetic laboratories through 15 years of study;
& Poiré 2009; Pecoits & Aubet, pers. comm. 2009) and, regardless, and several independent, quantitative methods to correct for
depositional palaeolatitudes for this unit are unconstrained. modest amounts of sedimentary inclination shallowing. Several
Within the southern Brazilian portion of the Dom Feliciano belt, lines of evidence demonstrate the effects of regionally cold cli-
Eerola (2006) identified isolated cobbles within schists of the mates near sea level during the Elatina ice age (Williams et al.
Brusque Group, which he correlated with the c. 750 Ma Chuos For- 2008). The conclusion appears inescapable: Neoproterozoic conti-
mation in Namibia. However, a much older age for the Brusque nental ice sheets extended deep within the tropics.
succession is possible, as it contains no detrital zircons younger Among the other high- to moderate-reliability data shown in
than 2000 Ma (Hartmann et al. 2003). A glaciogenic origin for Figure 7.1, the only other results obtained from sedimentary
this unit is not yet well demonstrated, but interestingly, Pazos rocks are the Puga, Nantuo and Walsh. Puga and Walsh magneti-
et al. (2008) suggest correlation of the Brusque Group with the zations are dominated by carbonate-associated haematite, so these
Lavalleja Group. This correlation suggests that scattered remnants results also may be less prone to inclination bias. The Nantuo result
of a pre-630 Ma ice age could be present throughout the region. represents siliciclastic rocks, so its remanence could be slightly
However, the name Lavalleja Group has been applied to various shallowed in inclination and hence, shallowed in palaeolatitude.
metamorphic rocks with protolith ages now found to range from However, estimates for the Konnarock, Rapitan, Mechum River,
Archaean to Ediacaran (Bossi & Cingolani 2009); thus, all corre- Grand Conglomerat, Upper Tindir and Gaskiers deposits are calcu-
lations should be treated with caution. lated indirectly from coeval igneous rocks on the same cratons.
Eerola (1995) also described diamictites and dropstone-bearing Those results are not affected by inclination shallowing.
laminated mudstones in the Camaquã basin, referred to as the Concerns that the Neoproterozoic geomagnetic field was
Passo da Areia Formation (see also Pazos et al. 2008) or, alterna- either nonaxial or nondipolar are valid in principle, but many
tively, the Picada das Graças Formation (Eerola 2006; Janikian palaeomagnetic results, including those from the Elatina-
et al. 2008). Any proposed glacial influence for that unit was dis- Nuccaleena succession, show typical patterns of a sporadically
counted by the latter study (Janikian et al. 2008), which provided reversing field with circularly symmetric secular variation about
a direct U –Pb SHRIMP zircon age of 580.0 + 3.6 Ma from an the time-averaged mean (e.g. Raub 2008; Schmidt et al. 2009;
interbedded tuff. Janikian et al. (2008) cite a palaeolatitude of Font et al. 2010). It should therefore be emphasized that a non-
23 þ 10/– 78 from the slightly older Hilario Formation volcanic actualistic geodynamo invoked to produce strictly polar or temper-
rocks, but as reviewed by Tohver et al. (2006), that study lacks ate glacial palaeolatitudes from the available palaeomagnetic
field tests and is only of moderate reliability. Nonetheless, database is as shockingly nonuniformitarian to geophysicists as
farther north in the Ribeira belt, Campanha et al. (2008) have near-equatorial glaciation is to palaeoclimatologists. Similarly,
determined a maximum age limit of c. 590– 580 Ma for diamictites hypotheses invoking rapid palaeolatitude shifts of continents to
of the highly deformed Iporanga Formation, which contain exotic generate erroneous palaeomagnetic latitudes, due to lax age con-
clasts and thus, could be of glaciogenic origin. Given the more con- straints in some instances, would require such motions at rates
vincing glacial influence on the Las Ventanas and Playa Hermosa beyond what is normally considered reasonable for plate tectonics.
formations of the same general age range on the proximal Rio de la Such shifts would need to be consistently biased to produce wholly
Plata craton, the Iporanga strata should be investigated further to low- to mid-palaeolatitudes at the expense of any polar results.
confirm or refute the influence of ice during sedimentation. Evans (2006) showed that Proterozoic evaporites produce sub-
Trindade & Macouin (2007) refer to glaciation at moderate to tropical palaeomagnetic latitudes, as expected for a uniformitarian
high palaeolatitudes in the Rio de la Plata craton, constrained by geodynamo, as well as low planetary obliquity. Williams (2008)
the ‘La Tinta’ palaeomagnetic pole (see below), and depicted as attempted to discredit the latter conclusion by appealing to uncer-
such in a 620 Ma global reconstruction. That citation probably tainties in geomagnetic polarity, to generate an equatorial mean for
refers to the Sierra del Volcán Formation, recognized as bearing the evaporites, as predicted by numerical simulations of high-
glaciogenic features with a possible Sturtian age (Pazos et al. obliquity palaeoclimate. However, that issue was already foreseen
2003). However, the unit lacks precise geochronological data and refuted by Evans (2006), who wrote (p. 53) ‘A zero mean could
(Pazos et al. 2008) and in fact could be as young as Ordovician be obtained if half of the palaeolatitudes were considered to rep-
(unpublished detrital zircon ages cited by Poiré & Gaucher resent the opposite hemisphere (geomagnetic polarity being gener-
2009). Regardless, the La Tinta pole is now considered to represent ally unknown in the Precambrian), but even then the total
a probable Cambrian overprint (Rapalini & Sánchez-Bettucci distribution would be significantly bimodal, in contrast with the
2008), leaving these deposits completely unconstrained in deposi- high-obliquity model predictions.’
tional palaeolatitude. Our summary analysis (Fig. 7.1) differs from the similar figure
in Evans (2000) in several minor ways; most importantly (i) a
slight increase in the estimated depositional palaeolatitude of the
Discussion Rapitan deposits, based on new geochronology that allows
precise application of the higher-reliability Franklin large
Figure 7.1 summarizes the results of our updated synthesis. igneous province pole that is immune to sedimentary inclination
Although several of the assessments of glacial palaeolatitudes shallowing, (ii) a slight increase in the estimated depositional
have changed in detail, owing to new geochronologic or palaeo- palaeolatitude of the Elatina deposits, based on new studies that
magnetic data, the general result is essentially the same as that quantify such inclination shallowing by several independent
determined by Evans (2000): low and moderate palaeolatitudes methods, (iii) exclusion of the Jbeliat palaeolatitude estimate due
are abundant, at all reliability levels, whereas few if any near-polar to its more conclusively pre-Cambrian age and thus inappropriate-
(.608) glacial deposits have thus far been identified in the Neo- ness of applying Early Cambrian palaeolatitudes from the Gond-
proterozoic rock record. From a total of 137 stratigraphic units wana Land apparent polar wander path, and (iv) inclusion of the
distinguished in the present analysis, only 22 deposits have deposi- somewhat reliable, near-polar palaeolatitude for the Fauquier For-
tional palaeolatitude constraints that are at least ‘somewhat’ mation as determined by palaeomagnetism on the overlying Catoc-
reliable, and among these, merely 10 are constrained moderately tin basalts and related intrusions. This last entry needs to be
well. Only three deposits (Upper Tindir, Nantuo and Elatina) are confirmed, but if so, it has the potential to strike down a pillar of
considered to have the highest degree of reliability, and all were the Precambrian high-obliquity Earth model (Williams 1993,
laid down in low to moderate palaeolatitudes. The lowest of 2008) – that Precambrian glaciogenic deposits are only known
these, Elatina, is supported by numerous field stability tests to from the palaeo-tropics, rather than the palaeo-poles. In general,
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 107

however, our update does not significantly affect the basic con- attractive model for evaluating the enigmatic geological record
clusion by Evans (2000) of a preponderance of low- to mid- of icehouse and hothouse environmental conditions just prior to
latitudes for Neoproterozoic glaciation. the dawn of animal life on our planet.
Our analysis is the most detailed of its kind in the published lit-
erature. Although our conclusions are nearly opposite to those of We thank the following individuals for constructive criticisms on drafts of this
Eyles & Januszczak (2004), we note that their analysis included manuscript: N. Aubet, M. Babinski, C. Gaucher, Z.-X. Li, A. Maloof, P. Pazos,
less than a fifth of the number of deposits considered herein. E. Pecoits, S. Pisarevsky, P. Schmidt, R. Van der Voo, and G. Williams. We
Among the deposits accepted by them as having ‘demonstrated’ would particularly like to thank E. Arnaud for her remarkable patience throughout
glacial influence, five have palaeolatitudes of at least somewhat the submission and revision process. Our research on Neoproterozoic ice ages has
reliable quality ranging between 13 and 378. One of the deposits been supported by the David and Lucile Packard Foundation, Agouron Institute
(Rapitan) has moderate palaeomagnetic reliability indicating tropi- for Geobiology, National Science Foundation and Australian Research Council’s
Tectonics Special Research Centre. This represents a contribution of the IUGS-
cal palaeolatitudes. Although we did not assess the evidence for or
and UNESCO-funded IGCP (International Geoscience Programme) Project #512.
against glacial influence on sedimentation in our present analysis,
we anticipate that future efforts in that regard will consider the full
extent of units listed in Table 7.1.
Trindade & Macouin (2007) also considered only a small References
fraction of the deposits listed in Table 7.1, although their con-
Abrajevitch, A. & Van der Voo, R. 2010. Incompatible Ediacaran
clusion is similar to ours. Considering only the subset of deposits
paleomagnetic directions suggest an equatorial geomagnetic dipole
discussed by Trindade & Macouin (2007), the most striking differ-
hypothesis. Earth and Planetary Science Letters, 293, 164– 170.
ences between their conclusions and ours are as follows: (i) our Ali, K. A., Stern, R. J., Manton, W. I., Johnson, P. R. & Mukherjee,
exclusion of the Scout Mountain palaeolatitude based on an S. K. 2009. Neoproterozoic diamictite in the Eastern Desert of Egypt
updated age for that unit, which differs from the palaeomagneti- and Northern Saudi Arabia: evidence of 750 Ma glaciation in the
cally studied rock units elsewhere in Laurentia, (ii) our inclusion Arabian– Nubian Shield? International Journal of Earth Sciences,
of a Gaskiers palaeolatitude applied indirectly from a palaeomag- doi: 10.1007/s00531-009-0427-3.
netic study of precisely coeval igneous rocks in Avalon, (iii) our Allen, P. A. 2007. The Huqf Supergroup of Oman: basin development
estimate of mid-latitude deposition for the Nantuo ice age, rather and context for Neoproterozoic glaciation. Earth-Science Reviews,
than near-equatorial, based on our selection of the direct Nantuo 84, 139– 185.
constraint, rather than the (in our opinion) lower-reliability Doush- Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to Snowball
antuo result, and (iv) our estimate of low-latitude deposition of the Earth. Nature Geoscience, 1, 817– 825.
Playa Hermosa Formation, preferred over the higher latitudes Alsharhan, A. S. & Nairn, A. E. M. 1997. Chapter 3: Infracambrian
implied by the now-obsolete La Tinta Formation. of the Middle East. In: Alsharhan, A. S. & Nairn, A. E. M. (eds)
The most interesting comparison to earlier palaeogeographic Sedimentary Basins and Petroleum Geology of the Middle East.
syntheses is between our present analysis and that of Hoffman & Elsevier, Amsterdam, 65 – 86.
Li (2009). The essence of our conclusions is similar: only low- Alvarenga, C. J. S. de, Boggiani, P. C., Babinski, M., Dardenne,
and mid-latitudes have been moderately or highly reliably deter- M. A., Figueiredo, M., Santos, R. V. & Dantas, E. L. 2009.
mined for most Neoproterozoic glacial deposits. The principal The Amazonian palaeocontinent. In: Gaucher, C., Sial, A. N., Hal-
exception to this rule stems from Ediacaran deposits such as verson, G. P. & Frimmel, H. E. (eds) Neoproterozoic– Cambrian
Tectonics, Global Change and Evolution. Elsevier, Amsterdam,
Loch na Cille, Squantum, Gaskiers and Serra Azul, all of which
Developments in Precambrian Geology, 16, 15 –28.
are estimated at .608 palaeolatitude by Hoffman & Li (2009).
Avigad, D., Stern, R. J., Beyth, M., Miller, N. & McWilliams, M. O.
Among those units, we only consider Squantum and Gaskiers as 2007. Detrital zircon U– Pb geochronology of Cryogenian diamictites
having direct palaeomagnetic constraints classified as even some- and Lower Paleozoic sandstone in Ethiopia (Tigrai): age constraints
what reliable, and the Gaskiers palaeolatitude, by our estimation, is on Neoproterozoic glaciation and crustal evolution of the southern
remarkably different at 31 þ 10/– 88. These discrepancies can be Arabian– Nubian Shield. Precambrian Research, 154, 88– 106.
attributed to the very distinct approaches between the two Azmy, K., Kendall, B., Creaser, R. A., Heaman, L. & de Oliveira,
studies. We took a bottom-up approach that considered ages and T. F. 2008. Global correlation of the Vazante Group, São Francisco
palaeolatitudes unconstrained unless demanded by strictly relevant Basin, Brazil: Re– Os and U– Pb radiometric age constraints. Pre-
data, whereas Hoffman & Li (2009) adopted a top-down approach cambrian Research, 164, 160– 172.
that assumed known glacial ages within a specified global plate- Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating of the
kinematic model. That model was based on a mostly independent Sete Lagoas cap carbonate (Bambuı́ Group, Brazil) and implications
palaeomagnetic database obtained primarily from well-dated for the Neoproterozoic glacial events. Terra Nova, 19, 401– 406.
igneous rocks with little relationship to glacial units. So, although Bailey, C. M., Peters, S. E., Morton, J. & Shotwell, N. L. 2007. The
both studies derive from the palaeomagnetic method, they are quite Mechum River Formation, Virginia Blue Ridge: a record of Neopro-
distinct in the assumptions and selection of data. The principal dis- terozoic and Paleozoic tectonics in southeastern Laurentia. American
cordance between our work and that of Hoffman & Li (2009) Journal of Science, 307, 1– 22.
occurs during the mid-Ediacaran interval, when the global palaeo- Beyth, M., Avigod, D., Wetzel, H.-U., Matthews, A. & Berhe, S. M.
geographic models are most uncertain. This period is associated 2003. Crustal exhumation and indications for Snowball Earth in the
with the fastest continental motions in Earth history, which may East African Orogen: north Ethiopia and east Eritrea. Precambrian
Research, 123, 187–201.
be attributed, in part, to global tectonic reorganization (Hoffman
Bingen, B., Griffin, W. L., Torsvik, T. H. & Saeed, A. 2005. Timing of
1999), rapid true polar wander (Raub et al. 2007b) and an isolated
Late Neoproterozoic glaciation on Baltica constrained by detrital
episode of nonuniformitarian geodynamo behaviour (Abrajevitch zircon geochronology in the Hedmark Group, south-east Norway.
& Van der Voo 2010). If such an isolated episode occurred, and Terra Nova, 17, 250–258.
it is far from proven, then it would appear to be limited to the mid- Bossi, J. & Cingolani, C. 2009. Extension and general evolution of the
Ediacaran interval rather than the late Cryogenian, Elatina or Puga Rı́o de la Plata Craton. In: Gaucher, C., Sial, A. N., Halverson,
ice ages (with uniformitarian symmetry in two-polarity palaeo- G. P. & Frimmel, H. E. (eds) Neoproterozoic –Cambrian Tectonics,
magnetic directions), or the older Cryogenian, Upper Tindir/ Global Change and Evolution. Elsevier, Amsterdam, Developments
Rapitan ice age (as judged by the highly stable Franklin palaeo- in Precambrian Geology, 16, 73– 85.
magnetic pole). Boudzoumou, F., Vandamme, D., Affaton, P. & Gattacceca, J. 2011.
In summary, the palaeomagnetic case for low-latitude Neopro- Neoproterozoic paleomagnetic poles in the Taoudeni basin (West
terozoic glaciation is strong, and Snowball Earth remains an Africa). Comptes Rendus Geoscience, 343, 284– 294.
108 D. A. D. EVANS & T. D. RAUB

Bowring, S. A., Landing, E., Myrow, P. & Ramezani, J. 2002. Abstract constraining block rotations in the Nares Strait region. Canadian
#13045-Geochronological constraints on terminal Neoproterozoic Journal of Earth Sciences, 46, 689–705.
events and the rise of metazoans. Astrobiology, 2, 457– 458. Eerola, T. 2006. Myöhäisproterotsooiset ilmastonmuutokset Tutkimuk-
Bowring, S. A., Grotzinger, J. P., Condon, D. J., Ramezani, J., sia Etelä-Brasiliassa [Neoproterozoic climate changes; research on
Newall, M. J. & Allen, P. A. 2007. Geochronologic constraints southern Brazil; Finnish with English abstract and bilingual figure
on the chronostratigraphic framework of the Neoproterozoic Huqf captions]. Geologi, 58, 164– 174.
Supergroup, Sultanate of Oman. American Journal of Science, 307, Embleton, B. J. J. & Williams, G. E. 1986. Low palaeolatitude of depo-
1097– 1145. sition for late Precambrian periglacial varvites in South Australia:
Calver, C. R., Black, L. P., Everard, J. L. & Seymour, D. B. 2004. implications for palaeoclimatology. Earth and Planetary Science
U –Pb zircon age constraints on late Neoproterozoic glaciation in Letters, 79, 419– 430.
Tasmania. Geology, 32, 893–896. Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic
Campanha, G. A. C., Basei, M. S., Tassinari, C. C. G., Nutman, A. P. & constraints upon the Neoproterozoic climatic paradox. American
Faleiros, F. M. 2008. Constraining the age of the Iporanga For- Journal of Science, 300, 347– 433.
mation with SHRIMP U– Pb zircon: implications for possible Edia- Evans, D. A. D. 2003. A fundamental Precambrian – Phanerozoic shift in
caran glaciation in the Ribeira Belt, SE Brazil. Gondwana Earth’s glacial style? Tectonophysics, 375, 353– 385.
Research, 13, 117– 125. Evans, D. A. D. 2006. Proterozoic low orbital obliquity and axial-dipolar
Caron, V., Ekomane, E., Mahieux, G., Moussango, P. & Ndjeng, E. geomagnetic field from evaporite palaeolatitudes. Nature, 444,
2010. The Mintom Formation (new): sedimentology and geochemis- 51 – 55.
try of a Neoproterozoic, Paralic succession in south-east Cameroon. Evans, D. A. D., Li, Z. X., Kirschvink, J. L. & Wingate, M. T. D. 2000.
Journal of African Earth Sciences, 57, 367– 385. A high-quality mid-Neoproterozoic paleomagnetic pole from the
Casquet, C., Fanning, C. M., Galindo, C., Pankhurst, R. J., Rapela, South China block, with implications for ice ages and the breakup
C. W. & Torres, P. 2010. The Arequipa Massif of Peru: new configuration of Rodinia. Precambrian Research, 100, 313– 335.
SHRIMP and isotope constraints on a Paleoproterozoic inlier in the Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for
Grenvillian orogen. Journal of South American Earth Sciences, 29, Neoproterozoic glaciations during the breakup of Rodinia after 750
128– 142. Ma. Earth-Science Reviews, 65, 1– 73.
Chakraborty, P. P., Dey, S. & Mohanty, S. P. 2010. Proterozoic plat- Fairchild, I. J. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the
form sequences of Peninsular India: implications towards basin evol- Earth System. Journal of the Geological Society of London, 164,
ution and supercontinent assembly. Journal of Asian Earth Sciences, 895–921.
39, 589– 607. Fang, W., Van der Voo, R. & Johnson, R. J. E. 1986. Eocambrian paleo-
Chew, D., Kirkland, C., Schaltegger, U. & Goodhue, R. 2007. Neo- magnetism of the Boston basin: evidence for displaced terrane. Geo-
proterozoic glaciation in the Proto-Andes: tectonic implications and physical Research Letters, 13, 1450–1453.
global correlation. Geology, 35, 1095–1098. Fanning, C. M. & Link, P. K. 2004. U –Pb SHRIMP ages of Neoproter-
Chumakov, N. M. 2009. The Baykonurian glaciohorizon of the Late ozoic (Sturtian) glaciogenic Pocatello Formation, southeastern Idaho.
Vendian. Stratigraphy and Geological Correlation, 17, 373– 381. Geology, 32, 881–884.
Cocks, L. R. M. & Torsvik, T. H. 2007. Siberia, the wandering northern Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian Gla-
terrane, and its changing geography through the Palaeozoic. Earth- ciation; data from the Adelaide Geosyncline, South Australia and
Science Reviews, 82, 29 –74. Pocatello Formation, Idaho, USA. Geological Society of Australia
Condon, D. J. & Prave, A. R. 2000. Two from Donegal: Neoproterozoic Abstracts, 91, 57 – 62.
glacial episodes on the northeast margin of Laurentia. Geology, 28, Fisher, N. I., Lewis, T. & Embleton, B. J. J. 1987. Statistical Analysis of
951– 954. Spherical Data. Cambridge, Cambridge University Press.
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005. Fitzsimons, I. C. W. 2000. Grenville-age basement provinces in East Ant-
U –Pb ages from the Neoproterozoic Doushantuo Formation, China. arctica: evidence for three separate collisional orogens. Geology, 28,
Science, 308, 95 –98. 879– 882.
Corkeron, M. 2007. ‘Cap carbonates’ and Neoproterozoic glacigenic Font, E., Trindade, R. I. F. & Nédélec, A. 2005. Detrital remanent mag-
successions from the Kimberley region, north-west Australia. Sedi- netization in haematite-bearing Neoproterozoic Puga cap dolostone,
mentology, 54, 871– 903. Amazon craton: a rock magnetic and SEM study. Geophysical
Corkeron, M. L. & George, A. D. 2001. Glacial incursion on a Journal International, 163, 491– 500.
Neoproterozoic carbonate platform in the Kimberley region, Austra- Font, E., Nédélec, A., Trindade, R. I. F. & Moreau, C. 2010. Fast or
lia. Geological Society of America Bulletin, 113, 1121– 1132. slow melting of the Marinoan Snowball Earth? The cap dolostone
Corsetti, F. A. & Kaufman, A. J. 2003. Stratigraphic investigations of record. Palaeogeography, Palaeoclimatology, Palaeoecology, 295,
carbon isotope anomalies and Neoproterozoic ice ages in Death 215– 225.
Valley, California. Geological Society of America Bulletin, 115, Frimmel, H. E. 2011. The Karoetjes Kop and Bloupoort Formations,
916– 932. Gifberg Group, South Africa, In: Arnaud, E., Halverson, G. P. &
Corsetti, F. A. & Kaufman, A. J. 2005. The relationship between the Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
Neoproterozoic Noonday Dolomite and the Ibex Formation: new Glaciations. Geological Society, London, Memoirs, 36, 233 –237.
observations and their bearing on ‘Snowball Earth’. Earth-Science Frimmel, H. E. & Jiang, S.-Y. 2001. Marine evaporites from an oceanic
Reviews, 73, 63– 78. island in the Neoproterozoic Adamastor ocean. Precambrian
Corsetti, F. A. & Lorentz, N. J. 2006. On Neoproterozoic cap carbon- Research, 105, 57– 71.
ates as chronostratigraphic markers. In: Xiao, S. & Kaufman, A. J. Frimmel, H. E., Fölling, P. G. & Eriksson, P. G. 2002. Neoproterozoic
(eds) Neoproterozoic Geobiology and Paleobiology. Springer, tectonic and climatic evolution recorded in the Gariep Belt, Namibia
New York, 273– 294. and South Africa. Basin Research, 14, 55 –67.
Corsetti, F. A., Stewart, J. H. & Hagadorn, J. W. 2007. Neopro- Gallet, Y., Pavlov, V. & Courtillot, V. 2003. Magnetic reversal fre-
terozoic diamictite-cap carbonate succession and d13C chemo- quency and apparent polar wander of the Siberian platform in the ear-
stratigraphy from eastern Sonora, Mexico. Chemical Geology, 237, liest Palaeozoic, inferred from the Khorbusuonka river section
129– 142. (northeastern Siberia). Geophysical Journal International, 154,
Dempster, T. J., Rogers, G. et al. 2002. Timing of deposition, orogen- 829– 840.
esis and glaciation within the Dalradian rocks of Scotland: constraints Gaucher, C. & Poiré, D. G. 2009. Palaeoclimatic events. Neoprotero-
from U– Pb zircon ages. Journal of the Geological Society of London, zoic– Cambrian evolution of the Rı́o de la Plata Palaeocontinent.
159, 83 – 94. In: Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel,
Denyszyn, S. W., Halls, H. C., Davis, D. W. & Evans, D. A. D. 2009. H. E. (eds) Neoproterozoic– Cambrian Tectonics, Global Change
Paleomagnetism and U– Pb geochronology of Franklin dykes in High and Evolution. Elsevier, Amsterdam, Developments in Precambrian
Arctic Canada and Greenland: a revised age and paleomagnetic pole Geology, 16, 123–130.
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 109

Gaucher, C., Frimmel, H. E. & Germs, G. J. B. 2009. Tectonic events Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary
and palaeogeographic evolution of southwestern Gondwana in the Science Letters, 258, 114–131.
Neoproterozoic and Cambrian. In: Gaucher, C., Sial, A. N., Hal- Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley,
verson, G. P. & Frimmel, H. E. (eds) Neoproterozoic–Cambrian J. L. 2004. U– Pb zircon date from the Neoproterozoic Ghaub Forma-
Tectonics, Global Change and Evolution. Elsevier, Amsterdam, tion, Namibia: constraints on Marinoan glaciation. Geology, 32,
Developments in Precambrian Geology, 16, 295–316. 817– 820.
Gladkochub, D. P., Wingate, M. T. D., Pisarevsky, S. A., Donskaya, Huang, B., Xu, B., Zhang, C., Li, Y. & Zhu, R. 2005. Paleomagnetism of
T. V., Mazukabzov, A. M., Ponomarchuk, V. A. & Stanevich, A. the Baiyisi volcanic rocks (ca. 740 Ma) of Tarim, Northwest China: a
M. 2006. Mafic intrusions in southwestern Siberia and implications continental fragment of Neoproterozoic Western Australia? Precam-
for a Neoproterozoic connection with Laurentia. Precambrian brian Research, 142, 83 –92.
Research, 147, 260– 278. Huckriede, R., Kürsten, M. & Venzlaff, H. 1962. Zur Geologie des
Goodge, J. W., Myrow, P., Williams, I. S. & Bowring, S. A. 2002. Gebiets zwischen Kerman und Saghand (Iran). Beihefte zum Geolo-
Age and provenance of the Beardmore Group, Antarctica: constraints gischen Jahrbuch, 51, 197.
on Rodinia supercontinent breakup. Journal of Geology, 110, Hyodo, H. & Dunlop, D. J. 1993. Effect of anisotropy on the paleomag-
393– 406. netic contact test for a Grenville dike. Journal of Geophysical
Gostin, V. A., McKirdy, D. M., Webster, L. J. & Williams, G. E. 2011. Research, 98, 7997–8017.
Mid-Ediacaran ice-rafting in the Adelaide Geosyncline and Officer Iglesia-Llanos, M. P., Tait, J. A., Popov, V. & Abalmassova, A. 2005.
Basin, South Australia. In: Arnaud, E., Halverson, G. P. & Palaeomagnetic data from Ediacaran (Vendian) sediments of the
Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Arkhangelsk region, NW Russia: an alternative apparent polar
Glaciations. Geological Society, London, Memoirs, 36, 673– 676. wander path of Baltica for the Late Proterozoic – Early Palaeozoic.
Grey, K., Hocking, R. M. et al. 2005. Lithostratigraphic nomenclature Earth and Planetary Science Letters, 240, 732– 747.
of the Officer Basin and correlative parts of the Paterson Orogen, Inglis, J. D., MacLean, J. S., Samson, S. D., D’Lemos, R. S., Admou, H.
Western Australia. Western Australia Geological Survey, Report, & Hefferan, K. 2004. A precise U –Pb zircon age for the Bleı̈da
93, 89. granodiorite, Anti-Atlas, Morocco: implications for the timing of
Haines, P. W., Hocking, R. M., Grey, K. & Stevens, M. K. 2008. Vines deformation and terrane assembly in the eastern Anti-Atlas.
1 revisited: are older Neoproterozoic glacial deposits preserved in Journal of African Earth Sciences, 39, 277–283.
Western Australia? Australian Journal of Earth Sciences, 55, Janikian, L., Almeida, R. P. et al. 2008. The continental record of Edia-
397– 406. caran volcano-sedimentary successions in southern Brazil and their
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice, global implications. Terra Nova, 20, 259– 266.
A. H. N. 2005. Toward a Neoproterozoic composite carbon-isotope Jenkins, R. J. F. 2011. Billy Springs glaciation, South Australia. In:
record. Geological Society of America Bulletin, 117, 1181–1207. Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A. logical Record of Neoproterozoic Glaciations. Geological Society,
2007. Evolution of the 87Sr/86Sr composition of Neoproterozoic London, Memoirs, 36 693– 699.
seawater. Palaeogeography, Palaeoclimatology, Palaeoecology, Kamo, S. L. & Gower, C. F. 1994. U –Pb baddeleyite dating clarifies age
256, 103– 129. of characteristic paleomagnetic remanence of Long Range dykes,
Hambrey, M. J. & Harland, W. B. (eds) 1981. Earth’s Pre-Pleistocene southeastern Labrador. Atlantic Geology, 30, 259–262.
Glacial Record. Cambridge University Press, Cambridge, 1004. Kamo, S. L., Gower, C. F. & Krogh, T. E. 1989. Birthdate for the Iapetus
Hamdi, B. 1992. Late Precambrian glacial deposits in central Iran. 29th Ocean? A precise U –Pb zircon and baddeleyite age for the Long
International Geological Congress, Abstracts, 2, 232. Range dikes, southeast Labrador. Geology, 17, 602–605.
Hannah, J., Yang, G., Bingen, B., Stein, H. & Zimmerman, A. 2007. Kamo, S. L., Krogh, T. E. & Kumarapeli, P. S. 1995. Age of the
560 and 300 Ma Re– Os ages constrain Neoproterozoic glaciation Grenville dyke swarm, Ontario-Quebec: implications for the timing
and record Variscan hydrocarbon migration on extension of Oslo rift. of Iapetus rifting. Canadian Journal of Earth Sciences, 32, 273– 280.
In: Redfield, T., Buiter, S. J. H. & Smethurst, M. A. (eds) geo- Karlstrom, K. E., Bowring, S. A. et al. 2000. Chuar Group of the
dynamics, Geomagnetism and Paleogeography: A 50 Year Celebra- Grand Canyon: record of breakup of Rodinia, associated change in
tion. NGU Report 2007.057 (Trondheim: Geological Survey of the global carbon cycle, and ecosystem expansion by 740 Ma.
Norway), 50. Geology, 28, 619– 622.
Hartmann, L. A., Bitencourt, M. F., Santos, J. O. S., McNaughton, Kaufman, A. J., Sial, A. N., Frimmel, H. E. & Misi, A. 2009. Neoproter-
N. J., Rivera, C. B. & Betiollo, L. 2003. Prolonged Paleoprotero- ozoic to Cambrian palaeoclimatic events in southwestern Gondwana.
zoic magmatic participation in the Neoproterozoic Dom Feliciano In: Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, H. E.
belt, Santa Catarina, Brazil, based on zircon U– Pb SHRIMP geochro- (eds) Neoproterozoic –Cambrian Tectonics, Global Change and
nology. Journal of South American Earth Sciences, 16, 477–492. Evolution. Elsevier, Amsterdam, Developments in Precambrian
Hebert, C. L., Kaufman, A. J., Penniston-Dorland, S. C. & Martin, Geology, 16, 369– 388.
A. J. 2010. Radiometric and stratigraphic constraints on terminal Kawai, T., Windley, B. F., Terabayashi, M., Yamamoto, H., Isozaki,
Ediacaran (post-Gaskiers) glaciation and metazoan evolution. Pre- Y. & Maruyama, S. 2008. Neoproterozoic glaciations in the mid-
cambrian Research, 182, 402–412. oceanic realm: an example from hemi-pelagic mudstones on Llandd-
Hodych, J. P. & Cox, R. A. 2007. Ediacaran U– Pb zircon dates for the wyn Island, Anglesey, U.K. Gondwana Research, 14, 105–114.
Lac Matapédia and Mt. St.-Anselme basalts of the Quebec Appala- Kempf, O., Kellerhals, P., Lowrie, W. & Matter, A. 2000. Paleomag-
chians: support for a long-lived mantle plume during the rifting netic directions in Late Precambrian glaciomarine sediments of the
phase of Iapetus opening. Canadian Journal of Earth Sciences, 44, Mirbat sandstone formation, Oman. Earth and Planetary Science
565– 581. Letters, 175, 181– 190.
Hoffman, P. F. 1999. The break-up of Rodinia, birth of Gondwana, true Kendall, B., Creaser, R. A., Calver, C. R., Raub, T. D. & Evans,
polar wander and the Snowball Earth. Journal of African Earth D. A. D. 2009. Correlation of Sturtian diamictite successions in
Sciences, 28, 17– 33. southern Australia and northwestern Tasmania by Re– Os black
Hoffman, P. F. 2009. Pan-glacial — a third state in the climate system. shale geochronology and the ambiguity of ‘Sturtian’-type diamic-
Geology Today, 25, 107–114. tite– cap carbonate pairs as chronostratigraphic marker horizons.
Hoffman, P. F. & Schrag, D. P. 2002. The Snowball Earth hypothesis: Precambrian Research, 172, 301– 310.
testing the limits of global change. Terra Nova, 14, 129– 155. Kent, D. V. & Dupuis, C. 2003. Paleomagnetic study of the Paleocene–
Hoffman, P. F. & Li, Z.-X. 2009. A palaeogeographic context for Neopro- Eocene Tarawan chalk and Esna shale: dual polarity remagnetizations
terozoic glaciation. Palaeogeography, Palaeoclimatology, Palaeo- of Cenozoic sediments in the Nile valley (Egypt). Micropaleontology,
ecology, 277, 158– 172. 49, 139– 146.
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., Key, R. M., Liyungu, A. K., Njamu, F. M., Somwe, V. & Banda, J.
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635 2001. The western arm of the Lufilian Arc in NW Zambia and its
110 D. A. D. EVANS & T. D. RAUB

potential for copper mineralization. Journal of African Earth South Africa and implications for the age of Neoproterozoic iron for-
Sciences, 33, 503– 528. mations. American Journal of Science, 310, 862– 888.
Khomentovsky, V. V. 1986. The Vendian System of Siberia and a stan- Mac Niocaill, C., Kilner, B., Stouge, S., Knudsen, M. F., Harper,
dard stratigraphic scale. Geological Magazine, 123, 333– 348. D. A. T. & Christiansen, J. L. 2008. The Neoproterozoic drift
Kilner, B., Mac Niocaill, C. & Brasier, M. 2005. Low-latitude glacia- history of Laurentia: a critical evaluation and new palaeomagnetic
tion in the Neoproterozoic of Oman. Geology, 33, 413– 416. data from Northern and Eastern Greenland. EOS Transactions of
Kirschvink, J. L. 1978. The Precambrian-Cambrian boundary problem: the American Geophysical Union, 89, S514–05.
paleomagnetic directions from the Amadeus Basin, Central Australia. Maloof, A. C., Schrag, D. P., Crowley, J. L. & Bowring, S. A. 2005.
Earth and Planetary Science Letters, 40, 91– 100. An expanded record of Early Cambrian carbon cycling from the Anti-
Kirschvink, J. L. 1992. Late Proterozoic low-latitude global glaciation: Atlas Margin, Morocco. Canadian Journal of Earth Sciences, 42,
the Snowball Earth. In: Schopf, J. W. & Klein, C. C. (eds) The Pro- 2195– 2216.
terozoic Biosphere: A Multidisciplinary Study. Cambridge University Martin, M. W., Grazhdankin, D. V., Bowring, S. A., Evans, D. A. D.,
Press, Cambridge, 51 –52. Fedonkin, M. A. & Kirschvink, J. L. 2000. Age of Neoproterozoic
Kirschvink, J. L. & Rozanov, A. Yu. 1984. Magnetostratigraphy of bilaterian body and trace fossils, White Sea, Russia: implications for
Lower Cambrian strata from the Siberian platform: a paleomagnetic metazoan evolution. Science, 288, 841– 845.
pole and a preliminary polarity time-scale. Geological Magazine, Mbede, E. I., Kampunzu, A. B. & Armstrong, R. A. 2004. Neoproter-
121, 189– 203. ozoic inheritance during Cainozoic rifting in the western and south-
Kravchinsky, V. A., Konstantinov, K. M. & Cogné, J.-P. 2001. western branches of the East African rift system: evidence from
Palaeomagnetic study of Vendian and Early Cambrian rocks of carbonatite and alkaline intrusions. Conference abstract, The East
South Siberia and Central Mongolia: was the Siberian platform African Rift System: Development, Evolution and Resources,
assembled at this time? Precambrian Research, 110, 61– 92. Addis Ababa, Ethiopia, June 20– 24, 2004.
Kravchinsky, V. A., Sklyarov, E. V., Gladkochub, D. P. & Harbert, McCausland, P. J. A., Van der Voo, R. & Hall, C. M. 2007.
W. P. 2010. Paleomagnetism of the Precambrian Eastern Sayan rocks: Circum-Iapetus paleogeography of the Precambrian – Cambrian tran-
implications for the Ediacaran –Early Cambrian paleogeography of sition with a new paleomagnetic constraint from Laurentia. Precam-
the Tuva-Mongolian composite terrane. Tectonophysics, 486, 65 – 80. brian Research, 156, 125–152.
Levashova, N. M., Kalugin, V. M., Gibsher, A. S., Yff, J., Ryabinin, McCausland, P. J. A., Smirnov, A., Evans, D. A. D., Izard, C. & Raub,
A. B., Meert, J. G. & Malone, S. J. 2010. The origin of the Baydaric T. 2009. Low-latitude Laurentia at 615 Ma: Paleomagnetism of the
microcontinent, Mongolia: constraints from paleomagnetism and Long Range Dykes and Coeval Lighthouse Cove Formation, Northern
geochronology. Tectonophysics, 485, 306–320. Newfoundland and SE Labrador. Abstracts, AGU-GAC/MAC-CGU
Li, Y., Li, Y., Sharps, R., McWilliams, M. & Gao, Z. 1991. Sinian paleo- Joint Assembly, Toronto, May 2009.
magnetic results from the Tarim block, western China. Precambrian McCausland, P. J. A., Hankard, F., Van der Voo, R. & Hall, C. M.
Research, 49, 61 –71. 2011. Ediacaran paleogeography of Laurentia: paleomagnetism and
40
Li, Z.-X. 2000. New palaeomagnetic results from the ‘cap dolomite’ of the Ar-39Ar geochronology of the 583 Ma Baie des Moutons syenite,
Neoproterozoic Walsh Tillite, northwestern Australia. Precambrian Quebec. Precambrian Research, 187, 58– 78.
Research, 100, 359– 370. McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial
Li, Z.-X. & Evans, D. A. D. 2011. Late Neoproterozoic 408 intraplate trinity: Neoproterzoic Earth history within the British-Irish Caledo-
rotation within Australia allows for a tighter-fitting and longer-lasting nides. Geology, 34, 909–912.
Rodinia. Geology, 39, 39– 42. McFadden, P. L. & Lowes, F. J. 1981. The discrimination of mean direc-
Linnemann, U., Gerdes, A., Drost, K. & Buschmann, B. 2007. The tions drawn from Fisher distributions. Geophysical Journal of the
continuum between Cadomian orogenesis and opening of the Rheic Royal Astronomical Society, 67, 19– 33.
Ocean: constraints from LA– ICP– MS U–Pb zircon dating and McNamara, A. K., Mac Niocaill, C., Van der Pluijm, B. A. & Van der
analysis of plate-tectonic setting (Saxo-Thuringian zone, northeastern Voo, R. 2001. West African proximity of the Avalon terrane in the
Bohemian Massif, Germany. In: Linnemann, U., Nance, R. D., latest Precambrian. Geological Society of America Bulletin, 113,
Kraft, P. & Zulauf, G. (eds) The Evolution of the Rheic 1161– 1170.
Ocean: From Avalonian –Cadomian Active Margin to Alleghe- McWilliams, M. & Schmidt, P. W. 2003. Paleomagnetism of Grassy
nian– Variscan Collision. Geological Society of America, Special Group rocks, King Island, Tasmania. European Geophysical
Paper, 423, 61 –96. Society, Geophysical Research Abstracts, 5, no. 12406.
Lund, K., Aleinikoff, J. N. & Evans, K. V. 2003. SHRIMP U –Pb geo- Meert, J. G. & Van der Voo, R. 1996. Paleomagnetic and 40Ar/39Ar
chronology of Neoproterozoic Windermere Supergroup, central study of the Sinyai Dolerite, Kenya: implications for Gondwana
Idaho: implications for rifting of western Laurentia and synchroneity assembly. Journal of Geology, 104, 131– 142.
of Sturtian glacial deposits. Geological Society of America Bulletin, Meert, J. G. & Van der Voo, R. 2001. Comment on ‘New palaeomag-
115, 349– 372. netic result from Vendian red sediments in Cisbaikalia and the
Macouin, M., Besse, J., Ader, M., Gilder, S., Yang, Z., Sun, Z. & problem of the relationship of Siberia and Laurentia in the
Agrinier, P. 2004. Combined paleomagnetic and isotopic data Vendian’ by S. A. Pisarevsky, R. A. Komissarova and A. N.
from the Doushantuo carbonates, South China: implications for the Khramov. Geophysical Journal International, 146, 867– 870.
“Snowball Earth” hypothesis. Earth and Planetary Science Letters, Meert, J. G., Van der Voo, R. & Payne, T. W. 1994. Paleomagnetism of
224, 387– 398. the Catoctin volcanic province: a new Vendian –Cambrian apparent
Macdonald, F. A., McClelland, W. C., Schrag, D. P. & Macdonald, polar wander path for North America. Journal of Geophysical
W. P. 2009a. Neoproterozoic glaciations on a carbonate platform Research, 99, 4625– 4641.
margin in Arctic Alaska and the origin of the North Slope subterrane. Meert, J. G., Van der Voo, R. & Ayub, S. 1995. Paleomagnetic investi-
Geological Society of America Bulletin, 121, 448–473. gation of the Neoproterozoic Gagwe lavas and Mbozi complex, Tan-
Macdonald, F. A., Jones, D. S. & Schrag, D. P. 2009b. Stratigraphic zania and the assembly of Gondwana. Precambrian Research, 74,
and tectonic implications of a newly discovered glacial 225– 244.
diamictite– cap carbonate couplet in southwestern Mongolia. Meert, J. G., Walderhaug, H. J., Torsvik, T. H. & Hendriks, B. W. H.
Geology, 37, 123–126. 2007. Age and paleomagnetic signature of the Alnø
Macdonald, F. A., Schmitz, M. D. et al. 2010a. Calibrating the Cryo- carbonatite complex (NE Sweden): additional controversy for the
genian. Science, 327, 1241–1243. Neoproterozoic paleoposition of Baltica. Precambrian Research,
Macdonald, F. A., Cohen, P. A., Dudás, F. Ö. & Schrag, D. P. 2010b. 154, 159– 174.
Early Neoproterozoic scale microfossils in the Lower Tindir Group of Meffre, S., Direen, N. G., Crawford, A. J. & Kamenetsky, V. 2004.
Alaska and the Yukon Territory. Geology, 38, 143–146. Mafic volcanic rocks on King Island, Tasmania: evidence for
Macdonald, F. A., Strauss, J. V., Rose, C. V., Dudás, F. Õ. & Schrag, 579 Ma break-up in east Gondwana. Precambrian Research, 135,
D. P. 2010c. Stratigraphy of the Port Nolloth Group of Namibia and 177– 191.
GLOBAL NEOPROTEROZOIC GLACIAL PALAEOLATITUDES 111

Melezhik, V. A., Kuznetsov, A. B., Fallick, A. F., Smith, R. A., Gor- Evolution. Elsevier, Amsterdam, Developments in Precambrian
okhov, I. M., Jamal, D. & Catuane, F. 2006. Depositional environ- Geology, 16, 87 –101.
ments and an apparent age for the Geci meta-limestones: constraints Pokrovskii, B. G., Melezhik, V. A. & Bujakaite, M. I. 2006. Carbon,
on the geological history of northern Mozambique. Precambrian oxygen, strontium, and sulfur isotopic compositions in late Precam-
Research, 148, 19– 31. brian rocks of the Patom Complex, central Siberia: communication
Metelkin, D. V., Belonosov, I. V., Gladkochub, D. P., Donskaya, 1. results, isotope stratigraphy, and dating problems. Lithology and
T. V., Mazukabzov, A. M. & Stanevich, A. M. 2005. Paleomag- Mineral Resources, 41, 450– 474.
netic directions from Nersa intrusions of the Biryusa terrane, Siberian Popov, V., Iosifidi, A., Khramov, A., Tait, J. & Bachtadse, V. 2002.
Craton, as a reflection of tectonic events in the Neoproterozoic. Paleomagnetism of Upper Vendian sediments from the Winter
Russian Geology and Geophysics, 46, 398– 413. Coast, White Sea region, Russia: implications for the paleogeography
Miller, N. R., Alene, M., Sacchi, R., Stern, R. J., Conti, A., Kröner, of Baltica during Neoproterozoic times. Journal of Geophysical
A. & Zuppi, G. 2003. Significance of the Tambien Group (Tigrai, Research, 107, doi: 10.1029/2001JB001607.
N. Ethiopia) for Snowball Earth events in the Arabian– Nubian Popov, V. V., Khramov, A. N. & Bachtadse, V. 2005. Palaeomagnet-
Shield. Precambrian Research, 121, 263– 283. ism, magnetic stratigraphy, and petromagnetism of the Upper
Misi, A., Kaufman, A. J. et al. 2007. Chemostratigraphic correlation of Vendian sedimentary rocks in the sections of the Zolotitsa River
Neoproterozoic successions in South America. Chemical Geology, and in the Verkhotina Hole, Winter Coast of the White Sea, Russia.
237, 143– 167. Russian Journal of Earth Sciences, 7, 1– 29.
Mitchell, R. N., Evans, D. A. D. & Kilian, T. M. 2010. Rapid Early Porter, S. M., Knoll, A. H. & Affaton, P. 2004. Chemostratigraphy of
Cambrian rotation of Gondwana. Geology, 38, 755–758. Neoproterozoic cap carbonates from the Volta Basin, West Africa.
Mukhopadhyay, J. & Chaudhuri, A. K. 2003. Shallow to deep-water Precambrian Research, 130, 99– 112.
deposition in a Cratonic basin: an example from the Proterozoic Pen- Prave, A. R. 1999. The Neoproterozoic Dalradian Supergroup of Scot-
ganga Group, Pranhita– Godavari Valley, India. Journal of Asian land: an alternative hypothesis. Geological Magazine, 136, 609– 617.
Earth Sciences, 21, 613–622. Ramezani, J. & Tucker, R. D. 2003. The Saghand region, central Iran:
Murthy, G. S. 1971. The paleomagnetism of diabase dikes from the U–Pb geochronology, petrogenesis and implications for Gondwana
Grenville Province. Canadian Journal of Earth Sciences, 8, tectonics. American Journal of Science, 303, 622–665.
802– 812. Rapalini, A. E. & Sánchez-Bettucci, L. 2008. Widespread remagneti-
Murthy, G., Gower, C., Tubrett, M. & Pätzold, R. 1992. Paleomag- zation of late Proterozoic sedimentary units of Uruguay and the
netism of Eocambrian Long Range dykes and Double Mer Formaa- apparent polar wander path for the Rio de La Plata craton. Geophysi-
tion from Labrador, Canada. Canadian Journal of Earth Sciences, cal Journal International, 174, 55 –74.
29, 1224– 1234. Raub, T. D. 2008. Prolonged deglaciation of Snowball Earth. PhD thesis,
Myrow, P. M., Pope, M. C., Goodge, J. W., Fischer, W. & Palmer, Yale University.
A. R. 2002. Depositional history of pre-Devonian strata and timing Raub, T. D., Evans, D. A. D. & Smirnov, A. V. 2007a. Siliciclastic
of Ross orogenic tectonism in the central Transantarctic Mountains, prelude to Elatina deglaciation: lithostratigraphy and rock magnetism
Antarctica. Geological Society of America Bulletin, 114, 1070– 1088. of the base of the Ediacaran System. In: Vickers-Rich, P. & Komar-
Nawrocki, J., Boguckij, A. & Katinas, V. 2004. New Late Vendian ower, P. (eds) The Rise and Fall of the Ediacaran Biota. Geological
palaeogeography of Baltica and the TESZ. Geological Quarterly, Society, London, Special Publication, 286, 53– 76.
48, 309– 316. Raub, T. D., Kirschvink, J. L. & Evans, D. A. D. 2007b. True polar
Pazos, P. J., Bettucci, L. S. & Loureiro, J. 2008. The Neoproterozoic wander: linking deep and shallow geodynamics to hydro- and bio-
glacial record in the Rı́o de la Plata Craton: a critical reappraisal. spheric hypotheses. In: Kono, M. (ed.) Treatise on Geophysics
In: Pankhurst, R. J., Truow, R. A. J., Brito Neves, B. B. & de Volume 5: Geomagnetism. Elsevier, Amsterdam, 565–589.
Wit, M. J. (eds) West Gondwana: Pre-Cenozoic Correlations Rieu, R., Allen, P. A., Cozzi, A., Kosler, J. & Bussy, F. 2007. A
Across the South Atlantic Region. Geological Society, London, composite stratigraphy for the Neoproterozoic Huqf Supergroup of
Special Publication, 294, 343– 364. Oman: integrating new litho-, chemo- and chronostratigraphic data
Pazos, P. J., Sánchez-Bettucci, L. & Tofalo, O. R. 2003. The record of the Mirbat area, southern Oman. Journal of the Geological
of the Varanger glaciation at the Rı́o de la Plata craton, Vendian- Society of London, 164, 997– 1009.
Cambrian of Uruguay. Gondwana Research, 6, 65 –77. Rieu, R., Allen, P. A., Etienne, J. L., Cozzi, A. & Wiechert, U.
Pecoits, E., Gingras, M., Aubet, N. & Konhauser, K. 2008. Ediacaran 2006. A Neoproterozoic glacially influenced basin margin succession
in Uruguay: palaeoclimatic and palaeobiological implications. Sedi- and ‘atypical’ cap carbonate associated with bedrock palaeovalleys,
mentology, 55, 689– 719. Mirbat area, southern Oman. Basin Research, 18, 471–496.
Pinna, P., Jourde, G., Calvez, J. Y., Mroz, J. P. & Marques, J. M. Roberts, D. 2003. The Scandinavian Caledonides: event chronology,
1993. The Mozambique Belt in northern Mozambique: Neoprotero- palaeogeographic settings and likely modern analogues. Tectonophy-
zoic (1100 –850 Ma) crustal growth and tectogenesis, and superim- sics, 365, 283–299.
posed Pan-African (800– 550 Ma) tectonism. Precambrian Rodrigues, J. B., Pimentel, M. M., Buhn, B., Dardenne, M. A. &
Research, 62, 1 – 59. Alvarenga, C. J. S. 2008. Provenance of Vazante Group — Prelimi-
Pisarevsky, S. A., Komissarova, R. A. & Khramov, A. N. 2001. Reply nary data. VI South American Symposium on Isotope Geology, San
to comment by J. G. Meert and R. Van der Voo on ‘New palaeomag- Carlos de Bariloche, Argentina, 4.
netic result from Vendian red sediments in Cisbaikalia and the Rooney, A. D., Selby, D., Houzay, J.-P. & Renne, P. R. 2010. Re– Os
problem of the relationship of Siberia and Laurentia in the geochronology of a Mesoproterozoic sedimentary succession, Taou-
Vendian’. Geophysical Journal International, 146, 871– 873. deni basin, Mauritania: implications for basin-wide correlations and
Pisarevsky, S. A., Wingate, M. T. D., Stevens, M. K. & Haines, Re– Os organic-rich sediments systematics. Earth and Planetary
P. W. 2007. Palaeomagnetic results from the Lancer 1 stratigraphic Science Letters, 289, 486–496.
drillhole, Officer Basin, Western Australia, and implications for Rose, C. V. & Maloof, A. C. 2010. Testing models for post-glacial ‘cap
Rodinia reconstructions. Australian Journal of Earth Sciences, 54, dolostone’ deposition: Nuccaleena Formation, South Australia. Earth
561– 572. and Planetary Science Letters, 296, 165–180.
Poidevin, J.-L. 2007. Stratigraphie isotopique du strontium et datation Rowan, C. J. & Tait, J. 2010. Oman’s Low Latitude ‘Snowball Earth’
des formations carbonatées et glaciogéniques néoprotérozoiques du Pole Revisited: Late Cretaceous remagnetisation of Late Neoproter-
Nord et de l’Ouest du craton du Congo. Comptes Rendus Geoscience, ozoic Carbonates in Northern Oman. American Geophysical Union,
339, 259– 273. Fall Meeting Abstract GP33C-0959.
Poiré, D. G. & Gaucher, C. 2009. Lithostratigraphy. Neoproterozoic – Sánchez-Bettucci, L. & Rapalini, A. E. 2002. Paleomagnetism of the
Cambrian evolution of the Rı́o de la Plata Palaeocontinent. In: Sierra de Las Animas Complex, southern Uruguay: its implications in
Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, H. E. the assembly of western Gondwana. Precambrian Research, 118,
(eds) Neoproterozoic– Cambrian Tectonics, global Change and 243– 265.
112 D. A. D. EVANS & T. D. RAUB

Schmidt, P. W. & Williams, G. E. 1995. The Neoproterozoic climatic Torsvik, T. H., Lohmann, K. C. & Sturt, B. A. 1995. Vendian glacia-
paradox: equatorial palaeolatitude for Marinoan glaciation near sea tions and their relation to the dispersal of Rodinia: Paleomagnetic
level in South Australia. Earth and Planetary Science Letters, 134, constraints. Geology, 23, 727–730.
107– 124. Trindade, R. I. F. & Macouin, M. 2007. Palaeolatitude of glacial depos-
Schmidt, P. W., Williams, G. E. & Embleton, B. J. J. 1991. Low palaeo- its and palaeogeography of Neoproterozoic ice ages. Comptes Rendus
latitude of Late Proterozoic glaciation: early timing of remanence in Geoscience, 339, 200–211.
haematite of the Elatina Formation, South Australia. Earth and Pla- Trindade, R. I. F., D’Agrella-Filho, M. S., Babinski, M., Font, E. &
netary Science Letters, 105, 355–367. Brito Neves, B. B. 2004. Paleomagnetism and geochronology of the
Schmidt, P. W., Williams, G. E. & McWilliams, M. O. 2009. Palaeo- Bebedouro cap carbonate: evidence for continental-scale Cambrian
magnetism and magnetic anisotropy of late Neoproterozoic strata, remagnetization in the São Francisco craton, Brazil. Precambrian
South Australia: implications for the palaeolatitude of late Cryogen- Research, 128, 83– 103.
ian glaciation, cap carbonate and the Ediacaran System. Precambrian Trindade, R. I. F., Font, E., D’Agrella-Filho, M. S., Nogueira,
Research, 174, 35 – 52. A. C. R. & Riccomini, C. 2003. Low-latitude and multiple geomag-
Shen, B., Xiao, S., Zhou, C., Kaufman, A. J. & Yuan, X. 2010. Carbon netic reversals in the Neoproterozoic Puga cap carbonate, Amazon
and sulfur isotope chemostratigraphy of the Neoproterozoic Quanji craton. Terra Nova, 15, 441–446.
Group of the Chaidam Basin, NW China: basin stratification in the Trompette, R. 1994. Geology of Western Gondwana (2000– 500 Ma):
aftermath of an Ediacaran glaciation postdating the Shuram event? Pan-African –Brasiliano Aggregation of South America and Africa.
Precambrian Research, 177, 241– 252. Balkema, Rotterdam.
Shields, G. A., Deynoux, M., Culver, S. J., Brasier, M. D., Affaton, Van der Voo, R. 1990. The reliability of paleomagnetic data. Tectonophy-
P. & Vandamme, D. 2007. Neoproterozoic glaciomarine and cap sics, 184, 1– 9.
dolostone facies of the southwestern Taoudéni Basin (Walidiala Vizan, H., Carney, J. N., Turner, P., Ixer, R. A., Tomasso, M.,
Valley, Senegal/Guinea, NW Africa). Comptes Rendus Geoscience, Mullen, R. P. & Clarke, P. 2003. Late Neoproterozoic to Early
339, 186– 199. Palaeozoic palaeogeography of Avalonia: some palaeomagnetic con-
Sial, A. N., Dardenne, M. A. et al. 2009. The São Francisco palaeocon- straints from Nuneaton, central England. Geological Magazine, 140,
tinent. In: Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, 685– 705.
H. E. (eds) Neoproterozoic– Cambrian Tectonics, Global Change Weil, A. B., Geissman, J. W. & Van der Voo, R. 2004. Paleomagnetism
and Evolution. Elsevier, Amsterdam, Developments in Precambrian of the Neoproterozoic Chuar Group, Grand Canyon Supergroup,
Geology, 16, 31 –69. Arizona: implications for Laurentia’s Neoproterozoic APWP and
Sohl, L. E., Christie-Blick, N. & Kent, D. V. 1999. Paleomagnetic Rodinia break-up. Precambrian Research, 129, 71 –92.
polarity reversals in Marinoan (ca. 600 Ma) glacial deposits of Aus- Williams, G. E. 1993. History of the Earth’s obliquity. Earth-Science
tralia: implications for the duration of low-latitude glaciation in Neo- Reviews, 34, 1– 45.
proterozoic time. Geological Society of America Bulletin, 111, Williams, G. E. 2008. Proterozoic (pre-Ediacaran) glaciation and the high
1120– 1139. obliquity, low-latitude ice, strong seasonality (HOLIST) hypothesis:
Sovetov, J. K. 2002. Vendian foreland basin of the Siberian cratonic principles and tests. Earth-Science Reviews, 87, 61 –93.
margin: Paleopangean accretionary phases. Russian Journal of Williams, G. E., Gostin, V. A., McKirdy, D. M. & Preiss, W. V. 2008.
Earth Sciences, 4, 363– 387. The Elatina glaciation, late Cryogenian (Marinoan Epoch), South
Sovetov, J. 2008. Marinoan Glaciation in the Siberian Craton: Locality, Australia: sedimentary facies and palaeoenvironments. Precambrian
Erosional Forms, Deposits and Constraints to Age. 33rd International Research, 163, 307– 331.
Geological Congress, Abstracts, Session CGC-04. Xu, B., Jian, P., Zheng, H., Zou, H., Zhang, L. & Liu, D. 2005. U–Pb
Sovetov, Yu. K. & Komlev, D. A. 2005. Tillites at the base of the Oselok zircon geochronology and geochemistry of Neoproterozoic volcanic
Group, foothills of the Sayan Mountains, and the Vendian lower rocks in the Tarim Block of northwest China: implications for the
boundary in the southwestern Siberian Platform. Stratigraphy and breakup of Rodinia supercontinent and Neoproterozoic glaciations.
Geological Correlation, 13, 337– 366. Precambrian Research, 136, 107– 123.
Stern, R. J., Avigad, D., Miller, N. R. & Beyth, M. 2006. Evidence for Xu, B., Xiao, S. et al. 2009. SHRIMP zircon U– Pb age constraints on
the Snowball Earth hypothesis in the Arabian – Nubian Shield and the Neoproterozoic Quruqtagh diamictites in NW China. Precambrian
East African Orogen. Journal of African Earth Sciences, 44, 1– 20. Research, 168, 247– 258.
Swanson-Hysell, N. L., Maloof, A. C., Weiss, B. P. & & Evans, Zhan, S., Chen, Y., Xu, B., Wang, B. & Faure, M. 2007. Late Neopro-
D. A. D. 2009. No asymmetry in geomagnetic reversals recorded terozoic paleomagnetic results from the Sugetbrak Formation of the
by 1.1-billion-year-old Keweenawan basalts. Nature Geoscience, 2, Aksu area, Tarim basin (NW China) and their implications to paleo-
713– 717. geographic reconstructions and the Snowball Earth hypothesis. Pre-
Tauxe, L., Kodama, K. P. & Kent, D. V. 2008. Testing corrections for cambrian Research, 154, 143–158.
paleomagnetic inclination error in sedimentary rocks: a comparative Zhang, Q.-R. & Piper, J. D. A. 1997. Palaeomagnetic study of Neoproter-
approach. Physics of the Earth and Planetary Interiors, 169, ozoic glacial rocks of the Yangzi Block: palaeolatitude and configur-
152– 165. ation of South China in the late Proterozoic Supercontinent.
Thomas, R. J., Chevallier, L. C. et al. 2002. Precambrian evolution of Precambrian Research, 85, 173– 199.
the Sirwa Window, Anti-Atlas Orogen, Morocco. Precambrian Zhang, Q.-R., Li, X.-H., Feng, L.-J., Huang, J. & Song, B. 2008. A
Research, 118, 1 –57. new age constraint on the onset of the Neoproterozoic glaciations
Thomas, R. J., Fekkak, A. et al. 2004. A new lithostratigraphic frame- in the Yangtze Platform, South China. Journal of Geology, 116,
work for the Anti-Atlas Orogen, Morocco. Journal of African Earth 423– 429.
Sciences, 39, 217– 226. Zhang, Q.-R., Chu, X. L. & Feng, L. J. 2009. Discussion on the Neopro-
Thompson, M. D. & Bowring, S. A. 2000. Age of the Squantum ‘Tillite’, terozoic glaciations in the South China Block and their related paleo-
Boston Basin, Massachusetts: U– Pb zircon constraints on terminal latitudes. Chinese Science Bulletin, 54, 1797– 1800.
Neoproterozoic glaciation. American Journal of Science, 300, Zhang, S., Evans, D. A. D. et al. 2006. New paleomagnetic results from
630– 655. the Nantuo Formation in south China and their paleogeographic
Thompson, M. D., Grunow, A. M. & Ramezani, J. 2007. Late Neopro- implications. EOS, Transactions of the American Geophysical
terozoic paleogeography of the southeastern New England Avalon Union, abstract GP34A-04.
zone: insights from U– Pb geochronology and paleomagnetism. Geo- Zhang, S., Jiang, G. & Han, Y. 2008. The age of the Nantuo
logical Society of America Bulletin, 119, 681–696. Formation and Nantuo glaciation in South China. Terra Nova, 20,
Tohver, E., D’Agrella-Filho, M. S. & Trindade, R. I. F. 2006. Paleo- 289– 294.
magnetic record of Africa and South America for the 1200–500 Ma Zhou, C., Tucker, R., Xiao, S., Peng, Z., Yuan, X. & Chen, Z. 2004.
interval, and evaluation of Rodinia and Gondwana assemblies. Pre- New constraints on the ages of Neoproterozoic glaciations in south
cambrian Research, 147, 193–222. China. Geology, 32, 437– 440.
Chapter 8

Biostratigraphy and stratigraphic subdivision of Cryogenian successions


of Australia in a global context

KATHLEEN GREY1*, ANDREW C. HILL2 & CLIVE CALVER3


1
Geological Survey of Western Australia, Department of Mines and Petroleum, 100 Plain Street, Western Australia, 6076
2
Centro de Astrobiologı́a (INTA-CSIC), Instituto Nacional de Técnica Aeroespacial, Ctra de Ajalvir, km 4,
28850 Torrejón de Ardoz, Madrid, Spain
3
Mineral Resources Tasmania, PO Box 56, Rosny Park, Tasmania 7018, Australia
*Corresponding author (e-mail: kath.grey@dmp.wa.gov.au)

Abstract: Cryogenian correlation in Australia is based on an extensive data set from the Centralian Superbasin and Adelaide Rift
Complex and integrates biostratigraphy and isotope chemostratigraphy to provide a three-dimensional interpretation based on outcrop
and drill holes. Studies are ongoing, but newer data are consistent with the distributions discussed here. From the chemostratigraphic
and biostratigraphic viewpoint, the first appearance of the acritarch Cerebrosphaera buickii, coupled with a large negative isotope excur-
sion at c. 800 Ma, supported by the first appearance of the stromatolite Baicalia burra, seems to have potential for boundary placement. It
is widely recognized across Australia and seems to have potential globally.

Australia has extensive Cryogenian successions that can be readily succession is probably no older than c. 830 Ma (Fig. 8.5), based
correlated using a variety of techniques. Cryogenian correlation is on U –Pb ages for the Gairdner Dyke Swarm, which either
based on both field sections and continuous core from drill holes. underlies or just intrudes the base of the succession in South
In Western Australia, this includes drill holes Empress 1 and 1A, Australia (see below), and global chemostratigraphic correlation
Lancer 1, Vines 1, other drill holes, and outcrops in the western (Macdonald et al. 2010). Early Cryogenian successions are wide-
Officer and Amadeus Basins and the Kimberley area. Results spread and have been identified in the Adelaide Rift Complex
from Western Australia are consistent with data from other drill (Callanna and Burra Groups), the Amadeus Basin (Bitter Springs
holes and field sections examined from across Australia, including Formation), the Georgina Basin (Yackah beds), the Ngalia Basin
those from the eastern Officer, Amadeus, Georgina, and Ngalia (Vaughan Springs Quartzite and Albinia Formation), the Officer
Basins, the Adelaide Rift Complex and Tasmania (Figs 8.1– Basin (Buldya Group and lateral equivalents in both Western
8.4). In general, biostratigraphic correlations (Figs 8.1 –8.3) corre- and South Australia, the Kimberley area (Ruby Plains Group),
spond to those obtained using lithostratigraphy, well-log and other parts of the Birrindudu Basin (Redcliff Pound Group), and
geophysical data and chemostratigraphy (d13C and 87Sr/86Sr; Tasmania (lower Togari Group and its equivalents). This part of
Walter et al. 1995, 2000; Morton & Drexel 1997; Hill & Walter the succession underlies the Sturt glaciation (Preiss et al. 2011),
2000; Hill et al. 2000a, b; Preiss 2000; Haines et al. 2004, 2008; named after the Sturt Tillite in the Adelaide Rift Complex), and
Grey et al. 2005; Mory & Haines 2005; Gorter et al. 2007; Hill, which includes the Areyonga Formation in the Amadeus Basin
unpublished data). and lateral equivalents in central Australia and Tasmania, but is
The Cryogenian was originally defined as a Neoproterozoic represented by a hiatus in Western Australia, except in Vines 1,
chronometric period extending from 850 Ma to 650 Ma. Since where evidence of two glaciations, represented by the older
the name Cryogenian is preoccupied, it is questionable whether Pirrilyungka Formation and the younger Wahlgu Formation, may
it should be used for any proposed chronostratigraphic units be present (Haines et al. 2008; Hill et al. 2011).
below the Ediacaran. However, for practical purposes, the term The interval between the Sturt Tillite and Elatina Formation and
Cryogenian is here used for the interval from the chronometrically their equivalents is not well represented across Australia. There is
defined base of the Cryogenian System and Period at 850 Ma to the a near-continuous succession in the Adelaide Rift Complex (the
chronometrically defined top of the Cryogenian at 650 Ma Umberatana Group) and a partial succession in the Amadeus
(Gradstein et al. 2004), but also incorporates the interval that lies Basin (Aralka Formation). In Western Australia, this interval is
between the 650 Ma boundary and the chronostratigraphically marked by a heavily karstified surface and disconformity in both
defined base of the succeeding Ediacaran System and Period with Empress 1A and Lancer 1. The succeeding Elatina glaciation
its Global Boundary Stratotype Section and Point (GSSP) at the (Williams et al. 2008) and its equivalents (previously referred
base of the Nuccaleena Formation, Flinders Ranges, South Australia to as the Marinoan glaciation) appear to be represented in all
(Knoll et al. 2006). This pragmatic approach facilitates discussion Australian Neoproterozoic basins.
pending decisions of the International Subcommission on Neopro- In Western Australia, the continuously cored Empress 1A and
terozoic (Cryogenian and Ediacaran) Stratigraphy. Only Australian Lancer 1 successions can be tied to each other and to other
Cryogenian successions are discussed here in detail because it is Officer Basin drill holes using well-log data, lithostratigraphy,
still difficult to correlate between Australia and successions else- seismic interpretation and chemostratigraphy (Grey et al. 2005).
where with any certainty. However, consistencies observed Acritarch and stromatolite ranges are consistent across the basin
across Australia auger well for eventual global correlation. and proposed biostratigraphic correlations correspond closely to
There are no unequivocal Tonian-age (.850 Ma) sedimentary non-biostratigraphic correlations. There is evidence to suggest
rocks in Australia, although several possible sections require that Australian successions can be similarly correlated to succes-
further investigation. The base of the recognized Cryogenian sions elsewhere in the world.

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 113– 134. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.8
114 K. GREY ET AL.

120° 130° 140° 150°

10° 10°

KIMBERLEY
REGION

Fig. 8.1. Locality map showing distribution


GEORGINA of Neoproterozoic basins in Australia and
20° NGALIA BASIN 20° localities sampled for microfossil and
AMADEUS BASIN
BASIN AS1 palynology studies in the Australian
Arunta Province
LD1
D1 O1
Cryogenian (adapted from Walter et al.
MW1
T6
H1 W1
Bb1 1995; Preiss 2000; Grey 2005b).
T9 MC1
Western
L1 Y2
Y3 Distribution data from Grey & Cotter
Musgrave Province
Officer Basin Lu1
K1A PEAKE & DENISON (1996); Cotter 1997, 1999; Grey 1999a,
Probable extent of V1 RANGES
Centralian Superbasin E1A
OFFICER B1 Curnamona 2005a, b; Grey et al. 2005b. Abbreviations:
BASIN Eastern G1 Province
Officer AS1, BMR Alice Springs 1; Bb1, Bluebush
Basin Gawler
Craton B4 1; B1, Birksgate 1; B4, Western Mining
STUART
SHELF
BLD 04; D1, Eagle Dragoon 1; E1A,
30° 30°
GSWA Empress 1/1A; G1, Comalco Giles
P12 ADELAIDE
RIFT COMPLEX 1; H1, Eagle Corporation Hussar 1; K1A,
Shell Kanpa 1A; L1, GSWA Lancer 1; LD1,
Lake Disappointment 1 (LDDH 1); Lu1,
Shell Lungkarta 1; MC1, Transoil Mount
Drillhole with Bitter Springs Formation Assemblage
Charlotte 1; MW1, Mount Winter 1; O1,
Drillhole with Cerebrosphaera buickii Assemblage Magellan Ooraminna 1; PP12; North
Drillhole with both assemblages
Outcrop with Bitter Springs Formation Assemblage KING ISLAND
Broken Hill PP 12; T6, GSWA TWB 6
(Trainor Water Bore 6); T9, GSWA TWB 9
1000 km
40°
TASMANIA
40° (Trainor Water Bore 9); V1, GSWA Vines
1; W1, Wallara 1; Y2, Shell Yowalga 2; Y3,
110° 120° 130° 140° 150° 160°
Shell Yowalga 3.

120° 130° 140° 150°

10° 10°

KIMBERLEY
REGION
R

GEORGINA

20° NGALIA BASIN 20°


AMADEUS BASIN
T BASIN BS
T Arunta Province BS P

T BS
S LB Y3 BS W1
LB
Western L1
LB Musgrave Province M5
Officer Basin LB PEAKE & DENISON
Probable extent of LB BG BG RANGES
E1A
Centralian Superbasin OFFICER Curnamona
BASIN Eastern BG Province Fig. 8.2. Locality map showing distribution of
Officer
Basin Gawler the Acaciella australica stromatolite
Craton
STUART assemblage in Australia. Only localities with
SHELF 30° identified specimens are shown (data from
30° BG
ADELAIDE Walter 1972; Preiss 1973b, 1976, 1987; Walter
RIFT COMPLEX
et al. 1979; Grey 1995, 1999b, 2005a,
unpublished data; Stevens & Grey 1997; Grey
& Blake 1999; Hill et al. 2000a; Grey et al.
W1
Drillhole Outcrop
2005). Abbreviations: T, Tarcunyah Group;
Stratigraphy
T Tarcunyah Group BS Bitter Springs Formation LB, Lower Buldya Group; R, Ruby Plains
LB Lower Buldya Group P Plenty Group
R Ruby Plains Group BG Burra Group Group; S, Sunbeam Group; BS, Bitter Springs
S Sunbeam Group KING ISLAND
Formation; P, Plenty Group; BG, Burra Group;
1000 km
40° 40° E1A, GSWA Empress 1A; L1, GSWA Lancer
TASMANIA
1; M5, Manya 5; W1, Wallara 1; Y3, Shell
110° 120° 130° 140° 150° 160°
Yowalga 3.
AUSTRALIAN CRYOGENIAN CORRELATIONS 115

120° 130° 140° 150°

10° 10°

KIMBERLEY
REGION

GEORGINA

20° NGALIA BASIN 20°


AMADEUS BASIN
BASIN
Arunta Province

UB L1 J
H1 W1
UB
Western UB Musgrave Province
Officer Basin UB UB PEAKE & DENISON
Probable extent of E1A RANGES
UB BG
Centralian Superbasin OFFICER Curnamona
N1 BASIN Eastern Province
Officer
Basin Gawler BG
Craton
BG
STUART
SHELF
Fig. 8.3. Locality map showing distribution
30° BG 30°
of the Baicalia burra stromatolite
BG BG ADELAIDE
RIFT COMPLEX assemblage in Australia. Only localities with
BG
identified specimens are shown (data from
Preiss 1972, 1974, 1976, 1985, 1987; Griffin
& Preiss 1976; Stevens & Grey 1997; Grey
Drillhole Outcrop
1999b, 2005a, unpublished data; Hill et al.
W1
Stratigraphy 2000a). Abbreviations: UB, Upper Buldyah
UB Upper Buldya Group BG Burra Group
J Johnnys Creek Beds Tg Togari Group KING ISLAND
Group; J, Johnnys Creek beds; BG, Burra
Group; Tg, Togari Group; E1A, GSWA
1000 km Tg
40°
TASMANIA
40° Empress 1A; H1, Eagle Corporation Hussar
1; L1, GSWA Lancer 1; N1, NJD 1; W1,
110° 120° 130° 140° 150° 160°
Wallara 1.

120° 130° 140° 150°

10° 10°

Fig. 8.4. Locality map showing


distribution of Neoproterozoic basins in
KIMBERLEY Australia and localities sampled for
REGION
chemostratigraphic studies in the
Australian Cryogenian (adapted from
GEORGINA Walter et al. 1995; Hill & Walter 2000;
20° 20° Preiss 2000; McKirdy et al. 2001; Grey
NGALIA BASIN
AMADEUS BASIN
BASIN
2005b). Drill hole abbreviations for
Arunta Province A1 chemostratigraphy: A1, AS 27; A2, AS
I LS
W1
D A2 A3 28; A3, AS 3; B1, BTD 1; B2, Blinman
L B
MC 2; E, Empress 1A; ET, ETM 5A-1; F,
Western
Officer Basin
E Musgrave Province
PEAKE & DENISON Forest 1; I, Illogwa Creek 6; L, Lancer
Probable extent of M5 C
RANGES 1; LN, LNM 10-1; M5, Manya 5; MC,
Centralian Superbasin OFFICER Curnamona
BASIN Eastern
UBPD
Province Mount Charlotte 1; S1, SAS 1; S3, SAR
Officer SR
Basin Gawler
DW
3; SC, SCYW 1a; SL, SLT 104; SR, SR
Craton B2
STUART
SC
S3 M
O LN 6; W1, Wallara 1, W2, Wokurna 2; W5,
ET
30°
SHELF SL EC
B1 30° Wokurna 5. Field section abbreviations
S1 DC
SD ADELAIDE for chemostratigraphy: A, Arthur
W2 UBMF RIFT COMPLEX River; B, Bluebush; C, Coominaree
W5
Dolomite type section; D, Dump; DC,
Depot Creek; DW, Doodney’s Well
Hills; EC, Enorama Creek; LS, Limbla
Syncline; M, Mallee Water;
Drillhole Outcrop
O, Oraparinna Diapir; SD, Skillogalee
KING ISLAND Dolomite in Port Germein Gorge;
F
1000 km A UBMF, Upper Burra Group,
40°
TASMANIA
40° mid-Flinders Ranges; UBPD, Upper
Burra Group, Peake and Denison
110° 120° 130° 140° 150° 160°
Ranges.
116 K. GREY ET AL.

ADELAIDE NW
KING KIMBERLEY
PER- AGE AMADEUS NGALIA GEORGINA RIFT TASMANIA REGION
ISLAND
IOD (Ma) SS NW Central Eastern Northeastern BASIN BASIN COMPLEX
C Cambrian Cambrian Cambrian Cambrian Cambrian Cambrian Cambrian
543
550 McFadden Lungkarta ‘lower’ Arumbera ‘lower’ Yuendumu
4 Fm Fm Sandstone Sandstone 12

LOUISA
Julie
Formation
EDIACARAN

volcanics
575 volcanics
CHD
MOPUNGA rhyodacite 11
UNGOOLYA GROUP WILPENA
Keppel

GRASSY GROUP
GROUP GROUP
Creek
600
3 Pertatataka
Newhaven Fm
Shale
Formation
Member

GROUP WDM W
625 Boondawari
Y

Fm W ? ? Pioneer/Olympic MDDM Black/Oorabra


UGE 9 10
650 2
6 7 8
Pirrilyungka 5
Fm Y illite S illite Member
675

700

T
Steptoe
Fm
‘CRYOGENIAN’

725 4
Kanpa
Fm BURRA
’ GROUP
750
BULDYA GROUP

1
?

775 Hussar
Fm

3 ? ?
800
2
CALANNA GROUP

Member Curdimurka
LM
Subgroup
Albinia
Yackah
WM CVE Fm
825 beds
Loves

SUNBEAM Units ROCKY CAPE


CDE GROUP
GROUP Arkaroola
Alinya Gillen Subgroup
Fm YCE Member
TQ V
850
1 1 BASEMENT 1

Key: Neoproterozoic glacial or Sequence boundary


partially glacial units
1 Geochronological age
Cap dolomite Volcanic
intrusion
Neoproterozoic sediments
Volcanics
Unconformity/disconformity

Fig. 8.5. Australian Neoproterozoic stratigraphy and correlations assuming a c. 635 Ma age (Hoffmann et al. 2004; Condon et al. 2005) for the Elatina Formation and
equivalents; however, see Williams et al. (2008) for an extended discussion of age constraints (adapted from Preiss 1987, 2000; Walter et al. 1995; Morton 1997; Grey
et al. 2005; Dunster et al. 2007). The length of time breaks at major unconformities and sequence boundaries are estimates only. Amadeus Basin successions were
described in Wells et al. (1970), Kennard et al. (1986), Korsch & Kennard (1991) and Lindsay (1993). Stratigraphic subdivision of the Adelaide Rift Complex and Stuart
Shelf follow Preiss (1987, 2000 and references therein), Haines (1988, 1990) and Reid & Preiss (1999). Officer Basin successions and correlations were detailed in
Jackson & van de Graaff (1981), Phillips et al. (1985), Townson (1985), Williams (1992), Preiss (1993), Walter & Gorter (1994), Walter et al. (1994), Bagas et al. (1995,
1999), Lindsay (1995), Lindsay & Reine (1995), Morton (1997), Hill et al. (2000a), Grey et al. (2005) and Haines et al. (2008). Georgina Basin successions were
described by Smith (1972), Walter (1980) and Dunster et al. (2007), the Ngalia Basin successions in Wells & Moss (1983), the Kimberley region in Dow & Gemuts
(1969), Blake et al. (1998), Corkeron & George (2001) and Corkeron (2007), and in Tasmania by Calver (1998), Calver & Walter (2000) and Calver et al. (2004).
Australia-wide correlations were discussed in Preiss et al. (1978), Coats & Preiss (1980), Walter et al. (1995, 2000), Calver & Walter (2000), Hill & Walter (2000), Hill
et al. (2000a, b), Calver et al. (2004) and Grey et al. (2005). SS, Supersequence number, Centralian Superbasin (Walter et al. 1995); CDE, Coominaree Dolomite
equivalent; CHD, Croles Hill Diamictite; CVE, Cadlareena Volcanics equivalent; LM, Lancer Member; MDDM, Mount Davenport Diamictite Member; TQ, Townsend
AUSTRALIAN CRYOGENIAN CORRELATIONS 117

Geochronological constraints (LA –ICPMS) on a porphyry that intrudes the Skillogalee Dolo-
mite and which is now thought to be ‘penecontemporaneous
Tonian (1000 –850 Ma) successions are probably not widespread with Skillogalee-age deposition’ (Drexel 2009). This conflicts
in Australia, although many units are inadequately dated. Units with an age on an underlying rhyolite from the Boucaut Volcanics
such as the Pindyin Sandstone and Alinya Formation in the at the base of the Burra Group cited as giving a U –Pb SHRIMP
eastern Officer Basin, and their presumed lateral equivalents, the zircon age of 777 + 7 Ma by Preiss (2000); however, the
Townsend Quartzite and Lefroy Formation, in the western Boucaut Volcanics date was recently considered unreliable
Officer Basin, the Kulail Sandstone (Close et al. 2005) and poss- by Fanning & Link (2008). A U –Pb zircon date on a thin volcani-
ibly the Dean Quartzite in the Bloods Range region of the clastic layer in the lower Wilyerpa Formation in the Adelaide
Amadeus Basin, the Heavitree Quartzite and lateral equivalents Rift Complex (deposited during the waning phase of the Sturt
in the Georgina and Amadeus Basin, and parts of the Callanna glaciation; Preiss et al. 2011), and previously referred to as
Group in the Adelaide Rift Complex could be either Tonian or being within the Merinjina Tillite, gave an age of 659 + 6 Ma
Cryogenian, depending on whether there is a stratigraphic break (Fanning & Link 2008), and indicates an age of c. 660 Ma for
between them and the overlying Bitter Springs Formation and its the Sturt glaciation. This is consistent with Re –Os ages from
lateral equivalents and whether they pre-date or were intruded black shales that closely succeed the Sturt glacials and their
by the Gairdner Dyke Swarm. The Alinya Formation in Giles 1 equivalents: 657.2 + 5.4 Ma in the Aralka Formation in the
(eastern Officer Basin), which seems to have an anomalous acri- Amadeus Basin, 643 + 2.4 Ma on the Tindelpina Shale Member
tarch assemblage (Zang 1995; see below), and the Lamil and in the basal Tapley Hill Formation in the Adelaide Rift Complex
Throssell Range Groups along the northeastern Pilbara margins (Kendall et al. 2006) and 641 + 5 Ma on a black shale above the
could be either younger than (Bagas et al. 2002) or older than Julius River Member in Tasmania (Kendall et al. 2007). A detrital
the Cryogenian Tarcunyah Group. In Tasmania, the Rocky Cape zircon U –Pb SHRIMP age of 725 + 11 Ma from the middle
Group (c. 1000–750 Ma) is probably at least in part of Tonian Kanpa Formation, .200 m below the Steptoe –Wahlgu Formation
age (Black et al. 2004). disconformity (Nelson 2002), indicates that the Western Austra-
The mafic Gairdner Dyke Swarm, which intrudes the Musgrave lian hiatus equivalent to the Sturt glaciation is also younger than
Province, Gawler Craton and Stuart Shelf (Parker et al. 1987; 725 Ma. There is no firm evidence in Australia for the Sturt glacia-
Preiss 1987; Cowley & Flint 1993) and cross-cuts underlying tion extending back much beyond 660 Ma. This is in contrast with
Mesoproterozoic successions, and its northwesterly extension, recent dating from NW Canada, where ages of 716.5 Ma were
the Amata Suite, probably relate to extension during the early tec- obtained for glacial units believed to be equivalents of the Sturt
tonic evolution of the Adelaide Rift Complex and the Centralian glaciation (Kendall et al. 2009; Macdonald et al. 2010).
Superbasin. It has not been determined whether the dykes underlie The age of the younger (Elatina) glaciation (Williams et al.
or intrude basal successions in these areas, but it is probable that 2008), and hence the base of the Ediacaran Period, remains uncer-
they slightly post-date the onset of deposition. The dykes are prob- tain. The Ediacaran boundary lies at the base of the Nuccaleena
able feeders for mafic lavas of the poorly dated Wooltana Volca- Formation, so it would be sensible for the underlying Elatina dia-
nics near the top of the Callanna Group and other volcanic rocks, mictite and lateral equivalents to be included with units of Cryo-
such as the Beda Volcanics and those in the Bitter Springs genian age, even if this part of the succession is younger than
Formation (Zhao et al. 1994); however, more precise dating of the previous upper chronometric boundary of 650 Ma (Gradstein
these formations is required before concluding they are all the et al. 2004).
same age. A Stuart Shelf dyke gave a SHRIMP U – Pb age of The Elatina Formation is currently correlated with the Ghaub
827 + 6 Ma on baddeleyite, and the Little Broken Hill Gabbro in Formation of Namibia, dated at 635.5 + 1.2 Ma (Hoffmann
the Willyama Inlier gave an age of 827 + 9 Ma (Wingate et al. et al. 2004) and the Nantuo Tillite of China, dated by an ash bed
1998). Sm –Nd isochrons of 867 + 47 and 802 + 35 Ma (Zhao within the cap dolostone above the tillite at 635.2 + 0.6 Ma
& McCulloch 1993; Zhao et al. 1994) were obtained from a dyke (Condon et al. 2005) because of the presence of similar cap carbon-
intruding the Stuart Shelf (Preiss 2000). Another of the dykes, ates. Lithologically, a succession on King Island (the Cottons
which cross-cuts Mesoproterozoic units in the Western Musgrave Breccia and overlying Cumberland Creek Dolostone) closely
Complex, gave a zircon 207Pb/206Pb age of 824 + 4 Ma resembles the Elatina and Nuccaleena Formations, and correlation
(Glikson et al. 1996). These ages constrain the maximum age of between the two successions appears robust (Calver & Walter
deposition of Australian Cryogenian successions to c. 830 Ma. 2000; Preiss 2000; Hoffman et al. 2009). The King Island succes-
Few Australian Cryogenian dates can be considered reliable sion is intruded by a 575 + 4 Ma subvolcanic sill, and overlain by
(Preiss 2000; Fanning & Link 2008). A thermal-ionization mass mafic volcanics with a Sm –Nd isochron age of 579 + 16 Ma
spectrometry (TIMS) U –Pb age of 802 + 10 Ma was reported (Meffre et al. 2004), although the timing of the intrusion relative
from the Rook Tuff, in the lower part of the Curdimurka Subgroup to deposition is unclear. In northwestern Tasmania, the Croles
of the Callanna Group in the Adelaide Rift Complex (Fanning et al. Hill Diamictite, upsection of the Sturt-equivalent Julius River
1986). Recent dating of a bedded tuff in the Fifteenmile Group Member, is locally underlain by rhyodacite with a SHRIMP
of the Coal Creek Inlier, Ogilvie Mountains, NW Canada at U –Pb age of 582 + 4 Ma (Calver et al. 2004), indicating that
c. 811.5 Ma (Macdonald et al. 2010) is consistent with the Rook this could be a Gaskiers-age glaciation, that is, c. 580 Ma. If the
Tuff age, and apparently lies at the same stratigraphic and Cottons Breccia and Elatina Formation are correlatives of the
chemostratigraphic level. Croles Hill Diamictite, then the base of the Ediacaran (as defined
Reid (2009) recently obtained an age of 795 + 5 Ma using by the base of the Nuccaleena Formation) could be c. 580 Ma,
laser ablation – inductively coupled plasma mass spectrometry rather than 635 Ma. Better age constraints are required on the

Fig. 8.5. (Continued) Quartzite; UGE, Umberatana Group (of the Adelaide Rift Complex) equivalents; WDM, Wanapi Dolomite Member; WM, Woolnough Member;
WV, Wantapella Volcanics; YCE, Younghusband Conglomerate equivalent. Geochronology: (1) 827 + 6 Ma (Wingate et al. 1998) and 824 + 4 Ma (Glikson et al.
1996), Gairdner Dyke Swarm; (2) 802 + 10 Ma, Rook Tuff (Fanning et al. 1986); (3) 797 + 5 Ma, Skillogalee Dolomite (Drexel 2009); (4) 725 + 11 Ma detrital age,
uppermost Kanpa Formation (Nelson 2002); (5) 659 + 6 Ma, Wilyerpa Formation (Sturt Tillite correlative) (Fanning & Link 2008); (6) 657.2 + 5.4 Ma, Aralka
Formation (Kendall et al. 2006); (7) 647.2 + 10 Ma and 645.1 + 4.8 Ma, Tindelpina Shale Member (Kendall et al. 2006); (8) 640.7 + 4.7 Ma (Kendall et al. 2007),
Black River Dolomite; (9) 657 + 17 Ma detrital age, Marino Arkose Member (Ireland et al. 1998); (10) 575 + 3 Ma (Calver et al. 2004), intrusive in Yarra
Creek Shale; (11) 582 + 4 Ma, rhyodacite flow beneath Croles Hill Diamictite (Calver et al. 2004); (12) 556 + 24 Ma detrital age, Bonney Sandstone (Ireland
et al. 1998).
118 K. GREY ET AL.

Elatina Formation and Nuccaleena Dolomite to confirm their cor- scope of this review. Moreover, very few of the records contain
relation with the Chinese and Namibian glaciations and determine adequate range charts, so it is very difficult to determine the rela-
whether only the Croles Hill Diamictite is a Gaskiers glaciation tive stratigraphic positions and ranges of each of the species.
equivalent, or whether other Australian Elatina glacial deposits
are also of this age.
Chronometric constraints on Ediacaran successions are poor Palynology
and restricted to a Rb –Sr whole-rock isochron of 588 + 35 Ma
on the postglacial Bunyeroo Formation, based on samples above In Australia, identical palynological successions are present in
and below the level of the Acraman impact ejecta layer in drill Empress 1A and Lancer 1 and are known from several other
hole SCYW 1a (Hill et al. 2008) and a single detrital zircon western Officer Basin drill holes, particularly LDDH 1, Kanpa
grain, dated at 556 + 24 Ma, from the lower part of the Rawnsley 1A and Yowalga 3 (Grey & Cotter 1996; Cotter 1997, 1999;
Quartzite (Preiss 2000). The ejecta layer, a significant stratigraphic Grey & Stevens 1997; Grey 1999a, 2005a; Hill et al. 2000a)
tie-line, has not been dated but was estimated to be c. 580 Ma (Fig. 8.1). Elements of the various assemblages have been
(Williams & Gostin 2005) based on a c. 635 Ma age for the recorded in other drill holes, but distributions and ranges have
Elatina glaciation. not yet been thoroughly documented (Grey, unpublished data).
The undated glaciogenic Egan Formation of the Kimberley area The oldest described Cryogenian assemblage is probably that
appears to correlate with carbonate units (the Julie Formation of from the Alinya Formation in Giles 1, eastern Officer Basin
the Amadeus Basin, the Elkera Formation of the Georgina Basin, (Zang 1995). This assemblage is puzzling because it contains Tra-
the basal Bonney Sandstone of the Adelaide Rift Complex and chyhystrichosphaera, Cymatiosphaeroides kullingii and Valeria
the Wilari Dolomite Member, and the Tanana Formation of the lophostriata (normally found in Tonian or Mesoproterozoic suc-
Officer Basin) that lie well above the cap carbonate and just cessions), vase-shaped microfossils, assigned to Melanocyrillium
below the first record of bilaterians in other basins (Grey & sp. (also usually found in early Cryogenian assemblages), Cryo-
Corkeron 1998). genian species similar to those in the Bitter Springs and Browne
Formations, and spiny forms resembling acanthomorphs recorded
from the Ediacaran (although none of the latter seems identical to
Biostratigraphy other Australian Ediacaran species). This strange combination of
taxa may indicate reworking of older rocks and possibly the pres-
Biostratigraphic correlation is increasingly significant for ence of two disconformable successions. This assemblage requires
Australian Neoproterozoic correlation, and is based on chert reassessment.
microfossils, palynology and stromatolite biostratigraphy. Paly- Vase-shaped microfossils are locally abundant in chert in the
nology provides good biostratigraphic control as well as indicating Black River Dolomite below the diamictitic Julius River
palaeoenvironment and thermal maturity. Microfossil assemblages Member in Tasmania (Saito et al. 1988), and are consistent with
consist predominantly of leiospheres, filaments and mat fragments, an age pre-dating the Sturt glaciation (Porter & Knoll 2000).
mostly of conservative, long-ranging species with simple mor- Elsewhere in Australia, acritarch and microfossil assemblages
phologies, but containing some more complex and time-diagnostic show consistent distribution patterns. In the western Officer
species (Schopf 1968; Schopf & Barghoorn 1969; Fairchild 1976; Basin, they are present in the Browne Formation and its lateral
Oehler 1976; Jackson & Muir 1981; Jackson & van de Graaff equivalents (Jackson & van de Graaff 1981; Grey & Cotter 1996;
1981; Zang & Walter 1992; Zang 1995; Grey & Cotter 1996; Cotter 1997, 1999; Grey & Stevens 1997; Grey 1999a, 2005a).
Cotter 1997, 1999; Grey & Stevens 1997; Grey 1999a, 2005a; Well-preserved microfossils from chert in the Browne Formation
Hill et al. 2000a, b). Data have been obtained from at least 1000 at the Madley Diapirs (Jackson & van de Graaff 1981, p. 34;
samples from about 30 drill holes across Australia, several consist- Cotter 1997) show low species diversity and mainly represent
ing of continuous core (Figs 8.1 & 8.6). Stromatolite biostratigra- permineralized cyanobacterial mat. Palynomorphs from the
phy is based on .200 samples collected from .80 localities, Browne Formation in Browne 1 and 2, Kanpa 1A, Dragoon 1
together with .70 horizons examined in drill core (Figs 8.2, 8.3 and Yowalga 2 and 3 include Myxococcoides sp., Siphonophycus
and 8.7, Table 8.1). Only localities with identified stromatolites sp., Leiosphaeridia spp. and Synsphaeridium spp., many of
are recorded here; there are numerous samples that have yet to which are long ranging and have been recorded from younger
be identified. Core data can be tied to field outcrops using both stro- and older successions (Cotter 1997, 1999). Some species of
matolite and acritarch biostratigraphy, and taxonomic ranges are Simia and Satka colonialica may have more restricted stratigraphic
consistent across the basins and correspond to the correlations distributions. Similar taxa were subsequently identified in other
determined by independent methods such as lithostratigraphy, Officer Basin drill holes, including LDDH 1, Empress 1A and
seismic and well-log data and chemostratigraphy. Lancer 1 (Grey & Cotter 1996; Grey 1999a, 2005a). This lower
Biostratigraphic correlation is more difficult globally because Buldya Group (Browne Formation) assemblage has many
so few successions have so far been adequately documented. species in common with assemblages from the Gillen Member of
Although there are probably at least two hundred assemblages the Bitter Springs Formation (Zang & Walter 1992).
recorded in the literature (Knoll 1996), ages are often poorly con- Palynomorphs from the upper Buldya Group (Hussar, Kanpa and
strained, so it is difficult to know precisely when in the Neoproter- Steptoe Formations) in Hussar 1, Lungkarta 1, Yowalga 2 and 3,
ozoic (or in some cases, in the Proterozoic) some of the successions Empress 1A and Lancer 1 (Grey & Cotter 1996; Cotter 1999;
occur. Several reviews have been published, most notably a com- Grey 1999a, 2005a; Hill et al. 2000a; Grey et al. 2005) include
pilation of c. 3500 occurrences of Proterozoic and Early Cambrian Cerebrosphaera buickii, C. ananguae, Chuaria sp. cf. circularis,
microfossils and microfossil-like objects, based on 470 strati- Coneosphaera sp., Eoentophysalis croxfordii, Eomicrocystis sp.
graphic units and 316 published papers by Mendelson & Schopf cf. elegans, Leiosphaeridia crassa, L. sp. cf. exsculpta, L. jacutica,
(1992). From the taxonomic point of view, the catalogue of acri- L. minutissima, L. tenuissima, L. ternata, Myxococcoides canta-
tarchs published by Fensome et al. (1990) is invaluable, but did brigiensis, Ostiana microcystis, Pterospermopsimorpha insolita,
not include microfossils other than those considered to be acri- Satka colonialica, Simia annulare, Stictosphaeridium sinapticuli-
tarchs. Many significant publications were overlooked in this ferum, ?Symplassosphaeridium sp., Synsphaeridium spp., Tasma-
catalogue and it is difficult to extract taxa that fall within the nites sp., as well as filamentous microfossils Calyptothrix sp. cf.
age constraints of interest. alternata, Clavitrichoides rugosus, Oscillatoriopsis amadeus,
Given the numerous revisions to stratigraphy of recent years, Oscillatoriopsis sp., Siphonophycus kestron, S. robustum, S. septa-
updating the records is a major undertaking and beyond the tum, S. solidum and S. typicum. Although taxa from the previous
AUSTRALIAN CRYOGENIAN CORRELATIONS 119

Fig. 8.6. (a– f) Cerebrosphaera buickii from Spitsbergen, Australia and the Grand Canyon. (a, b) Svanbergfjellet Formation, Spitsbergen: HUPC 62713B, P-2945-47M
(S-33-4), holotype (a); HUPC 62763, 86-G-33-2S, SEM (b). Photos in a and b are courtesy of N. Butterfield (specimens described in Butterfield et al. 1994). (c) Hussar
Formation, Officer Basin, Empress 1, 808.2 m/2, L-38-4. (d) Hussar Formation, Officer Basin, Lancer 1, 605.14 m/2, W-44-3, Officer Basin. (e) Pirilyungka Formation,
Vines 1, Officer Basin, 1685.33 m/2, L-57-2. (f ) BLD 4, Skillogalee Dolomite, Stuart Shelf, BLD 4, 1008.2 m, E-56-0. (g) PP12, Anama Siltstone, Stuart Shelf, PP12,
156.9 m, N-58-2. (h) ‘Finke beds’, Amadeus Basin, 1502.0 m, Q-33-0. (i) Lower Chuar Group (Galeros Fm), Grand Canyon, USA, SP14-63-14 (Nagy 2008, MSc thesis,
photo courtesy of S. Porter).

assemblage persist, the more elaborate forms tend to be fewer in Rhynie Sandstone, near the base of the Burra Group in drill hole
number, so assemblages are dominated by leiospheres, Siphono- PP 12 and the Skillogalee Dolomite (in drill hole BLD 4) in the
phycus spp., Synsphaeridium spp., Chuaria sp. cf. circularis and middle Burra Group (Hill et al. 2000a). The same species was
Cerebrosphaera buickii. recently identified in the ‘Finke beds’ between the Johnnys
In this assemblage, the distinctive acritarch Cerebrosphaera Creek Member of the Bitter Springs Formation and the Areyonga
buickii consistently first appears about the middle of the Hussar Formation in Wallara 1 in the Amadeus Basin (Grey, unpublished
Formation in the western Officer Basin (Grey & Cotter 1996; observations).
Cotter 1999; Hill et al. 2000a). In the Adelaide Rift Complex, C. buickii is a key species because it has been recognized
C. buickii is present in the Anama Siltstone Member of the at about the same stratigraphic level across Australia (Hill
120 K. GREY ET AL.

Cambrian
c. 695

c. 600
c. 800

c. 635
c. 660
Approximate age (Ma)

11
14
15

12
13

10
Stratigraphic package

8
9

6
7

1
5

2
3
PQ/RBb Parmites f. indet.
GM Tungussia erecta
LCM2/BrF Minjaria pontifera
LCM2/ERD Linella avis

Acaciella australica
LCM2/BrF/?CD equiv. Kulparia alicia

assemblage
LCM2/BrF Jurusania nisvensis
LCM2/BrF/WWF Inzeria intia

Sturt glaciation
CD Gymnosolen f.indet.
LCM2/BrF Boxonia pertaknurra
LCM2/BrF/SHF Basisphaera irregularis
LCM2/BrF/SHF/KF/WWF/Yb/CD Acaciella australica
BuG/UBG/UTG/JCb/WHF/JRM (clasts) Baicalia burra

Baicalia burra
assemblage
UBG Tungussia n. form
UBG Stomatolite n. form (pseudocolumnar)
Elatina glaciation
UBG/UTG Conophyton n. form
UBG/BuG Tungussia wilkatanna
THF Omachtenia n. form
BgL Acaciella augusta

assemblage
multiplex
Inzeria
BgL Inzeria conjuncta
BgL Inzeria multiplex

BgL Boxonia melrosa


AnF Katavia costata

AnF Tungussia etina


AnF Linella f. indet.

EF Inzeria f. indet.

assemblage
munyallina
Linella
EF Linella munyallina
EF Jurusania burrensis
EF Kulparia kulparensis
?AF Clast supported stromatolite
PF Divergent branching stromatolite
TF unnamed stromatolite
PiS & OF/WF (in cap carbonate) Kotuikania juvensis = Elleria minuta
upper carbonate BdF Tesca stewartii

upper carbonate BwF Stromatolite form 1 Walter et al.

upper carbonate BwF Eleonora boondawarica


upper carbonate BwF Acaciella savoryensis

Basal BnS Linella f. indet.

JF/Basal BnS/WDM/EkF/EgF Tungussia julia

EgF Stromatolite form Grey & Corkeron


EkF Georgina howchini

Fig. 8.7. Range chart of Neoproterozoic stromatolite distributions in Australia (see Table 8.1 for guide to abbreviations and additional stratigraphic information). Details
of the stratigraphic units mentioned here can be found by searching the Geoscience Australia Stratigraphic Names database on http://dbforms.ga.gov.au/pls/www/
geodx.strat_units.int.
AUSTRALIAN CRYOGENIAN CORRELATIONS 121

Table 8.1. Key to abbreviations in stromatolite range chart (Fig. 8.7)

Symbol Package Stratigraphic unit

AnF 8 Angepena Formation, Upalinna Subgroup, Umberatanna Group, Adelaide Rift Complex
AF 6 Aralka Formation (Mount Conner, Fenn Gap & Pioneer Ssst clast), Amadeus Basin
BdF 1 Boord Formation, western Amadeus Basin
BgL 9 Brighton Limestone, Upalinna Subgroup, Umberatanna Group, Adelaide Rift Complex
BnS 1 Bonney Sandstone, Pound Subgroup, Wilpena Group, Adelaide Rift Complex
BrF 13 Browne Formation, lower Buldya Group, western Officer Basin
BuG 12 Burra Group (Skillogalee Dolomiteand other units), Adelaide Rift Complex
BwF 1 Boondawari Formation, western Officer Basin
CD 13 Coominaree Dolomite, Callanna Group, Adelaide Rift Complex & eastern Officer Basin
EF 7 Etina Formation, Upalinna Subgroup, Umberatanna Group, Adelaide Rift Complex
EgF 1 Egan Formation, Louisa Downs Group, central Kimberley region
EkF 1 Elkera Formation, Mopunga Group, Georgina Basin
ERD 13 Eliot Range Dolomite, Ruby Plains Group, eastern Kimberley area
GM 14 Gillen Member, Bitter Springs Formation, Amadeus Basin
JCb 12 Johnnys Creek beds’, Amadeus Basin
JF 1 Julie Formation, Amadeus Basin
JRM 12 Julius River Member, Black River Dolomite, Togari Group, Smithton Basin
KF 13 Karara Formation, Tarcunyah Group, NE Pilbara
LCM2 13 Loves Creek Member unit 2, Bitter Springs Formation, Amadeus Basin
OF 2 Olympic Formation, Amadeus Basin
PF 5 Pertatataka Formation, Amadeus Basin
PiS 2 Pioneer Sandstone, Amadeus Basin
PQ 15 Paralana Quartzite, Arkaroola Subgroup, Callanna Group, Adelaide Rift Complex
RBb 15 River Broughton beds, Curdimurka Subgroup, Callanna Group, Adelaide Rift Complex
SHF 13 Skates Hills Formation, Sunbeam Group, western Officer Basin
TF 4 Trezona Formationm, Upalinna Subgroup, Umberatanna Group, Adelaide Rift Complex
THF 10 Tapley Hill Formation, Nepouie Subgroup, Umberatanna Group, Adelaide Rift Complex
UBG 12 Upper Buldya Group (upper Hussar, Kanpa & Steptoe Formations), western Officer Basin
UTG 12 Upper Tarcunyah Group Undifferentiated, Officer Basin, south east Pilbara
WDM 1 Wilari Dolomite Member, Tanana Formation, Ungoolya Group, eastern Officer Basin
WF 2 Wahlgu Formation, western Officer Basin
WHF 12 Wright Hill Formation, eastern Officer Basin
WWF 13 Waltha Woora Formation, Tarcunyah Group, NE Pilbara
Yb 13 Yackah beds, Plenty Group, Georgina Basin

Details of the stratigraphic units mentioned here can be found by searching the Geoscience Australia Stratigraphic Names
database: http://dbforms.ga.gov.au/pls/www/geodx.strat_units.int.

et al. 2000a) and could be an important index species for global Sparse, reworked fragments are present in the Chambers Bluff
correlation (Fig. 8.6). It was previously recorded from the Svan- Tillite in Nicholson 2, eastern Officer Basin (Eyles et al. 2007),
bergfjellet Formation and overlying Draken Formation, Akademi- another probable Sturt glacial equivalent. Elsewhere diamictite
kerbreen Group, Spitsbergen (Knoll et al. 1991; Butterfield et al. and interglacial successions appear barren.
1994) and the Tanner, Jupiter and Carbon Canyon Members of The overlying Ediacaran acritarch succession is very well con-
the Galeros Formation, lower and middle Chuar Group, Grand strained lithostratigraphically across Australia, and closely
Canyon, USA (Nagy 2008). Considering that there is good geo- matches correlations based on chemostratigraphy, seismic inter-
chronological constraint on the first appearance (younger than pretation and stromatolite biostratigraphy, even though absolute
802 + 10 Ma for the Rook Tuff and older than 795 + 5 Ma for dating remains poor (Grey 2005b; Willman et al. 2006; Willman
the Skillogalee Dolomite), and the coincidence or near coincidence & Moczydłowska 2008). Acritarch distributions are consistent
with three other key markers (d13C, 87Sr/86Sr and the first appear- and well constrained from above the Elatina glaciation up to the
ance of the stromatolite Baicalia burra), the first appearance of change of facies associated with the appearance of the bilaterian
C. buickii is a strong candidate for a boundary that could be corre- Ediacara fauna.
lated globally (see section ‘Chemostratigraphy’ below). Although biostratigraphic distributions are consistent across
Sturt glacial and post-glacial assemblages are poorly known Australia, they seem to be inconsistent with some of the distri-
throughout Australia. In the Adelaide Rift Complex only a few butions reported from NW Canada (Macdonald et al. 2010).
simple, poorly preserved leiospheres have been recorded There does not seem to be anything equivalent to the Wynniatt
(Timofeev 1966; Grey, unpublished data). Elsewhere, successions Group assemblage in Australia. Vase-shaped microfossils appear
appear barren. In Western Australia, a hiatus with a karstified sparse in the Australian successions, and there are no reports
surface characterizes the interval. In Vines 1, there may be a tran- of C. buickii at the appropriate stratigraphic levels in north
sition into the Pirrilyungka Formation, a newly defined diamictite America, with the exception of the Chuar Group.
unit that is probably a Sturt Tillite equivalent, and is directly
overlain by the Wahlgu Formation, a diamictite correlated with
the Elatina Formation (Haines et al. 2008; Hill et al. 2011). Stromatolites
C. buickii is present, possibly in situ, in the turbiditic Unit 1 at
the base of the Pirrilyungka Formation (Haines et al. 2008), but Stromatolites have been recorded at various levels throughout the
is only present as sparse, reworked fragments in the diamictite. Neoproterozoic of Australia (Walter 1972; Preiss 1972, 1973a, b,
122 K. GREY ET AL.

Fig. 8.8. (a– h) Taxa found in the Acaciella australica Stromatolite Assemblage: from Jay Creek, Amadeus Basin, Northern Territory (b, f, g); from Lancer 1, Officer
Basin, Western Australia (c, d, e). (a) Bioherms of A. australica in the Skates Hills Formation, Skates Hills, Officer Basin, Western Australia. (b) A. australica, polished
slab showing branching columns. (c) Basisphaera irregularis, drill core split face, 349 m. (d) Boxonia pertaknurra, drill core surface, 332.5 m. (e) Inzeria intia, drill core
split face 333 m. (f ) Kulparia alicia, GSWA 109264C, polished slab showing branching columns. (g) Linella avis, thin section, GSWA 109267. (h) Minjaria pontifera,
cut core face, Wallara 1, 1979 m, Amadeus Basin, Northern Territory.
AUSTRALIAN CRYOGENIAN CORRELATIONS 123

Fig. 8.9. Taxa found in the Baicalia burra Stromatolite Assemblage. All specimens from upper Buldya Group, Officer Basin, Western Australia; all core samples from
Empress 1A. (b– f, h, i) Drill core split faces. (a) B. burra, GSWA 84708, upper Buldya Group, near Lake Throssell. (b) B. burra, core face, c. 775 m, Kanpa Formation.
(c) B. burra, small columns, 510.8 m, Steptoe Formation. (d, e) Tungussia wilkatanna, Steptoe Formation: (d) 656.9 m (e), 513.5 m. (f ) Tungussia new form,
Empress 1A, 1077.4 m, Hussar Formation. (g, h) Conophyton new form, (g) GSWA 139573, upper Buldya Group, near Constance Headland; (h); 779.0 m, Kanpa
Formation. (i) Unnamed pseudocolumnar stromatolite, 520.7 m, Steptoe Formation.
124
120o 130
o
140o 150o ADELAIDE RIFT COMPLEX ADELAIDE RIFT COMPLEX
Peake and Denison Ranges Mid-North (Southern Flinders Ranges)
13
Thickness C (% ) 13
C (% )
Calthorinna (m) Thickness
-2 0 2 4 6 8 Belair (m)
KIMBERLEY Tillite
1528 Subgroup -2 0 2 4 6 8
REGION 1007

Dolomite
Member
Auburn
GEORGINA
20o 20
o
1400 900
AMADEUS BASIN
BASIN
Arunta Provin
ce 1300 800
Western
Officer W1 D A3
J
Basin
L B 1200 700
PEAKE & DENISON

Member
E
M5 C RANGES 1100 600
OFFICER

Kalachalp
Eastern
BASIN Officer
UBPD
o Basin 30o
30
Gawler
1000 500

W
Craton
STUART ADELAIDE 900 400
SHELF
SD
RIFT COMPLEX I
UBMF
800 300 J

Member
700 200

40
o
Drillhole
Field section KING ISLAND 40o
600 100 I
A
TASMANIA
500 0
Undalya
120o 130o 140o 150
o
Quartzite
400

s
ite
H

at y
om og
ol
Assemblage Str nol
300 Undalya

ly
2 4 6 8

Pa
Quartzite
578
200
WESTERN OFFICER BASIN 500 NW TASMANIA
100
Lancer 1 Empress 1A 400 Arthur River section
0 G
Unit A (mostly 300
s

13
ite

C (% )

s
m gy

ite
13 quartz sandstone)

m gy
ol

s
C (% )
ro lo

13

ol

ite
C (% ) Thickness
at

ro lo

m gy
St no

at
Depth 200 -2 0 2 4 6 8

ol
St no
Depth

ro lo
Member (m)
ly

at
(m) -2 0 2 4 6 8

ly

St no
Pa

-2 0 2 4 6 8
I (m)

Pa

ly
-2 0 2 4 6 8 ?F

Pa
act 100 500
500 H Steptoe J
1260
Keene 1200 0
Assemblage

Fm Assemblage Bungaree 400


Basalt 600 I Quartzite

K. GREY ET AL.
Kanp

600 1100
300
G
C.b.

Kanpa
700 H 1000 AMADEUS BASIN 200 ?E
700

s
Fm 900

ite
m gy
100

Assemblage
13

ol
C (% )

T
800

ro lo
at
St yno
800 800 Areyonga
Composite
-2 0 2 4 6 8 0

l
Pa
G 700
Formation
1049 ROCKY CAPE
900
G GROUP

C.b.
‘Finke
900 600 beds’ 1000 G
Hussar F

B.burra
Fm 1000
500
1000 Hussar
Fm
1100
F 400 900

1100 E 300 E
800
1200
E 200 ADELAIDE RIFT COMPLEX
1200
100 MTS
Peake and Denison Ranges
Lancer
Member Aeolian deposition 1300

s
700

ite
0
1300

W
act

ol
Ass.
D D

at
Depth
A.a.

om
Woolnough
A.a.

Woolnough
1400
Member
C Member
C Cadlareena (m)

r
D -2 0

St
600 Volcanics
1400
B 104
100

CALLANNA
B

Assemb.
EASTERN OFFICER BASIN B

Subgroup
Arkaroola
1500
A

GROUP

A.a.
A

CD
Basement Basement TD 1625 Manya-5 518
500
C or

Bluebush
50

outcrop
468 D
s
ite

YC
B
ol

Key
at

Depth 400
om

0
Cadlareena (m) -2 0 2 4 6 A weathered
r
St

Sandstone Stromatolite Volcanics


1139 372 outcrop
1150
Mudstone–siltstone Calcareous siltstone B 300
Assemblage

or
CDE

1200
Evaporite Dolomitic siltstone
D 200

Dolomite Missing section 1250


A
Limestone Basalt or 112
YCE

outcrop
100
Dump

1300
C
Unconformity Break in section
TD 1333 0
Heavitree Quartzite

Fig. 8.10.
AUSTRALIAN CRYOGENIAN CORRELATIONS 125

1974, 1976, 1985, 1987; Griffin & Preiss 1976; Walter et al. 1979, dominated by B. burra, which is geographically widely distributed.
1994; Grey 1995, 1999b, 2005a; Grey & Stevens 1997; Hill et al. In the Kanpa Formation, B. burra is accompanied by Conophyton
2000a; Grey et al. 2005, unpublished data; Williams et al. 2007; new form, and in both the Officer Basin and Adelaide Rift
Planavsky & Grey 2008), but discussion here is restricted mainly Complex by T. wilkatanna. The assemblage has been recorded in
to those of demonstrated stratigraphic significance (Figs 8.2, Empress 1A, Lancer 1, Kanpa 1A and Wallara 1, as well as
8.3 and 8.7–8.9, Table 8.1). Throughout Australia, pre-Sturt- being known from many Officer Basin field localities (Grey
glaciation Cryogenian successions are characterized by two wide- et al. 2005). B. burra and Conophyton form indet. have also
spread, apparently time-restricted stromatolite assemblages: the been collected from an unassigned unit near the top of the
lower Acaciella australica Stromatolite Assemblage (Figs 8.2, Tarcunyah Group near Constance Headland in the northwestern
8.7 and 8.8) and the upper B. burra Stromatolite Assemblage Officer Basin (Grey, unpublished data). A previously unknown
(Stevens & Grey 1997; Hill et al. 2000a; Figs 8.3, 8.7 and 8.9). form, tentatively identified as Tungussia? form indet., is present
The value of these two assemblages for biostratigraphy is clearly in both Empress 1A and Lancer 1.
established by the presence of both assemblages in the same rela- Both B. burra and T. wilkatanna are present in the Burra Group
tive stratigraphic relationships in three drill holes across two of the Adelaide Rift Complex (Preiss 1972, 1974, 1985, 1987),
basins. Two other assemblages, the Inzeria multiplex Stromatolite supporting correlation between the upper Buldya Group and
Assemblage and the Linella munyallina Stromatolite Assemblage at least part of the Burra Group. B. burra has also been recorded
(Fig. 8.7), are present in the interglacial succession, but so far are in Tasmania (Griffin & Preiss 1976), as erratics in Sturt glacial
known only from the Adelaide Rift Complex. Stromatolites are deposits (Julius River Member) of the Black River Dolomite
less common in the Ediacaran succession and seem to be of (Calver 1998; Grey, unpublished data). Unpublished data
more limited biostratigraphic application than the Cryogenian suggest that this assemblage is present in western USA and
ones (Fig. 8.7, Table 8.1). Canada in the Mackenzie Mountains, but none of the stromato-
The Acaciella australica Stromatolite Assemblage (Figs 8.2, 8.7 lites in this area or in other areas of NW Canada has been
and 8.8, Table 8.1), epitomized by Unit 2 of the Loves Creek formally identified. Testing whether or not the apparently abun-
Member, Bitter Springs Formation, Amadeus Basin, is character- dant stromatolites show similar distribution patterns to those
ized by Acaciella australica and 12 associated forms (Stevens & of Australia could provide a significant step forward in
Grey 1997; Hill et al. 2000a). In Western Australia, this assem- global correlation.
blage is known from the Browne Formation (including in continu-
ous core in Empress 1A and Lancer 1), the Skates Hills Formation
of the western Officer Basin (Hill et al. 2000a; Grey 1995, 2005a; Chemostratigraphy
Grey et al. 2005) and unassigned dolomite in the Tarcunyah Group
(Grey, unpublished data). In the last decade, advances in Neoproterozoic chemostratigraphic
The Acaciella australica Stromatolite Assemblage is wide- (d13C and 87Sr/86Sr) correlations throughout Australia have pro-
spread in the Amadeus Basin, including the Wallara 1 drillcore vided ties between the Australian Cryogenian and successions in
(Walter 1972; Walter et al. 1979; Grey, unpublished data). It is Canada, Svalbard, Namibia and Mongolia (Calver & Lindsay
also present in the Coominaree Dolomite of the Adelaide Rift 1998; Calver 2000; Calver & Walter 2000; Hill & Walter 2000;
Complex, equivalent successions in the eastern Officer Basin, Hill et al. 2000b; Walter et al. 2000; McKirdy et al. 2001; Hill
the Yackah beds of the Georgina Basin, and in the Eliot Dolomite 2005, unpublished data). Figures 8.10 and 8.11 show a compilation
in the eastern Kimberley (Preiss 1973b, 1987; Grey 1995; Grey & of pre-Sturt glaciation, and Sturt glaciation and post-Sturt glacia-
Blake 1999; Grey et al. 2005; Planavsky & Grey 2008). In South tion d13C stratigraphy of Australia, respectively. There are few
Australia, the A. australica Stromatolite Assemblage in the Coo- reliable Australian Cryogenian 87Sr/86Sr ratios, because dolomite
minaree Dolomite is older than the 802 + 10 Ma Rook Tuff. Its predominates over limestone. Despite minimal utility for
stratigraphic position relative to other biostratigraphic markers, Australia-wide correlations, they are extremely important globally
d13C and 87Sr/86Sr excursions and the Sturt glaciation is well- (see section ‘Discussion’).
documented in Australia (Hill & Walter 2000; Hill et al. 2000a, There is no official terminology in d13C or 87Sr/86Sr chemostra-
b; Hill 2005; unpublished data). One form, Linella avis, was first tigraphy. In the Neoproterozoic most authors focus on correlating
described from the Maly Karatau Ridge, Chichkan Formation, negative parts of d13C curves, because Neoproterozoic rocks pre-
Malokaroy Group and the Talass Range in Kazakhstan (Krylov dominantly have positive d13C values and any negative d13C
1967, 1975; Sergeev 1989; Grey & Blake 1999; Raaben et al. values occur only for relatively brief periods of time. This has
2001), which is possibly equivalent in age to the Bitter Springs led to the use of ‘excursion’ and ‘anomaly’, often preceded by
Formation (Sergeev 1989; pers. comm. 1996). Poorly documented ‘negative’, the connotation being that these are short-lived
forms from the Mackenzie Mountains of Canada appear to belong events. Even when not preceded by ‘negative’ it could confuse
to this assemblage and require further investigation (Grey, unpub- the reader into that interpretation. There are significant ‘positive’
lished data). anomalies or excursions, particularly in Cryogenian rocks, that
In Western Australia, the overlying B. burra Stromatolite are important for global correlation. The term ‘stage’ has been
Assemblage (Figs 8.3, 8.7 and 8.9, Table 8.1) ranges from about used to describe the long interval of negative d13C values in the
the middle Hussar Formation, through the Kanpa Formation and Loves Creek Member of the Bitter Springs Formation (Halverson
into the overlying Steptoe Formation. The assemblage is et al. 2005), but could equally be applied to long intervals of

Fig. 8.10. (Continued) Pre-Sturt glaciation d13C stratigraphy of Australia. Location of drill hole and outcrop sections shown in inset (see Fig. 8.4 for explanation of
abbreviations). Horizontal lines in palynology and stromatolites columns indicate first appearances: (A.a) Acaciella australica; (B) Baicalia burra Stromatolite
Assemblage; (C.b) Cerebrosphaera buickii. See text for an explanation of d13C correlation intervals A–J. Western Officer Basin: lithostratigraphy (Stevens & Apak
1999; Haines et al. 2004), d13C (Hill & Walter 2000; Hill 2005) and biostratigraphy (Grey 1999a, b, 2005a); eastern Officer Basin: lithostratigraphy (Morton 1997), d13C
(Hill & Walter 2000) and biostratigraphy (Preiss 1987; Grey 1995); Peake and Denison Ranges: lithostratigraphy (Fairchild 1976; Ambrose et al. 1981), d13C (Hill &
Walter 2000) and biostratigraphy (Preiss 1987; Hill et al. 2000a); southern Flinders Ranges: lithostratigraphy (Preiss 1987), d13C (Hill & Walter 2000) and
biostratigraphy (Hill et al. 2000a); Amadeus Basin: lithostratigraphy (Hill et al. 2000b, and references therein), d13C (Hill et al. 2000b) and biostratigraphy (unpublished
data); Tasmania: lithostratigraphy and d13C (Calver 1998). Abbreviations: CD, Coominaree Dolomite; CDE, Coominaree Dolomite equivalent; MTS, molar tooth
structure; TD, total depth of drill hole; WM, Woolnough Member; YC, Younghusband Conglomerate; YCE, Younghusband Conglomerate equivalent.
126 K. GREY ET AL.

positive values. The problem is that ‘stage’ has lithostratigraphic section ‘Discussion and Global Correlation’). The next youngest
and biostratigraphic connotations. In this manuscript we prefer well-preserved 87Sr/86Sr ratios in Australia are from within the
the use of ‘d13C interval’ or just ‘interval’ to describe those Sturt glacial deposits in northwestern Tasmania, where the
parts of d13C curves that are significant for Australian and global lowest ratio is 0.7063 (Calver 1998).
correlation and that may include positive or negative d13C excur- In Sturt glacial deposits and post-Sturt –pre-Elatina glaciation
sions, or both, or parts of d13C curves that do not have anomalously (interglacial) deposits there are lithological, geochemical (d13C
low or high values. This terminology is preferable for two reasons. intervals K – P) and geochronological similarities between the
The first is that it does not exclude those parts of d13C curves that Amadeus Basin, Stuart Shelf, Adelaide Rift Complex and north-
are not anomalous; a long interval of moderately positive d13C western Tasmania. The remainder of the interglacial succession,
values can be equally important for correlation. Second, in some excluding Elatina glaciation deposits, is only known in the Ade-
cases it minimizes focusing on every peak or trough in d13C laide Rift Complex, so no interbasinal comparison is possible.
curves that may be a result of local or regional factors, and However, the Adelaide Rift Complex d13C record compares
instead concentrates on the broader-scale d13C trends or patterns. closely with global records (see section ‘Discussion and Global
Ten distinct d13C intervals (A–J) have been identified below the Correlation’). A rise in 87Sr/86Sr ratios from 0.7063 within the
base of the Sturt glaciation in Australia based on correlations Sturt glacial deposits (northwestern Tasmania; Calver 1998) to
between at least two separate drill hole or outcrop sections (see 0.7071 in the interglacial succession (Adelaide Rift Complex;
also Hill & Walter 2000; Walter et al. 2000) (Fig. 8.10). McKirdy et al. 2001) also compares closely with global records.
Beneath interval A there is only a d13C stratigraphy in the For Elatina glaciation units there are data from dolomite beds –
Amadeus Basin (Gillen Member, Bitter Springs Formation) so originally primary precipitates – within the Olympic Formation
no Australia-wide correlations are possible; however, this interval (d13C, –1 to 3.5‰; Kennedy et al. 2001; Skotnicki et al. 2008)
is potentially important for global correlation. Intervals A –D are and Pioneer Sandstone (d13C, –2 to 2‰; Calver 1995) that
older than c. 800 Ma and coincide with the A. australica Stroma- define an upward increasing trend. These d13C data are comparable
tolite Assemblage described above. The palynological and micro- to those within the Ghaub Formation in Namibia (Kennedy et al.
fossiliferous records at this stratigraphic level across Australia are 2001) (see section ‘Discussion and Global Correlation’).
also consistent. The predominant chemostratigraphic feature at
this stratigraphic level is the presence of 13C-depleted d13C
values ( –4 to 0‰) (intervals B and D; best represented by Units Discussion and global correlation
1 & 2 of the Loves Creek Member, Bitter Springs Formation,
Amadeus Basin), which are separated by a comparatively brief From correlations shown in Figures 8.10 and 8.11, a composite
interval of positive d13C values (c. þ6‰) (interval C). There are Australian Cryogenian chemostratigraphy (d13C and 87Sr/86Sr)
no reliable 87Sr/86Sr ratios at this stratigraphic level, but there is compared with other global successions including the first
are in older sediments, in the Gillen Member of the Bitter records of C. buickii and the stromatolite B. burra (Fig. 8.12). In
Springs Formation. Spitsbergen and the USA, only Baicalia has been identified. To
Intervals E –J are younger than c. 800 Ma and older than avoid ambiguous d13C correlations between global successions,
c. 660 Ma, are characterized by predominantly positive d13C the tie-lines in Figure 8.12 (dark green dashed lines) are chosen
values (mostly þ2 to þ6‰), and coincide with the first appearance to coincide with the most extreme values, which avoids the more
of C. buickii and the B. burra Stromatolite Assemblage. First nuanced fluctuations that may be a result of more localized
records of these taxa occur in interval E in the Amadeus Basin (regional) heterogeneity. In this way it is possible to show simi-
and interval F in the western Officer Basin (Grey, unpublished larities in d13C and 87Sr/86Sr that extend globally.
data), and in the lowermost Burra Group in the Adelaide Rift Three ‘broad’ intervals of the global d13C curves are possible,
Complex (this chapter), which probably corresponds with interval which are consistent with radiometric dates, 87Sr/86Sr and
E (however, further d13C stratigraphy needs to be completed at this palaeontological data.
level in the Adelaide Rift Complex in order to verify this). d13C
values of between –3 and 0‰ in the middle Hussar Formation
may correlate with the Black River Dolomite in Tasmania (interval Australia – Canada– Svalbard: .800 Ma
F); however, this correlation is tentative because the Arthur River
section is not as thick as in Lancer 1 and Empress 1A in the western This correlation is based largely on an extended stratigraphic inter-
Officer Basin. It is probable that this part of the succession (interval val where d13C values fall between –4 and 0‰ (intervals B & D in
F) is missing from the central (Wallara 1 locality) and northeastern Australia; Fig. 8.10). The c. 830–800 Ma age for this anomaly is
(AS 27 and AS 28 localities) Amadeus Basin. Interval F has also based on radiometric dates from Australia (see section ‘Geochro-
not yet been identified in the Adelaide Rift Complex; further nology’) and is consistent with a c. 780 Ma age for the Little Dal
work needs to be completed to see whether it does exist, because Basalt at the top of the Mackenzie Mountains Supergroup in
it could potentially be important in strengthening Australian corre- Canada (Jefferson & Parrish 1989; Harlan et al. 2003) and a
lations. Within the ranges of C. buickii and B. burra, d13C values c. 811.5 Ma age for a bedded tuff in the Fifteenmile Group of
peak at about þ8‰ (interval H), immediately preceding an inter- the Coal Creek Inlier, Ogilvie Mountains, northwestern Canada
val of 13C-depleted d13C values (interval I). d13C values then return (Macdonald et al. 2010), which lies just beneath the extended
to between þ4 and þ6‰ (interval J) just prior to the Sturt glacial interval of negative d13C values. It also corresponds with a
deposits (best represented in the Kalachalpa Formation of the rise in seawater 87Sr/86Sr from 0.7055–0.7057 to 0.7062–
Peake and Denison Ranges. 0.7063. A third independent line of evidence for the uniqueness
The positive shift in d13C values from between – 4 and 0‰ of this part of the Cryogenian record is the first record of C.
(interval D) to between þ2 to þ6‰ (interval E) coincides with buickii. In Spitsbergen, C. buickii first occurs in the upper parts
a rise in 87Sr/86Sr from 0.7057 to 0.7063, as recorded in the of the Lower Dolomite Member of the Svanbergfjellet Formation
Bitter Springs Formation (Fanning 1986, reproduced in Hill (Butterfield et al. 1994), which is in the uppermost part of the
et al. 2000b, unpublished data). The positive shifts in d13C and extended interval of negative d13C values between c. 830 and
87
Sr/86Sr are followed by the first records of C. buickii and the 800 Ma (intervals B and D in Australia; Fig. 8.10). In Australia,
B. burra Stromatolite Assemblage (interval E). Therefore, there the first record of C. buickii is in slightly younger rocks but there
is an approximate coincidence between the first appearance of are no known suitable siltstone facies in Australia at this time
C. buickii and positive shifts in d13C and 87Sr/86Sr at c. 800 Ma, within which C. buickii is typically found. In the Grand Canyon,
which is consistent with data from Svalbard and Canada (see USA, the first record of C. buickii is in the lowermost Chuar
o
120o 130o 140o 150

KIMBERLEY
REGION

GEORGINA
ADELAIDE RIFT COMPLEX
o
20
o
20
AMADEUS BASIN
BASIN e
Western
Arunta Provinc
LS
Composite field and drillhole sections
Officer W1
Basin 13
Musgrave Province PEAKE & DENISON C (‰)
RANGES (m)
OFFICER
BASIN Eastern
Thickness
-10 -8 -6 -4 -2 0 2 4 6 8 10
Officer
30o Basin SC 30
o

Gawler
B2 El
M O
Craton 500
STUART
EC
ADELAIDE Y
DC
SHELF RIFT COMPLEX

Trezona
Fm.
1000 km
P
Drillhole 0
40
o
Field section KING ISLAND 40
o
Thickness -10 -8 -6 -4 -2 0 2 4 6 8 10
F
390

Enorama
TASMANIA

Shale
o
120o 130o 140o 150

NE O
0
AMADEUS BASIN Thickness -10 -8 -6 -4 -2 0 2 4 6 8 10 Key

AUSTRALIAN CRYOGENIAN CORRELATIONS


880
Limbla Syncline field section
Sandstone Glacial diamictite

Etina Formation
13
Thickness C (‰)
(m)
-6 -4 -2 0 2 4 500 Mudstone–siltstone Diamictite
562
N
Ringwood Member

Dolomite Ooids

UMBERATANA GROUP
500 ?
Limestone Missing section
0
Thickness -10 -8 -6 -4 -2 0 2 4 6 8 10
130

Brighton
Unconformity

Limest.
100
400
13
0 Cdol (‰ VPDB)
Depth -10 -8 -6 -4 -2 0 2 4 6 8 10
Aralka Formation

0 13
STUART SHELF WD
Ccarb (‰ VPDB)
13
estimated from + Corg
300 SCYW 1a drillhole
13
M Depth C (‰)
Whyalla

Tapley Hill Formation


(m) -2 0 2 4 6
Unamed member

Sandstone 500
1140

MCGM
BL
200 M
CENTRAL 1200
Appila Tapley Hill Formation

AMADEUS BASIN 1000


Wallara-1 drillhole M NW TASMANIA
100 Forest-1 drillhole
13
C (‰) 1300 13
Pioneer Depth C (‰)
Depth
L Sandstone (m) -4 -2 0 2 4 Kanunnah
(m) -6 -4 -2 0 2 4
Tillite TSM

1281 1500 Subgroup


790
Ar L L TSM L

TOGARI
0 1300 L 800

BRD
A

Formation

GP

JRM
K
Wilyerpa

900 K
Formation
Areyonga

1400 K 975

K 1450
2031
T.D. 1075m
1400
1425
Johnnys Creek
beds

Fig. 8.11. Sturt glaciation and post-Sturt glaciation d13C stratigraphy of the Australian Cryogenian. Location of drill hole and outcrop sections shown in inset (see Fig. 8.4 for explanation of abbreviations). See text
for an explanation of d13C correlation intervals K–P. Stratigraphy and d13C data from the Amadeus Basin and Stuart Shelf (Walter et al. 2000), the Adelaide Rift Complex (Walter et al. 2000; McKirdy et al. 2001), and
northwestern Tasmania (Calver 1995, 1998). In the Adelaide Rift Complex the following sections were used: drill hole Blinman 2 for the Wilyerpa (Sturt Tillite correlative) and Tapley Hill Formations; Depot Flat field section
for the Brighton Limestone; field section east of the Oraparinna Diapir for the Etina Formation; Mallee Water field section for the Enorama Shale; Doodney’s Well Hills field section for the Trezona Formation. Abbreviations: A,
Areyonga Formation; Ar, Aralka Formation; BL, Brighton Limestone; BRD, Black River Dolomite; El, Elatina Formation; JRM, Julius River Member; MCGM, Mount Caernarvon Greywacke Member; TSM, Tindelpina

127
Shale Member; WD, Wockerawirra Dolomite; Y, Yudnamutana Subgroup.
AUSTRALIA CANADA SVALBARD NAMIBIA MONGOLIA
87
87
Sr/86Sr (0.70..) 87
Sr/86Sr (0.70..)
87
Sr/86Sr (0.70..) 87
Sr/86Sr (0.70..) Sr/86Sr (0.70..)
60 70 80

128
13 13 60 70 80 13 13 13
AGE 60 70 80 60 70 80 60 70 80
(Ma) -8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8 -8 -6 -4 -2 0 2 4 6 8
635
T W

KEELE
0.7071

ELBO
0.7072 0.7071
650 0.7071
0.7071

TW
6 7 8 0.7068

T
5 0.7063 0.7068
S Rapit

,
675

CLG
0.7064
, 700

DRAKEN BACKLUNDTOPPEN ELBO


0.7067
USA
Wilyerp
13
725
illite 4
-8 -6 -4 -2 0 2 4 6 8
S

Kalachalp 0.7069
750

Kanp
775
* -8 -6 -4 -2 0 2 4 6 8

SVANBERGFJ

K. GREY ET AL.
0.7062 -8 -6 -4 -2 0 2 4 6 8
((

3
800
Johnnys
,
2
* * 60 70 80

0.7063

0.7062
13
Ccarb (‰)
GP

825 Siliciclastic
1 13
Carbonate/siliciclastic Corg (‰)
C. buickii 1st record
0.7057 Carbonate
Glacial * B. burra 1st record

((
850
-8 -6 -4 -2 0 2 4 6 8
-8 -6 -4 -2 0 2 4 6 8
60 70 80
60 70 80 0.7055

0.7055

-8 -6 -4 -2 0 2 4 6 8
60 70 80

Fig. 8.12. Global d13C, 87Sr/86Sr and biostratigraphic correlation of the Australian Cryogenian. The circled numbers on the time scale correspond to geochronological ages in Figure 8.5. Australia: 87Sr/86Sr ratios from Fanning
(1986) and reproduced in Hill et al. (2000b) (Gillen Member, unpublished data) (Unit 3, Loves Creek Member), Calver (1998) (Julius River Member, Black River Dolomite), and McKirdy et al. (2001) (Brighton Limestone
and Trezona Formation). Canada: d13C data from Halverson (2006) (Little Dal Group and Coates Lake Group), and Hoffman & Schrag (2002) and James et al. (2001) (Windemere Supergroup). 87Sr/86Sr ratios from Narbonne
et al. (1994), Kaufman et al. (1997) and Halverson et al. (2007b). c. 780 Ma age for the Little Dal Basalt (Jefferson & Parrish 1989; Harlan et al. 2003). Abbreviations: CLG, Coates Lake Group; Tw, Twitya Formation; Keele,
Keele Formation. Svalbard: d13C data from Halverson et al. (2005, 2007a). 87Sr/86Sr ratios from Halverson et al. (2005, 2007b). First record of Cerebrosphaera buickii, upper Lower Dolomite Member, Svanbergfjellet
Formation (Butterfield et al. 1994). Abbreviations: Elbo, Elbobreen Formation. Namibia: d13C data from Halverson et al. (2005). 87Sr/86Sr ratios from Yoshioka et al. (2003) and Halverson et al. (2007b). 746 Ma age from
Hoffman et al. (1996) and 760 Ma age from Halverson et al. (2005). Mongolia: d13C and 87Sr/86Sr data from Shields et al. (2002). USA: d13C data from Dehler et al. (2005). First record of Cerebrosphaera buickii, Tanner
Member, Galeros Formation, Chuar Group (S. Porter and Nagy, unpublished data). 770 Ma age from (Williams et al. 2003) and 742 Ma age from Karlstrom et al. (2000).
AUSTRALIAN CRYOGENIAN CORRELATIONS 129

Group, which has a maximum age of c. 770 Ma (Dehler et al. 2005, peaks in both d13C and 87Sr/86Sr within the stratigraphic range
and references therein) and is therefore consistent with the of C. buickii and B. burra.
approximate ages of the first records in Australia and Spitsbergen.
Only limited palynological studies have been carried out on the
Mackenzie Mountains successions, and have been mainly on Edia- Australia – Canada – Svalbard – Namibia – Mongolia:
caran to Cambrian samples (Baudet et al. 1989), but the Wynniatt
660– 635 Ma
Formation of the Shaler Supergroup in the Minto Inlier has 13 acri-
tarch species (Butterfield & Rainbird 1998). A synthesis of lithos-
The ‘interglacial’ part of the Cryogenian succession has a remark-
tratigraphic studies of arctic Canada reveals close similarities
ably consistent d13C and 87Sr/86Sr record. In immediately post-
between the Mackenzie Mountains and Shaler Supergroups, and
Sturt glaciation sediments, d13C values mostly fall within
more specifically between the Rusty Shale and Wynniatt For-
the range of –4 to –2‰, with the exceptions of Australia ( –6 to
mation (Rainbird et al. 1996). The absence of C. buickii in the
–4‰) and Svalbard ( –1 to 0‰). With the exception of Svalbard
Wynniatt Formation, whose probable correlative, the Rusty
(d13C between þ6 and þ7), d13C values then rise rapidly in all
Shale, is just beneath the extended interval of negative d13C
countries to between þ8 and þ10‰, and even higher in Mongolia.
values between c. 830 and 800 Ma, is consistent with the strati- 87
Sr/86Sr values rise concomitantly from 0.7068 to 0.7071. In
graphic level of C. buickii in Spitsbergen. Although further paly-
Australia, Canada and Namibia, d13C values then rapidly decline
nology in Canada above this stratigraphic level is necessary to
to between –6 and –10‰ (the Trezona ‘anomaly’; Halverson
determine exactly where C. buickii first appears, the first record
et al. 2005). These interglacial peaks and troughs are short-lived
of this acritarch is globally consistent. A fourth, and potentially
compared to the Ediacaran Shuram anomaly, and are more likely
more controversial line of evidence is the first record of the B.
to represent a global ocean signal (Bristow & Kennedy 2008,
burra Assemblage, which across Australia consistently first
and references therein). Unless evidence is presented to the con-
appears at about the same stratigraphic level as C. buickii. The stro-
trary, the stratigraphic level at which the maximum interglacial
matolite Baicalia has been reported in Svalbard (Halverson et al.
peaks in d13C and 87Sr/86Sr occur is a candidate for further subdi-
2007b) and the USA (Dehler et al. 2005), just above the first
vision of the Cryogenian Period, between c. 650 and 635 Ma. A
appearance of C. buickii, but has not been identified to form level.
stratigraphic level within or close to the lithological boundaries
Together, d13C, 87Sr/86Sr and the first records of C. buickii and
of the Sturt glaciation is not appropriate because of the apparent
B. burra have the potential to improve global correlation and poss-
range in radiometric ages for the Sturt glaciation globally
ibly inform a chronostratigraphic definition of the Cryogenian
(Fanning & Link 2004, 2008).
Period, and that c. 800 Ma could be a potential lower age limit
for the co-occurrence of significant geochemical and palaeontolo-
gical anomalies. The authors recognize, however, that more
research, in particular on the palaeontological record, needs to Conclusions
be completed to strengthen these arguments.
The thick and well-preserved Cryogenian successions in several
Australian basins can be correlated using a variety of means, and
Australia – Svalbard – Namibia – USA: 800– 700 Ma these techniques are most effectively used in combination.
However, poor preservation or lack of appropriate facies often
d13C values in this period are mostly moderately 13C-enriched with limits the applicability of biostratigraphy and chemostratigraphy,
several short-lived intervals to between –2 and 0‰. The dis- and the latter is often ambiguous owing to the non-uniqueness of
tinguishing feature of the d13C record is the occurrence of values patterns of variation. Chronometric constraints are particularly
of about þ8 to þ10‰ in Australia, Svalbard, Namibia, and poss- important, especially in testing global correlations, but such con-
ibly the USA. (The Chuar Group record is problematic because it is straints are regrettably few. Recent dating shows that the Sturt
predominantly based on d13Corg. None of the d13Ccarb values glaciation ended later (c. 660 Ma: Kendall et al. 2007; Fanning &
exceed þ4‰ and many of the ‘calculated’ (d13Ccarb ¼ 1TOC – Link 2008) than presumed correlatives on other continents. That
d13Corg) d13Ccarb values exceed þ10‰, the maximum known the same may be true of the Elatina glaciation is suggested (but
values in carbonate successions from Svalbard and Namibia.) A not proved) by dating from Tasmania (Calver et al. 2004). More
radiometric age of 760 Ma in Namibia places a maximum age con- dates are sorely needed, including a focus on the age of the Elatina
straint on the d13C peak. In Svalbard and Namibia there are two glacial, the close of which marks the end of the Cryogenian Period.
stratigraphic levels of þ8 to þ10‰ values, which makes precise As was the case with the base of the Ediacaran, it is probably not
correlation with Australia more ambiguous than lower in the suc- advisable to place the lower Cryogenian boundary at a horizon
cession. In the Coates Lake Group of Canada there are extremely associated with glaciation (in this case, the Sturt glaciation),
13
C-depleted d13C values that may correlate with the lower Polar- which is marked by major disconformities and unconformities.
isbreen Group of Svalbard, but it is difficult to be certain because in Additionally, although the Sturt glaciation appears to be consistent
the Coates Lake Group d13C values rapidly shift to þ8‰ and there in age across Australia, it may not be synchronous with glacial
is no record of this in Svalbard or elsewhere. An equivalent mag- units elsewhere that are normally correlated with it.
nitude d13C shift occurs between the two large Cryogenian glacia- The A. australica and B. burra Stromatolite Assemblages are
tions on all continents; however, 87Sr/86Sr values of 0.7064 in the Australia-wide markers, and unpublished data suggest they may
Coates Lake Group strongly suggest it is a pre-Sturt glacial unit. be present on other continents. However, the global distributions
The 87Sr/86Sr record in Svalbard suggests a rise in values from of these readily correlatable Australian taxa have yet to be demon-
0.7062 to 0.7069 concomitant with the rise in d13C values to strated. The first appearance of the acritarch C. buickii, coupled
their peak of þ8 to þ10‰. In Svalbard, the 87Sr/86Sr values with the first appearance of the stromatolite B. burra, appears to
then decrease to 0.7067. Ratios of 0.7064 in the Coates Lake be a significant marker, possibly globally, and should be con-
Group and 0.7063 in the Sturt glaciation in Tasmania suggest a sidered for determining the positioning of a boundary.
further decline in seawater 87Sr/86Sr. A similar rise then decline When global correlations are attempted with successions in
in seawater 87Sr/86Sr is known in Greenland (Fairchild et al. Canada, Svalbard, Namibia, the USA and Mongolia there are con-
2000) and, more recently, Scotland (Sawaki et al. 2010), so a sistencies between d13C and 87Sr/86Sr excursions beneath the Sturt
stratigraphic horizon at c. 760 Ma could be a strong candidate glaciation (c. 660 Ma in Australia) and between the Sturt and
for chronostratigraphic definition of the Cryogenian Period Elatina glaciations, and the approximate ages at which these
because of the coincidence of maximum pre-Sturt glaciation anomalies occur in other global successions are consistent with
130 K. GREY ET AL.

the Australian time scale. It thus appears that Cryogenian d13C and Ostiana microcystis Hermann 1976
87
Sr/86Sr excursions are of global extent and probably synchro- Pterospermopsimorpha insolita Timofeev 1969
nous and, together with the traditional tool of period definition, Satka colonialica Jankauskas 1979
first and last appearances of fossils, provides a way forward for Simia annulare Timofeev 1969 emend. Mikhailova in Jankauskas
further subdivision of the Neoproterozoic, at the base of the Cryo- et al. (1989)
genian Period. Stictosphaeridium sinapticuliferum
There are two stratigraphic levels where it may be possible to Symplassosphaeridium Timofeev 1959 ex Timofeev 1969
make a chronostratigraphic definition of the base of the Cryogen- Synsphaeridium Eisenack 1965
ian Period. The first is in rocks of c. 800 Ma (based on radiometric Tasmanites Newton 1875
ages in Australia and Canada), where there is an interval of Trachyhystrichosphaera Timofeev & German 1976 emend.
negative d13C values ( –4 to 0‰) that coincides with a rise in Jankauskas et al. 1989
87
Sr/86Sr ratios from 0.7055 to 0.7062 and the first appearance Valeria lophostriata (Jankauskas), Jankauskas 1982
of the distinctive acritarch C. buickii. Three independent lines of
evidence make this stratigraphic level a strong candidate for Filamentous microfossils
basal definition of the Cryogenian Period. The second is at
c. 760 Ma (a radiometric age from Namibia), where the pre-Sturt Calyptothrix sp. cf. Ca. alternata Jankauskas 1980
glaciation d13C record peaks at about þ8‰ in Australia, Svalbard, Clavitrichoides rugosus Mikhailova in Jankauskas et al. 1989
Namibia and the USA. 87Sr/86Sr ratios from Svalbard show a Oscillatoriopsis amadeus (Schopf & Blacic 1971) Butterfield 1994
concomitant increase from 0.7062 to 0.7069, and then a later Siphonophycus kestron Schopf 1968
decrease to 0.7064 (Coates Lake Group, Canada). There is no Siphonophycus robustum (Schopf 1968) Knoll et al. 1991
equivalent record of the 87Sr/86Sr rise from 0.7062 to 0.7069 in Siphonophycus septatum (Schopf 1968) Knoll et al. 1991
other countries due to incompatible lithologies, which weakens Siphonophycus solidum (Golub 1979) emend. Butterfield 1994
the case for boundary placement at this stratigraphic level. These Siphonophycus typicum (Hermann 1974) Butterfield et al. 1994
age rocks coincide with the stratigraphic ranges of C. buickii and
B. burra. The consistency of the data in numerous sections in Aus- Stromatolites
tralia, and limited reports of their occurrence elsewhere, indicate
that these two levels have potential for global boundary placement. Acaciella australica (Howchin 1914) Walter 1972
The synchroneity of glaciations in Australia and elsewhere in Baicalia burra Preiss 1972
the world remains unresolved, with new dates from NW Canada Basisphaera irregularis Walter 1972
conflicting with the limited data from Australia (Kendall et al. Boxonia pertaknurra Walter 1972
2009). Biostratigraphic data through this interval is also sparse, Conophyton Maslov 1937
but between the Sturt and Elatina glaciations (c. 660–635 Ma) Elleria minuta Walter and Krylov in Walter et al. 1979
there is the possibility for further subdivision of the Cryogenian Inzeria intia Walter 1992
Period at a stratigraphic level that coincides with a peak in Inzeria multiplex Preiss 1973
d13C of between þ8 and þ10‰ and 87Sr/86Sr of between Jurusania nisvensis Raaben 1964
0.7071 and 0.7072. Kulparia alicia (Cloud and Semikhatov 1969) Walter 1972
Linella avis Krylov 1967
Linella munyallina Preiss 1974
Appendix 1 Taxonomic citations Tungussia Semikhatov 1962
Tungussia wilkatanna Preiss 1974
The names of authors of scientific names have been omitted in the
text and text figures and instead are presented here. The names are The ideas and figures presented in this manuscript have benefited from discus-
the names of authors of taxa, not references, and therefore they are sions with numerous researchers, but in particular J. Gehling, P. Haines, R.
not necessarily cited in the reference list. Hocking, W. Preiss and G. Shields. Field samples and data have been collected
by many geologists for more than 100 years, especially M. Walter, W. Preiss, I.
Williams, N. Planavsky and M. Stevens. K.G. publishes with the permission of
Spheroidal microfossils and acritarchs the Executive Director of the Geological Survey of Western Australia and is
an Honorary Associate at Monash University and the University of Western
Cerebrosphaera ananguae Cotter 1999 Australia. This paper is a contribution to International Geological Correlation
Cerebrosphaera buckii Butterfield 1994 in Butterfield et al. 1994 Program 512: Neoproterozoic Ice Ages.
Chuaria sp. cf. Ch. circularis Walcott 1899; emend. Vidal and
Ford 1985
Coneosphaera sp. Luo 1991 References
Cymatiosphaeroides kullingii Knoll 1984, emend. Knoll et al.
1991 Ambrose, G. J., Flint, R. B. & Webb, A. W. 1981. Precambrian and
Eoentophysalis croxfordii (Muir 1976) Butterfield 1994 Paleozoic geology of the Peake and Denison Ranges. Geological
Eomicrocystis sp. cf. E. elegans Golovenok and Belova 1984 Survey of South Australia Bulletin 50.
Bagas, L., Grey, K. & Williams, I. R. 1995. Reappraisal of the Paterson
Leiosphaeridia crassa (Naumova 1949) Jankauskas 1989
Orogen and Savory Basin. Western Australia Geological Survey
Leiosphaeridia jacutica (Timofeev 1966) emend. Mikhailova and
Annual Review 1994–95, 55 – 63.
Jankauskas in Jankauskas et al. 1989 Bagas, L., Grey, K., Hocking, R. M. & Williams, I. R. 1999. Neoproter-
Leiosphaeridia minutissima (Naumova 1949) emend. Jankauskas ozoic successions of the northwestern Officer Basin: a reappraisal.
in Jankauskas et al. 1989 Western Australia Geological Survey Annual Review 1998– 99,
Leiosphaeridia sp. cf. L. exsculpta (Timofeev 1969) emend. 39 – 44.
Mikhailova in Jankauskas et al. 1989 Bagas, L., Camacho, A. & Nelson, D. R. 2002. Are the Neoproterozoic
Leiosphaeridia tenuissima Eisenack 1958 Lamil and Throssell Groups of the Paterson Orogen allochthonous?
Leiosphaeridia ternata (Timofeev 1966) emend. Mikhailova in Geological Survey of Western Australia Annual Review, 2000/01,
Jankauskas et al. (1989) 45 – 52.
Melanocyrillium Bloeser 1979 ex Bloeser 1985 Baudet, D., Aitken, J. D. & Vanguestaine, M. 1989. Palynology of
Myxococcoides cantabrigiensis Knoll 1982 uppermost Proterozoic and lowermost Cambrian Formations,
AUSTRALIAN CRYOGENIAN CORRELATIONS 131

central Mackenzie Mountains, northwestern Canada. Canadian Dunster, J. N., Kruse, P. D., Duffett, M. L. & Ambrose, G. J. 2007.
Journal of Earth Sciences, 26, 129–148. Geology and resource potential of the southern Georgina Basin.
Black, L. P., Calver, C. R., Seymour, D. B. & Reed, A. 2004. SHRIMP Northern Territory Geological Survey, Digital Information
U –Pb detrital zircon ages from Proterozoic and Early Palaeozoic Package, DIP007.
sandstones and their bearing on the early geological evolution of Eyles, C. H., Eyles, N. & Grey, K. 2007. Palaeoclimate implications
Tasmania. Australian Journal of Earth Sciences, 51, 885–900. from deep drilling of Neoproterozoic strata in the Officer Basin and
Blake, D. H., Tyler, I. M. & Sheppard, S. 1998. Geology of the Ruby Adelaide Rift Complex of Australia: a marine record of wet-based
Plains 1:100 000 sheet area (4460), Western Australia. Australian glaciers. Palaeogeography, Palaeoclimatology, Palaeoecology,
Geological Survey Organisation, Canberra. 248, 291–312.
Bristow, T. F. & Kennedy, M. J. 2008. Carbon isotope excursions and Fairchild, I. J., Spiro, B., Herrington, P. M. & Song, T. 2000. Controls
the oxidant budget of the Ediacaran atmosphere and ocean. on Sr and C isotope compositions of Neoproterozoic Sr-rich lime-
Geology, 36, 863–866. stones of East Greenland and North China. In: Grotzinger, J. P.
Butterfield, N. J. & Rainbird, R. H. 1998. Diverse organic-walled & James, N. P. (eds) Carbonate Sedimentation and Diagenesis in
fossils, including ‘possible dinoflagellates’, from the early Neoproter- the Evolving Precambrian World. SEPM Special Publication, 67,
ozoic of arctic Canada. Geology, 26, 963– 966. 297– 313.
Butterfield, N. J., Knoll, A. H. & Swett, K. 1994. Paleobiology of Fairchild, T. R. 1976. The geological setting and palaeobiology of a late
the Neoproterozoic Svanbergfjellet Formation, Spitsbergen. Fossils Precambrian stromatolitic microflora from South Australia. PhD
and Strata, 34, 84. thesis, University of California, USA.
Calver, C. R. 1995. Ediacarian Isotope Stratigraphy of Australia. PhD Fanning, C. M. 1986. 87Sr/86Sr of gypsum/anhydrite and carbonate
thesis, Macquarie University, Sydney, Australia. samples. Australian Mineral Development Laboratories Report G
Calver, C. R. 1998. Isotope stratigraphy of the Neoproterozoic Togari 6696/86 (unpublished).
Group, Tasmania. Australian Journal of Earth Sciences, 45, 865– 874. Fanning, C. M. & Link, P. K. 2004. U– Pb SHRIMP ages of the Neopro-
Calver, C. R. 2000. Isotope stratigraphy of the Ediacarian (Neoprotero- terozoic (Sturtian) glaciogenic Pocatello Formation, southeastern
zoic III) of the Adelaide Rift Complex, South Australia, and the Idaho. Geology, 32, 881–884.
overprint of water column stratification. Precambrian Research, Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian
100, 121– 150. Glaciation; data from the Adelaide Geosyncline, South Australia
Calver, C. R. & Lindsay, J. F. 1998. Ediacarian sequence and isotope and Pocatello Formation, Idaho, USA. In: Gallagher, S. J. &
stratigraphy of the Officer Basin, South Australia. Australian Wallace, M. W. (eds) Neoproterozoic Extreme Climates and the
Journal of Earth Sciences, 45, 513–532. Origin of Early Metazoan Life. Selwyn Symposium of the GSA
Calver, C. R. & Walter, M. R. 2000. The late Neoproterozoic Grassy Victoria Division, September 2008, Geological Society of Australia
Group of King Island, Tasmania: correlation and palaeogeographic Extended Abstracts, 91, 57– 62.
significance. Precambrian Research, 100, 299– 312. Fanning, C. M., Ludwig, K. R., Forbes, B. G. & Preiss, W. V. 1986.
Calver, C. R., Black, L. P., Everard, J. L. & Seymour, D. B. 2004. Single and multiple grain U– Pb zircon analyses for the early Adelai-
U –Pb zircon age constraints on late Neoproterozoic glaciation in dean Rook Tuff, Willouran Ranges, South Australia. Geological
Tasmania. Geology, 32, 893– 896. Society of Australia Abstracts, 15, 255– 304.
Close, D. F., Edgoose, C. J. & Scrimgeour, I. R. 2005. The Tjauwata Fensome, R. A., Williams, G. L., Barss, M. S., Freeman, J. M. & Hill,
Group: a late Mesoproterozoic rift succession underlying the south- J. M. 1990. Acritarchs and fossil prasinophytes: an index to genera,
western Amadeus Basin. Central Australian Basins Symposium: species and infraspecific taxa. American Association of Stratigraphic
Petroleum and Minerals Potential, 16 –18 August 2005, Northern Palynologists Foundation, Contributions Series 25, 771.
Territory Geological Survey, 39 – 40. GEOSCIENCE AUSTRALIA, Stratigraphic Names database. World
Coats, R. P. & Preiss, W. V. 1980. Stratigraphic and geochronological Wide Web address: http://dbforms.ga.gov.au/pls/www/geodx.
reinterpretation of Late Proterozoic glaciogenic sequences in the strat_units.int
Kimberley Region, Western Australia. Precambrian Research, 13, Glikson, A. Y., Stewart, A. J. et al. 1996. Geology of the western
181– 208. Musgrave Block, central Australia, with particular reference to the
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005. mafic– ultramafic Giles Complex. Australian Geological Survey
U –Pb ages from the Doushantuo Formation, China. Science, 308, Organisation Bulletin, 239, 206.
95 – 98. Gorter, J., Grey, K. & Hocking, R. 2007. The petroleum exploration
Corkeron, M. 2007. ‘Cap carbonates’ and Neoproterozoic glacigenic potential of the Australian Infracambrian (Ediacaran) of the
successions from the Kimberley region, north-west Australia. Sedi- Amadeus and Officer basins. Australian Petroleum Production and
mentology, 54, 871– 903. Exploration Association Journal, 47, 391–392 (extended abstract).
Corkeron, M. L. & George, A. D. 2001. Glacial incursion on a Neopro- Gradstein, F. M., Ogg, J. G., Smith, A. G., Bleeker, W. & Lourens,
terozoic carbonate platform in the Kimberley. Geological Society of A. J. 2004. A new geologic time scale, with special reference to
America Bulletin, 113, 1121– 1132. Precambrian and Neogene. Episodes, 27, 83 – 100.
Cotter, K. L. 1997. Neoproterozoic microfossils from the Officer Basin, Grey, K. 1995. Neoproterozoic stromatolites from the Skates Hills For-
Western Australia. Alcheringa, 21, 247–270. mation, Savory Basin, Western Australia, and a review of the distri-
Cotter, K. L. 1999. Microfossils from Neoproterozoic Supersequence 1 bution of Acaciella australica. Australian Journal of Earth Sciences,
of the Officer Basin, Western Australia. Alcheringa, 23, 63 – 86. 42, 123– 132.
Cowley, W. M. & Flint, R. B. 1993. Epicratonic igneous rocks and Grey, K. 1999a. Appendix 7: Proterozoic palynology of samples from
sediments. In: Drexel, J. F., Preiss, W. V. & Parker, A. J. (eds) Empress 1A. In: Stevens, M. K. & Apak, S. N. GSWA Empress 1
The Geology of South Australia, Volume 1, The Precambrian. and 1A well completion report, Yowalga Sub-basin, Officer Basin,
South Australian Geological Survey Bulletin, 54, 142– 147. Western Australia. Western Australia Geological Survey Record,
Dehler, C. M., Elrick, M., Bloch, J. D., Crossey, L. J., Karlstrom, K. 1999/4, 68 –69.
E. & Des Marais, D. J. 2005. High-resolution d13C stratigraphy of Grey, K. 1999b. Appendix 8: Proterozoic stromatolite biostratigraphy of
the Chuar Group (ca. 770– 742 Ma), Grand Canyon: Implications Empress 1A. In: Stevens, M. K. & Apak, S. N. GSWA Empress 1 and
for mid-Neoproterozoic climate change. Geological Society of 1A Well Completion Report, Yowalga Sub-basin, Officer Basin,
America Bulletin, 117, 32 –45. Western Australia. Western Australia Geological Survey Record,
Dow, D. B. & Gemuts, I. 1969. Geology of the Kimberley Region, 1999/4, 70 –72.
Western Australia: The East Kimberley. Geological Survey of Grey, K. 2005a. Preliminary report on the Proterozoic biostratigraphy,
Western Australia Bulletin, 120, 135. Lancer 1, Officer Basin, Western Australia. In: Mory, A. J. &
Drexel, J. F. 2009. Review of the Burra Mine Project, 1980– 2008 – a Haines, P. W. (eds) Lancer 1 Well Completion Report (interpretive
progress report. South Australia. Department of Primary Industries papers), Officer and Gunbarrel Basins, Western Australia. Western
and Resources Report Book, 2008/16, 76. Australia Geological Survey Record, 2005/4, 81.
132 K. GREY ET AL.

Grey, K. 2005b. Ediacaran palynology of Australia. Memoir of the Hill, A. C., Arouri, K., Gorjan, P. & Walter, M. R. 2000b. Geochem-
Association of Australasian Palaeontologists, 31, 439. istry of marine and non-marine environments of a Neoproterozoic
Grey, K. & Blake, D. 1999. Neoproterozoic (Cryogenian) stromatolites cratonic carbonate/evaporite: the Bitter Springs Formation, Central
from the Wolfe Basin, east Kimberley, Western Australia: correlation Australia. In: Grotzinger, J. P. & James, N. P. (eds) Carbonate
with the Centralian Superbasin. Australian Journal of Earth Sciences, Sedimentation and Diagenesis in the Evolving Precambrian. SEPM
46, 329– 341. Special Publication, 67, 327–344.
Grey, K. & Corkeron, M. 1998. Late Neoproterozoic stromatolites in Hill, A. C., Haines, P. W., Grey, K. & Willman, S. 2008. New records
glacigenic successions of the Kimberley region, Western Australia: of Ediacaran Acraman ejecta in drill holes from the Stuart Shelf and
evidence for a younger Marinoan glaciation. Precambrian Research, Officer Basin, South Australia. Meteoritics and Planetary Science,
92, 65 –87. 42, 1883– 1891.
Grey, K. & Cotter, K. L. 1996. Palynology in the search for Proterozoic Hill, A. C., Haines, P. W. & Grey, K. 2011. Neoproterozoic glacial
hydrocarbons. Western Australia Geological Survey Annual Review, deposits of central Australia. In: Arnaud, E., Halverson, G. P. &
1995– 96, 70– 80. Shields-Zhou, G. (eds) The Geological Record of Neopro-
Grey, K. & Stevens, M. K. 1997. Neoproterozoic palynomorphs of terozoic Glaciations. Geological Society, London, Memoirs, 36,
the Savory Sub-basin, Western Australia, and their relevance to 677– 691.
petroleum exploration. Western Australia Geological Survey Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
Annual Review, 1996– 97, 49– 54. 2004. U– Pb zircon date from the Neoproterozoic Ghaub
Grey, K., Hocking, R. M. et al. 2005. Lithostratigraphic nomenclature Formation, Namibia: constraints on Marinoan glaciation. Geology,
of the Officer Basin and correlative parts of the Paterson Orogen, 32, 817– 820.
Western Australia. Western Australia Geological Survey Report, Hoffman, P. & Schrag, D. 2002. The snowball Earth hypothesis: testing
93, 89. the limits of global change. Terra Nova, 14, 129–155.
Griffin, B. J. & Preiss, W. V. 1976. The significance and provenance of Hoffman, P., Hawkins, D., Isachsen, C. & Bowring, S. 1996.
stromatolitic clasts in a probable Late Precambrian diamictite Precise U–Pb zircon ages for early Damaran magmatism in the
in northwestern Tasmania. Proceedings of the Royal Society of Summas Mountains and Welwitschia Inlier, northern Damara belt,
Tasmania Papers, 110, 111–127. Namibia. Communications of the Geological Survey of Namibia,
Haines, P. W. 1988. Storm-dominated mixed carbonate/siliciclastic shelf 11, 47 – 52.
sequence displaying cycles of hummocky cross-stratification, Late Hoffman, P. F., Calver, C. R. & Halverson, G. P. 2009. Cottons
Proterozoic Wonoka Formation, South Australia. Sedimentary Breccia of King Island, Tasmania: glacial or non-glacial, Cryogenian
Geology, 58, 237–254. or Ediacaran? Precambrian Research, 172, 311– 322.
Haines, P. W. 1990. A late Proterozoic storm-dominated carbonate shelf Ireland, T. R., Flöttmann, T., Fanning, C. M., Gibson, G. M. &
sequence: the Wonoka Formation in the central and southern Flinders Preiss, W. V. 1998. Development of the Early Paleozoic Pacific
Ranges, South Australia. In: Jago, J. B. & Moore, P. S. (eds) The Margin of Gondwana from detrital zircon ages across the Delamerian
Evolution of a Late Precambrian –Early Palaeozoic Rift Complex: Orogen. Geology, 26, 243– 246.
Adelaide Geosyncline. Geological Society of Australia Special Jackson, M. J. & Muir, M. D. 1981. The Babbagoola Beds, Officer
Publication, 16, 177– 198. Basin, Western Australia: correlations, micropalaeontology, and
Haines, P. W., Mory, A. J., Stevens, M. K. & Ghori, K. A. R. 2004. implications for petroleum prospectivity. BMR Journal of Australian
GSWA Lancer 1 well completion report (basic data), Officer and Geology and Geophysics, 6, 81 – 93.
Gunbarrel Basins, Western Australia. Geological Survey of Jackson, M. J. & van de Graaff, W. J. E. 1981. Geology of the Officer
Western Australia Record, 2004/10, 39. Basin. Bureau of Mineral Resources Bulletin, 206, 1 – 102.
Haines, P. W., Hocking, R. M., Grey, K. & Stevens, M. K. 2008. Vines James, N., Narbonne, G. & Kyser, T. 2001. Late Neoproterozoic cap
1 revisited: are older Neoproterozoic glacial deposits preserved carbonates: Mackenzie Mountains, northwestern Canada: precipi-
in Western Australia? Australian Journal of Earth Sciences, 55, tation and global glaciation. Canadian Journal of Earth Science,
397– 406. 38, 1229– 1262.
Halverson, G. 2006. A Neoproterozoic chronology. In: Xiao, S. & Jefferson, C. & Parrish, R. 1989. Late Proterozoic stratigraphy, U/Pb
Kaufman, A. (eds) Neoproterozoic Geobiology and Paleobiology. zircon ages and rift tectonics, Mackenzie Mountains, northwestern
Topics in Geobiology 27. Springer, New York, 231–271. Canada. Canadian Journal of Earth Sciences, 26, 1784– 1801.
Halverson, G. P., Hoffman, P., Schrag, D. P., Maloof, A. C. & Karlstrom, K. E., Bowring, S. A. et al. 2000. Chuar Group of the
Rice, A. 2005. Towards a Neoproterozoic composite carbon Grand Canyon: Record of breakup of Rodinia, associated change in
isotope record. Geological Society of America Bulletin, 117, the global carbon cycle, and ecosystem expansion by 740 Ma.
1181– 1207. Geology, 28, 619–622.
Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A. Kaufman, A., Knoll, A. & Narbonne, G. 1997. Isotopes, ice ages and
2007a. Evolution of the 87Sr/86Sr composition of Neoproterozoic terminal Proterozoic Earth history. Proceedings of the National
seawater. Palaeogeography, Palaeoclimatology, Palaeoecology, Academy of Sciences, 95, 6600–6605.
256, 103– 129. Kendall, B., Creaser, R. A. & Selby, D. 2006. Re– Os geochronology
Halverson, G. P., Maloof, A. C., Schrag, D. P., Dudás, F. Ö. & of postglacial black shales in Australia: consequences for timing of
Hurtgen, M. 2007b. Stratigraphy and geochemistry of a ca. 800 the Sturtian glaciation. Geology, 34, 729–732.
Ma negative carbon isotope interval in northeastern Svalbard. Kendall, B., Creaser, R. A., Calver, C. R., Raub, T. D. & Evans,
Chemical Geology, 237, 23 –45. D. A. D. 2007. Neoproterozoic paleogeography, Rodinia breakup,
Harlan, S., Heaman, L., LeCheminant, A. & Premo, W. 2003. Gunbar- and Sturtian glaciation: constraints from Re– Os black shale ages
rel mafic magmatic event: a key 780 Ma time marker for Rodinia from southern Australia and northwestern Tasmania. Geological
plate reconstructions. Geology, 31, 1053–1056. Society of America Abstracts with Programs, 39, 335.
Hill, A. C. 2005. Stable isotope stratigraphy, GSWA Lancer 1, Officer Kendall, B. S., Creaser, R. A., Calver, C. R., Raub, T. D. & Evans,
Basin, Western Australia. In: Mory, A. J. & Haines, P. W. (eds) D. A. D. 2009. Correlation of Sturtian diamictite successions in
GSWA Lancer 1 Well Completion Report (Interpretive Papers), southern Australia and northwestern Tasmania by Re– Os black
Officer and Gunbarrel Basins, Western Australia. Geological shale geochronology and the ambiguity of ‘Sturtian’-type diamictite-
Survey of Western Australia Record, 2005/4, 1– 11. cap carbonate pairs as chronostratigraphic marker horizons. Precam-
Hill, A. C. & Walter, M. R. 2000. Mid-Neoproterozoic (830–750 Ma) brian Research, 172, 301–310.
isotope stratigraphy of Australia and global correlation. Precambrian Kennard, J. M., Nicoll, R. S. & Owen, M. 1986. Late Proterozoic and
Research, 100, 181– 211. Early Palaeozoic Depositional Facies of the Northern Amadeus
Hill, A. C., Cotter, K. L. & Grey, K. 2000a. Mid-Neoproterozoic Basin, Central Australia. 12th International Sedimentological
biostratigraphy and isotope stratigraphy in Australia. Precambrian Congress, Field Excursion 25B. Bureau of Mineral Resources,
Research, 100, 281– 298. Canberra, 125.
AUSTRALIAN CRYOGENIAN CORRELATIONS 133

Kennedy, M. J., Christie-Blick, N. & Prave, A. R. 2001. Carbon isoto- Australia: techniques and survey of preserved ultrastructure. Journal
pic composition of Neoproterozoic glacial carbonates as a test of of Paleontology, 50, 90 –106.
paleoceanographic models for snowball Earth phenomena. Parker, A. J., Rickwood, P. C. et al. 1987. Mafic dyke swarms of
Geology, 29, 1135– 1138. Australia. In: Halls, H. C. & Fahrig, W. F. (eds) Mafic Dyke
Knoll, A. H. 1996. Chapter 4: Archaean and Proterozoic palaeontology. Swarms. Geological Association of Canada, Special Papers, 34,
In: Jansonius, J. & McGregor, D. C. (eds) Palynology: Principles 401– 417.
and Applications. American Association of Stratigraphic Palynolo- Phillips, B. J., James, A. W. & Philip, G. M. 1985. The geology and
gists Foundation 1, Publishers Press, Salt Lake City, 51– 80. hydrocarbon potential of the northwestern Officer Basin. APEA
Knoll, A. H., Swett, K. & Mark, J. 1991. Paleobiology of a Neopro- Journal, 25, 52– 61.
terozoic tidal flat/lagoonal complex: the Draken Conglomerate Planavsky, N. & Grey, K. 2008. Stromatolite branching in the Neopro-
Formation, Spitsbergen. Journal of Paleontology, 65, 531–570. terozoic of the Centralian Superbasin, Australia: an example of shift-
Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blick, N. ing sedimentary and microbial control of stromatolite morphology.
2006. The Ediacaran Period: a new addition to the geologic time Geobiology, 6, 33– 45.
scale. Lethaia, 39, 13 – 30. Porter, S. M. & Knoll, A. H. 2000. Testate amoebae in the Neoproter-
Korsch, R. J. & Kennard, J. M. 1991. Geological and geophysical ozoic Era: evidence from vase-shaped microfossils in the Chuar
studies in the Amadeus Basin, central Australia. BMR Bulletin, 236, Group, Grand Canyon. Paleobiology, 26, 360–385.
511– 524. Preiss, W. V. 1972. The systematics of South Australian Precambrian and
Krylov, I. N. 1967. Rifeyskie i nizhne-kembriyskie stromatolity Tyan’- Cambrian stromatolites, part I. Transactions of the Royal Society of
Shanya i Karatau [Riphean and Lower Cambrian stromatolites of South Australia, 96, 67– 100.
Tien-Shan and Karatau]. Akademiya Nauk, SSSR, Trudy, Geologi- Preiss, W. V. 1973a. The Systematics of South Australian Precambrian
cheskii Institut Leningrad, 171, 76 (in Russian). and Cambrian Stromatolites, Part II. Transactions of the Royal
Krylov, I. N. 1975. Stromatolity Rifeya i Fanerozoya SSSR. [Riphean and Society of South Australia, 97, 91 –125.
Phanerozoic stromatolites in the USSR]. Akademiya Nauk SSSR, Preiss, W. V. 1973b. Early Willouran stromatolites from the Peake and
Trudy, Geologicheskii Institut, 274, 243 (in Russian). Denison Ranges and their stratigraphic significance. South Australian
Lindsay, J. F. 1993. Geological Atlas of the Amadeus Basin. Australian Department of Mines and Energy, Report Book, 73/208, 27.
Geological Survey Organisation, 25 plates. Preiss, W. V. 1974. The Systematics of South Australian Precambrian and
Lindsay, J. F. 1995. Geological Atlas of the Officer Basin South Australia. Cambrian Stromatolites, Part III. Transactions of the Royal Society of
Australian Geological Survey Organisation and Mines and Energy South Australia, 98, 185–208.
Department, South Australia, 30 plates. Preiss, W. V. 1976. Proterozoic stromatolites from the Nabberu and
Lindsay, J. F. & Reine, K. 1995. Well-log data, Officer Basin, South Officer Basins, Western Australia, and their stratigraphic signifi-
Australia. Australian Geological Survey Organisation Canberra cance. South Australia Geological Survey, Report of Investigations,
Record 1995/2, 13. 47, 51.
Macdonald, F. A., Schmitz, M. D. et al. 2010. Calibrating the Cryo- Preiss, W. V. 1985. Stratigraphy and tectonics of the Worumba Anticline
genian. Science, 327, 1241– 1243. and associated intrusive breccias. South Australia Geological Survey
McKirdy, D. M., Burgess, J. M. et al. 2001. A chemostratigraphic Bulletin, 52, 85.
overview of the late Cryogenian interglacial sequence in the Adelaide Preiss, W. V. 1987. The Adelaide Geosyncline –late Proterozoic strati-
Fold-Thrust Belt, South Australia. Precambrian Research, 106, graphy, sedimentation, paleontology and tectonics. South Australian
149– 186. Geological Survey Bulletin, 53, 438.
Meffre, S., Direen, N. G., Crawford, A. J. & Kamenetsky, V. S. Preiss, W. V. 1993. Neoproterozoic. In: Drexel, J. F., Preiss, W. V. &
2004. Mafic volcanic rocks on King Island, Tasmania: evidence for Parker, A. J. (eds) The Geology of South Australia, Volume 1, The
579 Ma break-up in east Gondwana. Precambrian Research, 135, Precambrian. South Australian Geological Survey Bulletin, 54,
177– 191. 171– 203.
Mendelson, C. V. & Schopf, J. W. 1992. Proterozoic and selected Early Preiss, W. V. 2000. The Adelaide geosyncline of South Australia and its
Cambrian microfossils and microfossil-like objects. In: Schopf, J. W. significance in Neoproterozoic continental reconstruction. Precam-
& Klein, C. (eds) Evolution of the Proterozoic Biosphere – A Multi- brian Research, 100, 21 –63.
disciplinary Study. Cambridge University Press, New York, Preiss, W. V., Walter, M. R., Coats, R. P. & Wells, A. T. 1978. Litho-
865– 895. logical correlations of Adelaidean glaciogenic rocks in parts of the
Morton, J. G. G. 1997. Lithostratigraphy and environments of depo- Amadeus, Ngalia and Georgina Basins. Bureau of Mineral Resources
sition. In: Morton, J. G. G. & Drexel, J. F. (eds) Petroleum Geology & Geophysics Australia Journal, 3, 43 –53.
Geology of South Australia, Volume 3: Officer Basin. South Preiss, W. V., Gostin, V. A., McKirdy, D. M., Ashley, P. M.,
Australian Department of Mines and Energy Resources Report Williams, G. E. & Schmidt, P. W. 2011. The glacial succession of
Book, 97/19, 47– 86. Sturtian age in South Australia – the Yudnamutana Subgroup. In:
Morton, J. G. G. & Drexel, J. F. (eds) 1997. Petroleum Geology of South Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The
Australia Volume 3: Officer Basin. South Australia, Department of Geological Record of Neoproterozoic Glaciations. Geological
Mines and Energy Resources Report Book, 97/19, 173. Society, London, Memoirs, 36, 701– 712.
Mory, A. J. & Haines, P. W. 2005. GSWA Lancer 1 well completion Raaben, M. E., Sinha, A. K. & Sharma, M. 2001. Precambrian Stroma-
report (interpretive papers), Officer and Gunbarrel Basins, Western tolites of India and Russia: A Catalogue of Type-Form-Genera.
Australia. Geological Survey of Western Australia Record, 2005/4, Birbal Sahni Institute of Palaeobotany, Lucknow, 125.
90. Rainbird, R. H., Jefferson, C. W. & Young, G. M. 1996. The early Neo-
Nagy, R. M. 2008. Microfossils from the Neoproterozoic Chuar Group, proterozoic sedimentary Succession B of northwestern Laurentia:
Grand Canyon, Arizona: taxonomy, paleoecological analysis and correlations and palaeogeographic significance. Geological Society
implications for life during the onset of Neoproterozoic glaciation. of America Bulletin, 108, 454–470.
PhD thesis, University of California, Santa Barbara. Reid, A. 2009. Appendix G: U– Pb zircon dating of the porphyry.
Narbonne, G., Kaufman, A. & Knoll, A. 1994. Integrated chemostra- Burra Mine. In: Drexel, J. F. (ed.) Review of the Burra Mine
tigraphy and biostratigraphy of the Windermere Supergroup, north- Project, 1980– 2008 – A Progress Report. South Australia. Depart-
west Canada: implications for Neoproterozoic correlations and the ment of Primary Industries and Resources Report Book, 2008/16,
early evolution of animals. Geological Society of America Bulletin, 68– 75.
106, 1281–1292. Reid, P. W. & Preiss, W. V. 1999. Parachilna map sheet (second edition).
Nelson, D. R. 2002. Compilation of geochronology data, 2001. Western South Australia Geological Survey, Geological Atlas 1:250 000
Australia Geological Survey Record, 2002/2, 282. Series, Sheet SH54-13.
Oehler, D. Z. 1976. Transmission electron microscopy of organic micro- Saito, Y., Tiba, T. & Matsubara, S. 1988. Precambrian and Cambrian
fossils from the Late Precambrian Bitter Springs Formation of cherts in northwestern Tasmania. Bulletin of the National Science
134 K. GREY ET AL.

Museum, Series C: Geology and Paleontology, National Science Walter, M. R., Veevers, J. J., Calver, C. R., Gorjan, P. & Hill, A. C.
Museum, Tokyo, 14, 59– 70. 2000. Dating the 840– 544 Ma Neoproterozoic interval by isotopes of
Sawaki, Y., Kawai, T. et al. 2010. 87Sr/86Sr chemostratigraphy of Neo- strontium, carbon and sulfur in seawater, and some interpretative
proterozoic Dalradian carbonates below the Port Askaig Glaciogenic models. Precambrian Research, 100, 371–433.
Formation, Scotland. Precambrian Research, 179, 150– 164. Wells, A. T. & Moss, J. F. 1983. The Ngalia Basin, Northern Territory:
Schopf, J. W. 1968. Microflora of the Bitter Springs Formation, Late Stratigraphy and structure. Australian Bureau of Mineral Resources
Precambrian, Central Australia. Journal of Paleontology, 42, Bulletin, 212, 88.
651– 688. Wells, A. T., Forman, D. J., Ranford, L. C. & Cook, P. J. 1970. Geology
Schopf, J. W. & Barghoorn, E. S. 1969. Microorganisms from the Late of the Amadeus Basin, central Australia. Bureau of Mineral
Precambrian of South Australia. Journal of Paleontology, 43, Resources Australia Report, 100.
111– 118. Williams, G. E. & Gostin, V. A. 2005. Acraman– Bunyeroo impact event
Sergeev, V. N. 1989. Microfossils from transitional Precambrian – (Ediacaran), South Australia, and environmental consequences:
Phanerozoic strata of central Asia. Himalayan Geology, 13, 269– 278. twenty-five years on. Australian Journal of Earth Sciences, 52,
Shields, G. A., Brasier, M. D., Stille, P. & Dorjnamjaa, D. 2002. 607– 620.
Factors contributing to high d13C values in Cryogenian limestones Williams, G. E., Jenkins, R. J. F. & Walter, M. R. 2007. No heliotrop-
of Western Mongolia. Earth & Planetary Science Letters, 196, ism in Neoproterozoic columnar stromatolite growth, Amadeus
99 –111. Basin, central Australia: geophysical implications. Palaeogeography,
Skotnicki, S. J., Hill, A. C., Walter, M. R. & Jenkins, R. 2008. Strati- Palaeoclimatology, Palaeoecology, 249, 80 – 89.
graphic relationships of Cryogenian strata disconformably overlying Williams, G. E., Gostin, V. A., McKirdy, D. M. & Preiss, W. V. 2008.
the Bitter Springs Formation, northeastern Amadeus Basin, Central The Elatina glaciation, late Cryogenian (Marinoan Epoch), South
Australia. Precambrian Research, 165, 243–259. Australia: sedimentary facies and palaeoenvironments. Precambrian
Smith, K. G. 1972. Stratigraphy of the Georgina Basin. Bureau of Mineral Research, 163, 307– 331.
Resources Australia Bulletin, 111, 156. Williams, I. R. 1992. Geology of the Savory Basin, Western Australia.
Stevens, M. K. & Apak, S. N. 1999. Empress 1 and 1A well completion Western Australia Geological Survey Bulletin, 141, 115.
report, Yowalga Sub-basin, Officer Basin, Western Australia. Williams, M., Crossey, L. J. et al. 2003. Dating sedimentary sequences:
Western Australia Geological Survey, Record, 1999/4, 110. in situ U/Th-Pb microprobe dating of early diagenetic monazite and
Stevens, M. K. & Grey, K. 1997. Skates Hills Formation and Tarcunyah Ar – Ar dating of marcasite nodules; case study from Neoproterozoic
Group, Officer Basin, carbonate cycles, stratigraphic position and black shales in the southwestern. U.S. Geological Society of America
hydrocarbon prospectivity. Western Australia Geological Survey Abstracts with Programs, 35, 595.
Annual Review for 1996–97, 55 – 60. Willman, S. & Moczydłowska, M. 2008. Ediacaran acritarch biota
Timofeev, B. V. 1966. Mikropaleofitologicheskoe issledovanie drevnikh from the Giles 1 drill hole, Officer Basin, Australia, and its potential
svit [Microphytological investigations of ancient formations]. Akade- for biostratigraphic correlation. Precambrian Research, 162,
miya Nauk SSSR, Isdatelskvo Nauka [Science Publishing House], 498– 530.
Moscow, 1 – 147. (Russian, published in English translation dated Willman, S., Moczydłowska, M. & Grey, K. 2006. Neoproterozoic
1974 by British Library – Lending Division, Yorkshire, England, (Ediacaran) diversification of acritarchs – a new record from the
214.) Murnaroo 1 drillcore, eastern Officer Basin, Australia. Review of
Townson, W. G. 1985. The subsurface geology of the western Officer Palaeobotany and Palynology, 139, 17 –39.
Basin – results of Shell’s 1980– 1984 petroleum exploration cam- Wingate, M. T. D., Campbell, I. H., Compston, W. & Gibson, G. M.
paign. APEA Journal, 25, 34 – 51. 1998. Ion microprobe U–Pb ages for Neoproterozoic basaltic mag-
Walter, M. R. 1972. Stromatolites and the biostratigraphy of the Austra- matism in south-central Australia and implications for the breakup
lian Precambrian and Cambrian. Palaeontological Association of Rodinia. Precambrian Research, 87, 135–159.
London, Special Papers in Palaeontology, 11, 190. Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003.
Walter, M. R. 1980. Adelaidean and Early Cambrian stratigraphy of the Systematic variations in C, O and Sr isotopes and elemental con-
southwestern Georgina Basin: correlation chart and explanatory centrations in Neoproterozoic carbonates in Namibia: implications
notes. Bureau of Mineral Resources Australia Report, 214, 21. for glacial to interglacial transition. Precambrian Research, 124,
Walter, M. R. & Gorter, J. D. 1994. The Neoproterozoic Centralian 69 – 85.
Superbasin in Western Australia: the Savory and Officer Basins. In: Zang, W. 1995. Early Neoproterozoic sequence stratigraphy and acritarch
Purcell, P. G. & Purcell, R. R. (eds) The Sedimentary Basins of biostratigraphy, eastern Officer Basin, South Australia. Precambrian
Western Australia. Petroleum Exploration Society of Australia, Research, 74, 119– 175.
West Australian Basins Symposium, Perth, WA, 1994, Proceedings, Zang, W. & Walter, M. R. 1992. Late Proterozoic and Early Cambrian
851– 864. microfossils and biostratigraphy, Amadeus Basin, central Australia.
Walter, M. R., Krylov, I. N. & Preiss, W. V. 1979. Stromatolites from Memoirs of the Association of Australasian Palaeontologists 12,
Adelaidean (Late Proterozoic) sequences in central and South 1 –132.
Australia. Alcheringa, 3, 287– 305. Zhao, J.-X. & McCulloch, M. T. 1993. Sm– Nd mineral isochron ages of
Walter, M. R., Grey, K., Williams, I. R. & Calver, C. R. 1994. Strati- Late Proterozoic dyke swarms in Australia: evidence for two distinc-
graphy of the Neoproterozoic to early Palaeozoic Savory Basin, tive events of mafic magmatism and crustal extension. Chemical
Western Australia, and correlation with the Amadeus and Officer Geology (Isotope Geoscience Section), 109, 341–354
Basins. Australian Journal of Earth Sciences, 41, 533– 546. Zhao, J.-X., McCulloch, M. T. & Korsch, R. J. 1994. Characterization
Walter, M. R., Veevers, J. J., Calver, C. R. & Grey, K. 1995. Neo- of a plume-related 800 Ma magmatic event and its implications for
proterozoic stratigraphy of the Centralian Superbasin, Australia. basin formation in central-southern Australia. Earth & Planetary
Precambrian Research, 73, 173–195. Science Letters, 121, 349– 367.
Chapter 9

A user’s guide to Neoproterozoic geochronology

DANIEL J. CONDON1* & SAMUEL A. BOWRING2


1
NERC Isotope Geoscience Laboratories, British Geological Survey, Keyworth NG12 5GS, UK
2
Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
*Corresponding author (e-mail: dcondon@bgs.ac.uk)

Abstract: Geochronology is essential for understanding Neoproterozoic Earth history. Here we review the types of rocks and minerals
that are used to date geologic events and the analytical protocols for the different radio-isotopic decay systems employed. We discuss the
limitations and potential of these methodologies for dating Neoproterozoic stratigraphy, highlighting the major sources and magnitudes
of uncertainties and the assumptions that often underpin them.

Crucial to understanding the nature and causes of glaciations, large directly to sedimentary chemical precipitates and organic-rich
perturbations of biogeochemical cycles and key evolutionary inno- components enriched in parent isotopes (U, Re, Lu) (see section
vations that took place during the Neoproterozoic is our ability ‘Whole-rock geochronometers (Re – Os, Lu –Hf, Pb/Pb)’).
to precisely correlate and sequence disparate stratigraphic
sections. Relative ages of events can be established within single
sections or by regional correlation using litho-, chemo- and/or Dating accessory minerals from volcanic rocks
bio-stratigraphy. However, relative chronologies do not allow for
testing of event synchroneity, robustness of correlations or calcu- Zircon (ZrSiO4) is a common accessory mineral in silicic volcanic
lation of rates and duration of change. At present, a major limit- rocks ranging from lavas to air-fall tuffs to volcaniclastic sedimen-
ation to our understanding of Neoproterozoic history is the tary rocks and is a nearly ubiquitous component of most clastic
dearth of precise and accurate radio-isotopic dates. However, the sedimentary rocks. The refractory and durable nature of zircon
dramatic increase in geochronological constraints over the past over a wide range of geological conditions means that it is likely
five years is cause for optimism. to retain its primary crystallization age even through subsequent
This chapter reviews the radio-isotopic techniques used for metamorphic events. Silicic air-fall tuffs are the most common vol-
dating of Neoproterozoic stratigraphy. It is aimed at the many geol- canic rocks in fossil-bearing sequences and are found in layers that
ogists, climate scientists, palaeobiologists and geophysicists who range in thickness from a millimetres to many metres. They are
use Neoproterozoic geochronology, especially those less familiar commonly preserved in marine settings. In most of these rocks
with the process of obtaining a date from a rock or mineral. We the primary volcanic ash has been altered, probably soon after
provide an outline of the strengths, weaknesses and limitations deposition, to clay minerals in a process that does not affect zircon.
of the different techniques, as well as critical evaluation of the What makes zircon ideal for U –Pb dating is the fact that,
assumptions that underpin the accuracy and precision of the calcu- because U has a similar charge and ionic radius to Zr, it substitutes
lated dates. We discuss specific examples where some of these readily into the zircon crystal structure (in modest amounts, typi-
assumptions are not valid, resulting in inaccurate dates assigned cally in the tens to hundreds of parts per million (ppm) range),
to certain stratigraphic levels. Overall, we emphasize the U –Pb whereas Pb has a different charge and larger ionic radius,
(zircon) geochronology of volcanic ash beds. However, we also leading to its effective exclusion from the crystal lattice. There-
review Re – Os, Lu –Hf and Pb/Pb geochronology and their appli- fore, at the time of crystallization (t0), there is effectively no Pb
cation to dating sedimentary rocks. Although the K – Ar and present in a crystal (although mineral and fluid inclusions may
40
Ar/39Ar methods are mainstays of Mesozoic and younger time contain Pb) and the present-day Pb is the direct product of in situ
scale calibration, there are few examples where they have been U decay since t0 (see section ‘Uranium – lead’ for further
successfully applied to dating eruption ages of Neoproterozoic details). An additional factor that makes zircon a robust chron-
volcanic rocks. This is most likely due to the susceptibility of ometer is its high closure temperature (.900 8C) for Pb diffusion
K-bearing minerals to alteration and open system behaviour (Cherniak & Watson 2003), or the temperature below which U and
related to metamorphism. Pb do not undergo significant thermally activated volume diffu-
sion. This means that zircons tend to preserve their primary ages,
even in volcanic rocks metamorphosed to amphibolites-facies
conditions.
What can be dated? The refractory and durable nature of zircon means that it is often
recycled through crustal processes of erosion, metamorphism and
The absolute age of Neoproterozoic sedimentary successions can magmatism, which, combined with its high closure temperature,
be determined via radio-isotopic dating of minerals separated means that it is possible to ‘inherit’ older zircon in newly
from volcanic rocks. These minerals, such as zircon, are highly formed igneous rocks. Silicic volcanic rocks are often the
enriched in the parent isotope (U in the case of zircon, see product of melting or assimilation of older zircon-bearing rocks
below) and can be dated via radio-isotopic decay schemes (such and often contain inherited zircon. This phenomenon is most com-
as U –Pb, Re – Os and K –Ar/40Ar/39Ar techniques). As these min- monly observed as older cores surrounded by younger, magmatic
erals crystallize (or become ‘closed’ to the loss of daughter zircon overgrowths. Such grains pose an analytical challenge, as
isotope) at the approximate time of the magmatic eruption, their the different domains should be analysed separately. This is best
age is assumed to approximate the depositional age of the volcanic achieved using either microbeam techniques (see section ‘U –Pb
rock. It is also possible in some cases to apply radio-isotopic dating microbeam techniques’), which employ the high spatial resolution

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 135– 149. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.9
136 D. J. CONDON & S. A. BOWRING

of focused ion beams or lasers to analyse different domains within closed-system behaviour, and it is common to use either a statisti-
a single crystal, or mechanical micro-sampling of single grains fol- cal measure of coherence (see section ‘Sources and types of uncer-
lowed by conventional analysis. In addition, silicic volcanic rocks tainty’) or pre-existing age constraints to aid in the interpretation.
often contain inherited whole grains that range in age from only However, although closed-system behaviour results in statistical
slightly older than the eruption to tens to hundreds of millions of coherence, the inverse is not always true (see section ‘Uncertain-
years older. These zircons are the refractory residue from assimila- ties as a result of geologic complexity’). Recent studies have
tion of older rocks in the magma or are incorporated during an employed combined textural and Sr isotope analyses of the carbon-
explosive eruption. A further complication arises when one con- ates as an independent proxy for disturbance during fluid flow
siders that the crystallization history of a magma is not necessarily events (Babinski et al. 2007). For a recent review of U –Pb
an instantaneous event and, in some circumstances (such as large- dating of carbonates see Rasbury & Cole (2009).
volume silicic eruptions), zircons (and other minerals) may crys-
tallize tens to hundreds of thousands of years prior to an eruption
Re– Os dating of organic-rich sediments. Both Re and Os are redox-
and/or over a protracted interval. Thus the zircons separated
sensitive metals with multiple oxidation states. Re decays to Os
from an air-fall tuff may record a range of dates from the eruption
with a half-life of c. 41 Ga (Table 9.1). Both metals can become
age to millions of years prior. In such situations, high-precision
concentrated within anoxic sediments via redox reactions near
single-grain analyses are required to isolate different populations
the sediment –water interface and also have been shown to be pre-
and assign an eruption age to the sample.
ferentially incorporated into the organic matter of shales (Creaser
It is also possible to date other U-bearing accessory phases from
et al. 2002). Following deposition, both metals are often relatively
volcanic ash beds such as monazite and titanite; however, these are
immobile and can behave as a closed system through burial, dia-
much less common in their occurrence than zircon. Monazite
genesis and low-grade metamorphism, allowing exploitation as a
[(LREE)PO4] is typical of peraluminous magmas and meta-
chronometer. In the past five years, the Re –Os geochronometer
morphic rocks, but it is also sometimes found in volcanic ash
has been applied to several organic-rich Neoproterozoic strati-
beds (e.g. Roden et al. 1990). Monazite has a similarly high
graphic intervals (Schaefer & Burgess 2003; Kendall et al. 2004,
closure temperature (.900 8C) to zircon and can incorporate
2006), as well as several Phanerozoic intervals (Cohen et al.
ThO2 at the percent level in addition to U into its crystal
1999; Creaser et al. 2002; Selby & Creaser 2005), and has given
lattice, making it ideally suited to U – Th –Pb dating. Titanite
new hope for dating thick sedimentary sequences devoid of volca-
(CaTiSiO5) is a common accessory phase in plutonic and meta-
nic rocks. For a recent review of Re – Os dating of organic-rich
morphic rocks, but somewhat less common in volcanic rocks.
sediments, see Kendall et al. (2009).
A potential advantage of titanite is that it has a lower closure temp-
erature than zircon and monazite (c. 650 8C) and thus does not
retain Pb until it cools below that temperature. Thus, in volcanic U–Pb (Pb/Pb) and Lu–Hf dating of phosphates. Phosphate minerals
rocks, pre-eruptive residence is not a limiting factor for its use in are enriched (relative to the fluid from which they precipitate) in
dating volcanic eruptions. However, titanite also incorporates rare earth elements (REE), Th and U, and therefore have potential
some initial Pb into its crystal structure that limits the precision for geochronology using the Lu – Hf, Th –Pb and U –Pb decay
of titanite U –Pb dates. schemes. Few studies have successfully exploited the Lu –Hf
and U – Pb system for the precise dating of sedimentary phosphates
(Barfod et al. 2002).
Dating chemical precipitates and organic residues In addition to dating ‘bulk’ phosphates, both monazite and xeno-
time (YPO4) are known to form during early diagenesis (Evans
Chemically precipitated rocks and organic residues from sedimen- et al. 2002; Rasmussen 2005). Both minerals have highly favour-
tary rocks can be exploited as chronometers using the isochron able U(Th)/Pb concentrations, but their occurrence as very small
approach, and calculated dates are interpreted as the time of pre- grains or overgrowths makes them difficult to isolate and
cipitation and/or early diagenesis. The most commonly dated analyse. Xenotime occurs as syntaxial overgrowths on zircons,
Neoproterozoic rocks using this approach include carbonates, but due to their small size (a few tens of micrometres) and textural
phosphates and organic-rich shales, as they contain high concen- complexity, an in situ isotopic technique with a spatial resolution
trations of the different parent nuclides (e.g. Re, U and Lu). of ,10 mm is required to successfully date xenotime; this has so
far only been achieved by ion microprobe. An alternative approach
U– Pb (Pb/Pb) dating of carbonates. At the time of formation, car- is the Th – U –total Pb electron microprobe technique (Williams
bonates incorporate U into their crystal lattice, typically with a et al. 2007). This technique has the disadvantage that it requires
concentration of several ppm (although in some cases this can high concentrations of Th, U and Pb for detection, limiting it to
be many tens of ppm), as well as (initial) Pb with concentrations phases such as monazite; however, it does have high spatial resol-
typically in the parts per billion (ppb) range, and therefore have ution (beam width down to ,5 mm) and therefore the ability to
the potential for U –Pb dating. Carbonates form in a variety of ter- analyse very small grains in their petrographic context, but with
restrial and marine environments, but most Neoproterozoic succes- uncertainties of several percent.
sions targeted for U –Pb dating are marine. Fluid-mediated
recrystallization of carbonates is a common process during burial
and diagenesis, and very few (if any) Neoproterozoic carbonates
Table 9.1. Radiometric decay systems used in geochronology (see text for
have escaped this process. The variable mobility of Pb and U
further details)
means that it is common for the two to become decoupled,
especially during recent exposure and weathering, in which case
Radioactive Radiogenic Half-life (years)
the U – Pb systematics are often unreliable. If the decoupling of
parent nuclide daughter nuclide
the U and Pb occurred recently, then the Pb/Pb systematics are
often relied upon for age information. Several studies have gener- 238
U 206
Pb 4.468  109
ated Pb/Pb isochrons from Neoproterozoic successions with vari- 235
U 207
Pb 7.038  108
able success (Babinski et al. 1999, 2007; Folling et al. 2000; 187
Re 187
Os 4.16  1010
Barfod et al. 2002; Chen et al. 2004), and most demonstrate evi- 176
Lu 176
Hf 3.71  1010
dence for disturbance of the Pb/Pb systematics during burial/ 87
Rb 86
Sr 4.944  1010
metamorphic events (Babinski et al. 1999; Folling et al. 2000). 40
K 40
Ar, 40Ca 1.25  109
A major issue related to Pb/Pb isochron dates is how to assess
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 137

Precipitation of xenotime also occurs during fluid and thermal


events, which introduces the potential for complexity through mul- ln (1+ 206Pb*/238U )

U
206Pb/238U date =

238
tiple episodes of growth (Rasmussen 2005). In contrast to xeno- λ238
950 Ma

Pb*/
time, diagenetic monazite tends to occur as impure nodules (up
to 2 mm in diameter) in shales. Studies of Palaeozoic diagenetic

206
monazite nodules have demonstrated the utility of the Pb/Pb and

ln (1+ 207Pb*/235U )
Th –Pb systems, although the U –Pb systematics have been per- 850 Ma
0.14
turbed, indicating recent U remobilization (Evans et al. 2002).

λ235
Neither (diagenetic) monazite or xenotime U – Pb geochronology
have been successfully applied to the dating of a Neoproterozoic
sedimentary succession, although recent Th –U –total Pb investi- 750 Ma
gation of monazite from the Adelaide Rift complex (Mahan 0.12
et al. 2010) does show potential.
650 Ma
In some cases the ‘whole-rock geochronometers’ may be useful
for obtaining temporal constraints on sedimentary successions
0.10 λ235t 235U
devoid of volcanic material. In general, best-case uncertainties 207Pb/206Pb date = e –1
associated with calculated isochron dates are in the 0.5 –1% λ238t 238U
e –1 today
level, comparable to those of U –Pb (zircon) microbeam dates. 207 235
Pb*/ U
Despite the relatively large uncertainties intrinsic to the dating of 0.9 1.1 1.3 1.5
these types of rocks, the Pb/Pb isochron method has been
applied and interpreted as an estimate of depositional ages (see Fig. 9.1. U–Pb (Wetherill) concordia diagram for the age range 542– 1000 Ma.
above). Problems associated with isochron geochronometers are The grey band represents the concordia curve plotted to reflect the uncertainties
centred around the lack of an independent check on open-system in the 235U and 238U decay constants.
behaviour.

zircon based on each individual decay scheme may be calculated


Maximum and minimum age constraints
and plotted on a concordia diagram (Fig. 9.1). On a conventional
(Wetherill) concordia diagram the X and Y axes are the ratio of
Not all sedimentary successions are amenable to direct dating via
radiogenic 207Pb to 235U and radiogenic 206Pb to 238U ratios,
radio-isotopic methods. In the absence of zircon-bearing volcanic
respectively, and the concordia curve represents the simultaneous
rocks, or chemical sediments for isochron dating, another approach
solution of the decay equations for a given age. A third date, the
is to obtain maximum age constraints by dating detrital zircons. 207
Pb/206Pb date, can be determined from only Pb isotopic
Detrital zircons in sedimentary rocks can range from being con-
measurements by knowing both the 235U and 238U decay constants
siderably older than the estimated depositional age they are con-
and the present-day 235U/238U ratio, which is assumed to be
tained within (often many hundreds of millions of years older) to
137.88 (Steiger & Jager 1977). Calculation of the U –Pb dates
close to the age of deposition, and in this latter case can provide
requires determination of the Pb* –U ratio (Pb* denotes radiogenic
useful constraints (Bingen et al. 2005). The youngest age deter-
Pb). See Figure 9.1 for age equations.
mined in a detrital population gives a maximum depositional age
The different half-lives of 238U and 235U (c. 4.5 and c. 0.7 Ga,
for the unit. Minimum age constraints can be provided by over-
respectively) mean that by the Neoproterozoic time, much
lying strata, which may contain age-diagnostic fossils, and/or
smaller amounts of 235U (relative to 238U) remained due to
cross-cutting igneous intrusions.
the higher decay rate, and smaller amounts 207Pb were produced
per increment of time relative to 206Pb. The advantage of two inde-
pendent chronometers in the same mineral is that it is possible to
Radio-isotopic geochronometers detect small amounts of open-system behaviour such as Pb loss
or the inheritance of older Pb. This is a major factor in our
A relatively small number of radioactive decay schemes are suit-
ability to make reliable, high-precision age determinations, as
able for dating Neoproterozoic rocks and are listed in Table 9.1.
we can evaluate whether a number of analyses represents a
Each of these is based upon the radioactive decay of a parent
single time of mineral growth. It has been known for several
nuclide to a stable daughter nuclide. Obtaining an accurate date
decades that zircons often show evidence of post-crystallization
using these decay systems requires that (i) the decay constant of
Pb loss. This has the effect of lowering the U –Pb ratios and the
the parent nuclide is accurately and precisely determined; (ii)
derived dates. The analyses plot below the concordia curve and
closed-system behaviour, which can be simply stated to mean
are termed normally ‘discordant’, with 207Pb/206Pb dates . 207
that the parent/daughter ratio has only changed by radioactive
Pb– 235U dates . 206Pb – 238U dates (Fig. 9.2). Silver & Duetsch
decay; and (iii) the initial daughter nuclide, if present, can be
(1963) demonstrated that Pb loss was correlated with radiation
precisely and accurately accounted for. In this section, we
damage. Since that time it has become widely appreciated that
outline the basic principles of the radio-isotopic geochronometers
Pb is not lost by thermally activated volume diffusion but rather
in Table 9.1, separating the U –Pb system applied to U-bearing
by fast-pathway diffusion from damaged parts of the crystal
accessory minerals from those that use an isochron approach
lattice. In order to minimize/eliminate the effects of post-
(Re – Os, Lu – Hf, Pb/Pb, etc.) applied to chemical precipitates
crystallisation Pb loss, it is possible to pre-treat zircons to prefer-
and organic residues.
entially remove domains that have lost Pb, leaving zircon that
has remained a closed system. One major approach involves phys-
Uranium – lead ically abrading away the exterior portions of the zircons (Krogh
1982a), based on the observation that the outer portions were
U –Pb geochronology is often regarded as the gold standard of richest in U and thus susceptible to radiation damage and Pb
geochronology because, unlike all other chronometers, it exploits loss. Krogh (Krogh 1982b) also demonstrated that the careful
two independent decay schemes, 235U to 207Pb and 238U selection of the least magnetic zircons (diamagnetic) often corre-
to 206Pb, and both the 238U and 235U decay constants are relatively sponded to those with lowest U contents, radiation damage, and
precise and accurate (Jaffey et al. 1971). Two separate dates for a least affected by Pb loss. These approaches were widely applied
138 D. J. CONDON & S. A. BOWRING

U Upper intercept or 207Pb/206Pb date: 720


238

713.65 ±0.70 (±4.5) Ma slope = e λt -1


Pb/
206

206Pb/238Udate:

D*/Dref
0.1165 711.65 ±0.58 (±1.4) Ma
t2

0.1155 705 t1
207Pb/235Udate:
712.35 ±0.83 (±1.9) Ma to
700
0.1145 Y-axis intercept = D0/Dref P/Dref
ss
-lo

695 Fig. 9.3. Schematic isochron diagram illustrating (i) the situation at t0 where
Pb

0.1135 different samples from the same stratigraphic interval record a spread in
parent(P)/daughter(Dref) ratios but a constant initial daughter isotopic
690 207 235 composition. D0 is the initial amount of daughter isotope and D* is the
0.975 0.99 1.00 1.01 1.02 Pb/ U
radiogenic daughter isotope (from decay and initial) such that the daughter
produced solely from decay (since t0) ¼ D* –D0.
Fig. 9.2. Schematic U–Pb concordia diagram illustrating the 206Pb– 238U,
207
Pb– 235U and 207Pb/206Pb dates that can be calculated. The data presented are
a subset of analyses from sample WM54 (Bowring et al. 2007). Uncertainties with a slope equal to e lt –1 (Fig. 9.3). A typical isochron is
are the 2s internal uncertainties, whereas those in parentheses are the 2s plotted P/Dref on the X axis and D*/Dref on the Y axis, where P
internal plus the systematic decay constant uncertainties: 0.11% for 238U and is the number of parent atoms in the sample, Dref is the number
0.14% for 235U (Jaffey et al. 1971). The grey band is the concordia line plotted of atoms of a stable reference isotope of the daughter element
to reflect the uncertainties in the U decay constants. Error ellipses (white) are and D* the total number of radiogenic daughter atoms plus
plotted with 2s internal uncertainties, and black error bars represent 2s internal initial atoms (D0) of the same isotope (amount of daughter atoms
plus systematic decay constant uncertainties. due to decay ¼ D* –D0) (Fig. 9.3). At the time of sample for-
mation (t0), all samples would define a horizontal line. However,
as the parent decays over time (t1, t2, etc.), each sample will
until the development of a new technique described as ‘chemical evolve along a slope of –1, and samples with higher initial
abrasion’ (CA-TIMS) (Mattinson 2005). This technique involves parent/daughter ratio (P/Dref) will be displaced most (Fig. 9.3),
annealing zircon grains at 800–900 8C followed by partial dissol- such that the isochron rotates to a positive slope. Assuming closed-
ution in hydrofluoric acid. This method effectively ‘mines out’ or system behaviour since the formation of the sample, a linear
preferentially dissolves the higher U parts of the zircon that have regression through the points allows calculation of the slope and
been damaged by radiation and are thus susceptible to Y-intercept from which the age and the initial isotopic composition
fast-pathway diffusion of Pb from the zircon crystal. In many of the daughter can be determined (Fig. 9.3).
cases these domains are irregular in shape and occur in the grain
interiors. This method seems to offer the promise of effective elim-
ination of open-system behaviour in most zircon. Microbeam tech- Analytical methodologies
niques (see section ‘U– Pb microbeam techniques’) have not
typically used pre-treatment techniques, as they assume that Pb The majority of age constraints for Neoproterozoic strata have
loss is restricted to the exterior portions of grains, which they been obtained by U –Pb zircon dating, so most of this section
attempt to avoid during the in situ analyses. In addition, grains is concerned with U –Pb analytical methods. The analytical
treated using this method are often extremely fragile and not amen- methods used for other chronometers have some similarities with
able to mounting and polishing. the ID-TIMS U –Pb method. The differences are discussed below.

Whole-rock geochronometers (Re – Os, Lu – Hf, Pb/Pb) U– Pb methodologies

When rocks or minerals do not have extreme enrichment in the There are two main approaches to U – Pb zircon geochronology: in
parent/daughter ratio but instead have a range, the isochron situ ‘microbeam’ techniques and isotope dilution thermal ioniz-
approach is often useful. The isochron method involves analyses ation mass spectrometry (ID-TIMS). The major difference is that
of multiple cogenetic samples (minerals or subsamples of a rock in ID-TIMS geochronology, zircon is dissolved and the U and
from an identical stratigraphic level) and is used for systems Pb separated from the other elements before analysis, but in
where initial daughter atoms are present and there is the possibility microbeam techniques the zircon is analysed using either a laser
of a range in the parent/daughter ratio. The Re –Os, Lu –Hf and that ablates the sample prior to injection into a plasma or a
U –Pb chronometers can be used to date carbonates, phosphates focused ion beam that sputters the sample and generates secondary
and organic-rich shales that incorporate both parent and daughter ions from a mineral mounted in epoxy and polished or a thin
isotopes at the time of their formation. Ideally, the isochron section.
approach allows determination of both an age and an initial isoto-
pic composition of the daughter element. U –Pb ID-TIMS. ID-TIMS analyses of zircon (either as multi-grain
For multiple cogenetic samples to preserve the time of system fractions, single grains or grain fragments) involve dissolution of
closure, samples must begin with (i) a homogeneous initial daugh- the zircon in the presence of tracer isotopes and are called
ter isotopic composition and (ii) a spread in the parent/daughter isotope dilution. Because the tracer is added before dissolution,
ratio such that, over time, different samples with different the tracer/sample isotope ratios stay constant, despite incomplete
parent/daughter ratios will evolve and will define a straight line recovery during chemistry and low ionization efficiency during
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 139

analysis. For U –Pb ID-TIMS analyses, the most common tracers SIMS and LA-ICP-MS techniques, the 207Pb/206Pb ratio is a
are 205Pb and 235U. 205Pb is an artificial isotope that does direct measurement; for SIMS mass-dependent fractionation
not occur in nature, whereas 235U occurs naturally, although an appears to be minimal and the measured ratio is commonly used,
enriched form is used as a tracer. Natural U in the zircon is whereas in LA-ICP-MS analyses mass-dependent fractionation is
assumed to have 238U/235U ¼ 137.88 (Steiger & Jager 1977), quantifiable and is corrected for, either using sample-standard
so the tracer can be used to determine the number of moles of bracketing and/or using a solution with known 205Tl/203Tl ratio
238
U and 235U in a sample. Following dissolution, the sample to correct for mass bias on Pb isotopic ratios. For further details
undergoes chemical purification using anion exchange chemistry of microbeam techniques see Ireland & Williams (2003) for a
that allows separation of Zr and REEs from Pb and U, and Pb review of SIMS U – Pb geochronology and Kosler & Sylvester
and U from one another. (2003) for a review of LA-ICP-MS geochronology.
Following purification, Pb and U are analysed separately by There is a tradeoff between the benefit of the high spatial
thermal ionization mass spectrometry, where the ratios of resolution provided by microbeam techniques and the precision
sample isotopes (204Pb, 206Pb, 238U, etc.) to the tracer isotopes of individual spot analyses. LA-ICP-MS and SIMS analyses are
(205Pb, 235U) can be measured. As the amount of tracer isotope approximately an order of magnitude less precise than ID-TIMS
added to the sample is known, the number of atoms of each (Ireland & Williams 2003; Kosler & Sylvester 2003). In situ tech-
naturally occurring isotope in the sample can be determined niques are without question essential tools for characterizing
based upon the measured ratio of sample isotope/tracer isotope complex (zoned) zircons from volcanic and metamorphic rocks
ratio. After corrections for mass-dependent isotope fractionation, and for characterizing detrital populations, which in some cases
and the minor contribution of common Pb and U from the can provide robust estimates of the maximum age of a sequence.
reagents, the tracer and labware, the sample 206Pb/207Pb,
206
Pb/238U and 207Pb/235U ratios can be determined and
206
Pb/207Pb, 206Pb/238U and 207Pb/235U dates calculated. See Isochron techniques
section ‘Calibrating tracers for isotope-dilution’ for further discus-
sion of the isotope dilution technique. Isochron techniques involve analysis of multiple samples assumed
Optimization of this technique means that it is now possible to be the same age, which have a spread in parent/daughter ratio
to date zircons with ,10 pg radiogenic Pb with a precision of and have remained closed systems. To maximize the probability
,0.1% on the U –Pb ratio for single-grain analyses. However, that the samples are the same age and have the same initial isotopic
it is a very labour-intensive technique, as each single U –Pb composition it is preferable to design an appropriate sampling
analysis takes several hours of mass spectrometry and chemical strategy. For example, Re – Os geochronology is often attempted
purification in an ultraclean laboratory environment, making it on laminated organic-rich shales, so it is essential to have multiple
time-consuming to develop high-n datasets. samples from the same restricted stratigraphic interval (e.g.
Kendall et al. 2004) to avoid samples that span an amount of
time much larger than uncertainties. In some cases, such as
U– Pb microbeam techniques working with core samples, this is not always possible (Schaefer
& Burgess 2003; Kendall et al. 2006). Samples that integrate a sig-
U –Pb geochronology by microbeam techniques has revolutio- nificant amount of time could introduce scatter if there is temporal
nized geochronology over the past two decades. The two major variation in the initial isotopic composition of the daughter
techniques are secondary ion mass spectrometry (SIMS), typified element, and/or the sample represents an amount of time that far
by the SHRIMP (sensitive high resolution ion microprobe) and exceeds measurement uncertainty, especially in condensed sec-
laser ablation inductively coupled plasma mass spectrometry tions. For U – Pb in carbonates, different cement domains may
(LA-ICP-MS). Both techniques (collectively termed ‘microbeam’ have very different U –Pb ratios, but is difficult to know a priori
techniques) offer high-spatial-resolution analyses using either a whether the cements are all exactly the same age and have the
focused ion beam to sputter a volume of zircon (SIMS) or a laser same initial ratios. In both cases, deviations from the assumption
that is used to vaporize a volume of zircon (LA-ICP-MS). of the same age and initial ratio lead to increased scatter and uncer-
Microbeam techniques allow in situ analysis of very small tainties in calculated dates.
volumes and thus high spatial resolution. A typical volume of Sample dissolution and purification techniques are similar to the
zircon analysed by an ion probe is cylindrical, 20 –30 mm in diam- procedures for U –Pb ID-TIMS. Before isotope ratio mass spec-
eter and several micrometres deep, with somewhat larger volumes trometry, samples undergo dissolution and chemical purification.
for LA-ICP-MS (Kosler & Sylvester 2003). In addition, the ana- For multi-element systems (such as Re –Os, U –Pb and Lu –Hf ),
lyses can be done relatively rapidly (many tens of analyses per isotopic tracers are added before dissolution for the isotope
day for LA-ICP-MS and tens for SIMS) and are amenable to dilution (see above), whereas for systems where only daughter
automation. Furthermore, these techniques also allow analysis of isotopes are measured (i.e. Pb/Pb), direct measurements of the
other isotopes/elements of interest (Hf, O, REEs) and can be isotope ratios are made. The isotopic composition is determined
made on the same zircon grains in close proximity to the volume via thermal ionization mass spectrometry, although it is also poss-
analysed for geochronology. ible to use solution-mode ICPMS for most elements.
Fundamental to the microbeam U – Pb zircon methods is use of a The accuracy and precision of isochron techniques is largely
primary standard against which the U – Pb ratio of the unknown controlled by having a sufficient spread in initial parent/daughter
zircon is calibrated. For SIMS techniques this calibration involves ratio, closed-system behaviour and an identical initial isotopic
analyses of standard zircon to develop a calibration curve for a composition for all samples. For precipitates such as carbonates
known U –Pb ratio (which is determined via ID-TIMS analyses) and phosphates there is often no significant detrital input;
and against which analyses of unknown zircons can be compared. however, this is not the case for organic-rich shales targeted for
This is achieved through analytical sessions where a standard Re –Os where there is potential for significant concentrations of
zircon is repeatedly analysed interspersed with analyses of initial Os from multiple sources. This has been demonstrated in
unknown zircons (this is termed sample-standard bracketing). several studies (Creaser et al. 2002; Kendall et al. 2004);
During the course of a single analytical session, the measured however, it is possible to limit the detrital Os contribution by
ratio/date of a standard can drift by several percent. For LA- selective dissolution of the organic component using a CrO3 –
ICP-MS the approach is somewhat similar in that sample-standard H2SO4 dissolution approach. Kendall et al. (2004) compared two
bracketing is applied to determine the inter-elemental fraction- dissolution methods (aqua regia v. CrO3 –H2SO4 dissolution) on
ation, which is then applied to the unknown zircons. In both greenschist facies organic-rich shale from the Old Fort Point
140 D. J. CONDON & S. A. BOWRING

Formation in Western Canada. Both dissolution techniques were These uncertainties are intrinsic to each analysis and represent
used on the same powders. However, the aqua regia method the minimum uncertainty that must be considered. Most of these
yielded scattered data and a resulting ‘isochron’ regression with sources of random uncertainty relate to the mass spectrometry
a mean square weighted deviation (MSWD) of 65 and a large measurements and our ability to measure and reproduce isotopic
age uncertainty (9%) in comparison to the CrO3 –H2SO4 dissol- ratios, such as corrections made for laboratory blank. Factors
ution method, which yielded an isochron with much less scatter such as the electronic noise of detectors place a theoretical limit
(MSWD ¼ 1.2) and a relatively low uncertainty (0.8% 2s) on the precision that can be achieved by detecting a certain
(Kendall et al. 2004). number of ions over a finite period of time (counting statistics).
However, for almost all geochronological applications, other
factors such as correction for mass-dependent fractionation that
Sources and types of uncertainty occurs during sample ionization, and correction for laboratory
blank and/or initial parent and daughter nuclide, dominate the
Without an accurate estimation of total uncertainty, the radio- analytical uncertainty budget, especially for small samples.
isotopic age of a given rock or mineral is of limited value. For The analytical uncertainties associated with U – Pb ID-TIMS
example, suppose a date of 618 Ma is reported for a detrital dates have decreased substantially over the past decade. This is
zircon from a unit that underlies a Marinoan-type glacial deposit. due in large part to a reduction in the common Pb levels introduced
If the date is relatively precise (and accurate), +2 Ma, then it in the laboratory (laboratory blank) and correction for mass frac-
could be inferred that the onset of glacial sediment accumulation tionation though use of a ‘double-spike’, where two tracer isotopes
post-dates deposition of the detrital component at 618 + 2 Ma. of the same element (202Pb/205Pb or 233U/235U, for example) are
On the other hand, if the 618 Ma date has an uncertainty of used for real-time mass fractionation correction, helping reduce
100 Ma then the detrital zircon could be as old as 718 Ma or the uncertainty in the mass-dependent fractionation. See Schmitz
young as 518 Ma, making it of limited use. Consider the case & Schoene (2007) and Bowring et al. (2006) for a more complete
where the 618 Ma is the 206Pb – 238U date with an uncertainty discussion of sources of uncertainty in U – Pb ID-TIMS analyses.
of 10 Ma, but the 207Pb/206Pb date is 650 + 400 Ma. Although In SIMS and LA-ICP-MS (in ‘dry’ mode) analyses, common Pb
the two dates overlap (and are therefore technically concordant), levels are intrinsically low and the uncertainty related to any cor-
the lack of precision on the 207Pb/206Pb date renders it impossible rection for common Pb is usually insignificant compared to the
to assess open-system behaviour (such as Pb loss or inheritance) uncertainty in the U –Pb normalization (see below). In ‘wet’
within the limits imposed by the precision of the 207Pb/206Pb mode, as is typical for LA-ICP-MS U –Pb analyses, the correction
date, so the accuracy of the 206Pb– 238U date is not known; if for Pb blank and isobaric interferences on the Pb and U peaks can
the zircon has lost Pb, the 206Pb– 238U date is a minimum date. be a significant source of uncertainty, requiring a correction to be
This latter example may seem extreme, but data of this type made (Horstwood et al. 2003).
have been published and are often uncritically cited. For
example, Ireland et al. (1998) published an extensive dataset of Microbeam U –Pb standardization. Microbeam U –Pb data are
SHRIMP U –Pb zircon dates on detrital zircons from the Kanman- acquired in analytical sessions where the analysis of unknown
too Group in Australia (it should be noted that it was not the inten- zircons are alternated with standards. The raw measured U –Pb
tion of this study to constrain the timing of sediment ratio of the standard varies or ‘drifts’ during an analytical
accumulation). The units sampled included the Marino Arkose, session due to slight changes in instrument parameters. Therefore
where 50 detrital zircons were analysed, the majority of which there is an uncertainty associated with the Pb–U standardization
were .1 Ga (n ¼ 48). Two grains yielded 206Pb– 238U dates of that must be considered. This uncertainty is on the order of 1%
649 + 17 and 655 + 17 Ma, which many researchers use to indi- (Stern & Amelin 2003), although the attributed magnitude is
cate that the Marino Arkose is c. 650 Ma (or younger) (Zhou et al. dependent on the frequency with which the standard is analysed.
2004; Halverson et al. 2005; Peterson et al. 2005). The 207Pb/206Pb There are differences of opinion on how the uncertainty related
dates associated with these two analyses are 470 + 440 Ma to the U –Pb standardization is factored into the total uncertainty of
and 666 + 307 Ma, respectively. Although it is possible that the a date. Some groups consider that the ‘standardization’ value is
two zircons are indeed c. 650 Ma, it is not legitimate to assume constant for a given session, so the uncertainty is systematic and
the 206Pb – 238U date is an accurate estimate of the age of the need only be considered when comparing data collected in differ-
zircons without considering both the 207Pb– 235U and 206Pb/207Pb ent analytical sessions. In this approach the session Pb– U uncer-
uncertainties. In practical terms these two detrital zircon dates tainty can be simply added to the weighted mean uncertainty in
provide no significant constraint on the timing of sedimentation. a manner analogous to the ID-TIMS tracer calibration uncertainty.
It is advisable for consumers of geochronological data to under- Conversely, other groups consider that the reproducibility of the
stand the various sources of error and when one must consider the standard is a reflection of the true external reproducibility of all
total uncertainty of a given date as opposed to its constituent parts. analyses and that it should be incorporated into individual analyses
Although the uncertainty of each date contains an internal/random of unknown zircon (Ireland & Williams 2003; Stern & Amelin
component in the total uncertainty, there are also components that 2003). Typically, larger uncertainties for each analysis result in a
are systematic, such as errors from the decay constants. When reduction of the MSWD for weighted mean dates from multiple
comparing ages determined by the same isotopic system these analyses, but has obvious implications for the identification of out-
can be ignored, offering a potential increase in resolving power. liers (see discussion in Ireland & Williams 2003). Information
In this section we review the different sources of uncertainties regarding the approach taken to the standardization uncertainty
and the assumptions that underlie the often quoted (or not) is often recorded in the footnotes to the data table or in the data
errors. For more detailed treatment of uncertainties in geochronol- repository, but can be crucial when trying to precisely sequence
ogy the following articles are recommended: Ireland & Williams rocks or calculate durations of events.
(2003), Stern & Amelin (2003), Schmitz & Schoene (2007), and Typical internal (2s) uncertainties for ID-TIMS 207Pb/206Pb
various papers by Ludwig (1980, 1991, 1998, 2003). and 206Pb – 238U zircon dates are c. 0.5 –0.2% and c. 0.1– 0.05%,
respectively, and for microbeam techniques internal (2s) uncer-
Random/internal uncertainties tainties on 207Pb/206Pb and 206Pb– 238U dates are c. 3 –5% and
c. 1 –2%, respectively. The difference is about an order of magni-
Random/internal uncertainties are those relating to the measure- tude, and for a 206Pb– 238U date of 600 Ma amounts to the differ-
ment of isotopic ratios of the sample, standards and blanks, and ence in uncertainty on a single date between 0.3 and 0.6 Ma for
are used in the derivation of errors of the radiogenic ratios. ID-TIMS and 6 and 12 Ma for microbeam U –Pb dates.
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 141

Systematic/external uncertainties respectively, where X is the analytical/internal uncertainty, Y is


the analytical uncertainty plus the systematic tracer calibration
Systematic uncertainties are those related to a parameter that has a uncertainty, and Z is the total uncertainty including X, Y and the
fixed value and is effectively constant for a suite of determinations, decay constant uncertainties. This permits use of the data with
and is usually determined by experiments. An example of such a the level of uncertainty that is appropriate to the problem being
parameter used in U –Pb geochronology with systematic errors addressed.
would include the 238U decay constant (Jaffey et al. 1971),
which will affect all 238U – 206Pb dates and thus can be ignored Age of primary standards for microbeam U –Pb dating. As discussed
when comparing any 238U derived dates. The level at which above, U –Pb microbeam techniques rely upon measurement of
these systematic errors have to be considered depends upon their the U – Pb ratio relative to standard minerals of known age. The
nature; if they are specific to a single laboratory (i.e. tracer cali- U –Pb dates of these minerals are determined via ID-TIMS ana-
bration) then the systematic uncertainty has to be considered lyses with typical total uncertainties of 0.1 –0.3% that should be
when comparing dates determined using a different tracer. Simi- propagated into the total uncertainty of the final U/Pb microbeam
larly, decay constant uncertainties affect all dates derived from date. Because the systematic uncertainty related to the age of
that decay constant equally, so they only need to be considered the primary standard is about an order of magnitude less than
when comparing dates derived from different decay constants the random errors related to the U – Pb determination of the
(i.e. 238U – 206Pb and Re –Os dates). They can otherwise be unknown mineral, it is often not considered significant.
ignored when calculating differences between dates based upon Intra- and intercrystal homogeneity is a fundamental require-
the same decay scheme. This distinction between random and ment of a zircon standard for microbeam U – Pb geochronology
systematic (or internal and external) types of uncertainty is impor- as the U – Pb ratio of the standard is considered invariant. Isotopic
tant, as in some circumstances where the same parameter (decay homogeneity is assessed by multiple ID-TIMS analyses on single
constant, lab tracer) is used for calculating a suite of dates the crystals and/or crystal fragments (Black et al. 2003; Schmitz et al.
effective uncertainty can be reduced. This results in increased tem- 2003) with variability assessed on a microgram scale. In contrast,
poral resolution at which synchronicity, rates and durations of microbeam techniques require standards that are homogeneous on
proxy record evolution can be assessed. the submicrometre scale. At present, zircon standards are either
chips of megacrysts (e.g. SL13 and 91500) or multicrystal
Decay constants. One source of systematic uncertainty that affects mineral separates from plutonic rocks (e.g. Temora, R33). In
all radio-isotopic dates is that related to uncertainty in the decay general the zircon standards are relatively homogeneous at the
constants (Table 9.1). Three approaches have been taken to deter- level that can be detected by either microgram ID-TIMS analyses
mine the decay constants (the probability that a given atom will or nanogram SIMS analyses. However there have been issues with
decay per unit of time) of the long-lived radionuclide: (i) direct at least one of the megacryst standards (SL13), which is hetero-
counting; (ii) in-growth; and (iii) geological comparison. Direct geneous at the micrometre-scale (Ireland & Williams 2003). The
counting involves the detection of alpha, beta or gamma activity fact that all zircon standards are natural means they are not
relative to the total number of radioactive atoms. In-growth perfect, as they are likely to be affected by zonation and/or Pb
relies upon the quantification of a decay product that is accumu- loss and/or other (matrix-related) differences. For most well-
lated from a quantity of high-purity parent nuclide over a well- characterized zircon standards this variation occurs well below
defined period of time. Geological comparison involves analyses the level of quantification/detection and is not significant.
of cogenetic rocks or minerals with multiple chronometers, assum-
ing that each chronometer should yield the same date. This Calibrating tracers for isotope-dilution. For isotopic analyses that
approach has the potential for relative intercalibration of the use the addition of isotopic tracers (isotope dilution), the accuracy
decay constants, but accurate intercalibration requires that at of the tracer calibration (isotopic composition and concentrations)
least one decay constant is accurate and known with some pre- has major control over the accuracy of the derived dates for a
cision (Mattinson 2000; Schoene et al. 2006). This is usually mineral or rock. Calibration of tracers is performed through admix-
assumed to be the 238U and 235U due to the precision with which ing the tracer with another solution of known isotopic composition
the decay constants have been determined (Jaffey et al. 1971) and, importantly, known purity. High-purity metals or salts (see
and the internal check provided by closed-system zircon analyses Selby et al. 2007 for details of a Re –Os tracer, Wasserburg
(Mattinson 2000; Schoene et al. 2006). et al. 1981 for Sm– Nd, Schoene et al. 2006 for details of a
The counting experiments of Jaffey et al. (1971) determined the U –Pb tracer calibration) are used as the basis of the gravimetric
238
U and 235U decay constants with uncertainties of 0.11% and reference solutions against which the concentration of the tracer
0.14%, respectively. These values have been adopted for use isotope can be determined. Therefore, the purity of the metal or
in geochronology (Steiger & Jager 1977). The 187Re and 176Lu salt, and the accuracy of the weighing prior to dissolution, controls
decay constants have been determined by both direct counting the precision and accuracy of the calibration. This total uncertainty
experiment and through geological comparison with the U –Pb is typically estimated at c. 0.1%. For multi-element tracers, the
system, and uncertainties are estimated at c. 0.4–0.5% (Scherer elemental (i.e. U –Pb) ratio is fixed, so the uncertainty in the
et al. 2001; Selby et al. 2007). tracer calibration is systematic and can be ignored for the practical
The incorporation of decay constant uncertainties is becoming purposes of age determinations generated using the same tracer.
increasingly important as the internal precision of dates is This is particularly useful when attempting to determine the
reduced and multiple geochronometers are being used to investi- relative time difference between samples such as determining sedi-
gate the same time intervals. The decay constant uncertainties ment accumulation rates (Bowring et al. 2007), or assessing syn-
for isochron dates are typically ,20% of the total uncertainty chroneity of events (Condon et al. 2005). At present, it is typical
budget; in contrast, the uncertainties in the U decay constants that each isotope laboratory has their own tracer. This means
are often .50% of the total uncertainty budget of U –Pb that the tracer uncertainty has to be considered when comparing
ID-TIMS dates (Fig. 9.2). The situation for the ID-TIMS U –Pb dates with other laboratories and other techniques. Recently, the
community is that 206Pb– 238U and 207Pb/206Pb dates often do U –Pb ID-TIMS community has made an effort to eliminate this
not overlap within analytical precision, and the U decay constant inter-laboratory uncertainty through the development and cali-
uncertainties must be considered (Ludwig 2000; Begemann et al. bration of a large amount of 205Pb– 233U – 235U tracer for commu-
2001; Schoene et al. 2006). As the ‘consumer’ often uses these nity use under the auspices of the EARTHTIME Initiative (Condon
dates interchangeably, we are now seeing 206Pb– 238U and et al. 2007). As analytical uncertainties are reduced it is becoming
207
Pb/206Pb age uncertainties presented as +X/Y/Z and +X/Z apparent that inter-laboratory (or inter-session) variations are
142 D. J. CONDON & S. A. BOWRING

comparable to, or greater than, internal precision. Although much in a magma chamber or the effects of very subtle open-system be-
of this can be eliminated by a common tracer solution for U –Pb haviour. Thus, high-precision analyses do not always transform
ID-TIMS, as discussed above, it is also desirable that along with into reduced uncertainties in calculated weighted mean dates.
analyses of unknown zircons, secondary standards (minerals or
synthetic solutions) are analysed in order to provide an accurate
assessment of long-term reproducibility and inter-laboratory Complex U –Pb zircon systematic. In the past decade, errors associ-
agreement. ated with ID-TIMS analyses have been reduced by almost an order
of magnitude. These reduced errors offer unprecedented precision,
but also expose geological complexity at the ,0.1% level, some-
Calculating an age from multiple dates times resulting in scatter that exceeds analytical uncertainties. It is
now common for a geochronologist to be faced with a population
A significant proportion of temporal constraints for Neoprotero- of zircon analyses that do not form a coherent cluster, and the
zoic strata are derived from U –Pb dates on zircons from volcanic crucial question is how to interpret the data to arrive at an eruption
rocks. The final reported date and associated uncertainty are often and/or depositional age. The advent of CA-TIMS pre-treatment
weighted mean dates derived from a number (n) of individual dates for the elimination of Pb loss has been extremely important, as it
on different zircons (or zircon sub-domains). This is the case for gives one confidence that in many cases Pb loss need not be con-
data acquired using both ID-TIMS and microbeam techniques. sidered as a cause of excess scatter. Furthermore, for Neoprotero-
The weighted mean weights each individual analysis (such as a zoic rocks, the concordia curve has a shallow enough slope, and the
207
single SIMS spot or single-grain ID-TIMS analyses) according Pb – 235U dates measured precisely enough to be able to evaluate
to its precision, so analyses with a low uncertainty contribute discordance at the per mil level. However, this is not the case for
more to the weighted mean than those with high uncertainty. microbeam U – Pb dates.
Importantly, the use of a weighted mean algorithm (or other aver- As outlined above, microbeam U – Pb dates on volcanic rocks
aging) is underpinned by the expectation of a single population rely upon the averaging of a relatively high-n dataset (10 –20) of
with normally distributed errors (i.e. there is no correlation relatively imprecise (c. 2– 4%) U –Pb determinations to get a
between precision of analyses and age). If the errors on the individ- weighted mean date with a precision of c. 1% or less. Underpinning
ual analyses are approximately equal (as is typical for microbeam these lower uncertainties is the assumption of a single population
U –Pb
pffiffiffi data), then the weighted mean uncertainty is proportional to with normally distributed errors. However, it is the low precision
1/ n. In this case, high-n datasets can be used to reduce the of each analysis combined with variability of the standard analyses
overall age uncertainty for data collected on a single population that bracket unknowns that often precludes the detection of subtle
with normally distributed errors. If the uncertainties on the individ- amounts of Pb loss or inheritance. Stated another way, if the
ual analyses are variable then the weighted mean uncertainty is amount of Pb loss or inheritance is less than the precision of a
controlled by the most precise analyses making the weighted single spot analysis, then it cannot be detected via normal statisti-
mean date less proportional to n. High-n datasets are critical for cal proxies (such as the MSWD), so the assumption of a normal
assessing analytical v. geological scatter; however, the real limit distribution may be invalid (Fig. 9.4). If Pb loss is the main
on the precision is the analytical uncertainty of single (spot or source of open-system behaviour, this will have the effect of
grain) analyses, as this controls our ability to resolve real variation lowering the 206Pb – 238U date on some analyses, as well as the
within a series of analyses due to ‘open-system’ behaviour. weighted mean 206Pb– 238U date. If unrecognized, the younger
A common measure of the ‘coherence’ of a dataset is a statistical dates could be misconstrued as the true crystallization age.
parameter called the MSWD (mean square of the weighted devi- To highlight the problem of assigning dates to samples from
ates; York 1966, 1967). A value of c. 1 indicates that the scatter multiple U –Pb analyses, we compare two published studies on
in the data can be explained by analytical uncertainties alone. zircons from the same ash bed from the lower Doushantuo For-
Values much less than 1 indicate that analytical uncertainties mation of China. Zhang et al. (2005) obtained a weighted-mean
206
have been overestimated. Values greater than 1 can indicate Pb – 238U age of 621 + 7 Ma (n ¼ 13, MSWD ¼ 1.13, does
either that the uncertainties have been underestimated or that not include the 0.48% error in U – Pb calibration) from concordant
another source of scatter, often called ‘geological’ scatter is U –Pb zircon analyses on their sample 04SC20A, which occurs c.
present. Furthermore, the actual MSWD value for which the 2 m above the cap carbonate in the Yangtze Gorge region. Condon
scatter of the data can be considered due to analytical factors et al. (2005) sampled the same ash bed (their sample YG04-2) and
alone is not restricted to a value of 1, but in fact varies according analysed the zircons by the ID-TIMS method. The ID-TIMS
to the number of data points in the calculation (Wendt & Carl dataset showed a spread in U –Pb dates that formed a linear
1991). So, to be 95% confident that the scatter of the data is due array below Concordia, indicating zero-age Pb loss. Three of the
to the analysis when n ¼ 5, an acceptable MSWD range would ten zircons analysed were ‘concordant’ and yield a weighted
be 0.2 –2.2, but for n ¼ 25 this would be 0.6– 1.5 (Wendt & Carl mean 206Pb– 238U date of 632.3 + 0.5 Ma (MSWD ¼ 1.15, does
1991). Although not often explicitly stated, an MSWD of 1 does not include the 0.1% error in U –Pb tracer calibration). The two
not necessarily mean there is a single (age) population. Rather, it dates are statistically different and do not overlap, even when esti-
indicates that if real (age) variation is present, it cannot be resolved mates of random and systematic error are considered; however,
within the precision of the individual analyses. It is not uncommon when compared at the precision of a single analysis, all the
to see both isochron and weighted mean dates reported with SHRIMP U –Pb dates overlap with the ID-TIMS dates (Fig. 9.4).
MSWDs much greater than 1. These dates likely have little geo- A likely interpretation of this dataset is that a number of SIMS ana-
logical significance and should not be used. lyses were located on domains that had lost a small amount of Pb,
resulting in a lower mean 206Pb– 238U date when all the analyses
are included. The higher precision of the ID-TIMS analyses
Uncertainties as a result of geologic complexity allows small amounts of Pb loss to be recognized and those data
points not included in calculating the final 206Pb– 238U date. A
Uncertainty as a result of geological complexity is the most diffi- third U –Pb date has been published for this same ash bed: Yin
cult to quantify. The most common cause of excess scatter is open- et al. (2005) obtained a weighted mean 206Pb – 238U age of
system behaviour resulting from either inheritance of older zircon 628.3 + 5.8 Ma (MSWD ¼ 0.86). This publication does not
and/or Pb loss. For U – Pb zircon analyses, reduced errors on include any details of the U –Pb data such as error in U –Pb nor-
single analyses often exposes fine-scale variability that may malization, but it is presumed this weighted mean uncertainty
reflect protracted or punctuated crystallization of zircon crystals does not include any systematic sources of uncertainty. This date
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 143

is indistinguishable from either the other SHRIMP date (Zhang


(a)
680 et al. 2005) or the ID-TIMS date (Condon et al. 2005).
As a second example of problems related to determining a
U
238

sample age from a population of U –Pb (zircon) dates we present


Pb/

new ID-TIMS data on zircon from another sample that was pre-
206

viously analysed by SHRIMP to derive a U – Pb date (Fig. 9.5).


Fanning and Link (Fanning & Link 2004) dated samples from the
640
0.104 Neoproterozoic Pocatello Formation, Idaho, using the SHRIMP
U –Pb method on zircon from a series of volcanic layers and
620 clasts within diamictite. We obtained new ID-TIMS U – Pb
0.100 zircon data on one of these samples, 06PL00, that had yielded a
weighted mean SHRIMP 206Pb/238U date of 708 + 8 Ma
600 (n ¼ 18, MSWD ¼ 1.7, includes additional 0.35% error in U/Pb
0.096 normalization). For a population of this size an MSWD of 1.7 is
at the upper limit of acceptability (Wendt & Carl 1991), possibly
indicating scatter due to non-analytical causes, that is real age vari-
0.092 ation in the population.
560
Six single zircon crystals have been analysed by ID-TIMS
207 235 (Table 9.2 & Fig. 9.5) at the Massachusetts Institute of Technology
0.5 0.6 0.8 0.9 Pb/ U
following the analytical procedure outlined in Bowring et al.
(b) (2007). Three zircons are from the same mineral separate used
for the SHRIMP study (06PL00) and three are from an attempted
U

recollection (36PL05) at the same stratigraphic level as sample


238

06PL00 in the Scout Mountain Member of the Pocatello For-


Pb/

635
mation, with the authors of the original study. One zircon
206

(00PL06 z2) is concordant at c. 705 Ma, whereas the remaining


five zircons form a population at c. 690 Ma. Within this population,
0.1030
four zircons give equivalent 206Pb– 238U dates with a weighted
mean of 687.4 + 0.6/1.3 Ma, and the fifth zircon (34PL05 z4)
0.1025 gives a similar 207Pb– 206Pb date but a slightly younger
206Pb– 208U date, which we interpret to reflect Pb loss.

This new ID-TIMS U – Pb date of 687.4 + 1.3 Ma (includes 1‰


625
tracer calibration error) is distinctly younger than the weighted
mean SHRIMP U –Pb date of 705 + 8 Ma (includes 0.35%
U –Pb calibration error) for the same sample and stratigraphic
level (Fanning & Link 2004). However, when we look at the
data in terms of single analysis data points, we can see that there
0.1010 are some similarities between the two datasets (Fig. 9.5): they
207 235
0.848 0.852 0.856 0.860 0.864 0.868
Pb/ U both contain zircon 206Pb– 238U dates of c. 705 Ma and c. 690 Ma.
The proportions of these different age populations is different,
with the ID-TIMS dataset being dominated by c. 690 Ma grains,
(c)
whereas this age population is subordinate in the SHRIMP
dataset, most likely reflecting the low-n nature of the ID-TIMS
660 dataset and perhaps differences in pre-selection of zircons prior
to analysis based upon morphology and so on. Importantly, the
relatively low precision of the SHRIMP U – Pb dates means that
640 it is not possible to discern different populations at the ,3%
level. The assumption of a single (age) population and resultant
weighted mean calculation results in a U –Pb date that is
620
c. 20 Ma older than the concordant ID-TIMS population, which
is considered the most robust estimate of the age of the rock.
These two case studies from the lower Doushantuo Formation,
China, and the Scout Mountain Member, Idaho, serve to illustrate
600
the problems related to the recognition of real scatter, be it due to
Pb loss and/or a mixed age population, at a resolution that is less
than the precision of single analysis. In such cases, exploiting
580 high-n datasets to derive weighted mean dates with lower uncer-
SHRIMP analyses (Zhang et al. 2005)
ID-TIMS analyses tainties can result in inaccurate age assignments for a given
(Condon et al. 2005) sample. This section is not meant to suggest all SHRIMP U –Pb
data are inaccurate but to demonstrate that beyond the level of a
Fig. 9.4. U –Pb zircon data on samples from the lower Doushantuo Formation
single analysis uncertainty (i.e. in the calculation of weighted
(2 m above the cap carbonate), Yangtze Gorge region, China. (a) U– Pb mean dates) assumptions are being made about the nature of the
concordia plot of SHRIMP U –Pb (zircon) dataset obtained on zircon from population. If these assumptions are valid then the weighted
(Zhang et al. 2005). (b) U– Pb concordia plot of ID-TIMS U–Pb (zircon) mean date and associated error will be accurate; however, if
dataset for the same ash bed (Condon et al. 2005). (c) Plot of single analysis these assumptions are not valid (i.e. it is not a single age population
206
Pb – 238U dates of both the SHRIMP and ID-TIMS datasets. The grey bar with normally distributed errors), then the calculated weighted
reflects the weighted mean SHRIMP U–Pb date (all points included) and mean date and associated error may be an inaccurate reflection
ID-TIMS date (three concordant analyses included). The error bar incorporates of the sample age. In situ techniques are without question essential
error in U–Pb calibration (SHRIMP and TIMS). tools for characterizing complex zircons from volcanic and
144 D. J. CONDON & S. A. BOWRING

metamorphic rocks and for rapid analysis of detrital populations.


Ideally, microbeam techniques can be used to rapidly characterize
a population of zircons by analysing a small volume of many
zircons, which could then be followed by conventional high-
precision geochronology of selected grains.

Linear arrays v. isochrons. As outlined above, a limitation of the


isochron approach is the lack of an independent check for open-
system behaviour, unlike the dual decay scheme of the U –Pb
system. Most studies use a combination of bracketing age con-
straints and/or a statistical measure of coherence (MSWD or
uncertainty) to assess whether the system has been perturbed;
however, precision and amount of scatter cannot be used as a
proxy for closed-system behaviour. For example, organic-rich
sediments from the Aralka Formation, Australia, have been ana-
lysed for Re –Os geochronology using both the aqua regia and
CrO3 – H2SO4 dissolution methods. Schaffer & Burgess (2003)
used the aqua regia dissolution method and obtained a three-point
isochron (samples integrated over 1.6 m stratigraphic thickness) of
592 + 14 Ma (MSWD  1). An expanded dataset collected over
10 m stratigraphic thickness yielded a nine-point regression and an
age of 623 + 18 Ma (MSWD ¼ 5.2). Subsequent Re –Os analyses
(on samples from a 2 m interval within the 10 m interval sampled
by Schaefer & Burgess (2003) using the CrO3 – H2SO4 dissolution
method yielded a 10-point isochron with an age of 657.2 + 5.4 Ma
(2s internal uncertainties, MSWD ¼ 1.2) (Kendall et al. 2006).
This difference is attributed to either a sampling and/or analytical
artefact(s) related to sample digestion and/or sample-spike equili-
bration (Kendall et al. 2006, 2009).
It is clear from the Aralka case study that one cannot use
coherence of a dataset as a means to assess accuracy, especially
with isochrons based upon low-n datasets (or sub-sets). The
bottom line is that a suite of samples with the same initial ratio
and a range of parent/daughter ratios that evolve in a closed
system can yield an isochron; however, a suite of samples that
define a linear array alone cannot be inferred to represent closed-
system behaviour, as simple mixing of two reservoirs can yield
linear arrays.

Dating the Neoproterozoic

Geochronology plays an important role in the independent inte-


gration of geographically disparate stratigraphic sections and
therefore plays a key role in addressing questions pertaining to
the correlation of glaciogenic intervals and their synthesis in
terms of number of glacial events, their duration and geographic
extent. Several recent studies have assessed the published age con-
straints for Neoproterozoic glaciations in order to address such
questions, and we briefly summarize the current state of play in
light of what we consider the key datasets that pertain to the
tempo of glaciation, the development of early animals and other
interesting events.

How many glacial intervals?

In the past ten years, the number and quality of geochronological


data have increased considerably and (at least) three distinct
glacial ‘intervals’ have been geochronologically delineated.
These temporally distinct glacial intervals are also broadly
Fig. 9.5. U– Pb zircon data on samples from the Scout Mountain Member,
distinct in terms of the physical characteristics of associated
Pocatello Formation, southern Idaho. (a) U– Pb concordia plot of SHRIMP
U–Pb (zircon) dataset obtained on sample 00PL00 (Fanning & Link 2004).
strata (Kennedy et al. 1998; Hoffman & Schrag 2002; Halverson
(b) New U –Pb concordia plot of ID-TIMS U–Pb (zircon) dataset on zircons
et al. 2005).
from the 00PL00 mineral separate and a recollection from the same
stratigraphic level (see Table 9.2). (c) Plot of single analysis 206Pb– 238U dates I. Ediacaran glaciations. Ediacaran-age (c. 635 to 542 Ma) glacial
of both the SHRIMP and ID-TIMS datasets. The grey bar reflects the weighted deposits are preserved on at least three palaeocontinents. The
mean SHRIMP U–Pb date (all points included) and ID-TIMS date (00PL06 z2a Gaskiers Formation on the eastern Avalonian Peninsula has been
and 34PL05 z4 excluded) and error incorporates error in U–Pb calibration. directly dated by U –Pb (zircon) ID-TIMS dates on ash beds that
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 145

Table 9.2. ID-TIMS U-Pb data on zircons from 06PL00 and 34PL05

Concentrations Age (Ma)


206 208 207 207 206 206 207 207
U Pb Th Pb* Pb Pb Pb Pb corr. Pb Pb Pb
Fractions1 (ppm)2 (ppm)2 U3 Pb*/Pbc4 204
Pb5 206
Pb6 206
Pb6 % err7 235
U6 % err7 238
U6 % err7 coef. 238
U8 235
U8 206
Pb8

34PL05 z19 55 7 0.69 33.5 1935.4 0.2160 0.0625 0.13 0.96940 0.18 0.11254 0.12 0.696 687.49 688.15 690.35
34PL05 z29 33 4 0.50 35.8 2174.6 0.1559 0.0626 0.24 0.97056 0.31 0.11253 0.19 0.655 687.44 688.75 693.04
34PL05 z49 67 9 1.07 74.3 3915.9 0.3331 0.0625 0.09 0.96463 0.12 0.11203 0.08 0.690 684.51 685.69 689.56
06PL00 z1 13 2 0.49 10.9 672.6 0.1541 0.0625 0.35 0.96970 0.48 0.11252 0.30 0.676 687.37 688.31 691.35
06PL00 z2 42 5 0.78 22.5 1278.0 0.2433 0.0630 0.13 1.00320 0.21 0.11551 0.16 0.779 704.68 705.43 707.82
06PL00 z3 52 7 0.78 18.4 1052.3 0.2426 0.0625 0.15 0.96939 0.25 0.11248 0.19 0.795 687.13 688.15 691.53

1
z1, z2, etc. are fractions composed of single grains. Grains were annealed and leached prior to dissolution. See Bowring et al. (2007) for analytical procedures.
2
Volumes estimated from images, absolute uncertainty is c. 50%.
3
Model Th–U ratio calculated from radiogenic 208Pb/206Pb ratio and 207Pb/206Pb age.
4
Ratio of radiogenic Pb (including 208Pb) to common Pb.
5
Measured ratio corrected for spike and fractionation only. Mass fractionation corrections were based on analysis of NBS-981 and NBS-983. Corrections of 0.25 + 0.04%/
amu (atomic mass unit) and 0.07 + 0.04%/amu were applied to single-collector Daly analyses and dynamic Faraday– Daly analyses, respectively.
6
Corrected for fractionation, spike and blank. All common Pb was assumed to be procedural blank.
7
Errors are 2s, propagated using the algorithms of Ludwig (1980).
8
Calculations are based on the decay constants of Jaffey et al. (1971).
9
34PL05 was collected within a few hundred metres of the original 06PL00 sample locality and at the same stratigraphic level within the Scout Mountain Member. See
Bowring et al. (2007) for details of the analytical methods.

occur below, within and above the glacial Gaskiers Formation, Formation is dated at c. 712 Ma (Bowring et al. 2007), although
constraining its age to c. 584 to 582 Ma with a maximum duration this glacial deposit, where dated, is not associated with a cap car-
of 2.6 Ma (Bowring et al. 2003). Similar-age (,600–542 Ma) bonate. U –Pb (zircon) SHRIMP dates from the Pocatello For-
glacial successions occur elsewhere on Avalonia (Thompson & mation, Idaho (Fanning & Link 2004), are commonly cited as
Bowring 2000), the NE margin of Laurentia (Condon & Prave age constraints for Sturtian glaciation. However, as we have out-
2000), Baltica (Bingen et al. 2005) and the Tarim Block, NW lined above, there are accuracy issues with some of these dates
China (Xu et al. 2009); however, the age of these units is less and, perhaps more importantly, their geological context has been
well constrained, making it impossible to assess the synchronicity questioned (Fanning & Link 2008). Similarly, commonly cited
of Ediacaran glaciation(s). age constraints from the Edwardsburg Formation preserved in
roof pendants of the Idaho Batholith (Lund et al. 2003) lack
robust sedimentological and stratigraphic constraint in order to
II. ‘Marinoan’ glaciations. Marinoan-type glacial deposits are con-
assess their relevance for the Sturtian glacial episode.
sidered those associated with cap carbonates with distinctive litho- Relevant age constraints from post-Sturtian and pre-Marinoan
logical and isotopic similarities (Kennedy et al. 1998; Hoffman glacial deposits come from both Australia and China. In Australia,
et al. 2007) and they occur on nearly every palaeocontinent. A a Re –Os date of 657.2 + 5.4 Ma (Kendall et al. 2006) for the basal
maximum age constraint for the Marinoan glaciation comes Aralka Formation provides a minimum constraint on the age of
from China and Australia. In China, the top of pre-glacial the underlying Areyonga glacial deposits in the Amadeus Basin,
Datangpo Formation is constrained by a U – Pb (zircon) SHRIMP which is correlated with the type-Sturtian glacial deposits
date of 654.5 + 3.8 Ma (Shihong et al. 2008). In Australia, a (Walter et al. 2000). In China, a U –Pb (zircon) ID-TIMS date of
Re –Os isochron age of 643 + 2.4 Ma (Kendall et al. 2006) was 663 + 4 Ma from the Datangpo Formation (Zhou et al. 2004)
obtained from black shale of the Tindelpina Member at the base directly overlies the glacial Tiesiao Formation and its associated
of the pre-Marinoan Tapley Hill Formation. Synglacial age con- Mn-rich cap carbonate, again providing a minimum age constraint
straints come from 635.5 + 1.2 Ma U– Pb (zircon) ID-TIMS on a Sturtian-type glacial deposit. Thus, Sturtian-type glacial
date from an ash bed within the basinal facies of the Ghaub For- deposits are constrained to the interval c. 746– 663 Ma.
mation (Hoffmann et al. 2004) and a similar 636.3 + 4.9 Ma However, the nature of the current dataset, with issues remaining
U –Pb (zircon) SHRIMP date from the base of the Nantuo For- about the glacial affinity/stratigraphic context of certain succes-
mation (Shihong et al. 2008). Termination of this glacial episode sions as well as those relating to the radio-isotopic dates them-
is constrained by an ash bed within the cap carbonate of the selves, means that it is not possible to infer anything about
Nantuo Tillite (Yangtze Platform), which has been dated at duration and/or synchronicity of the glacial deposits and, inferen-
635.2 + 0.6 Ma (Condon et al. 2005). Combined, these data indi- tially, the nature of the glacial episode(s) they reflect.
cate a duration of ,10 Ma and globally synchronous termination
of the Marinoan glaciation at c. 635 Ma. It should be noted that
there still remains a lack of consensus over the age of the actual IV. Pre-Sturtian? glaciations. Evidence for a pre-Sturtian glaciation
Marinoan glacial in its type section and inferentially the base of (c. 740 –760 Ma) comes from a small number of sections in
the Ediacaran Period. Namibia (Frimmel et al. 1996), Zambia (Key et al. 2001) and
NW China (Xu et al. 2009); however, in each case, either the
III. Sturtian glaciation(s). Sturtian-type glacial deposits are typi- exact relationship of the dates to the glacial unit is inferred and a
cally defined as being stratigraphically lower than Marinaon-type complicated local stratigraphy lessens the confidence that the
deposits and are commonly associated with dark, organic-rich dated unit directly constrains a glacial deposit (Frimmel et al.
cap carbonates and sedimentary ironstones. On the Congo 1996; Key et al. 2001) and/or the glacial nature of the dated unit
craton, the Chuos Formation overlies the Naauwpoort volcanics remains equivocal (Frimmel et al. 1996; Xu et al. 2009). As
dated at 746 + 2 Ma (Hoffman et al. 1996), providing a robust such, the community should remain circumspect about an older
maximum age constraint. In Oman, the glaciogenic Ghubrah Cryogenian glaciation until more equivocal evidence is presented.
146 D. J. CONDON & S. A. BOWRING

Dating the earliest fossils Other Neoproterozoic events

There is growing consensus on the development of Ediacaran-age In addition to the climatic and biological events/developments
organisms based upon the sequencing of key assemblages via that took place during the Neoproterozoic, there were also a
high-precision U –Pb dating. We briefly review the nature of, series of other events that took place, including impact events
and age constraints for, key assemblages/Fauna. and major isotope excursions. Age constraints for most of these
events are based upon correlation-based inference and as such
I. Avalon assemblage. The Avalon assemblage is preserved in we do not discuss them here. One event, the Shuram/Wonoka
relatively deep-water marine settings within the Avalon zone of carbon isotope excursion, is directly constrained by radio-isotopic
Newfoundland and England, and is characterized by rangemorphs, dating and is discussed further below.
frond- and bush-shaped colonies, which lacked mobility. The ear-
liest Ediacaran (macroscopic) fossil (Charnia, C. masoni, I. Shuram/Wonoka carbon isotope excursion. The Shuram/Wonoka
C. wardi), fronds some 2 m in length, occur in the Drook For- carbon isotope excursion (CIE) is a pronounced large-magnitude
mation (Narbonne & Gehling 2003) and are dated at c. 577 Ma (c. 15‰) d13C excursion that is recorded in Oman, Australia,
(Bowring et al. 2003), some 5 Ma after the end of the Gaskiers China, Siberia, Namibia and several other locations (Burns &
glaciation. This assemblage extends in range to c. 560 Ma in Matter 1993; Calver 2000; Condon et al. 2005; Melezhik
Newfoundland and England; however, this upper limit requires et al. 2005, 2008). Termination of this event has been dated at
more robust constraint. c. 550 Ma in the Doushantuo Formation, where d13C cross
from negative to positive (Condon et al. 2005). The onset (and
II. White Sea assemblage. The White Sea assemblage is preserved in inferentially the duration) has yet to be satisfactorily constrained,
a shallow marine setting in the Vendian sections of the White Sea with detrital zircons from the base of the CIE providing
and the Ediacaran Member of Australia (Narbonne 2005). This maximum age constraints dated at c. 620 Ma (Bowring et al.
assemblage includes fronds similar to those of the Avalon assem- 2007, 2009). The duration of this unique CIE is a matter of some
blage, but also includes segmented and discoid fossils (such as debate, with estimates varying from c. 50 Ma (Le Guerroue et al.
Kimberella) as well as horizontal traces indicating mobile bilater- 2006) to c. 5– 11 Ma (Bowring et al. 2007). Constraining
ians were present, marking the advent of muscularity recorded in the timing of onset of the Shuram/Wonoka CIE is critically
the fossil record. In the White Sea sections, this assemblage is con- important to assessing hypotheses that implicate this event in
strained to c. 555 Ma (Martin et al. 2000), probably extending to oxygenation of oceans and early animal evolution (Condon et al.
close to the base of the Cambrian. 2005; Fike et al. 2006).
III. Nama assemblage. Nama assemblage fossils, typified by those
of the Nama Group, Namibia, include rangemorphs (mainly multi- Future directions
frondate fronds) and horizontal traces similar to the White Sea
assemblage, with an age c. 550 Ma to the base of the Cambrian. Clearly, research on the Neoproterozoic Earth System would
An important addition to the Nama assemblage are the weakly greatly benefit from more and, perhaps more importantly, better
calcified metazoans, Cloudina and Namacalathus, representing a dates. The number of possible dates is limited by the identification
major evolutionary development (calcification) and providing a of suitable sections/materials for analyses: ash bed bearing succes-
record of predation (Hua et al. 2003). These weakly calcified sions for U– Pb (zircon) dating, organic-rich sediments for Re – Os
metazoans occur in many terminal Ediacaran successions analyses, and so on. There is undoubtedly a wealth of material to be
(Oman, Namibia, China and Canada), and their age range is con- analysed from intervals of interest. Timely identification of targets
strained from c. 550 Ma to the base of the Cambrian (Grotzinger among the background of similar-looking material, which will not
et al. 1995; Amthor et al. 2003; Condon et al. 2005). yield meaningful age constraints, requires a change in approach,
with the development of better screening protocols, in order to
IV. Doushantuo fauna. The Doushantuo fauna comprise phospha-
maximize efficiency
tized eggs and embryos (Xiao et al. 1998, 2000), as well as micro-
From the U –Pb zircon perspective we propose there is great
scopic sponges (Li et al. 1998), and thus provides critical
utility in breaking down the microbeam v. ID-TIMS barrier. We
constraint upon the earliest animals. The age of the Doushantuo
have attempted to highlight in this chapter that these two analytical
Formation is constrained to c. 635 –550 Ma (Condon et al.
methods have complementary attributes: the microbeam U –Pb
2005). These fossils occur in the middle of the formation and are
techniques are capable of high sample throughput, a factor of
therefore likely c. 590 to 570 Ma, although this remains to be con-
c. 10 –20 greater than U –Pb ID-TIMS, whereas the ID-TIMS
strained by direct dating. The Miaohe biota of the upper Doushan-
U –Pb analyses are c. 10– 20 times more precise. Sample through-
tuo Formation contains diverse macroalgal assemblages (Xiao
put for ID-TIMS analyses is an issue, and what may not be appreci-
et al. 2002). Although not part of the White Sea assemblage, the
ated by the non-practitioner is that for every dated sample that
Miaohe biota is similar in age, being slightly older than the
makes it into the literature, there will have been many more that
550.1 + 0.6 Ma ash from the top of the Doushantuo Formation
have been analysed and yielded no meaningful chronological
(Condon et al. 2005), perhaps indicating a temporal coincidence
information, most likely due to a predominance of inherited/xeno-
between the development of macroalgae and bilaterians (White
crystic material. An optimized approach for an enhanced U –Pb
Sea assemblage, see above).
(zircon) based chronology for the Neoproterozoic would involve
Recently Love et al. (2009) documented an abundance of a a first-phase microbeam U – Pb dating, where samples are effec-
Demospongiae-specific biomarker from Cryogenian (c. 710– tively ‘screened’, prior to a second phase involving ID-TIMS
635 Ma) sedimentary rocks in Oman. These data indicate a U –Pb dating of samples that yield microbeam U –Pb dates of inter-
pre-Ediacaran age for the advent of multicellularity. est. Where samples contain mixed-age zircon populations it would
Finally, the disappearance of these Ediacaran fossil assemblages be possible to either remove zircons from a resin mount of
(Rangemorphs, weakly calcified metazoans) is coincident with ID-TIMS analyses or, preferably, to simply mount the zircons on
the base of the Cambrian and a negative d13C isotope excursion double-sided sticky tape, making removal for ID-TIMS analysis
(Grotzinger et al. 1995; Amthor et al. 2003). The base of the Cam- straightforward. This approach is not suitable for SIMS U – Pb ana-
brian is dated at c. 541 Ma (Amthor et al. 2003; Bowring et al. lyses, but is easily adaptable for LA-ICP-MS analyses. Combining
2007). The Early Cambrian fossil record documents a period of the efficiency of the microbeam techniques with the precision of
enhanced biomineralization and increased muscular development. ID-TIMS would be the most effective plan of attack for improving
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 147

the chronology of Neoproterozoic strata, and this would require a considered. An additional systematic uncertainty arises from
collaborative approach between the various groups involved. decay constant uncertainties. These uncertainties have been deter-
mined experimentally or assessed via geological comparison with
another decay scheme.
Conclusions There is considerable variation in the published literature
regarding the treatment of the constituent parts of the total uncer-
There are a number of radio-isotopic dating techniques tainty. In some cases, such as when using ID-TIMS 206Pb– 238U
that can be used to constrain the age of Neoproterozoic dates generated using a single isotopic tracer solution, certain
components can be ignored (in this case tracer calibration uncer-
sedimentary successions tainty and 238U decay constant uncertainty). This can be useful
when attempting to determine sediment accumulation rates
These include U –Pb dating of minerals (predominantly zircon) (Bowring et al. 2007) or assess the synchronicity of events
from interlayered volcanic rocks, and Re – Os or Pb/Pb isochron (Condon et al. 2005).
dating of organic-rich sediments or carbonates. Most of the Neo- The constituent of the total uncertainty that is most difficult to
proterozoic time scale is calibrated by U –Pb zircon dates from quantify is that related to the geology. This can be considered at
air-fall tuffs, but this dataset is augmented by a growing number the scale of relating a given mineral to a stratigraphic level (i.e.
of Re –Os and Pb/Pb isochron dates. do the magmatic zircons date the exact age of a given air-fall
Each of the radio-isotopic systems has different strengths and tuff?). However, a first-order uncertainty is often the relationship
weaknesses. U –Pb (zircon) dating is considered the premier geo- of a dated unit to the section of interest. Regardless of how
chronometer; however, it is of limited use in successions devoid of robust the radio-isotopic data are, if the geological context of a
zircon-bearing volcanics. Realizing the goal of a robust and highly dated sample is poorly constrained, the date is of little use.
resolved temporal framework for the Neoproterozoic will require
exploitation of all these different methodologies and refined
searches for volcanic air fall deposits. A robust, highly resolved chronology for the Neoproterozoic
is yet to be fully realized

All dates involve some subjective interpretation Although there has been a marked increase in the number of radio-
isotopic age constraints for Neoproterozoic sedimentary succes-
Dates interpreted as depositional ages are usually based on mul- sions in the past decade, there is still much we do not know, limit-
tiple analyses from the same stratigraphic horizon. In the case of ing our ability to test correlations and rate dependent hypotheses.
U –Pb (zircon) dates from volcanic rocks, the reported age is This is in part due to the lack of suitable material to date in key sec-
usually based on a weighted mean 206Pb – 238U date in which a vari- tions, although it is also due to the reproducibility of some of the
able number of single 206Pb– 238U dates (either single-grain or radio-isotopic techniques and systems employed. Accelerated pro-
single-spot analyses) are weighted (based upon their associated gress could be made through increased collaboration within the
uncertainty) and a mean calculated. Underpinning this mean date, community, so the most appropriate and efficient methodologies
and its lower uncertainty, is the assumption of a single population are applied to the intervals of interest.
with normally distributed errors, an assumption that is not always
valid. Isochron dates are similarly based upon the linear regression We thank E. Arnaud, M. Babinbski, J. Etienne, M. Horstwood, S. Kamo, B.
of a number of data points that are assumed to be cogenetic, have a Kendall, A. Maloof and T. Prave for constructive reviews. This chapter is an
common initial daughter isotopic composition, and the samples outgrowth of a continuing research programme supported by the NSF, the
analysed have acted as a ‘closed system’ since their formation. NASA Astrobiology Program and NERC. This represents a contribution of the
Although most attempt to produce age constraints with the IUGS- and UNESCO-funded IGCP (International Geoscience Programme)
lowest possible uncertainty, the best measure of high-precision Project #512.
age constraints is the precision of a single analysis. It is this pre-
cision that controls the evaluation of ‘open-system’ behaviour References
and thus the accuracy of the final date. Although the coherence
of a dataset is often used as a proxy for closed-system behaviour, Amthor, J. E., Grotzinger, J. P., Schroder, S., Bowring, S. A.,
the coherence is limited by the precision of the individual analyses. Ramezani, J., Martin, M. W. & Matter, A. 2003. Extinction of
Cloudina and Namacalathus at the Precambrian – Cambrian boundary
in Oman. Geology, 31, 431– 434.
The uncertainty of a date is as important as the date itself Babinski, M., Van Schmus, W. R. & Chemale, F. 1999. Pb/Pb dating
(Ludwig 2003) and Pb isotope geochemistry of Neoproterozoic carbonate rocks
from the Sao Francisco basin, Brazil: implications for the mobility
The total uncertainty of a radio-isotopic date comprises random (or of Pb isotopes during tectonism and metamorphism. Chemical
internal) and systematic (or external) components. The random/ Geology, 160, 175– 199.
internal uncertainties are related to the measurement of the isotopic Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating of
ratios and the corrections applied. Systematic uncertainties are the Sete Lagoas cap carbonate (Bambui Group, Brazil) and impli-
cations for the Neoproterozoic glacial events. Terra Nova, 19,
those related to the uncertainty in absolute value of various con-
401– 406.
stant parameters used in the calculation of either an isotopic
Barfod, G. H., Albarede, F., Knoll, A. H., Xiao, S. H., Telouk, P.,
ratio or in the calculation of the date itself. Analytical uncertainties Frei, R. & Baker, J. 2002. New Lu– Hf and Pb/Pb age constraints
should reflect the ability to reproduce a given isotopic ratio and on the earliest animal fossils. Earth and Planetary Science Letters.
represent the minimum uncertainty that may be considered. For 201, 203–212.
microbeam U –Pb dates, the standard calibration is best considered Begemann, F., Ludwig, K. R. et al. 2001, Call for an improved set of
as a non-systematic uncertainty and should be incorporated into decay constants for geochronological use. Geochimica et Cosmochi-
each individual spot U –Pb date uncertainty (Ireland & Williams mica Acta, 65, 111– 121.
2003). If comparing dates generated using different techniques Bingen, B., Griffin, W. L., Torsvik, T. H. & Saeed, A. 2005. Timing of
or using different calibration materials (such as mineral standards Late Neoproterozoic glaciation on Baltica constrained by detrital
for microbeam dates or isotopic tracer for ID-TIMS), then the zircon geochronology in the Hedmark Group, south-east Norway.
systematic uncertainties related to these calibrations must be Terra Nova, 17, 250–258.
148 D. J. CONDON & S. A. BOWRING

Black, L. P., Kamo, S. L., Allen, C. M., Aleinikoff, J. N., Davis, Grotzinger, J. P., Bowring, S. A., Saylor, B. Z. & Kaufman, A. J.
D. W., Korsch, R. J. & Foudoulis, C. 2003. TEMORA 1: a new 1995. Biostratigraphic and geochronological constraints on early
zircon standard for Phanerozoic U– Pb geochronology. Chemical animal evolution. Science, 270, 598–604.
Geology, 200, 155–170. Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
Bowring, S., Myrow, P., Landing, E. & Ramezani, J. 2003. Geochro- Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
nological constraints on terminal Neoproterozoic events and the rise isotope record. Geological Society of America Bulletin, 117,
of metazoans. Geophysical Research Abstracts, 5, 219. 1181– 1207.
Bowring, S. A., Schoene, B., Crowley, J. L., Ramezani, J. & Condon, Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
D. J. 2006. High-precision U –Pb zircon geochronology and the testing the limits of global change. Terra Nova, 14, 129–155.
stratigraphic record: progress and promise. The Paleontological Hoffman, P. F., Hawkins, D. P., Isachsen, C. E. & Bowring, S. A. 1996.
Society Papers, 12, 25– 46. Precise U–Pb zircon ages for early Damaran magmatism in the
Bowring, S. A., Grotzinger, J. P., Condon, D. J., Ramezani, J., Summas Mountains and Welwitschia Inlier, northern Damara Belt.
Newall, M. & Allen, P. A. 2007. Geochronologic constraints of Communications of the Geological Survey of Namibia, 11, 47– 53.
the chronostratigraphic framework of the Neoproterozoic Huqf Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
Supergroup, Sultanate of Oman. American Journal of Science, 307, Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635
1097– 1145. Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary
Bowring, S. A., Grotzinger, J. P., Condon, D. J., Ramezani, J. & Science Letters, 258, 114– 131.
Newall, M. J. 2009. Reply to Comment: Oman Chronostratigraphy. Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
American Journal of Science, 309, 91– 96. 2004. U– Pb zircon date from the Neoproterozoic Ghaub Formation,
Burns, S. J. & Matter, A. 1993. Carbon isotopic record of the latest Pro- Namibia: constraints on Marinoan glaciation. Geology, 32, 817– 820.
terozoic from Oman. Eclogae Geologicae Helvetiae, 86, 595–607. Horstwood, M. S. A., Foster, G. L., Parrish, R. R., Noble, S. R. &
Calver, C. R. 2000. Isotope stratigraphy of the Ediacaran (Neoproterozoic Nowell, G. M. 2003. Common-Pb corrected in situ U– Pb accessory
III) of the Adelaide Rift Complex, Australia, and the overprint of water mineral geochronology by LA-MC-ICP-MS. Journal of Analytical
column stratification. Precambrian Research, 100, 121–150. Atomic Spectrometry, 18, 837– 846.
Chen, D. F., Dong, W. Q., Zhu, B. Q. & Chen, X. P. 2004. Pb/Pb ages of Hua, H., Pratt, B. R. & Zhang, L.-Y. 2003. Borings in Cloudina shells:
Neoproterozoic Doushantuo phosphorites in South China: constraints complex predator– prey dynamics in the terminal Neoproterozoic.
on early metazoan evolution and glaciation events. Precambrian Palaios, 18, 454– 459.
Research, 132, 123– 132. Ireland, T. R. & Williams, I. S. 2003. Considerations in zircon geochro-
Cherniak, D. J. & Watson, E. B. 2003. Diffusion in Zircon. Reviews in nology by SIMS. Reviews in Mineralogy and Geochemistry, 53,
Mineralogy and Geochemistry, 53, 113– 143. 215– 241.
Cohen, A. S., Coe, A. L., Bartlett, J. M. & Hawkesworth, C. J. 1999. Ireland, T. R., Flottmann, T., Fanning, C. M., Gibson, G. M. &
Precise Re–Os ages of organic-rich mudrocks and the Os isotope Preiss, W. V. 1998. Development of the early Paleozoic Pacific
composition of Jurassic seawater. Earth and Planetary Science margin of Gondwana from detrital-zircon ages across the Delamerian
Letters, 167, 159–173. orogen. Geology, 26, 243–246.
Condon, D. J. & Prave, A. R. 2000. Two from Donegal: Neoproterozoic Jaffey, A. H., Flynn, K. F., Glendenin, L. E., Bentley, W. C. &
glacial episodes on the northeast margin of Laurentia. Geology, 28, Essling, A. M. 1971. Precision measurement of half-lives and
951– 954. specific activities of 235U and 238U. Physics Reviews, C4, 1889– 1906.
Condon, D., Zhu, M. Y., Bowring, S., Wang, W., Yang, A. H. & Jin, Kendall, B. S., Creaser, R. A., Ross, G. M. & Selby, D. 2004. Con-
Y. G. 2005. U– Pb ages from the Neoproterozoic Doushantuo straints on the timing of Marinoan ‘Snowball Earth’ glaciation by
187
Formation, China. Science, 308, 95– 98. Re/187Os dating of a Neoproterozoic, post-glacial black shale in
Condon, D., Schoene, B., Bowring, S., Parrish, R., McLean, N., Western Canada. Earth and Planetary Science Letters, 222,
Noble, S. & Crowley, Q. 2007. EARTHTIME; isotopic tracers 729– 740.
and optimized solutions for high-precision U–Pb ID-TIMS geochro- Kendall, B., Creaser, R. A. & Selby, D. 2006. Re– Os geochronology
nology. Eos, Transactions, American Geophysical Union, 88, Fall of postglacial black shales in Australia: constraints on the timing of
Meeting Supplement, Abstract V41E-06. ‘Sturtian’ glaciation. Geology, 34, 729–732.
Creaser, R. A., Sannigrahi, P., Chacko, T. & Selby, D. 2002. Further Kendall, B., Creaser, R. A. & Selby, D. 2009. 187Re– 187Os Geochro-
evaluation of the Re– Os geochronometer in organic-rich sedimentary nology of Precambrian Organic-Rich Sedimentary Rocks. Geological
rocks: a test of hydrocarbon maturation effects in the Exshaw Society, London, Special Publications, 326, 85 –107.
Formation, Western Canada Sedimentary Basin. Geochimica et Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K. H. &
Cosmochimica Acta, 66, 3441– 3452. Arthur, M. A. 1998. Two or four Neoproterozoic glaciations?
Evans, J. A., Zalasiewicz, J. A., Fletcher, I., Rasmussen, B. & Geology, 26, 1059–1063.
Pearce, N. J. G. 2002. Dating diagenetic monazite in mudrocks: Key, R. M., Liyungu, A. K., Njamu, F. M., Somwe, V., Banda, J.,
constraining the oil window? Journal of the Geological Society, Mosley, P. N. & Armstrong, R. A. 2001. The western arm of the
159, 619– 622. Lufilian Arc in NW Zambia and its potential for copper mineraliz-
Fanning, C. M. & Link, P. K. 2004. U –Pb SHRIMP ages of Neoproter- ation. Journal of African Earth Sciences, 33, 503.
ozoic (Sturtian) glaciogenic Pocatello Formation, southeastern Idaho. Kosler, J. & Sylvester, P. J. 2003. Present trends and the future of zircon
Geology, 32, 881–884. in geochronology: laser ablation ICPMS. Zircon, 53, 243– 275.
Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian gla- Krogh, T. E. 1982a. Improved accuracy of U– Pb zircon ages by the
ciation: data from the Adelaide Geosyncline, South Australia and creation of more concordant zircon systems using an air abrasion
Pocatello Formation, Idaho, USA, Selwyn Symposium Vol. 91. Geo- technique. Geochimica et Cosmochimica Acta, 46, 637– 649.
logical Society of Australia Abstracts, Melbourne, Geological Krogh, T. E. 1982b. Improved accuracy of U –Pb zircon dating by
Society of Australia, 57 – 62. selection of more concordant fractions using a high-gradient mag-
Fike, D. A., Grotzinger, J. P., Pratt, L. M. & Summons, R. E. 2006. netic separation technique. Geochimica et Cosmochimica Acta, 46,
Oxidation of the Ediacaran Ocean. Nature, 444, 744–747. 631– 635.
Folling, P. G., Zartman, R. E. & Frimmel, H. E. 2000. A novel Le Guerroue, E., Allen, P. A., Cozzi, A., Etienne, J. L. & Fanning,
approach to double-spike Pb/Pb dating of carbonate rocks: examples M. 2006. 50 Myr recovery from the largest negative d13C excursion in
from Neoproterozoic sequences in southern Africa. Chemical the Ediacaran ocean. Terra Nova, 18, 147– 153.
Geology, 171, 97 –122. Li, C. W., Chen, J. Y. & Hua, T. E. 1998. Precambrian sponges with
Frimmel, H. E., Klotzli, U. S. & Siegfried, P. R. 1996. New Pb/Pb cellular structures. Science, 279, 879– 882.
single zircon age constraints on the timing of Neoproterozoic glacia- Love, G. D., Grosjean, E. et al. 2009. Fossil steroids record the appear-
tion and continental break-up in Namibia. Journal of Geology, 104, ance of Demospongiae during the Cryogenian period. Nature, 457,
459– 469. 718– 721.
USER’S GUIDE TO NEOPROTEROZOIC GEOCHRONOLOGY 149

Ludwig, K. R. 1980. Calculation of uncertainties of U– Pb isotope data. Schoene, B., Crowley, J. L., Condon, D. J., Schmitz, M. D. &
Earth and Planetary Science Letters, 46, 212–220. Bowring, S. A. 2006. Reassessing the uranium decay constants for
Ludwig, K. R. 1991. Isoplot – a plotting and regression program for geochronology using ID-TIMS U– Pb data. Geochimica et Cosmochi-
radiogenic isotope data. USGS Open File Report, 91– 445. mica Acta, 70, 426– 445.
Ludwig, K. R. 1998. On the treatment of concordant uranium –lead ages. Selby, D. & Creaser, R. A. 2005. Direct radiometric dating of the
Geochimica et Cosmochimica Acta, 62, 665– 676. Devonian-Mississippian time-scale boundary using the Re –Os
Ludwig, K. R. 2000. Decay constant errors in U– Pb concordia-intercept black shale geochronometer. Geology, 33, 545– 548.
ages. Chemical Geology, 166, 315–318. Selby, D., Creaser, R. A., Stein, H. J., Markey, R. J. & Hannah, J. L.
Ludwig, K. R. 2003. Mathematical-statistical treatment of data and errors 2007. Assessment of the Re-187 decay constant by cross
for 230Th/U geochronology. Reviews in Mineralogy and Geochemis- calibration of Re –Os molybdenite and U– Pb zircon chronometers
try, 52, 631– 656. in magmatic ore systems. Geochimica et Cosmochimica Acta, 71,
Lund, K., Aleinikoff, J. N., Evans, K. V. & Fanning, C. M. 2003. 1999–2013.
SHRIMP U–Pb geochronology of Neoproterozoic Windermere Shihong, Z., Ganqing, J. & Yigui, H. 2008. The age of the Nantuo
Supergroup, central Idaho: implications for rifting of western Lauren- Formation and Nantuo glaciation in South China. Terra Nova, 20,
tia and synchroneity of Sturtian glacial deposits. Geological Society 289– 294.
of America Bulletin, 115, 349–372. Silver, L. T. & Duetsch, S. 1963. Uranium – lead isotopic variations in
Mahan, K. H., Wernicke, B. P. & Jercinovic, M. J. 2010. Th–U –total zircon: a case study. Journal of Geology, 71, 721– 758.
Pb geochronology of authigenic monazite in the Adelaide rift Steiger, R. H. & Jager, E. 1977. Subcommission on geochronology –
complex, South Australia, and implications for the age of the type convention on use of decay constants in geochronology and
Sturtian and Marinoan glacial deposits. Earth and Planetary cosmochronology. Earth and Planetary Science Letters, 36,
Science Letters, 289, 76– 86. 359– 362.
Martin, M. W., Grazhdankin, D. V., Bowring, S. A., Evans, D. A. D., Stern, R. A. & Amelin, Y. 2003. Assessment of errors in SIMS zircon
Fedonkin, M. A. & Kirschvink, J. L. 2000. Age of Neoproterozoic U–Pb geochronology using a natural zircon standard and NIST
bilatarian body and trace fossils, White Sea, Russia: implications for SRM 610 glass. Chemical Geology, 197, 111–142.
metazoan evolution. Science, 288, 841– 845. Thompson, M. D. & Bowring, S. A. 2000. Age of the Squantum ‘tillite’,
Mattinson, J. M. 2000. Revising the ‘gold standard’ – the uranium decay Boston Basin, Massachusetts: U –Pb zircon constraints on terminal
constants of Jaffey et al. 1971. EOS, AGU Fall Meeting Supplement Neoproterozoic glaciation. American Journal of Science, 300,
Abstact V61A-02. 630– 655.
Mattinson, J. M. 2005. Zircon U –Pb chemical abrasion (‘CA-TIMS’) Walter, M. R., Veevers, J. J., Calver, C. R., Gorjan, P. & Hill, A. C.
method: combined annealing and multi-step partial dissolution analy- 2000. Dating the 840– 544 Ma Neoproterozoic interval by isotopes of
sis for improved precision and accuracy of zircon ages. Chemical strontium, carbon, and sulfur in seawater, and some interpretative
Geology, 220, 47 –66. models. Precambrian Research, 100, 371–433.
Melezhik, V. A., Fallick, A. E. & Pokrovsky, B. G. 2005. Enigmatic Wasserburg, G. J., Jacobsen, S. B., Depaolo, D. J., Mcculloch, M. T.
nature of thick sedimentary carbonates depleted in 13C beyond the & Wen, T. 1981. Precise determination of Sm/Nd ratios, Sm and
canonical mantle value: the challenges to our understanding of the Nd isotopic abundances in standard solutions. Geochimica et Cosmo-
terrestrial carbon cycle. Precambrian Research, 137, 131–165. chimica Acta, 45, 2311– 2323.
Melezhik, V. A., Roberts, D., Fallick, A. E. & Gorokhov, I. M. 2008. Wendt, I. & Carl, C. 1991. The statistical distribution of the mean
The Shuram– Wonoka event recorded in a high-grade metamorphic squared weighted deviation. Chemical Geology: Isotope Geoscience,
terrane: insight from the Scandinavian Caledonides. Geological 86, 275– 285.
Magazine, 145, 161– 172. Williams, M. L., Jercinovic, M. J. & Hetherington, C. J. 2007.
Narbonne, G. M. 2005. The Ediacara biota: Neoproterozoic origin of Microprobe monazite geochronology: understanding geologic pro-
animals and their ecosystems. Annual Review of Earth and Planetary cesses by integrating composition and chronology. Annual Review
Sciences, 33, 421– 442. of Earth and Planetary Sciences, 35, 137–175.
Narbonne, G. M. & Gehling, J. G. 2003. Life after snowball: the oldest Xiao, S. H., Zhang, Y. & Knoll, A. H. 1998. Three-dimensional preser-
complex Ediacaran fossils. Geology, 31, 27– 30. vation of algae and animal embryos in a Neoproterozoic phosphorite.
Peterson, K. J., McPeek, M. A. & Evans, D. A. D. 2005. Tempo and Nature, 391, 553– 558.
mode of early animal evolution: inferences from rocks, Hox, and Xiao, S. H., Yuan, X. L. & Knoll, A. H. 2000. Eumetazoan
molecular clocks. Paleobiology, 31, 36– 55. fossils in terminal Proterozoic phosphorites? Proceedings of the
Rasbury, E. T. & Cole, J. 2009. Directly dating geologic events: U–Pb National Academy of Sciences of the United States of America, 97,
dating of carbonates. Reviews of Geophysics, 47, RG3001, doi: 13 684– 13 689.
10.1029/2007RG000246. Xiao, S. H., Yuan, X. L., Steiner, M. & Knoll, A. H. 2002. Macroscopic
Rasmussen, B. 2005. Radiometric dating of sedimentary rocks: the appli- carbonaceous compressions in a terminal Proterozoic shale: a sys-
cation of diagenetic xenotime geochronology. Earth-Science tematic reassessment of the Miaohe biota, south China. Journal of
Reviews, 68, 197. Paleontology, 76, 347– 376.
Roden, M. K., Parrish, R. R. & Miller, D. S. 1990. The absolute age of Xu, B., Xiao, S. et al. 2009. SHRIMP zircon U– Pb age constraints on
the Eifelian Tioga Ash Bed, Pennsylvania. Journal of Geology, 98, Neoproterozoic Quruqtagh diamictites in NW China. Precambrian
282– 285. Research, 168, 247–258.
Schaefer, B. F. & Burgess, J. M. 2003. Re– Os isotopic age constraints Yin, C. Y., Tang, F. et al. 2005. U–Pb zircon age from the base of
on deposition in the Neoproterozoic Amadeus Basin: implications for the Ediacaran Doushantuo Formation in the Yangtze Gorges,
the ‘Snowball Earth’. Journal of the Geological Society, 160, South China: constraint on the age of Marinoan glaciation. Episodes,
825– 828. 28, 48– 49.
Scherer, E., Munker, C. & Mezger, K. 2001. Calibration of the lute- York, D. 1966. Least squares fitting of a straight line. Canadian Journal
tium – hafnium clock. Science, 293, 683– 687. of Physics, 44, 1079– 1086.
Schmitz, M. D. & Schoene, B. 2007. Derivation of isotope ratios, York, D. 1967. The best isochron. Earth and Planetary Science Letters, 2,
errors, and error correlations for U–Pb geochronology using 479– 482.
205
Pb-235U-(233U)-spiked isotope dilution thermal ionization Zhang, S. H., Jiang, G. Q., Zhang, J. M., Song, B., Kennedy, M. J. &
mass spectrometric data. Geochemistry Geophysics Geosystems, 8, Christie-Blick, N. 2005. U –Pb sensitive high-resolution ion
Q08006, doi: 10.1029/2006GC001492. microprobe ages from the Doushantuo Formation in south China:
Schmitz, M. D., Bowring, S. A. & Ireland, T. R. 2003. Evaluation of constraints on late Neoproterozoic glaciations. Geology, 33,
Duluth Complex anorthositic series (AS3) zircon as a U– Pb geochro- 473– 476.
nological standard: new high-precision isotope dilution thermal ion- Zhou, C., Tucker, R., Xiao, S., Peng, Z., Yuan, X. & Chen, Z. 2004.
ization mass spectrometry results. Geochimica et Cosmochimica New constraints on the ages of Neoproterozoic glaciations in south
Acta, 67, 3665–3672. China. Geology, 32, 437– 440.
Chapter 10

Modelling the Snowball Earth

YVES GODDÉRIS1*, GUILLAUME LE HIR2 & YANNICK DONNADIEU3


1
LMTG, CNRS-Observatoire Midi-Pyrénées, 14 avenue Edouard Belin, 31400 Toulouse, France
2
IPGP, 4 place Jussieu, 75252 Paris, France
3
LSCE, CEA-CNRS, Orme des Merisiers, 91191Gif-sur-Yvette, France
*Corresponding author (e-mail: godderis@lmtg.obs-mip.fr)

Abstract: We review most of the modelling studies performed to date to understand the initiation and melting of a Snowball Earth, as
well as to describe the glacial environment during the glaciation itself. All the described scenarios explaining the onset of glaciation rely
on a sufficient decrease in the concentrations of atmospheric greenhouse gases (GHGs), typically resulting from the equatorial palaeo-
geography of the late Proterozoic. It is still heavily debated whether or not the oceanic ice cover was thick during the glaciation itself.
However, a consensus has arisen that the most climatically stable scenarios imply the existence of a globally frozen ocean, with a thick ice
cover caused by the flowing of high-latitude sea-ice glaciers towards the equator. Depending on the characteristics of the ice, a thin ice
layer may have persisted along the equator, but this numerical solution is rather fragile. During the snowball event itself, model results
suggest the existence of wet-based continental glaciers. Some parts of the continents may have remained ice-free. From the modelling
perspective, the most significant problem in the snowball hypothesis, particularly in its ‘hard snowball’ version (the most stable numeri-
cally), is the melting phase. With improved modelling, the CO2 threshold required to melt the snowball is much higher than initially
thought, significantly above 0.29 bar. Indeed, because of the very cold conditions prevailing at the surface of the Earth during the
glacial event, the atmosphere becomes vertically isothermal, strongly limiting the efficiency of the greenhouse effect. This melting
problem is further highlighted by geochemical modelling studies that show that weathering of the oceanic crust might be an active
sink of CO2 during the glacial event, limiting the rise in atmospheric CO2. The solution might be found by considering the input
of dark dust from catastrophic volcanic eruptions that would efficiently decrease the albedo of the ice. Finally, modelling studies
also explore the aftermath of the glaciation. The world might have been drier than initially anticipated, resulting in the persistence of
the supergreenhouse effect for at least one million years after the melting phase.

The Snowball Earth hypothesis is an exciting research field numerical models – that is, models not restricted to first-order
for climate modellers. First, it requires the use of diverse models mass balance calculations, but rather accounting for the physics
to explore the salient aspects of the theory, from global geochem- of a snowball event. For this reason, we do not attempt to
ical models to ice sheet and climate models (Table 10.1). Second, discuss all the hypotheses proposed to explain multiple equatorial
coupling of these models is often required, which results in the glaciations.
construction of new numerical tools that can be applied to other
palaeoclimate problems. Finally, the modelling of such an
extreme environment leads scientists to push their numerical How to initiate a snowball glaciation?
models to their physical limits, which is always instructive in
terms of the numerical and physical behaviour of these models. Thus far, three scenarios have been proposed and numerically
The first modelling studies on the ice-albedo instability that tested to explain the onset of a snowball glaciation. They all rely
underpins the Snowball Earth hypothesis were performed inde- on large drops in the partial pressure of greenhouse gases (GHGs).
pendently by William Sellers and Mikhail Budyko (1969), with
a global Energy Balance Model (EBM, or 0D model). With their Scenario 1: It has been suggested that the partial pressure of atmos-
respective models, they demonstrated that reducing the current pheric methane might have been quite high during the Proterozoic
solar constant by even a few percent drives a rapid transition of (Pavlov et al. 2003), reaching up to 100– 300 ppmv because of
the Earth climate into an ice-covered state, characterized by a intense activity of methanogenic bacteria in an anoxic ocean.
high albedo (0.85) and global mean temperature of –100 8C The progressive oxidation of the atmosphere might have led to
(Sellers 1969). Modelling of the nonlinear response of the ice the collapse of the atmospheric methane stock in the Neopro-
line as a function of the incoming absorbed energy demonstrates terozoic, leaving the atmosphere with a low CO2 content, thus
the existence of three stables states of the Earth climate system: triggering global glaciations before pCO2 could self-adjust via
(1) ice-free (e.g. the warm Cretaceous), (2) partially ice-free the silicate-weathering feedback (Walker et al. 1981). This mech-
(e.g. the Phanerozoic glaciations and the present-day state of the anism has not been fully tested to date and requires specific but
Earth system) and (3) globally ice-covered (e.g. the Neoprotero- unvalidated assumptions: the rise in O2 and subsequent collapse
zoic Snowball Earth). Historically, the first two of these states in CH4 must have occurred during the Neoproterozoic, and the
have been the focus of most climatic modelling. However, since global glaciation must have followed this event within no more
the publication of Hoffman et al. (1998), significant efforts have than a few million years, based on the inferred residence time of
been made to understand the third stable state with respect to the carbon in the ocean – atmosphere system (François & Goddéris
Neoproterozoic glaciations. 1998). Should this time span be longer, the imbalance between
Below we present a summary of the results and the main solid Earth degassing and continental silicate weathering would
modelling efforts carried out to date to understand the causes of force CO2 to rise and compensate for the methane collapse.
Neoproterozoic Snowball glaciations, the environment during gla- The two following scenarios rely on the palaeogeographic
ciation, and the mechanisms able to trigger deglaciation. We will setting of the Neoproterozoic. This is still a matter of intense
focus on hypotheses that have been quantified using complex debate, but there is now a general consensus about the dominantly

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 151– 161. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.10
Table 10.1. A summary of all modelling studies of the Neoproterozoic ice ages (with type of model employed and reference)

Tools Authors Note

Snowball Earth inception


EBM Rose & Marshall (2009) Ice-albedo instability
EMIC Micheels & Montenari (2008) Slushball conditions
GEO Goddéris & Donnadieu (2008) Comment on Peltier (2007)
AGCM Ishiwatari et al. (2007) Ice-albedo instability
EMIC Lewis et al. (2007) Sea-ice dynamics
AGCM-GEO Goddéris et al. (2007) Continental drift and CO2 decrease
AGCM-GEO Peltier et al. (2007) CO2 scenario for a slushball
AGCM Romanova et al. (2006) Test of the climatic factors
AGCM Pavlov et al. (2005) Impact of interstellar dust
EMIC Donnadieu et al. (2004b) Sea-ice dynamics
EMIC-GEO Donnadieu et al. (2004a) Continental drift and CO2 decrease
EMIC Lewis et al. (2004) Test of the climatic factors
OAGCM Poulsen & Jacob (2004) Sea-ice dynamics
AGCM-EMIC Ramstein et al. (2004) Review
EMIC-GEO Donnadieu et al. (2004c) Continental drift and CO2 decrease
AGCM-ISM Pollard & Kasting (2004) Glacial deposits
OAGCM Peltier et al. (2004) Ice-albedo instability
AGCM Jenkins (2004) Ice-albedo instability
EMIC Stone & Yao (2004) Ice-albedo instability
EBM-GEO Goddéris et al. (2003) Igneous provinces and CO2 decrease
EMIC Lewis et al. (2003) Ocean dynamics
OAGCM Poulsen (2003) Climatic feedbacks
AGCM-ISM Baum & Crowley (2003) Climatic feedbacks
EMIC Bendtsen (2002) Climatic feedbacks
EMIC Bendtsen & Bjerrum (2002) Climatic feedbacks
AGCM Donnadieu et al. (2002) Testing the high obliquity
OAGCM Poulsen et al. (2002) Climatic feedbacks
GEO Schrag et al. (2002) Methane and CO2 decrease
OAGCM Poulsen et al. (2001) Climatic feedbacks
AGCM Baum & Crowley (2001) Climatic feedbacks
EBM-ISM/AGCM Hyde et al. (2000) Slushball solution
AGCM Chandler & Sohl (2000) Climatic feedbacks
AGCM Jenkins (2000) Climatic feedbacks
AGCM Jenkins & Smith (1999) Climatic feedbacks
AGCM Jenkins & Frakes (1998) Climatic feedbacks
EBM Crowley & Baum (1993) Climatic feedbacks
EBM-GEO Marshall et al. (1988) Ice-albedo instability
EBM Budyko (1969) Ice-albedo instability
EBM Sellers (1969) Ice-albedo instability
During the Snowball Earth
GEO Le Hir et al. (2008b) CO2 evolution, carbon cycling
GEO Le Hir et al. (2008a) Seawater composition
EBM þ ice-shelf flow Warren & Brandt (2006) Sea-ice thickness
EBM þ ice-shelf flow Pollard & Kasting (2006) Sea-ice thickness
EBM þ ice-shelf flow Goodman (2006) Comment
EBM þ ice-shelf flow Pollard & Kasting (2005) Sea-ice thickness
GCM-ISM Donnadieu et al. (2003) Continental ice dynamics
Complex 1D ice model Goodman & Pierrehumbert (2003) Sea-ice thickness
Spectral model Warren et al. (2002) Sea-ice thickness
Sea ice model McKay (2000) Sea-ice thickness
Melting the Snowball Earth
AGCM-dust model Le Hir et al. (2010) Explicit effect of volcanic ash on snow/ice albedo
AGCM Abbot & Pierrehumbert (2010) Explicit effect of volcanic ash on snow/ice albedo
AGCM Le Hir et al. (2007) Inversion of the atmospheric vertical thermal gradient
EMIC Lewis et al. (2006) Sensitivity to snow and ice albedo values
AGCM Pierrehumbert (2005) Inversion of the atmospheric vertical thermal gradient
AGCM Pierrehumbert (2004) Inversion of the atmospheric vertical thermal gradient
EBM-ISM Crowley et al. (2001) Near-Snowball Earth
EBM with CO2 ice clouds Caldeira & Kasting (1992) CO2 ice cloud effect
Tools Year/ Authors / Journal Note

Aftermath of the Snowball Earth


GCM-GEO/weathering model Le Hir et al. (2009) Weathering rates, climate, CO2 restoring
GEO Higgins & Schrag (2003) Carbon isotopic signal

Energy balance models (EBMs) only compute the radiative energy budget of the atmosphere. The transport is implicitly included as well as the water
cycle. Radiative convective models (RCMs) include a description of the radiative energy budget of the atmosphere, and of the upward transport of
energy by convection. General circulation models (GCMs) resolve explicitly the transport of energy and matter within the atmosphere (AGCM) and
the ocean (OGCM) on a three-dimensional grid and physically compute the water cycle. Earth system Models of Intermediate Complexity (EMIC)
bridge the gap between the simple RCM, the EBM and the highly complex GCMs. They couple simplified atmospheric and oceanic models and are
generally designed for long-term simulations (104 to 105 years). GEO stands for global biogeochemical cycle models and ISM for ice-sheet model.
MODELLING SNOWBALL EARTH 153

low-latitude location of most of the continental blocks from


800 Ma to 600 Ma (Evans 2000; Meert & Powell 2001; Torsvik
et al. 2001; Trindade et al. 2003; Macouin et al. 2004; Meert &
Torsvik 2004; Hoffman & Li 2009).

Scenario 2: This scenario (Schrag et al. 2002) was developed to


explain the unusual relationship between the onset of glaciation
and a large negative d13C (carbonate) anomaly that preceded it
(Halverson et al. 2002). High pre-glacial d13C values (.þ5‰)
were assumed to result from a combination of high phosphorous
flux to the oceans and efficient burial of organic carbon in the
deltas of large equatorial watersheds from continents clustered
in the tropics. Analogous present-day deltas include those of the
Amazon and southeastern Asian rivers such as the Irrawady,
Ganges and Salween. However, it is worth noting that the
organic matter being buried in the Amazon and Bengal fans is
today partly of terrestrial origin, which would not have been the
case in the Neoproterozoic. Furthermore, the high sedimentation
rates and resulting efficient burial of organic matter in southeastern
Asia (up to 100% of the amount of organic carbon reaching the
sediment) is very high, not because of the low-latitude location
of these catchments, but because of their source in the Himalayas
(Galy et al. 2007). Thus, it is not clear that Neoproterozoic palaeo-
geography alone could account for the unusually efficient organic
carbon burial.
The subsequent decrease in d13C prior to the onset of snowball
glaciation is interpreted to be the result of a protracted release of
CH4 derived from methanogenesis within the previously buried
organic carbon pool. In the Schrag et al. (2002) model, after avail- Fig. 10.1. CO2 consumption by continental weathering rate for a
able reservoirs for methane clathrate on continental margins were supercontinental (Rodinia) and a dispersed configuration in 104 mol of Ca2þ or
filled, CH4 began to leak to the anoxic water column and then to the Mg2þ/km2/a, for a fixed 6.5 PAL of CO2. Source: Donnadieu et al. (2004a).
atmosphere, where it drove an increase in CH4 partial pressure.
The gradual substitution of CO2 by CH4 then generated a severe
vulnerability in global climate because CH4 has much shorter resi- low latitudes, in contrast to the latitudinal break-up of Pangaea,
dence times than CO2 and is not similarly regulated through sili- which did not trigger widespread glaciation.
cate weathering. Thus, if this situation persisted for several However, Rodinia break-up alone is not sufficient to initiate a
hundreds of thousands of years, then analogous to the previous snowball. Indeed atmospheric CO2 stabilizes above the required
scenario, a sharp decrease in the CH4 flux would precipitate a threshold (a value dependent on the model used, see next
sudden cooling event. section). However, the possible solution lies in the geodynamics
This model for triggering a Snowball Earth has the advantage of of supercontinental break-up, which is heralded by the onset of
being the only one that can account for the major negative d13C large magmatic provinces as a result of the accumulation of heat
anomalies preceding Cryogenian glaciations (Halverson et al. below large continental assemblages (Courtillot et al. 1999) and
2005; Prave et al. 2009). However, this scenario has not been subsequent release through mafic eruptions. Specifically, a
thoroughly modelled, and some aspects remain unresolved, such mantle superplume, starting at 830 Ma and lasting some 85
as what triggered the collapse in the CH4 flux and whether O2 million years, is associated with the early stages of Rodinia
partial pressure was sufficiently low to allow the accumulation break up (Li et al. 1999; Li et al. 2003). Tholeiitic magmatism
of high concentrations of atmospheric methane. has been identified between 825 and 755 Ma, especially in
Australia (Wingate et al. 1998; Wingate & Giddings 2000) and
Scenario 3: The prominent tectonic event of the Neoproterozoic northwestern Laurentia (Park et al. 1995), but also in south
was the break-up of the Rodinia supercontinent. The breakup China (Li et al. 1999) and on the Congo craton (Key et al.
started c. 800 Ma ago, when continental masses were distributed 2001). Collectively, these magmatic events appear to define a
in the tropics (Li et al. 2003). Donnadieu et al. (2004a) calculated plume time-cluster based on the database of Ernst and Buchan
the climatic and geochemical impact of this break-up, through the (2002). The most widespread flood basalt event probably occurred
development of a numerical model coupling a 2.5D climate model on Laurentia, with a minimum size of 3  106 km2 and possibly
(Petoukhov et al. 2000) and a global biogeochemical cycle model equivalent in area to the original Siberian Traps.
(Goddéris & Joachimski 2004). The use of an explicit climate The onset of continental flood basalts results first in a massive
model allows the subsequent increase in rainfall and runoff but short-lived release of CO2 into the atmosphere, leading to an
above the continents to be captured as the breakup proceeds. As enhanced greenhouse effect. However, a few million years after
a result, CO2 consumption through continental weathering is the end of the eruptive phase, this CO2 excess is rapidly consumed
stimulated, driving atmospheric CO2 levels towards the threshold by intensified silicate weathering. In about 4 million years, the CO2
required to initiate a snowball glaciation (Fig. 10.1). This hypoth- level stabilizes at a value lower than the pre-eruption level
esis therefore allows atmospheric CO2 to decrease to the threshold because highly weatherable basalt on the surface consumes CO2
level without invoking a change in CO2 degassing. Donnadieu (Dessert et al. 2001). The efficiency of this long-term cooling
et al. (2004a) computed a decrease in atmospheric CO2 from effect depends on the surface area of the large igneous province
1830 ppmv down to 510 ppmv during the breakup of Rodinia as well as on the precipitation and temperature where the traps
and a concomitant drop in global mean annual temperature from occur (Goddéris et al. 2003). Donnadieu et al. (2004a) have
10.8 8C to 2 8C, assuming a solar constant reduced by 6% relative demonstrated that the weathering of freshly erupted basaltic
to the present. Importantly, the drastic cooling produced in this surface as a disaggregating Rodinia drifts towards the humid equa-
model is the result of dominantly longitudinal break-up in the torial regions is sufficient to trigger a snowball glaciation. Strictly
154 Y. GODDÉRIS ET AL.

speaking, this scenario can only apply to the first snowball glacia- the ice and the final collapse into a snowball state is mainly
tion (i.e. Sturtian episode) and does not explicitly account for the driven by the atmosphere, and the ocean plays only a minor role.
negative d13C anomaly that precedes this glaciation (Prave et al. The snowball CO2 thresholds were computed to be 90 ppmv for
2009). Nevertheless, this study indicates that a globally cool a supercontinental configuration and 150 ppmv for the dispersed
climate is a predictable consequence of the low-latitude break-up configuration along the equator. These levels are much lower
of Rodinia. than initially expected from atmospheric model simulations
(Jenkins & Smith 1999). Similar conclusions regarding the critical
role played by the Hadley transport once the ice line reaches 308
The ice and climate of the Snowball Earth latitude were obtained recently with a general circulation model
(GCM) coupled to a thermodynamic sea-ice model, assuming an
The onset of the snowball: ice-albedo instability idealized supercontinent located at the equator (Poulsen & Jacob
2004).
Decrease in atmospheric GHGs will always cool global climate.
However, even with very low GHG levels, is the Snowball Earth
a plausible climatic end-member? A first set of numerical studies During the Snowball Earth: the slushball theory as a solution
focused on the possibility of sea ice reaching the equator. Simple to explain the survival of photosynthetic life?
energy balance models all display a climatic instability once the
ice line reaches about 308 latitude (Budyko 1969). This runaway How can climate models that predict a totally ice-capped ocean be
ice-albedo feedback occurs once the surface radiation entering reconciled with palaeontological evidence that establishes the per-
the open ocean regions cannot balance the heat loss by reflection sistence of photosynthetic activity throughout the snowball event?
of sunlight from low-latitude ice-covered regions, resulting in Hyde et al. (2000) proposed a scenario where the equatorial ocean
rapid ice growth and global cooling. However, initial coupled remains ice-free during snowball events. This ‘tropical oasis’ sol-
ocean – atmosphere general circulation models did not follow ution is colloquially referred to as the ‘slushball’ hypothesis and
this behaviour prescribed by EBMs. Using the coupled Fast has gained traction amongst geologists who see evidence for
Ocean Atmosphere Model (FOAM), which accounts for ocean open ocean conditions during the presumed snowball glaciations
dynamics, Poulsen (2003) performed a simulation with a unique (e.g. Leather et al. 2002; Allen & Etienne 2008). This result was
atmospheric pCO2 fixed at 140 ppm, a solar constant reduced by obtained by coupling a simple energy balance model and an ice-
7%, an idealized equatorial supercontinent and a sea-ice thickness sheet model under a solar constant reduced by 6% and assuming
of 1000 m. He forced the ice line to be located at 108 latitude. Once a very low CO2 level of 140 ppmv. Under such conditions,
the prescribed condition for the location of the ice line was released, massive icecaps develop on continents, as well as thick ice
it did not continue to the equator but rather retreated to higher lati- sheets on the global ocean. Then, the calculated icecaps were
tudes, .408N and S. This result was used to argue that a full snow- specified as boundary conditions of a GCM, in which the CO2
ball climate is climatically untenable (Poulsen et al. 2001; Poulsen partial pressure was prescribed at 700 ppmv. The GCM output
2003). However, Lewis et al. (2004) demonstrated that this instan- showed that the ocean remains free of ice below 258 latitude.
taneous sea-ice retreat was caused by the absence of sea-ice However, this climatic solution is not self-consistent: if the gla-
dynamics and was an artefact of the way the energy required to ciers were allowed to respond to the 700 ppm pCO2, they would
melt the sea ice was treated. Using a similar model, Lewis et al. retreat. And if CO2 was reduced to 140 ppmv, the equatorial
(2007) then obtained the surprising result that the ocean completely ocean would freeze. This slushball simulation is therefore not
froze over when pCO2 fell below 3000 ppmv. robust, because the threshold for the onset of a Snowball Earth
The potential of the onset of a full snowball glaciation in a simulated with a GCM is generally far below 700 ppmv (Poulsen
coupled dynamic model has been confirmed by a study performed 2003; Donnadieu et al. 2004a) and the stability range of the
with an Earth Model of Intermediate Complexity (EMIC) that took slushball solution.
into account sea-ice thermodynamics and ocean dynamics, where An additional argument against the slushball model comes from
two realistic continental configurations were tested (superconti- sea-ice dynamics. As demonstrated by Lewis et al. (2007), a
nent Rodinia and a disaggregated Rodinia along the equator; sea-ice margin reaching low latitudes without driving a total
Donnadieu et al. 2004b). Probably the most important conclusion collapse of the ocean is only obtained in simulations omitting
arising from this study was that climatic instability exists for both the sea-ice dynamic process. In oceanic models including sea-ice
continental configurations, meaning that ocean dynamics cannot dynamics, the ice-albedo instability feedback does not allow a
prevent the onset of a Snowball Earth. However, they also found stable slushball solution.
that the CO2 threshold required to trigger a snowball is (i) highly
dependent on the continental configuration, being higher for the
dispersed continental configuration, and (ii) below the threshold During the Snowball Earth: a clear equatorial thin-ice
values calculated with an atmospheric model, confirming the as a solution for life?
important role of the ocean dynamics on the onset of a Snowball
Earth, but also emphasizing the role of atmospheric meridional Based on observations of lakes in the Antarctic Dry Valley,
heat transport. Indeed, the Hadley cells strengthen as the ice McKay (2000) formulated an alternative scenario to explain the
moves equatorward in response to decreasing atmospheric CO2 survival of photosynthesis. Using a simple method to reconstruct
pressure, bringing more and more heat from the equator to the the energy balance of the ice cover, he showed that if the mean
ice front, thus slowing its advance. But this strong negative feed- annual temperature at the equator was around –30 8C, then the
back collapses once the ice line reaches 308 latitude. Indeed, as equilibrium thickness of the bare sea ice in this area might have
the ice line is now located on the descending branch of the been as little as a few metres, assuming an ice albedo of ,0.75
Hadley cell, the cold air in contact with the ice is now efficiently (McKay 2000). This result was obtained assuming low absorption
transported back towards the equator along the lower meridional efficiency of the ice within the visible part of the solar spectrum, so
branch of the cell (Bendsten 2002). Air temperatures at the that heat can be transferred to the water below the ice, thus limiting
equator then rapidly decrease to zero, shutting down the climatic the growth of the ice sheet through seawater freezing. Such a low
negative feedback, allowing the onset of snowball glaciation. absorption coefficient for the ice requires a slow rate of freezing,
Heat transport in the ocean is more complex and largely depends which prevents the incorporation of air bubbles and keeps the ice
on the continental configuration. Nevertheless, Donnadieu et al.’s relatively clear. However, such clear ice is inconsistent with
(2004b) key finding is that heat transport during the advance of high albedo, which requires a high bubble density. The calculated
MODELLING SNOWBALL EARTH 155

thickness is critically dependent on the ice albedo, exceeding Furthermore, Pollard & Kasting (2005) argue that, in any case,
several tens of metres if the ice albedo is .0.8. low-latitude large lakes and confined seas like the modern Mediter-
The thin-ice model has been challenged by Warren et al. (2002), ranean, protected from planetary-scale sea-glacier flow by sur-
who described the absorption of solar energy as a function of wave- rounding continents, would maintain areas with thin ice cover.
length and depth, accounting for the scattering of radiation by air This model has been challenged by Goodman (2006) and
bubbles inside the ice. They found that the absorption of incident Warren & Brandt (2006). Goodman (2006) noted that the atmos-
energy within the first centimetres of the ice sheet was much pheric and oceanic water cycle must be internally balanced,
higher than previously expected (McKay 2000), particularly in because the glaciation lasts for millions of years, to avoid any con-
the near-infrared wavelength band (0.7 –3 mm). Furthermore, the centration of the water mass into the atmosphere or the ocean.
visible part of the spectrum is largely scattered by air bubbles, Because Pollard & Kasting (2005) calculated that the ice sheet
brine inclusions and cracks. Heat is thus not efficiently transferred was formed by freezing of seawater and that net sublimation trans-
to the water below the ice. As a consequence, their results sug- fers water from the ice sheet to the atmosphere in the tropics, either
gest that the ocean was everywhere capped by at least several some meteoric ice would have to return water to the ocean (which
hundred metres of ice. In Warren et al.’s (2002) model, very low did not happen in the model) or else marine ice could not reach the
albedos (,0.4) of the ice would be required to allow a thin-ice surface, and all of the ice-sheet surface would consist of meteoric
solution, implying a very low air bubble density. Such albedo ice. However, meteoric ice is bubbly and opaque, with a high
values are not realistic under the tropical sun, because it would albedo, thus counteracting the internal melting process proposed
warm the ice above the melting point. Another argument supports by Pollard & Kasting (2005). In conclusion, whether the thin-ice
the thick-ice hypothesis. The latitudinal air-temperature gradient solution can be maintained or not is still a matter of debate. We
in a snowball environment would produce thicker sea ice at can only conclude that very clear thin ice is a fragile solution,
higher latitudes than around the equator. This is not dynamically and can potentially be formed only where meteoric ice is absent.
stable, so ice would flow from high to low latitudes as sea glaciers To date, model studies have reconstructed the Snowball Earth as
reducing the thickness gradient, but without forming a uniform a hostile environment for photosynthetic marine algae and associ-
thickness everywhere due to slow flow rates (Goodman & ated food chains, with no ice-free equatorial ocean, nor tropical
Pierrehumbert 2003). thin-ice. However, the first description of the snowball theory
The question of thick v. thin ice remains open. A recent model- suggested the existence of cracks in the sea ice, whether it was
ling study coupling an EBM with an ice-shelf model demonstrates thin or not (Hoffman et al. 1998). In a cold environment, the
the possibility of the existence of areas of thin ice (,3 m thick) equatorial sea ice would form cracks if its speed exceeds
below a latitude of 208 (Pollard & Kasting 2005) (Fig. 10.2). 50 m/a. Newly formed sea ice in cracks will contain brine chan-
Climate simulations of the Snowball Earth conditions predict net nels, which today host a variety of organisms including photo-
positive surface accumulation (precipitation –evaporation) of a trophs. Moreover, if the ice was thin, winds and tidal forces
few mm/a in mid to high latitudes. Snow thus accumulates on were probably strong enough to generate leads (Pollard &
the ice, strongly increasing the albedo and allowing the growth Kasting 2005). Also, active hydrothermal areas in shallow water
of thick sea ice. But in low latitudes, the net surface accumulation should maintain open waters in their direct vicinity. Hence the
is negative, the sea ice remains free of snow, and the thin-ice existence of several types of refugia required to explain the per-
solution is valid. sistence of photosynthesis cannot be ruled out from numerical
Pollard & Kasting (2005) calculated that the snow limit is simulations alone. Indeed, large populations of microscopic
located around a latitude of 128. Below this latitude, the sea ice organisms can populate a small volume in which light remains
is a marine ice accumulated through basal freezing, much less available. They can preserve a relative diversity even if the remain-
bubbly and clearer than the snow-derived glacial ice. The high- ing oasis is small (Knoll 2003).
latitude thick ice (150 m thick) flows from high latitude at a rate
of 800 m/a and eventually crosses the 128 latitude limit.
Because the precipitation –evaporation (P– E) budget becomes During the Snowball Earth: glacial deposits and continental
negative, the snow accumulated on this ice starts to sublimate or ice behaviour
melt. As a result, up to 5 W m22 is absorbed by the ice, and internal
melting occurs between latitudes of 128 and 108. Thin ice may thus The behaviour of the terrestrial glacial regime during the Neopro-
be maintained at low latitude, where the albedo can be as low as terozoic glaciations is still a matter of debate. Some Snowball Earth
0.45, despite the flowing of high-latitude sea-ice glaciers. detractors claim that the glacial sequences cannot be explained
with the Snowball Earth scenario. Indeed, the near shutdown of
the hydrological cycle simulated by climatic models, once the
Earth is entirely glaciated, stands in contrast to the requirements
for active, wet-based continental ice sheets to produce the
observed thick glacial deposits. The extent and dynamics of the
continental ice has been estimated through the forcing of an ice-
sheet model with the output of a GCM in snowball conditions
(Donnadieu et al. 2003). The GCM (LMDz) was run assuming
1 PAL (preindustrial atmospheric level, or 280 ppmv) of CO2, a
solar constant reduced by 6%, and a realistic palaeogeographic
configuration including an estimate of the continental relief
(Greenvillian-aged mountains, with a prescribed elevation of c.
2000 m, relics of the last major orogenic cycle). With those for-
cings, the ice line rapidly reaches 308 latitude, reaching the
instability threshold and resulting in the onset of a snowball
state. The ice-sheet model (Ritz et al. 2001) is then run using the
climate fields generated by the GCM. The three main conclusions
of this study were as follows. First, thick continental ice-sheets
Fig. 10.2. Sea-ice thickness during a snowball in the Pollard & Kasting (2005) build up within a few hundred thousand years, with a mean of
model, as a function of latitude. The dashed line indicates bubbly ice, and the 2500 m and a maximum thickness of 5000 m above Laurentia
solid line clear ice. Flowing of sea glaciers is accounted for. and Antarctica. The ice sheets nucleate on high relief and in less
156 Y. GODDÉRIS ET AL.

than 50 ka spread across the continental interior. Second, the insolation. This threshold rises to 0.29 bar with the same model
model predicts a final relative fall in sea level of 200 m. if a realistic decrease of 6% in the solar constant is used. It would
However, the drop in sea level is highly dependent on several com- then take 8 million years to melt the hard snowball, still assuming
peting phenomena. Glacioeustasy is the dominant control; for the present-day degassing rate, without accounting for possible
example, with an average thickness of ice on all continents of dissolution of carbon into the ocean (Higgins & Schrag 2003),
2000 m, the fall will be c. 500 m, assuming an emerged continental which would increase the time required to accumulate threshold
surface of one-quarter the area of the oceans. However, CO2 levels. Furthermore, an energy balance model is probably
glacio-isostatic rebound, hydro-isostatic rise of the seafloor (due not the best tool with which to explore high CO2 climate, because
to a reduction in water load) and the tectonic subsidence balance too many parameterizations used in EBMs (such as cloud
the initial glacio-eustatic fall, reducing the relatively low sealevel forcing) are calibrated on the present climate. Working with an
fall despite a very significant land ice volume. Performing similar atmospheric GCM (the FOAM GCM) and increasing CO2 partial
GCM – ice sheet model simulations, but without prescribing any pressure in a fully glaciated world, Pierrehumbert (2004) showed
relief on the continents, Pollard & Kasting (2004) also show that that 0.29 bar is strictly a minimum value for the threshold, and
thick ice sheets build up on continents in the tropics, but the most likely it is much higher. Through an extrapolation of the
location of the foci of the ice sheets is determined by the dynami- CO2 radiative forcing function, he showed that 2 bar of CO2
cally controlled regions of positive P– E. might have been required to melt the snowball, corresponding to
Third a large area of the calculated continental ice sheet is wet- an unrealistic glacial duration of .50 million years.
based, so basal sliding is expected to occur. It is particularly active These simulations were carried out assuming an idealized
in the equatorial area where the calculated sliding speed can reach rectangular supercontinent and present-day orbital parameters.
several tens of m/a. If the sea ice in the equatorial area was thin The greenhouse efficiency G can be defined as the difference
(see above discussion), such sliding of the ice sheets may between the amount of energy emitted by the Earth surface and
explain the main features of glacial deposits accumulated during the amount of energy lost as long-wave radiation at the top of
the Snowball Earth event (Donnadieu et al. 2003). Some parts of the atmosphere, the outgoing long-wave radiation (OLR):
the continents remain ice-free, even after the ice sheet reaches a
steady-state volume after 400 ka in the ice-sheet model. Such G ¼ sTS4  OLR (1)
ice-free areas may act as dust sources, darkening the snow over
large surfaces through aeolian deposition, helping to maintain where Ts is the average surface temperature. OLR is a function
areas of thin ice on the ocean (Pollard & Kasting 2005). of the effective planetary temperature at the top of the atmosphere,
Te, and is equal to sTe4 (for the modern Earth, Te ¼ 255 K). Today,
the atmosphere is not isothermal, and the effective temperature is
Melting the Snowball Earth
lower than the temperature at the surface. Consequently G is posi-
tive. During the snowball event, in the winter hemisphere, the
The commonly accepted scenario for the melting of the Snowball
atmosphere is isothermal with height, because of inefficient solar
Earth is from Kirschvink (1992) and Hoffman et al. (1998).
heating of the surface (high albedo) and convection due to
However, the solution for escape from a snowball was first pro-
absence of a warm ocean. The air-temperature profile as a function
posed by Walker et al. (1981) in their seminal paper on silicate
of height even displays an inversion at ground level, due to the
weathering and the long-term stabilization of global temperatures.
extremely cold conditions at the Earth surface. As a result, Te is
Before the publication of the Snowball Earth hypothesis, they had
almost equal to Ts, and the greenhouse efficiency approaches
already proposed a solution to the melting problem:
0. Adding CO2 to the system will thus not efficiently warm the
‘If global glaciations were to occur, the rate of silicate weathering should fall Earth, as long as the temperature profile remains isothermal. The
very nearly to zero, and carbon dioxide should accumulate in the atmosphere outgoing long-wave radiation might even exceed the black-body
at whatever rate it is released from volcanoes. Even the present rate of release radiation at the Earth surface, and the net greenhouse effect is
would yield 1 bar of carbon dioxide in only 20 million years. The resultant
large greenhouse effect should melt the ice cover in a geological short slightly negative for the winter hemisphere. In the summer hemi-
period of time.’ sphere, the low tropopause limits the vertical temperature gradient,
reducing the efficiency of the greenhouse effect, which is com-
This general prescription for the recovery from a snowball
pounded by the virtual lack of water vapour feedback at such
remains essentially valid. Indeed, the snowball hypothesis funda-
cold temperatures. Furthermore, the cloud greenhouse effect is
mentally relies on the strong reduction of the carbon sink during
reduced, because the high-altitude clouds (contributing the most
the snowball. Organic carbon burial was certainly maintained at
to the greenhouse effect) are thin or absent because of the very
a very low level during the event, because most of the ocean was
low water vapour concentration in the cold upper troposphere in
covered by thick ice. Also, silicate weathering was strongly inhib-
the summer hemisphere, and at any level in the winter hemisphere
ited by a drastic decrease in the continental water cycle and the low
(Pierrehumbert 2004).
temperatures. As a result, in the absence of rains to scrub the atmos-
A solution to this problem may be found in reduction of the
phere of CO2, the carbon originating from the degassing of the solid
albedo. Decreasing the albedo would increase the amount of
Earth should accumulate in the atmosphere to very high levels, until
solar flux absorbed at the surface and promote the onset of a
a threshold for deglaciation is reached. Caldeira & Kasting (1992)
normal vertical temperature profile of the atmosphere in the
calculated the oft-cited figure of 450 times the present-day (i.e.
winter hemisphere and increase the efficiency of the greenhouse
1992) CO2 level (or 0.126 bar) as the threshold where CO2 levels
effect. Causes of the reduction in albedo are not yet defined, but
counteract the strong ice-albedo feedback and initiate melting. At
the thin-ice solution or the presence of dirty snow generated by
the present-day degassing rate of 6.8  1012 mol/a (a value
volcanic eruptions or windblown dust from deserts are reasonable
chosen to balance the present-day continental silicate weathering;
mechanisms (Pierrehumbert 2004).
Gaillardet et al. 1999), it would take only 3.5 million years to
Expanding on the work of Pierrehumbert (2004, 2005), Le Hir
reach this threshold. However, as in all facets of modelling of the
et al. (2010) explored the impact of a large volcanic eruption in
Snowball Earth, the situation is more complicated.
the final stage of the snowball glacial event when atmospheric
CO2 had already reach 0.2 bar. Huge eruptions, similar to the
The melting problem: the climatic modelling argument Toba eruption at 74 ka, occur once or twice per million years
and release up to 2  1016 g of dust into the atmosphere. Given
The threshold value of 0.126 bar CO2 for initiating snowball the expected duration of the glacial event (close to 10 million
melting was calculated with an EBM assuming the present-day years), it seems plausible that a major volcanic eruption occurred
MODELLING SNOWBALL EARTH 157

during the final glacial stage. A GCM (LMD) was used that included
an improved calculation of the albedo of the snow, including
the effect of aging of the snow and dust input. Le Hir et al. (2010)
showed that the sudden release of 2  1016 g of dust into the atmos-
phere decreases the albedo of the surface along the equator by about
0.2 units, but actually increases the albedo by about 0.1 units
between 208 and 408 latitude. The decrease in albedo along
the equator is due to dust accumulation in an area where net ablation
of the snow cover is predicted. Temperature rises along the equator,
activating Hadley cell circulation and leading to an increase in
fresh snow precipitation above 208 latitude. Dust is buried
under the fresh snow and albedo increases. Overall, the mean
annual temperature between latitudes of –10 and þ108 rises by
about 6 8C, reaching values as high as –2 8C. Although a volcanic
eruption does not seem to be sufficient to generate the dramatic
melting of the Snowball Earth, dust input does seem to be a plausible
mechanism to reduce the required CO2 threshold.

Fig. 10.3. Snowball duration as a function of the CO2 threshold needed to


The melting problem: the geochemical initiate melting. The stars correspond to estimated levels with various models
modelling argument and boundary conditions (stars), assuming accumulation of CO2 in the
atmosphere, and no diffusion in the ocean. S is the solar constant normalized to
Problems also arise from geochemical modelling. Using a model the present-day value. The triangle indicates a simulation where atmospheric
describing the global carbon and alkalinity cycles during the CO2 is allowed to dissolve into the ocean through cracks, assuming the 0.29 bar
Snowball Earth event, Le Hir et al. (2008b) calculated that an threshold (EBM simulation with a solar constant reduced by 6%). The square
open water area of 3000 km2 during glaciation is large enough to further assumes consumption of CO2 by seafloor weathering, buffering the rise
in CO2 and hence delaying the melting.
allow efficient diffusion of atmospheric carbon into the ocean
such that the atmosphere and the ocean are in equilibrium with
respect to carbon at the million-year timescale. Present-day
emerged geothermal environments occupy a surface with an area understood, and no investigations have been performed with
of c. 1.5  106 km2 (Dessert et al. 2003). Hence, during global gla- complex numerical models to date. Three studies explore the geo-
ciations, some small open-water areas, easily in excess of chemical global cycles in the aftermath of the snowball. Two of
3000 km2, may be maintained by heat production and lava flows. them are extremely simple (box models) (Higgins & Schrag
For open water to be maintained, two conditions must be fulfilled. 2003; Le Hir et al. 2008a). In both cases, the principle difficulty
First, the hydrothermal zone must not be too deep, otherwise their in simulating the post-glacial environment arises from the quanti-
heat would be diffused over an area so large that it cannot melt the fication of the weathering rates of the continental surfaces under a
overlying sea ice. Second, the hydrothermal zone must not be too very warm climate. Weathering during the melting phase is crucial
shallow on the continental shelf where the heat flux would only for at least two reasons (Higgins & Schrag 2003): (i) cap-carbonate
melt the sea ice in the immediate vicinity, resulting in a lake accumulations are thought to be the consequence of the transfer of
above the thick sea ice which may be disconnected from the alkalinity to the ocean from the extremely fast dissolution of con-
global ocean under the thick sea ice (Goodman 2006). tinental carbonates exposed to acid rain under several hundred
If only 2% of geothermal environments had fulfilled these con- times the present-day CO2 atmospheric pressure, and (ii) the rate
ditions, then massive dissolution of CO2 should have occurred, at which continental silicate rocks dissolve controls the time
acidifying the ocean. Even assuming the existence of a buffering span needed to restore normal climatic conditions, because silicate
mechanism through seafloor carbonate dissolution, the pH of the weathering consumes atmospheric CO2 and stores it as carbonate
water drops to about 6 within 3 to 4 million years after the onset sediments on the seafloor. However, weathering is a function of
of the Snowball glaciation, with atmospheric CO2 levels reaching air temperature and continental runoff. It can be safely assumed
more than 0.1 bar (Le Hir et al. 2008b). Such acidification of the that temperatures are high under 0.3 bar of atmospheric CO2
waters enhances the dissolution rate of the oceanic basaltic crust once the melting is completed, but global precipitation patterns
by a factor of more than 4, even assuming a glacial thermohaline in such a climate are not easy to predict. Lee Kump (personal com-
circulation fixed at 1% of its present-day value. Seafloor weather- munication, cited by Higgins & Schrag 2003) suggested a
ing consumes exospheric carbon, trapping it as carbonates in the maximum sevenfold increase in continental runoff compared to
veins of the oceanic crust (Alt & Teagle 1999). This means that the present-day global value. However, a recent extensive numeri-
all carbon sinks are not drastically reduced during the glacial cal study in which a GCM was coupled to a transport-reactive
event, thus partly counteracting the rise in atmospheric CO2 model simulating weathering rates demonstrated that such an
thought to be required to melt the snowball. Le Hir et al. increase is not realistic (Le Hir et al. 2009). Indeed, the total
(2008b) calculated that within 30 million years (thought to be amount of energy that can be used to evaporate waters from the
the maximum duration of a snowball event), atmospheric CO2 ocean and to produce continental rainfall in a supergreenhouse
cannot rise above 0.24 bar, a level below the predicted melting environment (post-snowball) is limited by the total incoming
threshold determined by Pierrehumbert (2004). Thus, these solar energy. Le Hir et al. (2009) showed that above 0.11 bar of
results further exacerbate the problem of whether a snowball CO2, all incoming solar radiation is used as latent heat. Accounting
could actually be melted (Fig. 10.3). for a realistic geographical distribution of the continents and an
explicit modelling of the water cycle, the model predicts that the
continental runoff cannot exceed 1.2 times the present-day
Aftermath of the Snowball Earth runoff under more than 0.11 bar of CO2.
Nevertheless, assuming a sevenfold increase in the continental
Few studies have dealt with the global environment and climate runoff, Higgins & Schrag (2003) suggested that continental weath-
during the melting of a snowball. The extreme greenhouse atmos- ering (carbonate þ silicate rocks) may have increased by a factor
phere expected to be typical of the melting phase is not well of 50 directly after melting, relative to present-day values. As a
158 Y. GODDÉRIS ET AL.

result, normal values of atmospheric CO2 (around several hundred preceded by negative d13C anomalies (Prave et al. 2009),
ppmv) would be restored in 200 ka, starting from 0.126 bar. In supporting the idea that snowball events are based on
their study, Le Hir et al. (2009) found that CO2 consumption major perturbations to the carbon cycle, most likely very
through continental weathering cannot exceed 10 times the low pCO2. The general tectonic context of the Neoprotero-
present-day value, even accounting for the high surface area of zoic is probably the main precondition for the onset of
rock flour covering the continents, because weathering rates are global glaciation, facilitating atmospheric CO2 pumping by
inhibited by the moderate increase in precipitation. Consequently, continental weathering. The break-up of the Rodinia super-
it takes at least one million years for atmospheric CO2 to return to continent triggered an increase in continental weathering
pre-glacial values in response to silicate weathering in the through enhanced humidity above continental surfaces that
post-snowball environment. led to a generally cool climate during the Neoproterozoic.
One of the most striking features of the post-snowball environ- As long as the continents were dispersed along the equator,
ment is the accumulation of up 18.5 m (on average) of cap dolo- climate remained cool. Within this cool context, additional
stones on continental shelves, characterized by very low d13C triggers, such as the weathering of magmatic provinces
(c. –4‰, Pee Dee Belemnite Standard (PDB)). Heretofore, mod- erupted during the initial phase of the break-up, or atmos-
elling this peculiar phase of the snowball sequence has proven pheric methane collapse, were required to initiate snowball
challenging, because modelling the melting of huge icecaps glaciations.
under very high CO2 levels is not straightforward. For that (2) During the snowball. First, sea ice is allowed to reach the
reason, no explicit modelling has been performed and significant equator once the CO2 threshold for initiating a runaway
simplification is required. Regarding the d13C signature of cap car- ice-albedo feedback is reached. Whether this sea ice was
bonates, Higgins & Schrag (2003) calculated that the nadir of the thick or thin is still debatable, but the thin-ice solution
negative d13C excursion is observed 10 ka after the resumption of cannot be excluded, thus potentially allowing a vast refuge
continental weathering. This decrease in d13C directly after for primary producers. Like modern ice sheets, snowball
melting initiation is linked to the decrease in the carbon isotopic ice sheets on the continents grew to several kilometres in
fractionation between CO2 and carbonates, itself forced by a pre- thickness and were wet-based, consistent with the thick
scribed rise in sea surface temperature of 30 8C. In their model, accumulation of glacial deposits during the glaciations.
with a very high prescribed weathering flux, cap-dolostone Finally, massive dissolution of atmospheric CO2 into the
accumulation occurs in a few ka. In a similar study, coupling a ocean occurred if even a small surface area of the ocean
radiative –convective climate model to a global biogeochemical (3000 km2) remained open. In this scenario the ocean may
cycle model, Le Hir et al. (2008a) found that the most negative have been largely acidified during the snowball events,
d13C value was attained 100 ka after the resumption of continental with a pH as low 6, promoting the increase in the consump-
weathering, suggesting a longer time span for cap-dolostone tion of carbon by the oceanic crust by a factor of 4. Hence,
accumulation. This longer time span was also an outcome of the with such an efficient carbon sink, the CO2 accumulation
more extensive modelling in Le Hir et al. (2009). Accounting for would not be linear, but asymptotic, approaching around
the total amount of water released by the melting of the huge 0.25 bar of atmospheric CO2.
icecap (200  103 km3; Donnadieu et al. 2003), and estimating (3) Melting the snowball. Many unresolved questions remain
the total flux of Mg2þ and Ca2þ released by weathering during concerning snowball modelling, but the most important of
the transgressive phase with a reactive-transport model, Le Hir them concerns the initiation of melting. A significant con-
et al. (2009) calculated that only 50 cm of pure dolostone would clusion of the modelling studies is that a very high CO2
accumulate on the shelves in the first 10 ka, far below the estimated level is required to melt the snowball, probably .0.29 bar.
worldwide average thickness of 18.5 m (Hoffman et al. 2007). This level might never be reached, because carbon is con-
This slow rate of accumulation suggests that cap-dolostone depo- sumed by the enhanced dissolution of the basaltic oceanic
sition may have spanned longer than 10 ka, collaborating argu- crust. Efforts should focus on this problem, exploring pro-
ments for a longer timescale for cap-dolostone deposition based cesses that might reduce the albedo of the Earth and enable
on the existence of multiple magnetic reversal events measured melting at lower CO2 levels.
in various cap dolostones (Trindade et al. 2003; Raub & Evans (4) In the aftermath of the snowball glaciations. The alkalinity
2006). However, a corollary of this conclusion is that, because required to accumulate the cap dolostone results from the
the cap dolostone was strictly deposited during the post-glacial weathering of continental carbonate during the supergreen-
melting phase (Hoffman et al. 2007), it requires that deglaciation house following the snowball event. This post-snowball
was also protracted, and it is not clear if this is physically feasible. environment needs to be explored in more detail. It seems
Therefore, the timescale of melting and cap-dolostone precipi- that the most prominent and rapid environmental pertur-
tation remains an unresolved question, awaiting careful modelling bations occurred during the onset and subsequent relaxation
of the melting phase. of the supergreenhouse in the direct aftermath of the snow-
ball. Modelling studies contradict earlier assertions of a
highly vigorous hydrological cycle during the supergreen-
Discussion and conclusions house event, with runoff only as much as 20% higher com-
pared to its present-day value. Consequently, weathering
Do the climate models teach us something about the causes and rates were also lower, and rather than a few hundred thousand
consequences of the Neoproterozoic climatic events? The answer years (Hoffman et al. 1998; Higgins & Schrag 2003), it most
is surely yes if the models have a physical basis. They allow us likely took .1  106 years for atmospheric CO2 to return
to explore the space of parameters of the snowball glacial to pre-glacial values. These results also suggest that cap-
events, the variability of these parameters being fixed by field dolostone accumulation endured on the order of 100 ka.
studies or by the known physical behaviour of the climate. The
answer is no if they are mere mathematical manipulations per- If the conditions for initiation of the Snowball Earth seem to be
formed without heed to realistic physical parameters and geologi- reasonably well understood (Table 10.1), much work remains to
cal constraints. The most important results of the modelling studies resolve the melting and climate throughout the deglaciation. Cor-
can be summarized as follows: rectly simulating the duration of deglaciation with fully coupled
climate and ice-sheet models could provide important constraints
(1) Onset of snowball glaciations. At least two Cryogenian gla- on global cap-dolostone precipitation. However, nature is more
ciations were global in extent (Macdonald et al. 2010) and complex than models, and it is unrealistic to expect that models
MODELLING SNOWBALL EARTH 159

will ever capture the full complexity of the Neoproterozoic ice Donnadieu, Y., Ramstein, G., Fluteau, F., Roche, D. & Ganopolski,
ages. Nevertheless, models provide important clues and some A. 2004b. The impact of atmospheric and oceanic heat transport on
quantitative constraints on the controversial question of whether the sea ice-albedo instability during the Neoproterozoic. Climate
these catastrophic events really did occur. Dynamics, 22, 293– 306.
Donnadieu, Y., Ramstein, G., Goddéris, Y. & Fluteau, F. 2004c.
This represents a contribution of the IUGS- and UNESCO-funded IGCP (Inter- Global tectonic setting and climate of the Late Neoproterozoic: a
national Geoscience Programme) Project #512. climate–geochemical coupled study. In: Jenkins, G. S., McMena-
min, M. A. S., McKay, C. P. & Sohl, L. (eds) The Extreme Protero-
zoic: Geology, Geochemistry, and Climate. Geophysical Monograph
146. American Geophysical Union, Washington, DC, 79 – 89.
References Ernst, R. E. & Buchan, K. L. 2002. Maximum size and distribution
in time and space of mantle plumes: evidence from large igneous
Abbott, D. S. & Pierrehumbert, R. T. 2010. Mudball: surface dust and provinces. Journal of Geodynamics, 34, 309–342.
snowball Earth deglaciation. Journal of Geophysical Research, 115, Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic
doi: 10.1029/2009JD012007. constraints upon the Neoproterozoic climatic paradox. American
Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to Snowball Journal of Science, 300, 347–433.
Earth. Nature Geoscience, 1, 817– 825. François, L. M. & Goddéris, Y. 1998. Isotopic constraints on
Alt, J. C. & Teagle, D. A. H. 1999. The uptake of carbon during the the Cenozoic evolution of the carbon cycle. Chemical Geology,
alteration of ocean crust. Geochimica Cosmochimica Acta, 63, 145, 177–212.
1527– 1535. Gaillardet, J., Dupré, B., Louvat, P. & Allègre, C. J. 1999. Global
Baum, S. K. & Crowley, T. J. 2001. GCM response to late Precambrian silicate weathering and CO2 consumption rates deduced from the
(590 Ma) ice-covered continents. Geophysical Research Letters, chemistry of the large rivers. Chemical Geology, 159, 3– 30.
28, doi: 10.1029/2000GL011557. Galy, V., France-Lanord, C., Beyssac, O., Faure, P., Kudrass, H. &
Baum, S. K. & Crowley, T. J. 2003. The snow/ice instability as a mech- Palhol, F. 2007. Efficient organic carbon burial in the Bengal fan
anism for rapid climate change: a Neoproterozoic snowball Earth sustained by the Himalayan erosional system. Nature, 450, 407– 410.
model example. Geophysical Research Letters, 30, doi: 10.1029/ Goddéris, Y. & Joachimski, M. M. 2004. Global change in the late
2003GL017333. Devonian: modelling the Frasnian–Famennian short-term carbon
Bendtsen, J. 2002. Climate sensitivity to changes in the solar insolation isotope isotope excursions. Palaeogeography, Palaeoclimatology,
in a simple coupled climate model. Climate Dynamics, 18, 595– 609. Palaeoecology, 202, 309– 329.
Bendtsen, J. & Bjerrum, C. J. 2002. Vulnerability of climate on Earth Goddéris, Y. & Donnadieu, Y. 2008. Carbon cycling and snowball
to sudden changes in insolation. Geophysical Research Letters, 29, Earth. Nature, 456, doi: 10.1038/nature07653.
doi: 10.1029/2002GL014829. Goddéris, Y., Donnadieu, Y. et al. 2003. The Sturtian glaciation: fire
Budyko, M. I. 1969. The effect of solar radiation variations on the climate and ice. Earth and Planetary Science Letters, 211, 1 –12.
of the Earth. Tellus, 21, 611–619. Goddéris, Y., Donnadieu, Y. et al. 2007. Coupled modeling of global
Caldeira, K. & Kasting, J. F. 1992. Susceptibility of the early Earth to carbon cycle and climate in the Neoproterozoic: links between
irreversible glaciation caused by carbon dioxide clouds. Nature, 359, Rodinia breakup and major glaciation. Comptes Rendus Geoscience,
226– 228. 339, 212–222.
Chandler, M. A. & Sohl, L. E. 2000. Climate forcings and the initiation Goodman, J. C. 2006. Through thick and thin: marine and meteoric ice
of low-latitude ice sheets during the Neoproterozoic Varanger glacial in a ‘Snowball Earth’ climate. Geophysical Research Letters, 33,
interval. Journal of Geophysical Research – Atmospheres, 105, doi: 10.1029/2006GL026840.
20 737– 20 756. Goodman, J. C. & Pierrehumbert, R. T. 2003. Glacial flow of floating
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005. marine ice in ‘Snowball Earth’. Journal of Geophysical Research,
U –Pb ages from the Neoproterozoic Doushantuo Formation, China. 108, doi: 10.1029/2002JC001471.
Science, 308, 95 – 98. Halverson, G. P., Hoffman, P. F., Schrag, D. P. & Kaufman, A. J.
Courtillot, V., Jaupart, C., Manighetti, I., Tapponnier, P. & 2002. A major perturbation of the carbon cycle before the Ghaub
Besse, J. 1999. On causal links between flood basalts and continental glaciation (Neoproterozoic) in Namibia: prelude to snowball
breakup. Earth and Planetary Science Letters, 166, 177–195. Earth? Geochemistry, Geophysics, Geosystems, 3, doi: 10.1029/
Crowley, T. J. & Baum, S. K. 1993. Effect of decreased solar luminosity 2001GC000244.
on Late Precambrian ice extent. Journal of Geophysical Research, Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
98, 16 723– 16 732. Rice, A. H. N. 2005. Towards a Neoproterozoic composite carbon
Crowley, T. J., Hyde, W. T. & Peltier, W. R. 2001. CO2 levels required isotope record. Geological Society of America Bulletin, 117,
for deglaciation of a ‘Near-Snowball’ Earth. Geophysical Research 1181–1207.
Letters, 28, 283– 286. Higgins, J. A. & Schrag, D. P. 2003. Aftermath of a snowball Earth. Geo-
Dessert, C., Dupré, B., François, L. M., Schott, J., Gaillardet, J., chemistry Geophysics Geosystems, 4, doi: 10.1029/2002GC000403.
Chakrapani, G. J. & Bajpai, S. 2001. Erosion of Deccan Traps Hoffman, P. F. & Li, Z.-X. 2009. A palaeogeographic context for Neopro-
determined by river geochemistry: impact on the global climate and terozoic glaciation. Palaeogeography, Palaeoclimatology, Palaeo-
the 87Sr/86Sr ratio of seawater. Earth and Planetary Science ecology, 277, 158– 172.
Letters, 188, 459– 474. Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P.
Dessert, C., Dupré, B., Gaillardet, J., François, L. M. & Allègre, 1998. A Neoproterozoic Snowball Earth. Science, 281, 1342– 1346.
C. J. 2003. Basalt weathering laws and the impact of basalt weather- Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
ing on the global carbon cycle. Chemical Geology, 202, 257– 273. Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635
Donnadieu, Y., Ramstein, G., Fluteau, F., Besse, J. & Meert, J. 2002. Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary
Is high obliquity a plausible cause for Neoproterozoic glaciations? Science Letters, 258, 114–131.
Geophysical Research Letters, 29, doi: 10.1029/2002GL015902. Hyde, W. T., Crowley, T. J., Baum, S. K. & Peltier, W. R. 2000.
Donnadieu, Y., Fluteau, F., Ramstein, G., Ritz, C. & Besse, J. 2003. Neoproterozoic ‘snowball Earth’ simulations with a coupled
Is there a conflict between the Neoproterozoic glacial deposits and climate/ice sheet model. Nature, 405, 425– 429.
the snowball Earth interpretation: an improved understanding with Ishiwatari, M., Nakajima, K., Takehiro, S. & Hayashi, Y.-Y. 2007.
numerical modeling. Earth and Planetary Science Letters, 208, Dependence of climate states of gray solar atmosphere on solar con-
101– 112. stant: from the runaway greenhouse to the snowball states. Journal of
Donnadieu, Y., Goddéris, Y., Ramstein, G., Nédelec, A. & Meert, Geophysical Research, 112, doi: 1029/2006JD007368.
J. G. 2004a. Snowball Earth triggered by continental break-up Jenkins, G. S. 2000. Global climate model high-obliquity solutions to
through changes in runoff. Nature, 428, 303– 306. the ancient climate puzzles of the Faint-Young Sun Paradox and
160 Y. GODDÉRIS ET AL.

low-altitude Proterozoic Glaciation. Journal of Geophysical Marshall, H. G., Walker, J. C. G. & Kuhn, W. R. 1988. Long-term
Research, 105, 7357– 7370. climate change and the geochemical cycle of carbon. Journal of
Jenkins, G. S. 2004. High obliquity as an alternative hypothesis to early Geophysical Research, 93, 791– 801.
and late Proterozoic extreme climate conditions. In: Jenkins, G. S., McKay, C. P. 2000. Thickness of tropical ice and photosynthesis on a
McMenamin, M. A. S., McKay, C. P. & Sohl, L. (eds) The snowball Earth. Geophysical Research Letters, 27, 2153–2156.
Extreme Proterozoic: Geology, Geochemistry, and Climate. Geophy- Meert, J. G. & Powell, C. M. 2001. Assembly and break-up of Rodinia:
sical Monograph 146. American Geophysical Union, Washington, introduction to the special volume. Precambrian Research, 110, 1– 8.
DC, 73– 78. Meert, J. G. & Torsvik, T. H. 2004. Paleomagnetic constraints on
Jenkins, G. S. & Frakes, L. A. 1998. GCM sensitivity test using increased Neoproterozoic ‘Snowball Earth’ continental reconstructions. In:
rotation rate, reduced solar forcing and orography to examine low Jenkins, G. S., McMenamin, M. A. S., McKay, C. P. & Sohl, L.
latitude glaciation in the Neoproterozoic. Geophysical Research (eds) The Extreme Proterozoic: Geology, Geochemistry, and
Letters, 25, 3525–3528. Climate. Geophysical Monograph 146. American Geophysical
Jenkins, G. S. & Smith, S. R. 1999. GCM simulations of snowball Earth Union, Washington, DC, 5 –11.
conditions during the Late Proterozoic. Geophysical Research Micheels, A. & Montenari, M. 2008. A snowball Earth versus a slush-
Letters, 26, 2263–2266. ball Earth : results from Neoproterozoic climate modeling sensitivity
Key, R. M., Liyungu, A. K., Njamu, F. M., Somwe, V., Banda, J., experiments. Geosphere, 4, 401–410.
Mosley, P. N. & Armstrong, R. A. 2001. The western arm of the Park, J. K., Buchan, K. L. & Harlan, S. S. 1995. A proposed giant
Lufilian Arc in NW Zambia and its potential for copper mineraliz- dyke swarm fragmented by the separation of Laurentia and
ation. Journal of African Earth Science, 33, 503–528. Australia based on paleomagnetism of ca.780 Ma mafic intrusions
Kirschvink, J. L. 1992. Late Proterozoic low-latitude global glaciation: in western North America. Earth and Planetary Science Letters,
the snowball earth. In: Schopf, J. W. & Klein, C. (eds) The Proter- 132, 129– 139.
ozoic Biosphere. Cambridge University Press, Cambridge, 51 –52. Pavlov, A. A., Hurtgen, M. T., Kasting, J. F. & Arthur, M. A. 2003.
Knoll, A. H. 2003. Life on a Young Planet. Princeton University Press, Methane-rich Proterozoic atmosphere. Geology, 31, 87 –90.
New Jersey. Pavlov, A. A., Toon, O. B., Pavlov, A. K., Bally, J. & Pollard, D.
Leather, J., Allen, P. A., Brasier, M. D. & Cozzi, A. 2002. Neoproter- 2005. Passing through a giant molecular cloud: ‘snowball’ glaciations
ozoic snowball Earth under scrutiny: evidence from the Fiq glaciation produced by interstellar dust. Geophysical Research Letters, 32, doi:
of Oman. Geology, 30, 891–894. 10.1029/2004GL021890.
Le Hir, G., Ramstein, G., Donnadieu, Y. & Pierrehumbert, R. T. Peltier, W. R., Tarasov, L., Vettoretti, G. & Solheim, L. P. 2004.
2007. Investigating plausible mechanisms to trigger a deglaciation Climate dynamics in deep time: modeling the ‘snowball bifurcation’
from a hard snowball Earth. Comptes Rendus Geoscience, 339, and assessing the plausibility of its occurrence. In: Jenkins, G. S.,
274– 287. McMenamin, M. A. S., McKay, C. P. & Sohl, L. (eds) The
Le Hir, G., Goddéris, Y., Donnadieu, Y. & Ramstein, G. 2008a. A geo- Extreme Proterozoic: Geology, Geochemistry, and Climate. Geophy-
chemical modelling study of the evolution of the chemical compo- sical Monograph 146. American Geophysical Union, Washington,
sition of seawater linked to a snowball glaciation. Biogeosciences, DC, 107–124.
5, 253–267. Peltier, W. R., Liu, Y. & Crowley, J. W. 2007. Snowball Earth pre-
Le Hir, G., Goddéris, Y., Ramstein, G. & Donnadieu, Y. 2008b. vention by dissolved organic carbon remineralization. Nature, 450,
A scenario for the evolution of the atmospheric pCO2 during a 813– 818.
snowball Earth. Geology, 36, 47 –50. Petoukhov, V., Ganopolski, A., Brovkin, V., Claussen, M., Eliseev,
Le Hir, G., Donnadieu, Y. et al. 2009. The snowball Earth aftermath: A., Kubatzki, C. & Rahmstorf, S. 2000. CLIMBER-2: a
exploring the limits of continental weathering processes. Earth and climate system model of intermediate complexity. Part I: model
Planetary Science Letters, 277, 453–463. description and performance for present climate. Climate Dynamics,
Le Hir, G., Donnadieu, Y., Krinner, G. & Ramstein, G. 2010. Toward 16, 1 – 17.
the snowball Earth deglaciation. Climate Dynamics, 35, 285–297. Pierrehumbert, R. T. 2004. High levels of atmospheric carbon dioxide
Lewis, J. P., Weaver, A. J., Johnston, S. T. & Eby, M. 2003. Neopro- necessary for the termination of global glaciation. Nature, 429,
terozoic ‘snowball Earth’: dynamic sea ice over a quiescent ocean. 646– 649.
Paleoceanography, 18, doi: 10.1029/2003PA000926. Pierrehumbert, R. T. 2005. Climate dynamics of a hard Snowball Earth.
Lewis, J. P., Eby, M., Weaver, A. J., Johnston, S. T. & Jacob, R. L. Journal of Geophysical Research, 110, doi: 10.1029/2004JD005162.
2004. Global glaciation in the Neoproterozoic: reconciling previous Pollard, D. & Kasting, J. F. 2004. Climate-ice sheet simulations of
modelling results. Geophysical Research Letters, 31, doi: 10.1029/ Neoproterozoic glaciation before and after the collapse to snowball
2004GL019725. Earth. In: Jenkins, G. S., McMenamin, M. A. S., McKay, C. P. &
Lewis, J. P., Weaver, A. J. & Eby, M. 2006. Deglaciating the snowball Sohl, L. (eds) The Extreme Proterozoic: Geology, Geochemistry,
Earth: sensitivity to surface albedo. Geophysical Research Letters, and Climate. Geophysical Monograph 146. American Geophysical
33, doi: 10.1029/2006GL027774. Union, Washington, DC, 91 –105.
Lewis, J. P., Weaver, A. J. & Eby, M. 2007. Snowball versus slushball Pollard, D. & Kasting, J. F. 2005. Snowball Earth: a thin-ice solution
Earth: dynamic versus nondynamic sea ice? Journal of Geophysical with flowing sea glaciers. Journal of Geophysical Research, 111,
Research, 112, doi: 10.1029/2006JC004037. doi: 10.1029/2004JC002525.
Li, Z. X., Li, X. H., Kinny, P. D. & Wang, J. 1999. The break up of Pollard, D. & Kasting, J. F. 2006. Reply to comment by Stephen
Rodinia: did it start with a mantle plume beneath South China? G. Warren and Richard E. Brandt on ‘Snowball Earth: A thin-ice sol-
Earth and Planetary Science Letters, 173, 171–181. ution with flowing sea glaciers’. Journal of Geophysical Research
Li, Z. X., Li, X. H., Kinny, P. D., Wang, J., Zhang, S. & Zhou, H. 2003. 111, C09017, doi: 10.1029/2006JC003488.
Geochronology of Neoproterozoic syn-rift magmatism in the Yangtze Poulsen, C. J. 2003. Absence of a runaway ice-albedo feedback in the
Craton, South China and correlations with other continents: evidence Neoproterozoic. Geology, 31, 115– 118.
for a mantle superplume that broke up Rodinia. Precambrian Poulsen, C. J. & Jacob, R. L. 2004. Factors that inhibit snowball Earth
Research, 122, 85 – 109. simulation. Paleoceanography, 19, doi: 10.1029/2004PA001056.
Macdonald, F. A., Schmitz, M. D. et al. 2010. Calibrating the Cryo- Poulsen, C. J., Pierrehumbert, R. T. & Jacob, R. L. 2001. Impact of
genian. Science, 327, 1241–1243. ocean dynamics on the simulation of the Neoproterozoic ‘snowball
Macouin, M., Besse, J., Ader, M., Gilder, S., Yang, Z., Sun, Z. & Earth’. Geophysical Research Letters, 28, 1575–1578.
Agrinier, P. 2004. Combined paleomagnetic and isotopic data Poulsen, C. J., Jabob, R. L., Pierrehumbert, R. T. & Huynh, T. T.
from the Doushantuo carbonates, South China: implications for the 2002. Testing paleogeographic controls on a Neoproterozoic snow-
‘snowball Earth’ hypothesis. Earth and Planetary Science Letters, ball Earth. Geophysical Research Letters, 29, doi: 10.1029/
224, 387– 398. 2001GL014352.
MODELLING SNOWBALL EARTH 161

Prave, A. R., Fallick, A. E., Thomas, C. W. & Graham, C. M. 2009. A Torsvik, T. H., Carter, L. M., Ashwal, L. D., Bhushan, S. K., Pandit,
composite C-isotope profile for the Neoproterozoic Dalradian Super- M. K. & Jamtveit, B. 2001. Rodinia refined or obscured: paleomag-
group of Scotland and Ireland. Journal of the Geological Society, netism of the Malani igneous suite (NW India). Precambrian
London, 166, 1 –13. Research, 108, 319–333.
Ramstein, G., Donnadieu, Y. & Goddéris, Y. 2004. Proterozoic glacia- Trindade, R. I. F., Font, E., D’Agrella-Filho, M. S., Nogueira,
tions. Comptes Rendus Geoscience, 336, 639–646. A. C. R. & Riccomini, C. 2003. Low-latitude and multiple geomag-
Raub, T. D. & Evans, D. A. D. 2006. Magnetic reversals in basal Edia- netic reversals in the Neoproterozoic Puga cap carbonate, Amazon
caran cap carbonates: a critical review. Eos, Transactions American craton. Terra Nova, 15, 441–446.
Geophysical Union, 87, abstract GP41-02. Walker, J. C. G., Hays, P. B. & Kasting, J. F. 1981. A negative
Ritz, C., Rommelaere, V. & Dumas, C. 2001. Modeling the evolution of feedback mechanism for the long-term stabilization of Earth’s
Antarctic ice sheet over the last 420 000 years: implications for alti- surface temperature. Journal of Geophysical Research, 86,
tude changes in the Vostok region. Journal of Geophysical Research, 9776–9782.
106, 31 943– 31 964. Warren, S. G. & Brandt, R. E. 2006. Comment on ‘Snowball Earth:
Romanova, V., Lohmann, G. & Grosfeld, K. 2006. Effect of land a thin-ice solution with flowing sea glaciers’ by D. Pollard and
albedo, CO2, orography, and oceanic heat transport on extreme cli- J.F. Kasting. Journal of Geophysical Research, 111, C09016, doi:
mates. Climate of the Past, 2, 31 –42. 10.1029/2005JC003411.
Rose, B. E. J. & Marshall, J. 2009. Ocean heat transport, sea ice, and Warren, S. G., Brandt, R. E., Grenfell, T. C. & McKay, C. P.
multiple climate states: insights from energy balance models. 2002. Snowball Earth: ice thickness on the tropical ocean. Journal
Journal of Atmospheric Sciences, 66, 2828– 2843. of Geophysical Research, 107, doi: 10.1029/2001JC001123.
Schrag, D. P., Berner, R. A., Hoffman, P. F. & Halverson, G. P. 2002. Wingate, M. T. D., Campbell, I. H., Compston, W. & Gibson, G. G.
On the initiation of a Snowball Earth. Geochemistry, Geophysics, 1998. Ion microprobe U– Pb ages for Neoproterozoic basaltic mag-
Geosystems, 3, doi: 10.1029/2001GC000219. matism in south-central Australia and implications for the breakup
Sellers, W. D. 1969. A global climatic model based on the energy of Rodinia. Precambrian Research, 87, 135–159.
balance of the Earthatmosphere system. Journal of Applied Meteorol- Wingate, M. T. D. & Giddings, J. W. 2000. Age and paleomagnetism
ogy, 8, 392– 400. of the Mundine Well dyke swarm, Western Australia: implications
Stone, P. H. & Yao, M. S. 2004. The ice-covered Earth instability in a for an Australia– Laurentia connection at 755 Ma. Precambrian
model of intermediate complexity. Climate Dynamics, 22, 815– 822. Research, 100, 335–357.
Chapter 11

The record of Neoproterozoic glaciation in the Taoudéni Basin, NW Africa

G. A. SHIELDS-ZHOU1*, MAX DEYNOUX2 & LAWRENCE OCH1


1
Department of Earth Sciences, University College London, Gower Street, London, WC1E 6BT, UK
2
Laurélie, 12270 Bor et Bar, France
*Corresponding author (e-mail: g.shields@ucl.ac.uk)

Abstract: The Taoudéni Basin covers over 1 000 000 km2 of the West African Craton, bounded by Pan-African orogenic belts. Four
supergroups separated by craton-scale unconformities are recognized, with Neoproterozoic glaciogenic deposits occurring at the base
of Supergroup 2. The Jbéliat Group occurs along a continuous, 1300-km-long, narrow belt from the Adrar region of Mauritania to
the eastern limit of the Hank in Algeria and comprises thin glacial drift capped widely by periglacial polygonal structures, with more
complex glacial sequences preserved in palaeo-depressions. A thicker, variously marine and continental glaciogenic succession can
be found in southern parts, while fully marine, glacially influenced successions are only known from the extreme SW of the basin.
The ‘triad’ sequence of diamictites overlain by barite-bearing ‘cap’ dolostones and then by green shales and/or bedded cherts (silexites)
is ubiquitous and has long been used to correlate the Supergroup 1/2 boundary across the basin and into the surrounding orogenic belts.
The bedded cherts commonly show a volcanic influence and are cemented by early marine calcite at their base at Adrar, Mauritania.
Although fossil-based age constraints are scarce and ambiguous, regional tectonic events indicate that ‘triad’ deposition occurred
between the Bassaride (665– 655 Ma) and Dahomeyide (610–580 Ma) orogens. Recent U –Pb zircon studies of ignimbrite tuffs
provide a minimum age for the glaciation of c. 600 Ma. Correlation of supergroup 2 glacial deposits with the c. 635 Ma end-Cryogenian
(‘Marinoan’) glaciation is likely and is supported by limited carbon and strontium isotope data. Barite is commonly found within the cap
carbonate and may relate to methane seepage and/or unusual oceanographic conditions after deglaciation. Several studies have attributed
sequence complexity within the post-glacial succession to isostatic reequilibration. The Taoudéni Basin represents a rare Neoproterozoic
example of terrestrial tillites and associated periglacial facies.

On the West African Craton, the Neoproterozoic and Palaeozoic average), continuous tabular blanket over distances of 1000–
sedimentary cover can be subdivided into four supergroups (or 1500 km (Fig. 11.1). The Taoudéni Basin was possibly a less
megasequences; e.g. Deynoux et al. 2006) bounded by craton- rigid part of the platform, and is rimmed by the Mauritanide and
scale unconformities (Trompette 1973). The lithostratigraphic Hoggar-Iforas fold belts. No substantial sediment thickening
‘triad’ association (diamictite-cap dolostone-bedded chert), which occurs in the central part of the basin, but extra subsidence has
forms the base of Supergroup 2, has long been used as a marker caused some thickening in peripheral troughs and sub-basins,
horizon (Zimmermann 1960; Leprun & Trompette 1969). The such as the Gourma Basin to the east and the Madina-Kouta and
first detailed studies of the glacial sedimentology of this region Bové basins to the west, which were involved in the Pan-African
were conducted during the 1970s, and what follows relies and/or Hercynian fold belts.
heavily on over three decades of study by French researchers With the exception of the area fringing the peripheral fold belts,
who recorded the extraordinary preservation of glacial and perigla- the Taoudéni Basin has been affected by regional-scale (epeiro-
cial features across vast areas of arid North Africa (e.g. Deynoux genic) tilting, which has generated low angular unconformities,
1982; Deynoux et al. 2006). This chapter mainly focuses on and the reactivation of basement faults frequently intruded by
some better known sections of the Taoudéni Basin that have diabase sills and dykes. The sedimentary deposits are generally
recently been the subject of isotopic studies (Alvaro et al. 2007; devoid of tectonism, while metamorphism is limited to the
Shields et al. 2007a, b). Owing to the large size of the Taoudéni thermal effects of diabase intrusions in the proximal host material
Basin, these descriptions are necessarily oversimplified and that took place when the Atlantic Ocean began opening during
the reader is referred to the published literature for more detailed the Jurassic Period. In this regard, some sedimentary rocks may
information on these and other (e.g. Bassaride, Rokelide, Anti- have experienced locally high-grade, hydrothermal alteration. In
Atlas, Hoggar-Iforas and Dahomeyide belts, and the Tindouf spite of their age, the sediments have undergone only modest
Basin) Neoproterozoic glaciogenic successions of the West burial diagenesis because the sedimentary pile is relatively thin
African craton (Deynoux 1980, 1985; Deynoux & Trompette, and the platform has remained above sea level since the end of
1981; Proust & Deynoux 1994; Deynoux et al. 2006). Glaciogenic the Carboniferous Period.
deposits of the neighbouring Volta Basin (Fig. 11.1) are presumed The Volta Basin thickens eastward, where it becomes gently
to be correlative with those of the Taoudéni Basin Supergroup 2 folded before disappearing beneath the Dahomeyide thrust belt.
and will also be considered here. According to sparse contextual information (Deynoux et al.
2006), the Volta Basin can be considered either to be a single fore-
land basin or as part of the Pan-African passive margin overlain by
Structural framework and basin setting a foreland basin made up of the Tamale Supergroup, lateral equiva-
lents of which can be found in intermontane, molasse-filled
The West African Craton has been a tectonic entity since at least grabens of the Dahomeyide belt (Affaton et al. 1991).
1600 Ma and is rimmed by mobile belts that became active
during late Proterozoic and Palaeozoic times and are related to
the Transaharan suture zone (Fig. 11.1). On the craton, the sedi- Stratigraphy
mentary cover is preserved in widespread, variably connected
basins containing upper Proterozoic and Palaeozoic deposits of On the West-African platform, the Neoproterozoic and Palaeozoic
which the largest is the Taoudéni Basin, forming a thin (3 km on sedimentary cover can be subdivided into four supergroups

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 163– 171. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.11
164 G. A. SHIELDS-ZHOU ET AL.

Fig. 11.1. Simplified geological map of the


Taoudéni Basin and adjacent areas in NW
Africa (Source: Deynoux et al. 2006).

(megasequences) separated by craton-scale unconformities strata of the Char, Atar and Assabet el Hassiane Groups
(Trompette 1973). Supergroup 1 rests with major unconformity (Fig. 11.2). The Atar Group and overlying siliciclastic rocks of
upon metamorphic and granitic basement across the Taoudéni the Assabel-el-Hassiane Group in Mauritania experienced minor
Basin and contains sedimentary strata of Mesoproterozoic age tilting during Pan-African events before being peneplained by sub-
(Clauer et al. 1982; Rooney et al. 2010). Supergroup 2 begins sequent uplift and erosion to form a largely flat, sub-glacial sub-
with glacial deposits resting with an erosional and slightly strate (Trompette 1973; Deynoux, 1980). Glacial deposits and
angular unconformity upon Supergroup 1 or directly upon the their ‘cap dolostone’ form the Jbéliat Group, which outcrops
basement. The glacial deposits are capped by a thin, but continuous nearly continuously for 1300 km along a narrow belt from the
unit of calcareous dolomite, which contains barite in places. Adrar region to the eastern limit of the Hank in Algeria (Fig. 1).
Bedded cherts and green shales with a thin but continuous lime- The Adrar cap carbonate package consists of one or two dolostone
stone bed at their base generally overlie this cap-dolostone unit units, with an intervening siliciclastic package of up to 40 metres
and form the uppermost part of the so-called ‘triad’ of diamictite- thickness, and a laterally extensive, thin limestone bed that discon-
cap dolostone-bedded chert. Supergroup 3 consists of Late Ordo- formably overlies the uppermost dolostone forming the base of the
vician glacial deposits and post-glacial Silurian shales, whereas bedded cherts and shales of the Téniagouri Group.
Supergroup 4 corresponds to unconformably overlying Devonian In the Walidiala Valley, which straddles the Guinea-Senegal
and Carboniferous strata. border at the extreme southwestern margins of the Taoudéni
In the Adrar region of Mauritania, glacial deposits of Super- Basin, Supergroups 1 and 2 overlie basement rocks of the
group 2 are unconformably underlain by early Neoproterozoic Kenieba Inlier and are intruded and protected from erosion by
NEOPROTEROZOIC GLACIATION IN THE TAOUDÉNI BASIN 165

Fig. 11.2. Stratigraphy of the northern


part of the Taoudéni Basin in the Adrar area
and the Hank-Fersiga area (after Deynoux
et al. 2006). Ages are from Rooney et al.
2010 and Lahondère et al. 2005.

plateau-forming diabase sills of Jurassic age. Supergroup 1 con- which represents the glacially related deposits, and the Nandou-
sists of intertidal to supratidal, coarse-grained, red wackestones mari Formation, up to 130 m thick, which comprises at its base
and intervening siltstone forming the Ségou Group. The lower quartz arenites overlain by dolostone, siltstone and bedded chert
part of Supergroup 2 is represented by the Mali Group, which (Shields et al. 2007a).
rests on the Mesoproterozoic Ségou Group with deep erosional The southern region of the Taoudéni Basin stands out due to the
unconformity. The Mali Group (Fig. 11.3) consists of siliciclastic, presence below the triad and associated craton-wide unconformity
mostly siltstone units and comprises in its basal part the regionally of a 400–500-m-thick complex succession showing glacial influ-
correlative triad of glaciogenic strata, dolostone and bedded chert. ence (Fig. 11.4). Both triad and underlying glacially influenced
The Mali Group has been subdivided into two formations (Culver strata were initially assigned to the ‘Bakoye Group’ (Simon 1979);
& Hunt 1991): the Hassanah Diallo Formation, 50 –120 m thick, however, the marked erosional and angular unconformity below

Fig. 11.3. Schematic geological map and


stratigraphic column of the Mali Group of
the southwestern Taoudéni Basin in the
Walidiala Valley (modified after Shields
et al. 2007a).
166 G. A. SHIELDS-ZHOU ET AL.

of the Reguibat Shield (Fig. 11.1) with inferred glacial movement


southward on the platform and laterally towards oceanic troughs or
basins located at the present position of the Pan-African belts.
Accordingly, continental glaciation is recorded in the northern
part of the Taoudéni Basin by a thin, irregular (0 –50 m thick)
veneer of terrestrial tillites with subordinate proglacial outwash
deposits preserved in limited shallow depressions. Towards the
south of the platform, the glacial drift thickens (150 –200 m),
showing marine influence in small intracratonic basins in western
Mali (Kayes area) and becoming wholly marine at the margins
of oceanic troughs bordering the Pan-African Bassaride Belt in
eastern Senegal and Guinea.
Along the northern margin of the Taoudéni Basin, glacial drift
comprises just a few metres of terrestrial tillite, with polygenic
and often striated pebbles to boulders overlying striated pave-
ments. Locally, a complex facies association is preserved in
smooth large-scale palaeo-depressions such as in the Jbéliat area
in Adrar, Mauritania (Fig. 11.1). In these palaeo-depressions, the
glacial deposits thicken up to 50 m and comprise distinct glacial
sequences suggesting at least three glacial advances and retreats,
with terrestrial tillites, outwash sandstones and various glacial fea-
tures such as striated clasts and pavements, roches moutonnées,
lacustrine varves with dropstones, and remarkable polygonal struc-
tures and sand wedges related to permafrost just below the cap
dolostone (Deynoux & Trompette 1976; Deynoux 1980, 1982,
1985). These deposits represent probably one of the only well-
preserved, purely continental records of a Pre-Pleistocene ice
sheet (Eyles & Januszczak 2004). Although true tillites are rarer
in the southern part of the Taoudéni Basin, diamictites of the
Koniakari Group and associated sandstones and shales form an
up to 200-m-thick marine to continental glacial succession well
exposed in the Kayes area of Mali (Rossi et al. 1984; Deynoux
Fig. 11.4. Stratigraphy and main depositional environments of the glacially
et al. 1991). In areas of only marine sedimentation like the Wali-
related deposits in the southwestern part of the Taoudéni Basin in Mali diala Valley of Senegal, glacial influence can only be recognized
(Tambaoura plateaux) (modified from Deynoux et al. 2006). due to the occurrence of dropstones between debris flow deposits
of the Pelel Member and in shales of the overlying Diagoma
Member (Culver & Hunt, 1991; Shields et al. 2007a).
the diamictites of the triad suggests that the underlying succession The post-glacial cap-carbonate succession in Adrar has been
records a distinct history. Deynoux et al. (1989, 2006) subsequently described recently by Shields et al. (2007b) and Alvaro et al.
subdivided the Bakoye Group into a Koniakari Group, correspond- (2007), and only a brief summary is given here. The generally buff-
ing to the diamictites and post-glacial carbonate unit of the triad, and coloured cap dolostone is either well-bedded or brecciated with
a Wassangara Group, corresponding to the underlying, at least thin and discontinuous intervening stringers of black chert.
partly glacially influenced succession. The Wassangara Group Bedding surfaces are commonly erosional at all scales, that is,
forms most of the Tambaoura and Mandingue Plateaux in western scoured, rather than gradational or crinkled as with microbialites,
Mali and part of the Afollé Massif in southern Mauritania and show crossbedding only rarely. When passing to the brecciated
(Fig. 11.1). It rests with an erosional and locally angular unconfor- facies, beds become laterally and vertically folded and disrupted.
mity upon the Souroukoto Group. Fissures are common and are generally filled with ferroan
The Volta Basin (Fig. 11.5) succession rests with major dolomite, black chert, white quartz or barite. A second generation
unconformity on the Eburnean basement of the Leo Shield. The of fissures is filled with very coarse ferroan calcite spar. Micro-
lower (Bombouaka) Supergroup is up to 1000 m thick, comprising scopically, the Jbéliat dolostone comprises alternating grain sizes
largely siliciclastic rocks interspersed with thin calcareous beds in from dolomicrite to dolosparite with poorly developed or residual
its middle part, and is in general reminiscent of the lower part of cement made up of a mixture of clay, cryptocrystalline silica, side-
Supergroup 1 in the southern part of the Taoudéni Basin. The rite and pyrite.
middle (Pendjari or Oti) Supergroup is much thicker (2500 – The cap dolostone splits locally into two distinct stratigraphic
4000 m) and begins with glaciogenic deposits at its base, resting horizons providing space for 40 (Adrar) to 80 m (Hank) of onlap-
unconformably on various parts of the Bombouaka Supergroup ping green to purple shales and siltstones including sandy dolos-
or directly on basement. The triad of diamictite, dolostone and tone intercalations and passing upward in the Hank region into
chert passed upwards into green turbiditic siltstones and shales, sandy tidal deposits and eolian sandstone (Moussine-Pouchkine
which contain phosphorite deposits. The upper (Tamale) Super- & Bertrand-Sarfati 1997). Along the Atar cliff, the upper dolostone
group (500 m thick) is mainly exposed in Ghana and lies uncon- horizon becomes progressively sandy until it forms a bank up to
formably on the Pendjari Supergroup. It comprises terrestrial 7 m thick made up entirely of coarse-grained to granular, trough
coarse-grained siliciclastic deposits with intercalated fine-grained crossbedded, dolomitic sandstone with isolated lenses of finer
and calcareous horizons. dolostone. Barite is associated with the brecciated facies at
Jbéliat, Mauritania (mostly in the upper cap-dolostone unit) and
at Pont de Kabaté in Mali (Shields et al. 2007b), and appears as
Glaciogenic deposits and associated strata palisadic crystals filling cracks or cavities, as well as monocrystal-
line geodes and laminated domes up to 30 cm across.
Palaeogeographic reconstructions of the triad-associated glacia- In the Jbéliat area and part of the Atar cliff area, over
tion of NW Africa imply an ice sheet centred towards the north c. 30– 40 km of the cap-dolostone exposure, a persistent
NEOPROTEROZOIC GLACIATION IN THE TAOUDÉNI BASIN 167

2°W 0° 2°E Panafrican internal units


100 km
12° Suture zone

Kara struct. Unit


Panafrican
Atacora struct. Unit external units
Buem struct. Unit
10°

Tamale Megasequence
A
Flyschoid succession
Peniari or Oti
Tillite and cap dolostone Megasequence
B

KFZ

Yemboure Gr.

Fosse aux lions Gr. Bambouaka


Megasequence
Dapaong Gr.

6° Eburnean basement
c Ocean
Atlanti
Thrust

A B

Fig. 11.5. Simplified geological map and a


synthetic cross-sections of the Volta Basin
50 km and Dahomeyide belt (modified after
Deynoux et al. 2006).

30– 50-cm-thick horizon of limestone is intercalated between Individual beds have sharp or erosional bases and fine upwards
the cap dolostone and bedded cherts/siltstones of the overlying from a surface of detrital dolomite/quartz lag of c. 100 mm size
Téniagouri Group. It is characterized by its purple to grey-green to dolomicrite of ,10 mm in diameter (Shields et al. 2007a).
colour and its brecciated texture defined by oriented, millimetre- The Volta Basin tillite fills an erosional and slightly angular
to centimetre-scale angular flakes of detrital, largely volcanogenic unconformity of glacial origin with striated glacial pavements
material and authigenic barite (Hoffman & Schrag 2002; Shields along NW –SE trending palaeo-valleys. Its modest thickness
et al. 2007b), cemented by early marine calcite. At both Jbéliat (c. 2 m) and geological context suggests that this deposit represents
and the Atar Cliffs area, this limestone horizon may comprise a ‘true’ terrestrial tillite, which is overlain by a ,6 m cap dolos-
reverse-graded, calcite nodules or spherulites, which exhibit tone (Porter et al. 2004). At Bwipe (Ghana), the cap dolostone
cores of volcanogenic chert. The contact of the purple limestone unit is thicker, comprising up to 12 m of finely laminated,
unit with the underlying cap dolostone is generally sharp and ‘microbial’ dolostone (Nedelec et al. 2007).
marked by an erosional hardground surface; however, its contact
with the overlying cherts is transitional.
In the marine succession at Walidiala Valley, the Bowal
Member dolostone similarly consists of centimetre-to-decimetre Boundary relations with overlying and underlying
bedded, internally laminated dolostone and dolomitic siltstone non-glacial units
beds, overlying similar siltstone containing abundant lonestones
(dropstones). The dolostone unit shows widespread slumping Tilting during Pan-African orogenic events and largely pre-glacial
associated with the arrival of debris comprising volcanogenic peneplanation caused glacial and glacially influenced units of
diamict that underwent pervasive dolomitization after redeposi- Supergroup 2 to be deposited unconformably upon marine sedi-
tion. The bedded dolostone unit is characterized by various mentary rocks of Supergroup 1 throughout the Taoudéni Basin,
forms of brittle deformation: fractures, tepee-like buckling and and apparently across NW Africa, including in the Volta Basin.
discontinuous fitted brecciation with created space being filled In the Adrar region of Mauritania, the top of the glacial drift is
by early chert and/or by dolospar. In undeformed beds, this fine- marked by highly indurated, black, coarse-grained sandstone,
grained dolostone exhibits millimetre-scale layering defined by which forms patterned ground with polygonal structures and
mechanical laminations with no textural evidence for any sand wedges typical of permafrost (Deynoux 1982). Laminated
microbial influence. Microscopically, the Bowal Member dolo- dolostone drapes the immediately post-glacial topography or
stone comprises equigranular grains of dolomite (generally rests directly upon the underlying and commonly striated perigla-
.90%) with accessory quartz and biotite grains of similar size. cial substrate in both Taoudéni and Volta basins.
168 G. A. SHIELDS-ZHOU ET AL.

In the Walidiala Valley section in Senegal, glaciomarine influ- sandwiched between terrestrial strata showing fluvial, eolian and
ence wanes up section as dropstones become both smaller and glacial influence (Deynoux et al. 1991).
rarer in mudstones of the Diagoma Member, the top 10 m of Simple discoidal impressions were reported from strata below
which seem to be devoid of any dropstones. In places, channels glaciogenic deposits by Bertrand-Sarfati et al. (1995) in the
cut into these mudstones and are filled with coarse-grained, cross- Hank region of northern Mauritania; however, these are not
bedded sandstone and gravels of the Tanagué Member. A 2– 7 m necessarily of Ediacaran age, as simple circular impressions are
thick, silty cap dolostone, the Bowal Member, caps the glaciogenic also known from Cryogenian strata in Canada and older Protero-
sequence and rests erosionally on the underlying siltstone due to zoic strata elsewhere (e.g. Cruse & Harris 1994). The discovery
slumping or possibly conformably on coarse-grained siliciclastics of Cambrian-type small shelly fossils, including the proto-
(Shields et al. 2007a). gastropod Aldanella attleborensis in dolostone scree in Walidiala
Valley, Senegal (Culver et al. 1988) caused some to infer a Cam-
brian age for the Taoudéni Basin glaciation (Bertrand-Sarfati et al.
Chemostratigraphy 1995; Evans 2000). Although the block was believed to derive
from the post-glacial Bowal Member cap dolostone, subsequent
Only limited isotope work has been carried out in Neoprotero- investigations have not been able to confirm the findings and
zoic strata of the Taoudéni and Volta basins. Earlier studies recent publications tend to play down their significance (Porter
reported Sr- and C-isotope data from the type section of the uncon- et al. 2004; Deynoux et al. 2006; Alvaro et al. 2007; Shields
formably underlying Atar Group, in Mauritania (Veizer et al. et al. 2007a, b).
1983; Fairchild et al. 1990), but only recently have studies
focused on post-glacial strata across the basin (Alvaro et al.
2007; Shields et al. 2007a, b). Despite pervasive recrystallization Palaeolatitude and palaeogeography
and multiple phases of cementation, which makes primary
carbon-isotope trends (e.g. the Jbéliat section, Mauritania, and There are no firm palaeomagnetic constraints on the position of
the Goumaré section, Mali) hard to recognize, published results the West African craton during the Neoproterozoic. Existing data
reveal a consistent picture of a post-glacial cap dolostone unit are consistent with a position at mid to high latitudes; however,
marked by low d13CPDB (averaging – 4‰) trending up-section to the primary nature of all published results can be questioned
more negative values; this is most clearly seen in the Walidiala (Evans 2000).
Valley sections of Senegal and the Koniakari section, Mali.
Lowermost d13C values (– 6‰) are reported from a limestone
unit at the base of the Téniagouri Group, which overlies post- Geochronological constraints
glacial cap-dolostone strata in the Atar Cliffs area of Mauritania
(Alvaro et al. 2007). Shields et al. (2007b) trace a trend to more Early geochronological studies of the 1970s and 1980s applied
positive values through this limestone unit with d13C as high as K –Ar and Rb –Sr isotope techniques to mudstone units on the
þ3.7‰ in places. High-resolution isotopic data from four sections West African craton and demonstrated a late Neoproterozoic age
of the Volta Basin (Burkina Faso: Arli, Kodjari, Koundjouari; and for the ‘Infracambrian’ glaciogenic units of Supergroup 2
Ghana: Bwipe) show a similarly decreasing upward trend (Porter (Clauer et al. 1982; Clauer & Deynoux 1987). Subsequent work
et al. 2004; Nedelec et al. 2007). Positive correlation between has largely confirmed this viewpoint, although some confusion
d13C and d18O appears to be characteristic (Porter et al. 2004; was injected by the discovery (Culver et al. 1988) of Cambrian-
Shields et al. 2007a). type fossils in dolostone scree of the Walidiala Valley, Senegal
Barite samples from Mali and Mauritania have been analysed (see discussion above). The most robust age constraints derive
for Sr and S isotopes (Shields et al. 2007b). 87Sr/86Sr ratios are from U – Pb studies of zircons from bedded tuffs within the
consistent between barite deposits (0.70773 –0.70814) with the lower Téniagouri Group, Mauritania (Lahondère et al. 2005).
results of leaching experiments indicating that Sr in the barite These new age constraints (609.7 + 5.5 Ma, U –Pb TIMS zircon
was derived from seawater with 87Sr/86Sr ¼ 0.7077 –8. In con- and 606 + 6 Ma, U –Pb SHRIMP zircon) provide a minimum
trast, barite d34S compositions range widely, exhibiting values age for the glacial Jbéliat Group and its cap dolostone of about
between þ20‰ and þ46‰. This indicates that barite sulphate 600 Ma, which is consistent with correlation with the comparably
derives from a marine source with d34S of about þ20‰, and well-dated c. 635 Ma glaciogenic successions of South China
that it has undergone considerable isotopic fractionation due to sul- (Condon et al. 2005) and Namibia (Hoffmann et al. 2004).
phate reduction (Shields et al. 2007b). Sr isotope compositions Deynoux et al. (2006) discuss existing age constraints in more
have also been reported for the cap dolostone at Bwipe (Volta detail and conclude that triad deposition across the Taoudéni
Basin), where they range from 0.7061 to 0.7073 (Nedelec et al. Basin, and likely also in the Volta Basin, took place between the
2007). Such low values are atypical for basal Ediacaran dolo- Bassaride orogen (Pan-African I, 665– 655 Ma) and the Daho-
stones, which generally exhibit variably radiogenic ratios (e.g. meyide orogen (610 – 580 Ma).
Yoshioka et al. 2003) that are unlikely to be representative of Isotope studies of post-glacial dolostone units across the basin
ambient seawater. (Shields et al. 2007a, b; Alvaro et al. 2007) support this interpret-
ation (see Discussion below), and fossils are scarce and do not
provide additional age constraints. Although some glaciogenic
Other characteristics successions of the Taoudéni Basin may show multiple glacial
advances and retreats, for example, in Mali (Deynoux et al.
Economic deposits associated with the glaciogenic strata are 1991), there is at present no firm evidence for any older, mid-
limited. Barite deposits are commonly associated with the cap Cryogenian (‘Sturtian’) glaciation. According to Villeneuve
dolostone and are of high purity but are too localized to be of econ- (1988) and Villeneuve & Cornée (1994), the glacially influenced
omic interest. The Atar and Assabet el Hassiane groups and their Wassangara Group, which lies unconformably below the triad in
equivalents below the glacial level in Mauritania and Mali (as the Kayes area of Mali (Proust & Deynoux 1994), also postdates
well as the Silurian shales above the Late Ordovician glacial depos- the collision between the West African craton and a western Sene-
its) have attracted interest as petroleum source rocks. Despite a gas galese block at 660 Ma (Pan-African 1 orogeny). The above con-
show in the 1970s, only four wells are known to have been drilled to straints permit regional correlation of Neoproterozoic to Cambrian
date (Craig et al. 2009). Banded haematite cherts (BIFs) are found strata of the West African craton and surrounding Pan-African
within shales of the Wassangara Group (Taoudeni Basin, Mali), belts (Deynoux et al. 2006, fig. 14).
NEOPROTEROZOIC GLACIATION IN THE TAOUDÉNI BASIN 169

Discussion The transgressive unit consists of a regionally extensive, 2 –


7-m-thick, silty dolostone, the Bowal Member, which is isotopi-
The glacial formations on the rim of the northern Taoudéni Basin cally and petrographically indistinguishable from c. 635 Ma cap-
consist of terrestrial tillites and proglacial outwash sandstones, dolostone units elsewhere in NW Africa and worldwide (Shields
which are capped in the Adrar region of northern Mauritania by et al. 2007a). A large volcaniclastic debris flow has caused slump-
a persistent horizon of large-scale polygonal structures and sand ing and soft-sediment deformation within the cap dolostone of the
wedges. These polygonal structures are evidence that a prolonged Bowal Member. The widespread association of pyroclastic depos-
periglacial period with temperatures low enough for the develop- its with cap dolostone in geographically distant successions of the
ment of permafrost (Deynoux 1982) followed the retreat of gla- Taoudéni Basin and neighbouring Hoggar-Iforas Belt (Deynoux
ciers in this part of the craton. Analogous polygonal structures et al. 2006) implies that volcanism and deglaciation were roughly
and sand wedges can also be found 500 km farther south in the contemporaneous across a large area. Early cementation is demon-
Afollé Massif, south of Kiffa, indicating that permafrost was not strated in this case by pervasive dolomitization of the debris flow
restricted to the Adrar region but extended over a large part of matrix and supports the notion of unusually high carbonate satur-
the Taoudéni Basin. Post-glacial cap dolostones of the northern ation levels during the aftermath of latest Cryogenian glaciation
Taoudéni Basin mark an abrupt change in the regional depositional (Hoffman & Schrag 2002).
environment from terrestrial to marine. The origin of the barite mineralization in the post-glacial
The cap dolostones and overlying shales (with calcitic cherts at carbonate succession was specifically addressed by Shields et al.
the base) of the Téniagouri Group are considered to relate to degla- (2007b), who considered that barite formed during the second
cial to post-glacial eustatic transgression, respectively. However, marine transgression but before deposition of the overlying
the presence of intervening shales and sandstones, levelling Téniagouri Group shales and bedded cherts. Isotopic evidence
palaeoreliefs (see above) as observed in the Adrar and Hank suggests that barite precipitated due to the mixing of sulphate-rich
regions, highlights the complexity of this transgression. According seawater with a more reducing Ba-rich fluid. The consistent
to Aı̈t-Kaci Ahmed & Moussine-Pouchkine (1994), and Bertrand- association of barite occurrences throughout the basin with terres-
Sarfati et al. (1997), in the Hank Algerian border region (Guetta- trial periglacial deposits (Deynoux 1980, 1982) and the sporadic
tira-Grizim-Fersiga area), the cap carbonates and intervening occurrence of the barite permit the suggestions that barite mineral-
shales, which contain intercalations of phosphatic grainstones, ization was controlled by local seepage of permafrost methane
were deposited during the glacioeustatic transgression. Owing following deglaciation. Such a scenario implies a relatively short
to the balance between glacioeustasy and isostatic rebound, the time interval between deglaciation and barite genesis, which is
cap carbonates were locally subjected on topographic highs to sub- somewhat supported by the general agreement between barite
aerial exposure with the development of a complex phosphatic Sr- and S-isotope values and those of other basal Ediacaran
profile, including mini-stromatolites, while sedimentation contin- cap-carbonate successions (Shields et al. 2007b). The common
ued in the depressions. Intervening green shales pass progressively association of capdolostone to limestone transitions with barite
upward into sandy tidal deposits and eolian sandstones in the mineralization could, however, point to a global oceanographic
Hank. These sandstones are overlain by a transgressive surface origin for the barite, whereby deeper Ba-rich seawater beneath a
covered by a few metres of conglomeratic and phosphatic sand- persistent pycnocline mixed with sulphate-rich surface seawater
stones that preceded the large-scale development on the whole after deglaciation.
platform of shales equivalent to the Téniagouri Group in Maurita- The negative d13C trend shown in Taoudéni Basin post-glacial
nia. Accordingly, the post-glacial cap-carbonate succession across dolostones is matched by identical trends in correlative sections
the Taoudéni Basin may represent a condensed deposit that relates of the neighbouring Volta Basin (Porter et al. 2004; Nedelec
to the onset of the glacioeustatic transgression, with the transgres- et al. 2007) and worldwide (Kennedy 1996; Halverson et al.
sive and highstand system tracts preserved only in palaeo- 2004), all from dolostone units immediately overlying glaciogenic
depressions. strata interpreted to belong to the latest Cryogenian (commonly
As in the Hank, the cap dolostone in the Adrar reveals a referred to as ‘Marinoan’) glaciation. The affiliation of the Taou-
complex, relative sea-level record, with the post-glacial ‘purple’ déni Basin triad with other diamictite – cap dolostone associations
limestone representing a condensation horizon that marks the from around the world is consistent with the abovementioned geo-
beginning of a second, marine transgression, apparently coinciding chronological constraints and Sr- and S- isotope data, thus lending
with regional volcanism. Mineral relationships confirm that further support to the notion of a global deglaciation at c. 635 Ma.
regional volcanism occurred after palaeo-topographic highs had In this regard, the limestone unit reported from above post-glacial
for a time become emergent, allowing karst dissolution to occur. cap dolostones in Adrar, Mauritania (Alvaro et al. 2007; Shields
Carbonate saturation levels were high enough to grow calcite crys- et al. 2007b) may be equivalent to isotopically similar post-glacial
tals rapidly on and in caverns immediately beneath the seafloor limestones around the world (cf. Hoffman & Schrag 2002). The
before glassy debris could disintegrate. The transition from dolo- anomalously low 87Sr/86Sr ratios reported by Nedelec et al.
mite to calcite precipitation is similar to other such transitions (2007) remain inexplicable; they could relate to diagenetic isotopic
around the world (Hoffman & Schrag 2002), and could be correla- exchange with juvenile volcanic minerals or to the fact that
tive based on the co-occurrence of barite and isotopic trends. contemporaneous seawater 87Sr/86Sr was significantly lower
At the base of the Neoproterozoic-age Mali Group of the south- than previously believed.
western Taoudéni Basin, debris flows and turbidite-like, sandy
units of the Pelel Member pass upward into siltstone and shale This represents a contribution of the IUGS- and UNESCO-funded IGCP (Inter-
of the Diagoma Member. These two units represent the progressive national Geoscience Programme) Project #512.
evolution from some portion of a fan delta fed by a nearby ice shelf
to a more distal environment disturbed only by the occasional
References
fallout from passing icebergs. The appearance of coarse-grained,
crossbedded sandstone beds and gravels of the overlying Affaton, P., Rahaman, M. A., Trompette, R. & Sougy, J. 1991. The
Tanagué Member heralds a return to a shallower, fluvially influ- Dahomeyide orogen: tectonothermal evolution and relationships
enced environment before abrupt transgression caps the glacio- with the Volta Basin. In: Dallmeyer, R. D. & Lécorché, J. P.
genic succession. The succession in the Walidiala Valley could (eds) The West African Orogens and Circum-Atlantic Correlatives.
represent therefore the stratigraphic expression of a glacial Springer-Verlag, New York, 107–122.
retreat in a proximal glaciomarine environment affected by gla- Aı̈t-Kaci Ahmed, A. & Moussine-Pouchkine, A. 1994. Les formation
cioeustasy and isostatic rebound (Shields et al. 2007a). cambriennes de Fersiga (Sud-Ouest du Tanezrouft): nouvelle
170 G. A. SHIELDS-ZHOU ET AL.

interprétation de la sédimentation glaciaire et post-glaciaire sur le Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for
Craton Ouest Africain. Bulletin du Service géologique d’Algérie, Neoproterozoic glaciations during the breakup of Rodinia after 750
5, 3 –21. Ma. Earth Science Reviews, 65, 1– 73.
´ lvaro, J. J., Macouin, M., Bauluz, B., Clausen, S. & Ader, M.
A Fairchild, I. J., Marshall, J. D. & Bertrand-Sarfati, J. 1990. Strati-
2007. The Ediacaran sedimentary architecture and carbonate pro- graphic shifts in carbon isotopes from Proterozoic stromatolitic car-
ductivity in the Atar Cliffs, Adrar, Mauritania: palaeoenvironments, bonates (Mauritania): influences of primary mineralogy and
chemostratigraphy and diagenesis. Precambrian Research, 153, diagenesis. American Journal of Science, 290A, 46– 79.
236– 261. Halverson, G. P., Maloof, A. C. & Hoffman, P. 2004. The Marinoan
Bertrand-Sarfati, J., Moussine-Pouchkine, A. Q., Amard, B. & Aı̈t glaciation (Neoproterozoic) in northeast Svalbard. Basin Research,
Kaci Ahmed, A. 1995. First Ediacaran fauna found in western 16, 297– 324.
Africa and evidence for an Early Cambrian glaciation. Geology, 23, Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
133– 136. testing the limits of global change. Terra Nova, 14, 129–155.
Bertrand-Sarfati, J., Flicoteaux, R., Moussine-Pouchkine, A. & Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
Aı̈t Kaci Ahmed, A. 1997. Lower Cambrian apatitic stromatolites 2004. U– Pb zircon date from the Neoproterozoic Ghaub Formation,
and phospharenites related to the glacio-eustatic cratonic rebound Namibia: constraints on Marinoan glaciation. Geology, 32, 817– 820,
(Sahara, Algeria). Journal of Sedimentary Research, 67, 957– 974. doi:10.1130/G20519.l.
Clauer, N. & Deynoux, M. 1987. New information on the probable Kennedy, M. J. 1996. Stratigraphy, sedimentology, and isotopic geo-
isotopic age of the Late Proterozoic glaciation in West Africa. Pre- chemistry of Australian Neoproterozoic postglacial cap dolostones:
cambrian Research, 37, 89– 94. deglaciation, d13C excursions and carbonate precipitation. Journal
Clauer, N., Caby, R., Jeanette, D. & Trompette, R. 1982. Geochro- of Sedimentary Research, 66, 1050– 1064.
nology of sedimentary and metasedimentary Precambrian rocks of Lahondère, D., Roger, J. et al. 2005. Notice explicative des cartes géo-
the West African Craton. Precambrian Research, 18, 53– 71. logiques à 1/200,000 et 1/500,000 de l’extrème sud de la Mauritanie.
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005. DMG, Ministère des mines et de l’industrie, Nouakchott, Rapport
U –Pb ages from the Neoproterozoic Doushantuo Formation, China. BRGM/RC-54273-FR, 610.
Science, 308, 95 –98. Leprun, J.-C. & Trompette, R. 1969. Subdivision du Voltaı̈en du
Craig, J., Thurow, J., Thusu, B., Whitham, A. & Abuttaruma, Y. massif de Gobnangou (République de Haute-Volta) en deux séries
2009. Global Neoproterozoic Petroleum Systems: The Emerging discordantes séparées par une tillite d’âge éocambrien probable.
Potential in North Africa. Geological Society, London, Special Comptes Rendus de l’Académie des Sciences, Paris, 269,
Publications, 326. 2187– 2190.
Cruse, T. & Harris, L. B. 1994. Ediacaran fossils from the Stirling Range Moussine-Pouchkine, A. & Bertrand-Sarfati, J. 1997. Tectonosedi-
Formation, Western Australia. Precambrian Research, 67, 1 –10. mentary subdivisions in the Neoproterozoic to Early Cambrian
Culver, S. J. & Hunt, D. 1991. Lithostratigraphy of the Precambrian – cover of the Taoudenni Basin (Algeria, Mauritania, Mali). Journal
Cambrian boundary sequence in the southwestern Taoudeni Basin, African Earth Sciences, 24, 425– 443.
West Africa. Journal of African Earth Sciences, 13, 407–413. Nedelec, A., Affaton, P., France-Lanord, C., Charriere, A. &
Culver, S. J., Pojeta, J. & Repetski, J. E. 1988. First record of Early Alvaro, J. 2007. Sedmentology and chemostratigraphy of the
Cambrian shelly microfossils from West Africa, Geology, 16, Bwipe Neoproterozoic cap dolostones (Ghana, Volta Basin): a
695– 599. record of microbial activity in a peritidal environment. Comptes
Deynoux, M. 1980. Les formations glaciaires du Précambrien terminal Rendus Geoscience, 339, 223– 239.
et de la fin de l’Ordovicien en afrique de l’Ouest. Deux exemples Porter, A. M., Knoll, A. H. & Affaton, P. 2004. Chemostratigraphy of
de glaciation d’inlandsis sur une plate-forme stable. Travaux du Lab- Neoproterozoic cap carbonates from the Volta Basin, West Africa.
oratoire des Sciences de la Terre St-Jerome, Marseille (B), 17, 554. Precambrian Research, 130, 99 – 112.
Deynoux, M. 1982. Periglacial polygonal structures and sand wedges in Proust, J. N. & Deynoux, M. 1994. Marine to non-marine sequence
the late Precambrian glacial formations of the Taoudeni Basin in architecture of an intracratonic glacially related basin. Late Protero-
Adrar of Mauritania (West Africa). Palaeogeography, Palaeoclima- zoic of the West African platform in western Mali. In: Deynoux,
tology, Palaeoecology, 39, 55– 70. M., Miller, J. M. G., Domack, E. W., Eyles, N., Fairchild, I. J.
Deynoux, M. 1985. Terrestrial or waterlain glacial diamictites? Three & Young, G. M. (eds) Earth’s Glacial Record. Cambridge University
case studies from the Late Precambrian and Late Ordovician glacial Press, Cambridge, 121– 145.
drifts in West Africa. Palaeogeography, Palaeoclimatology, Rooney, A. D., Selby, D., Houzay, J.-P. & Renne, P. R. 2010.
Palaeoecology, 51, 97– 141. Re –Os geochronology of Mesoproterozoic sediments from
Deynoux, M. & Trompette, R. 1976. Late Precambrian mixtites: the Taoudeni basin, Mauritania: implications for basin-wide corre-
glacial and/or nonglacial? A discussion dealing especially with the lations, supercontinent reconstruction and Re– Os systematics of
mixtites of West Africa. American Journal of Science, 276, organic-rich sediments. Earth and Planetary Science Letters, 289,
1302– 1315. 486– 496.
Deynoux, M. & Trompette, R. 1981. Late Precambrian tillite of the Rossi, P., Deynoux, M. & Simon, B. 1984. Les formations glaciaires
Taoudeni Basin, West Africa. In: Hambrey, M. J. & Harland, du Précambrien terminal et leur contexte stratigraphique
W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge (formations pré et post-glaciaires et dolérites permiennes(?) du
University Press, New York, 123– 134. massif du Kaarta (dans le bassin de Taoudéni au Mali occidental
Deynoux, M., Marchand, J. & Proust, J. N. 1989. Notice explicative de (Afrique de l’Ouest). Sciences Géologiques Bulletin, Strasbourg,
la carte géologique du Mali occidental az 1/200.000. Feuilles Kan- 37, 91 – 106.
kossa, Kayes, Kossanto. République du Mali, Direction Nationale Shields, G. A., Deynoux, M., Culver, S. J., Brasier, M. D., Affaton,
de la Géologie et des Mines, Bamako. Klöckner Industrie-Anlagen, P. & Vandamme, D. 2007a. Neoproterozoic glaciomarine and cap
Duisburg, 54– 81. dolostone facies of the southwestern Taoudéni Basin (Walidiala
Deynoux, M., Proust, J. N. & Simon, B. 1991. Late Proterozoic glacially Valley, Senegal/Guinea, NW Africa). Comptes Rendus de l’Acadé-
controlled shelf sequences in western Mali (West Africa). Journal mie des Sciences: Geosciences, 339, 186– 199.
African Earth Sciences, 12, 181–198. Shields, G. A., Deynoux, M., Strauss, H., Paquet, H. & Nahon, D.
Deynoux, M., Affaton, P., Trompette, R. & Villeneuve, M. 2006. 2007b. Barite-bearing cap carbonates of the Taoudéni Basin, north-
Pan-African tectonic evolution and glacial events registered in west Africa: sedimentary and isotopic evidence for methane
Neoproterozoic to Cambrian cratonic and foreland basins of West seepage from permafrost after a Neoproterozoic glaciation. Precam-
Africa. Journal of African Earth Sciences, 46, 397– 426. brian Research, 153, 209–235.
Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic Simon, B. 1979. Essai de synthèse sur les formations sédimentaires de la
constraints upon the Neoproterozoic climatic paradox. American partie occidentale du Mali. Rapport Inédit Laboratoire Géologie
Journal of Science, 300, 347– 433. Dynamique Univ. Aix-Marseille III, 133.
NEOPROTEROZOIC GLACIATION IN THE TAOUDÉNI BASIN 171

Trompette, R. 1973. Le Précambrien supérieur et le Paléozoique inférieur Villeneuve, M. & Cornée, J. J. 1994. Structure, evolution and palaeo-
de l’Adrar de Mauritanie (bordure occidentale du bassin de Taoudeni, geography of the West African craton and bordering belts during
Afrique de l’Ouest). Un exemple de sédimentation de craton, Etude the Neoproterozoic. Precambrian Research, 69, 307– 326.
stratigraphique et sédimentologique. Travaux du Laboratoire des Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003.
Sciences de la Terre St-Jerome, Marseille (B), 7. Systematic variations in C, O, and Sr isotopes and elemental
Veizer, J., Compston, W., Clauer, N. & Schidlowski, M. concentrations in Neoproterozoic carbonates in Namibia: impli-
1983. 87Sr/86Sr in Late Proterozoic carbonates: evidence for a cations for glacial to interglacial transition. Precambrian Research,
‘mantle’ event at 900 Ma ago. Geochimica et Cosmochimica 124, 69– 85.
Acta, 47, 295– 302. Zimmermann, M. 1960. Nouvelle subdivision des séries antégothlan-
Villeneuve, M. 1988. Evolution comparée du bassin de Taoudéni et de la diennes de l’Afrique occidentale (Mauritanie, Soudan, Sénégal),
chaine des Mauritanides en Afrique de l’Ouest. Comptes Rendus Aca- Rap. 21st International Geological Congress, Copenhagen, 8,
démie des Sciences, Paris, 307, 663– 668. 26– 36.
Chapter 12

Neoproterozoic glaciogenic diamictites of the Katanga Supergroup, Central Africa

SHARAD MASTER1* & MAREK WENDORFF2


1
Economic Geology Research Institute, School of Geosciences, University of the Witwatersrand, P. Bag 3, WITS
2050, Johannesburg, South Africa
2
AGH University of Science and Technology, Faculty of Geology, Geophysics and Environmental Protection,
al. A. Mickiewicza 30, 30-059 Krakow, Poland
*Corresponding author (e-mail: sharad.master@wits.ac.za)

Abstract: Glaciogenic sediments of the Katanga Supergroup are represented by two units. The syn-rift Grand Conglomerat Formation
(,765 + 5 Ma to .735 + 5 Ma) occurs within the Nguba Group, and the Petit Conglomerat Formation defines the base of the Kunde-
lungu Group deposited in the earliest foreland basin of the Lufilian orogenic belt located between the Congo and Kalahari cratons. Their
glacial origin is inferred on the basis of the following features: the common and widespread occurrence of thick polymictic conglomerates
and diamictites with faceted and striated clasts, massive structure, abundant poorly sorted fine-grained matrix, and the presence of planar-
laminated shales (laminites) with dropstones. Glaciomarine facies associations prevail over most of the geographic extent of both units,
but at the northern periphery of the depository, continental glacial facies are present. The glaciomarine units are succeeded by carbonates:
the Kakontwe Limestone and ‘Calcaire Rose’ respectively. The clasts in the glaciogenic units are of extrabasinal and intrabasinal pro-
venance. Lower boundaries, conformable in the basin centre, evolve to unconformities in the marginal areas to the N and S. The palaeo-
magnetic evidence suggests deposition in low latitudes.

The Neoproterozoic – Lower Palaeozoic Katanga Supergroup origin of these diamictites based on a number of characteristics,
of Central Africa (Katanga Province of Democratic Republic of such as the great lateral continuity and thickness of these beds,
Congo, Zambia, and eastern Angola) is exposed in the Lufilian and the extrabasinal origin and faceted and striated nature of
belt (part of the continental system of the Pan-African orogenic the clasts.
belts of Africa), and also forms a less deformed plateau The Grand Conglomerat Formation has also been designated
molasse/foreland sequence over the Congo Craton (Fig. 12.1). as unit Ki1.1 (François 1973) and as the Mwale Formation, or
Katangan sediments were initially deposited in intracratonic Ng1.1 (Batumike et al. 2007), while the Petit Conglomerat has
rifts related to early Neoproterozoic extension along the southern been referred to as the Ks1.1 (François 1973), and the Kyandamu
margin of the Congo Craton. Continental break-up accompanied Formation or Ku1.1 (Batumike et al. 2007). The two diamictites
by mafic volcanism led to the formation of major unconformities, are developed over a very extensive area, covering at least
and development of a passive margin (Nguba Group). Subsequent 65 000 km2, occurring roughly in the region between latitudes 88
deformation in late Neoproterozoic to early Palaeozoic orogenic and 138S, and longitudes 248 and 298E (Figs 12.2 & 12.3).
belts (Lufilian-Zambezi), and syntectonic sedimentation in associ-
ated foreland basins, was related to the collision of the Congo and
Kalahari Cratons during the amalgamation of the Gondwana
Supercontinent (Hanson 2003; Wendorff 2005a, b; Master et al. Structural and stratigraphic framework
2005).
Regional mapping of Katanga (at a scale of 1:500 000) first The Katanga Supergroup was deposited on a basement comprising
commenced with the work of Franz Edward Studt, Jules Cornet the Palaeoproterozoic Lufubu Metamorphic Complex and the
and Henri Buttgenbach (Studt et al. 1908; Studt 1908, 1913). Bangweulu Block (Rainaud et al. 2005a; De Waele et al. 2006),
The first identification of glacial diamictites in Katanga was and the Mesoproterozoic Kibaran Belt (Kokonyangi et al. 2006).
made by Stutzer (1911, 1913a) and by Grosse (1912), but both Traditionally, the Katanga Supergroup has been subdivided into
authors misidentified these Katangan diamictites as belonging to the Roan, Lower and Upper Kundelungu Groups (Cailteux et al.
the late Carboniferous Dwyka glaciation. Robert (1912a, b) first 1994; Kampunzu & Cailteux 1999). According to the recently
identified glacial diamictites belonging to the Kundelungu revised stratigraphy (Wendorff 2003, 2005b, c; Wendorff & Key
System in the Katanga region on the basis of striated pebbles. 2009), the Katanga Supergroup is subdivided into the following
Hennig (1915) regarded the glacial beds of Central Africa ident- five groups: Roan (basal group), Nguba, Kundelungu, Fungurume
ified by Stutzer (1911, 1913a) and Grosse (1912) as being much and Plateau (Fig. 12.2). Not all researchers in the region have
older than the Dwyka. Different facies were identified in the accepted the revised stratigraphy (e.g. Cailteux et al. 2007;
Katangan diamictites by Delhaye (1920). As a consequence of Batumike et al. 2006, 2007).
more detailed regional mapping, at a scale of 1:200 000, by van The Roan and Nguba groups record two distinct rifting stages
Doorninck (1928) and by Maurice Robert from 1926 to 1931, in resulting from early Neoproterozoic extension within Rodinia.
which diamicitites were used extensively as easily recognized The Kundelungu, Fungurume and Plateau groups were deposited
regional marker horizons, it was finally established by Robert in the succeeding foreland basins related to collision of the
(1933), Gysin (1934) and Grosemans (1935) that there are two dia- Congo and Kalahari cratons during Gondwana assembly. The
mictites, the Grand Conglomerat and Petit Conglomerat, which Grand Conglomerat Formation occurs within the Nguba Group
were regarded as forming, respectively, the base of the Lower and records syn-rift glaciation after initial basaltic volcanism.
and Upper ‘Series’ (‘Serie Inférieur’ and ‘Serie Supérieur’), con- The succeeding Petit Conglomerat Formation formed after the
stituting the upper part of the Kundelungu ‘System’. Van Door- first orogenic event to affect the southern part of the Lufilian belt
ninck (1928) and Robert (1940a, b, 1947) argued for a glacial in what is now Zambia.

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 173– 184. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.12
174 S. MASTER & M. WENDORFF

(siliciclastic unit, or the Mindola Subgroup) at the base, grading


upwards into a mixed association of siliciclastic and carbonate
strata, the Kitwe Subgroup, succeeded by a carbonate platform
sequence, the Bancroft Subgroup, which prograded from the
south (Binda 1994).
Major uplift (765 Ma) in the southern part of the Roan rift
basin (Zambia) terminated deposition of the Roan Group platform
carbonates and led to the opening of the Nguba rift (Wendorff
2005b). Wendorff (2005b, c) has interpreted the breccias of the
Mufulira Formation (lower Mwashya) as syn-rift olistostromes,
derived from uplifted Roan strata as products of mass-wasting
and deposited by sediment-gravity flows at the base of the
Mwashya Subgroup. Northward expansion of the Nguba rift,
beyond the northern margin of the Roan rift, resulted in prograda-
tion of the olistostromes and their nonconformable deposition
upon pre-Katangan basement in what is now the fold-thrust
belt region in the Democratic Republic of Congo (DRC). The suc-
ceeding Mwashya strata are composed of terrigenous siliciclastic
rocks, silicified oolitic/pisolitic grainstones, algal dolomites and
ironstones (middle Mwashya), overlain by shales with siltstones
grading upwards to black shales deposited under increasingly
Fig. 12.1. Position of the Lufilian arc in the Pan-African orogenic belts system
anoxic conditions (upper Mwashya). Basaltic volcaniclastic
in central and southern Africa. Box outlines the area shown in Fig. 12.2.
rocks and lavas mostly form subordinate interbeds in the
Mwashya Subgroup, but are locally significant. For example, a
The Roan Group (Fig. 12.4a), deposited in the first Katangan rift broad belt of basaltic lava can be traced in a NE direction
basin (,880 Ma; Armstrong et al. 2005), nonconformably over- from eastern Angola across NW Zambia and into southern DRC
lies variably eroded pre-Katangan basement and forms a continu- (Unrug 1987, 1988; Kampunzu et al. 1993, 2000; Tembo et al.
ous transgressive succession from terrigenous clastic sediments 1999; Key et al. 2002). In the centre of the basin, the upper

Fig. 12.2. Regional geology of the Lufilian belt (modified from Porada 1989; Porada & Berhorst 2000; Wendorff 2003). Localities discussed in the text: L, Lwaio
Mission; M, Mwinilunga; Ks, Kansanshi; MF, Musonda Falls; Ma, Mansa; M-K, Makonga-Kibambale. 1– 10, positions of stratigraphic logs shown in Fig. 12.5.
NEOPROTEROZOIC GLACIOGENIC DIAMICTITES OF THE KATANGA SUPERGROUP 175

Fig. 12.3. Distribution of the Grand


Conglomerat and Petit Conglomerat
diamictites in Katanga (Democratic
Republic of Congo) and in adjacent areas of
Zambia (modified after Cahen &
Lepersonne 1979).

Fig. 12.4. (a) Simplified version of revised stratigraphy of the Katangan succession based upon syntectonic conglomerate complexes and unconformity-bounded
megasequences (Wendorff 2005b; Wendorff & Key 2009). Ages after Key et al. (2002) and Master et al. (2005), and references therein. Wavy lines indicate major
unconformities. (b) Previous/traditional subdivisions of the Katangan and lithostratigraphic correlation between the DRC and Zambia (simplified from Cailteux et al.
1994). Ore horizons: *Cu in Lower Roan in Zambia; **Cu– Co in the Mines Group in the DRC. Glaciogenic horizons: GC, Grand Conglomerat; PC, Petit Conglomerat.
Note that the Mines Group in the DRC consists of allochthonous megablocks, which resulted from dismemberment of nappes/thrust sheets with the Roan Group strata
and gravity-driven emplacement of the mineralized Roan megablocks into the foreland basin (the nappes originated in the south of the orogen (now Zambia) and were
thrust towards the foreland region in the north (today’s DRC) during the Pan-African orogenesis). Therefore the Mines Group does not exist as an individual stratigraphic
unit (Wendorff 2003, 2005b), and the megablocks are embedded in the Fungurume Group strata, as shown in (a).
176 S. MASTER & M. WENDORFF

boundary of the Mwashya Subgroup is transitional into the Grand Lufilian arc, and in the northern parts of the Katangan basin, the
Conglomerat (Binda & Van Eden 1972; Bodiselitsch et al. 2005) metamorphic grade is very low, and sedimentary structures are
and is defined by the appearance of dropstones in massive or lami- very well preserved (Master et al. 2005).
nated black shales typical of the underlying upper division of
Mwashya Subgroup. Mafic igneous rocks that occur within the
Roan, Mwashya and Grand Conglomerat have been interpreted Glaciogenic deposits and associated strata
as indicators of rifting continuing throughout the deposition of the
Roan and Nguba groups (Tembo et al. 1999; Kampunzu et al. Grand Conglomerat Formation, Nguba Group
2000). The northward propagation of the rifting is attributed
to the formation of the Kundelungu Aulacogen, which was The diamictite of the Grand Conglomerat has long been interpreted
filled with up to 7 km of Neoproterozoic sediments, including the as glaciogenic (Cahen 1978). The evidence for a glacial origin rests
two glacial diamictites, the Grand and Petit Conglomerates on the common and widespread occurrence (over an area of around
(Dumont & Hanon 1997; Master 2007). Sedimentary strata that 65 000 km2) of thick (up to a maximum of 1200 m) polymictic
overlie the Grand Conglomerat include the Kakontwe Limestone deposits containing subrounded to subangular faceted clasts
Formation. Haematitic and jaspilitic banded Fe-formations are (ranging from gravel-sized fragments to boulders .1 m3 in
known to occur both in the Mwashya Subgroup below the Grand volume) with striations, sometimes in multiple sets (in more than
Conglomerat and in Nguba Group strata above it (Robb et al. 20 localities); the generally massive, poorly sorted nature of the dia-
2003, 2004). mictite with a fine-grained matrix supporting the coarser clasts; and
Inversion from an extensional to a compressional tectonic the presence of associated planar-laminated shales with dropstones
regime occurred in the south of the Lufilian belt after the depo- (van Doorninck 1928; vanden Brande 1936; Cahen 1954, 1963,
sition of the Grand Conglomerat but before the Petit Conglomerat 1978; François 1973; Binda & van Eden 1972; Dumont & Cahen
was deposited (Wendorff 2005b). West of the Kafue Anticline in 1977; Wendorff & Key 2009). Striated clasts have been described
Zambia, the Petit Conglomerat rests unconformably on a folded by many authors and illustrated by Stutzer (1911), Studt (1913),
succession of the Roan-Nguba groups (Wendorff & Key 2009). Grosse (1918), Robert (1946) and Cahen (1963).
It defines the base of the Kundelungu Group, which is an infill of The clasts in the Grand Conglomerat diamictite are overwhel-
the first foreland basin (Wendorff 2005b), itself superposed on mingly composed of quartzite and vein quartz, but also include a
the former Kundelungu aulacogen (Master 2007), and is composed variety of other lithologies, such as granites, gneisses, mica-schists,
of marine sandy shales, shales and dolomites, with thick proximal porphyries and basic rocks (van Doorninck 1928; Robert 1940a;
conglomerate complexes occurring in the south of the Lufilian belt. Master et al. 2005). François (1973), working in the western end
The Petit Conglomerat Formation is overlain by a pink dolomite of the Lufilian arc, recorded clasts in the Grand Conglomerat
(the Calcaire Rose). made up predominantly (.75%) of quartzites (fine-grained,
The succeeding Fungurume Group fills the second foreland coarse-grained, rarely micro-conglomeratic) derived from the
basin formed in the northern part of the fold-thrust region of the Kibaran Belt. In the Makonga and Kibambale regions (Fig. 12.3)
Lufilian belt in the DRC (Wendorff 2003, 2005b) and contains in the NW Katangan Basin, adjacent to the Kibaran Belt, Dumont
synorogenic conglomerates and megablocks derived from nappes & Cahen (1977) recorded clasts of quartzites, biotite granite, phyl-
composed of the older Katangan strata uplifted to the south and lite, graphic pegmatite, tourmalinite, quartz, calcic marbles, jasper,
thrust northwards (compare Fig. 12.4a, b). The lower boundary carnelian agate, black silicified pisolites, black chert, amygdaloidal
is unconformable on folded Kundelungu strata, and the succession spilitic lavas and dolerite. Clasts made of oolitic cherts were
evolves from sedimentary olistostromes through transitional recorded by Cornet (1897), Studt (1913), Stutzer (1913b), Guille-
shallow marine and continental redbeds to shallow marine main (1913), Grosse (1918) and Dumont & Cahen (1977),
sequences of siliciclastic and carbonate strata. The degree of defor- especially in the vicinity of Mwashya in the northern part of the
mation gradually decreases between the northern marginal part of Lufilian arc.
the external fold-thrust belt and the succeeding undeformed In the central Lufilian arc, most clasts have rounded edges, but
Plateau Group (also known as the Biano Group) further to the completely rounded clasts are absent (van Doorninck, 1928). In
north. Continental arkoses and shales of the latter unit were depos- the Luapula beds, clasts in clast-supported cobble and boulder
ited in the youngest foreland basin, extending to the north of conglomerates are well rounded, but clasts in matrix-supported
the Fungurume Group foreland (Figs 12.2 & 12.4). 40Ar/39Ar poorly sorted pebbly sandstones and conglomerates are subangular
dating of detrital muscovite from the Plateau Group shows that to subrounded (Abraham 1959). A study of clast shapes in the
these sedimentary rocks were deposited after 573 + 5 Ma Grand Conglomerat diamictites from two localities in the Likasi
(Master et al. 2005). district was made by Museu (1987), who found that the quartzite
The deformation of the Katanga Supergroup to form the Lufilian clasts (c. 70% of the population) show a very poor sorting, but
arc occurred during the collision of the Kalahari and Congo the median clast size diminished from NE towards the SW.
cratons, in the time interval from c. 590 to c. 512 Ma, followed More than 80% of the quartzite clasts were rounded or subrounded,
by a protracted period of post-orogenic uplift and cooling extend- and a flattening index showed values typical for torrential environ-
ing to c. 483 Ma (Porada & Berhorst 2000; John et al. 2003, 2004; ments; hence these diamictites were interpreted by Museu (1987)
Rainaud et al. 2005b). During the Lufilian Orogeny, up to 150 km to have been deposited in fluviatile or glaciofluvial environments.
of crustal shortening occurred (Porada & Berhorst 2000; Jackson Petrographic studies of the Grand Conglomerat were initiated
et al. 2003). Thick-skinned thrusting resulted in the formation of by Beck and Wagner, as reported by Stutzer (1911). Grosse
basement-cored domes in the Domes Region (Daly et al. 1984) (1918) and van Doorninck (1928) described the tillite as having
(Fig. 12.2), while thin-skinned thrusting resulted in the external a matrix of extremely fine-grained, ash-grey, almost opaque
fold-thrust belt (Jackson et al. 2004). Former Roan rift basins suf- clayey material, with small (up to 0.75 mm), splinters and more
fered tectonic inversion during basin deformation (Selley et al. rounded fragments of predominantly quartz, with sporadic rare
2005). In the folded Lufilian fold-thrust belt, the metamorphic fine-grained fragments of granite, graphic granite, quartz por-
overprint ranges from greenschist to amphibolite grade, with phyry, feldspars, muscovite, chlorite, quartzite, magnetite, horn-
local development of talc-kyanite whiteschists (John et al. 2004), blende and other minerals. The petrography of the Grand
and the diamictites are deformed and cleaved, with recrystalliza- Conglomerat at Kipushi indicates that it contains numerous
tion of matrix constituents to biotite-bearing schists (Gysin 1934; clasts derived from a mixed plutonic (granitic and amphibolitic)
Key et al. 2002; Broughton et al. 2002; Master et al. 2005; and metavolcanic (quartz porphyry schist) terrain, such as the
Rainaud et al. 2005b). However, in the external parts of the Lufubu Metamorphic Complex (Master et al. 2005). Detailed
NEOPROTEROZOIC GLACIOGENIC DIAMICTITES OF THE KATANGA SUPERGROUP 177

modern petrographic studies of the Grand Conglomerat diamictites Proximal facies of a .400-m-thick gravelly fan-delta succes-
have found microfabrics and microstructures in these diamictites sion at Mushishima, west of the Kafue Anticline in Zambia
that are comparable with those found in modern glaciogenic tills (Figs 12.5 & 12.6), contain two interlayers of the glaciogenic sedi-
(Delpomdor 2007; Delpomdor et al. 2008). ments (Wendorff & Key 2009). The fan-delta conglomerates are
massive, matrix-supported and form thick amalgamated beds
that locally contain dish structures testifying to rapid deposition
Facies in the Grand Conglomerat and early post-depositional fluid escape. Clast-supported conglom-
erate layers are rare. Together with pebbles, cobbles and solitary
The Grand Conglomerat sediments represent a broad range of con- boulders of quartzite, arkose and siltstone, there occur clasts of car-
tinental, shallow marine and deep marine facies controlled by the bonate rocks derived from the Roan Group.
sedimentary environment, ongoing tectonic evolution of the A more distal facies association of a proglacial fan delta occurs
depository, climatic variations and changes in ice dynamics. at Kansanshi to the west (Fig. 12.2), where a c. 200-m-thick

Fig. 12.5. Lithostratigraphic sections


discussed in this chapter (from Wendorff &
Key 2009, and references therein). Sections
2– 5 are modified from Wendorff (2005b),
section 6 from Andersen & Unrug (1984),
section 7 after Key et al. (2002), sections 8
and 9 from François & Cailteux (1981), and
sections 10 and 11 after Dumont & Cahen
(1977). The thicknesses of
lithostratigraphic units in section 7 are
approximate due to variable dips and
poor exposure.
178 S. MASTER & M. WENDORFF

Fig. 12.6. Schematic SW to NE


cross-sections through the Katangan Basin
during deposition of the Grand
Conglomerat, showing the Nguba Rift
superposed on the earlier Roan Rift (see
Fig. 12.2 for locations). (a) Eastern region –
Kafue Anticline and Bangweulu Block
(Fig. 12.5), from Mushishima to
Mansa-Lueba on the Luapula River. (b)
Western region – Zambia and DRC
(Fig. 12.5), from Lwaio-Mwinilunga to
Kibambale-Makonga.

glaciogenic succession is composed of coarse glacial rain-out turbidites (109-m-thick) of the underlying Mwashya Subgroup.
(dropstones up to 10 cm across) of Roan and Mwashya derivation Diamictites of the Grand Conglomerat are interbedded with turbi-
and muddy-silty suspension deposits laid down simultaneously dites, and are interpreted to have formed by sediment-gravity flow
with intermittent sediment-gravity flows ranging from gravelly processes in a glaciomarine basin (Master et al. 2005).
debris flows to sandy turbidites, and associated with subordinate In the Itawa area near Ndola, Zambian Copperbelt, there are
traction currents (Wendorff & Key 2009). Bed thickness usually great thickness variations of the glaciomarine facies of the Grand
ranges from a few centimetres to a few tens of centimetres. There Conglomerat (Binda & Van Eden 1972). On an east –west section
are also rare sequences of c. 1-m-thick matrix- to clasts-supported line the thickness varies from 6 m to a maximum of 67 m, and back
conglomerate beds grading upwards to intervals of pebbly mud- to 21 m, over a distance of just 1.5 km. Isopach contours of the
stone over 1 m thick, which may represent very thick pebbly turbi- Grand Conglomerat, based on its thickness in the 14 boreholes
dites/megaturbidites (Postma et al. 1988). Reworking of bottom given in fig. 1 of Binda & Van Eden (1972), indicate the existence
sediment by intermittent weak traction currents is expressed by of a .5-km-long, c. 1-km-wide, trough-like feature trending
solitary silty current ripplemarks embedded in massive or lami- SSE –NNW, which may have been a submarine canyon. In
nated mudstone beds. Dropstones may occur in any of these borehole section IT 28 at Itawa (Fig. 12.5, Section 5) the Grand
sedimentary facies; their origin as ice-rafted melt-out debris is Conglomerat is c. 150 m thick and contains subangular to sub-
especially obvious when they occur as outsized clasts loaded into rounded rain-out clasts varying from granules to boulders of silici-
very thin beds and finely laminated siltstones and mudstones, the clastic and dolomitic sedimentary rocks derived from the
clast diameter often exceeding the thickness of the encompassing Roan-Mwashya succession. Characteristic for this section is a
strata. The textural and structural sedimentary features of this rain-out diamictite composed of grey and black silty and sandy
facies association suggest subaqueous deposition as a proglacial mudstone often modified by remobilization and redeposition
fan delta or apron (Eyles & Eyles 1992; Wendorff & Key 2009). within the basin as debris flows (Wendorff & Key 2009). Numer-
The facies associations and regional facies gradients between ous slump-folds (Binda & Van Eden 1972) and slump-generated
Mushishima and Kansanshi suggest the presence of a prominently debris flows that occur in both dropstone-rich and dropstone-
uplifted southern margin of the Nguba rift in what is now Zambia devoid intervals suggest deposition on a palaeoslope unstable
(Wendorff 2005b). This uplifted margin is thought to have been because of rapid deposition or syndepositional tectonic move-
the source area supplying the basin adjacent to the north with ments of the basin floor, or both processes combined (Eyles &
large amounts of detritus produced by glacial erosion of the Januszczak 2004; Wendorff & Key 2009). The uppermost part
Roan and Mwashya lithologies (Wendorff & Key 2009). of the Grand Conglomerat Fm. at Itawa is a C –U sequence
In the Chambishi Basin, in borehole MJZC/9, the 26-m-thick from laminite to thin turbidites representing Ta,e and Tc,e
Grand Conglomerat is conformable upon black shales and sequences of Bouma intervals. It is interpreted as distal facies
NEOPROTEROZOIC GLACIOGENIC DIAMICTITES OF THE KATANGA SUPERGROUP 179

deposited in a non-ice contact region of the basin (Wendorff & the basin, grade into carbonates of the Kakontwe Limestone
Key 2009). A 26-m-thick massive diamictite represents the Formation. This pattern suggests that the marginal marine strata
Grand Conglomerat Fm. in borehole Ks 17 at Mokambo, to the originated in a low-topography graded shelf during the syn-
east of Itawa. and immediately post-glacial period. On the other hand, the
At the NE margin of the depository, in the Luapula River valley maximum thicknesses of the Grand Conglomerat in the NW
(Bangweulu Block region; Fig. 12.5, Section 6), glaciofluvial region of the Nguba rift at Tombolo suggest subsidence rates
sandstone and pebbly sandstone with large-scale cross-bedded three times higher than in the south and several times higher
sets, indicating palaeocurrent flow towards the SW and south, is than in the rifted margin adjacent to the north. This situation
interbedded within a massive diamictite. suggests extremely high sedimentation rates, which must have
In the NW part of the external fold-thrust belt in Zambia balanced the high subsidence to maintain a low-gradient palaeo-
between Mwinilunga and Lwaio Mission (Fig. 12.5, Section 7; topography of the shelf. Such high sedimentation rates may be
Fig. 12.6), a sequence of marine conglomerates, laminites and explained by extremely effective supply of the clastic material
dropstone-bearing siltstones and shales occurs. The clasts rep- during deglaciation, derived from the Kibaran Belt further to the
resent basement granite and gneiss and fragments derived from NW, where considerable topographic relief remained (Cahen &
the Roan units (Wendorff & Key 2009). Lepersonne 1979).
At Lufunfu, in the NW part of the Lufilian Belt in the DRC
(Fig. 12.5, Section 8; Fig. 12.6), the glaciomarine massive diamic-
tite with a carbonaceous, slightly calcareous mudstone matrix Petit Conglomerat Formation, Kundelungu Group
reaches a thickness of 950 m (François 1973). The thickness of
the glaciogenic strata decreases southwards, as dominant clast The Petit Conglomerat Formation forms the base of the Kunde-
size decreases from pebble to granule and a conglomerate and lungu Group, and may overlie the Nguba Group with an erosional
several sandstone interbeds wedge out. These facies trends indi- unconformity (Wendorff 2003). The Petit Conglomerat diamicitite
cate that the source of the glaciogenic material supplied to the is, like the Grand Conglomerat, thought to be of glacial origin,
northern part of the basin was located to the north of the Nguba based on the abundant and widespread presence of faceted and
rift. At Tombolo, some 100 km north of Lufunfu (Fig. 12.5, striated clasts of both intrabasinal and extrabasinal origin
Section 9; Fig. 12.6), the diamictite reaches a maximum thickness (Vanden Brande 1936; Cahen 1978). It is overlain consistently
of 1200 m. by a carbonate unit, the ‘Calcaire Rose’ or ‘Dolomie Rose de
In the Makonga-Kibambale area located within the Plateau Lusele’, which consists of a 5–10-m-thick finely and regularly
region, which was a graben/aulacogen basin during deposition of bedded pink dolomite (François 1973; Dumont & Cahen 1977).
the Nguba Group (Unrug 1987; Wendorff 2005b; Master 2007; The Petit Conglomerat Formation has also been called the Kayan-
Fig. 12.3, Fig. 12.5 (Sections 10 and 11), Fig. 12.6), c. 150 km damu Formation (Ku1.1; Batumike et al. 2007).
north of the Lufilian fold-thrust belt, the Grand Conglomerat is rep- Cahen (1978) distinguished two facies in the Petit Conglomerat:
resented by a massive diamictite with interbeds of coarse-grained, a southern diamictite facies with small (,2 cm) clasts, and a north-
cross-bedded, feldspathic wacke (Dumont & Cahen 1977). Faceted ern mixed diamictite and conglomerate facies, with large clasts
and striated pebbles and cobbles of igneous and metamorphic rocks (up to 1 m granite clasts described by Grosemans 1935) of
are derived from the Kibaran basement to the north and clasts of varied compositions. The thickness of the Petit Conglomerat also
sedimentary rocks originated locally by scouring of the underlying varies from north to south: it is thickest in the north central part
Mwashya Subgroup. The glacial strata are intercalated with and of the Katangan basin (up to 80 m in the Lukafu area; Vanden
overlain by basaltic pillow lavas and volcanic breccias, and Brande 1936). In the northern part of the western Lufilian Arc,
intruded by dolerite sills. The pillow lavas imply a generally sub- François (1973) records its thickness varying from 50 to 35 m,
aqueous environment, but it is not certain whether the glacial depo- while in the southern part of the same area, its thickness varies
sition between Makonga and Kibambale occurred in a marine or from 40 to 35 m. In the Kipushi area further southeast, it is only
continental setting (Dumont & Cahen 1977). 24 m (Master et al. 2005). The Petit Conglomerat formation
In summary, the nature of the sedimentary facies and their trends thins towards the northwestern margin of the Katangan basin: it
suggest that the Grand Conglomerat glaciogenic sediments were is 25 m thick in the Makonga area, adjacent to the Kibaran Belt.
derived from both the southern and northern margin of the basin. On the northeastern margin of the basin, where they onlap
Stratigraphy in the southern region (at the southern margin of the against the basement rocks of the Bangweulu Block, the equivalent
Nguba rift) is characterized by interbeds of the glacial sediments rocks of the Petit Conglomerat Formation in the Luapula Beds are
within a coarse-clastic, high-energy fan-delta association and up to 100 m thick, but here they consist of a series of alternating
abundance of clasts derived from the Roan and Mwashya units conglomerates and pebbly sandstones (Abraham 1959). The con-
(Fig. 12.6). The presence of the fan-delta facies both below and glomerates have a coarse, poorly sorted angular to subangular
above the glaciogenic strata, and the deposits of a proglacial fan arkosic sandy matrix, with well-rounded cobbles and boulders
delta, imply a strong uplift of the southern shoulder of the (up to 40 cm in diameter) made of granites and rhyolite porphyries
Nguba rift. This is consistent with the existence of a possible sub- from the Bangweulu Block basement, as well as quartzites and
marine channel and considered here as the reason why the Grand shales from the Mporokoso Group. The pebbly sandstones are
Conglomerat in the southern region has characteristics of a deep trough crossbedded, having a poorly sorted subangular arkosic
marine deposit adjacent to a steep palaeoslope, for example, a matrix, with scattered angular to subangular pebbles, many of
considerable proportion of mass-flow phenomena, ranging which exhibit a triangular form and faceted appearance, similar
from coarse clastic debris flows to megaturbidites to small-scale to glaciogenic clasts described by Von Engelin (1930). Because
low-density turbidites, as well as abundant slump beds and of their textural and compositional immaturity, and the evidence
slump-initiated debris flows composed entirely of fine-grained that the pebbles have undergone little transport, these beds were
intrabasinal components. interpreted by Abraham (1959) as periglacial outwash gravels.
At the northern margin, on the other hand, the continental gla- Immature, feldspathic, pebbly sandstones, which are interbedded
ciogenic deposits grade southward towards the open marine within the conglomerate complex are trough crossbedded with
basin into glaciomarine melt-out sediments (Fig. 12.6). These ripple marks and peculiar cracks with polygonal outlines in plan
interfinger laterally with, and pass upwards into, proximal sand- view, resemble patterned ground typical of periglacial regions
stone interbeds of marginal marine origin. The proximal, postgla- (Washburn 1969). These were interpreted by Daily & Cooper
cial clastic (terrigenous) facies become progressively finer within (1976) to be a relict three-dimensional network of ice veins in a
the marine basin further to the south, and in the distal regions of permafrost environment.
180 S. MASTER & M. WENDORFF

In the northern regions, clasts recorded in the Petit Conglom- Chemostratigraphy


erat consist of quartz, granites, basic rocks, agates, amygdaloidal
lavas, rhyolites, quartzites, siliceous oolites, sandstones and Reconnaissance carbon and oxygen isotope studies of Katangan
shales, including many faceted and striated clasts (Grosemans carbonate rocks have been made by Master & Verhagen (unpub-
1935; Vanden Brande 1936; Batumike et al. 2006). Many of lished data, 1993), and by Bodiselitsch (2004). Carbon isotopic
the clasts originated from the adjacent Kibaran Belt to the values (d13C) in the Roan Group recorded by Master & Verhagen
NW and from the Kibambale volcanic complex to the north (unpublished data, 1993) range between þ2.2 and þ4.2‰ Vienna
(Dumont & Cahen 1977). In the Kapulo area of NE Katanga, Pee Dee Belemnite (VPDB), while the values in the Mwashya Sub-
just north of Lake Mweru, a distinctive and heterogeneous group show a negative trend with stratigraphic height, ranging
suite of clast types (maximum diameter, 30 cm) includes quart- from þ2.8 to – 1.2‰ VPDB. A sample of carbonate from the
zite, rhyolite, porphyries, alaskites and rare clasts of gneiss, matrix of the Grand Conglomerat had a d13C value of –4.6‰
mica schists, metaconglomerates and pisolitic black cherts VPDB. In the Kakontwe carbonate overlying the Grand Conglom-
(André 1976; Cahen 1978) derived from the adjacent Bang- erat, the d13C values show a stratigraphic upward positive trend,
weulu Block to the east (Master et al. 2005). Recent petro- increasing from þ2.7 to þ5.5, while in the upper Nguba Group,
graphic and geochemical studies on the Petit Conglomerat and values are between –0.2 to þ1.8‰ VPDB (Master & Verhagen,
other sedimentary rocks of the Nguba Group (Batumike et al. unpublished data, 1993). These isotopic values correlate well
2006, 2007) support the north –south facies variations, and a with the secular trends for the composite Neoproterozoic d13C
derivation from the Kibaran Belt and Bangweulu Block to the record both preceding and following the Sturtian glaciation
NW and NE of the Katangan basin. (Halverson et al. 2005). The carbon and oxygen isotopic data of
Petrographic studies of the Petit Conglomerat from Kipushi Bodiselitsch (2004) remain unpublished; however, Bodiselitsch
Mine were made by Master et al. (2005). Here the Petit et al. (2005) reported that d13C values in the Calcaire Rose and
Conglomerat consists of a fine-grained biotitic siltstone with a other carbonate rocks overlying the Petit Conglomerat at Kipushi
few scattered clasts, averaging about 0.5 mm across, but ranging (DRC) were consistently in the range of –2 to –4‰ VPDB.
up to a maximum size of 5 mm. The clasts consist of quartz, car- Bodiselitsch et al. (2005) obtained detailed major and trace
bonate, shale, chert and altered orthoclase. These clasts are element geochemical profiles from drill cores that intersect the
mainly of intrabasinal derivation, with some contribution from Grand and Petit Conglomerat diamictites and overlying carbonate
basement granitoids (orthoclase). The presence in this rock of acri- rocks. They found significant Ir anomalies at or near the base of the
tarchs (of planktonic origin), similar to acritarchs previously carbonate rocks immediately overlying both the Grand Conglom-
described from Kundelungu beds, indicates that the rock was erat and Petit Conglomerat diamictites at Kipushi, and also above
deposited in glaciomarine environment (Master et al. 2005). the Grand Conglomerat at Chambishi. They interpreted the Ir
Recent detailed petrographic studies of the Petit Conglomerat anomaly to be due to platinum group elements (PGE)-enriched
diamictites from the Kiaka anticline indicate that they show meteoritic debris (mainly in the form of interplanetary dust par-
microfabrics and microstructures comparable to those found ticles, or IDPs), which had accumulated on ice caps on a frozen
in modern glaciogenic tills (Delpomdor 2007; Delpomdor earth during prolonged Cryogenian glaciations (i.e. during the
et al. 2008). Snowball Earth model of Kirschvink 1992; Hoffman et al.
1998), and which were released into the sediment during melting
of the ice. They used the magnitude of the sharp Ir anomaly
Boundary relations with overlying and underlying above the Petit Conglomerat diamictite to estimate a minimum
non-glacial units time of 12 Ma for the duration of the ice cap, assuming a meteoritic
flux similar to that which prevailed in the later Phanerozoic.
The lower contact between the Grand Conglomerat and the
underlying strata of the Mwashya Subgroup varies from being
completely conformable, for example, in the Itawa area (Binda Other characteristics
& Van Eden 1972), in the central part of the Chambishi Basin
(Bodiselitsch et al. 2005) and in the central Lufilian arc (van Economic deposits
Doorninck 1928), to disconformable onto the lower or middle
Mwashya Subgroup in the western part of the Chambishi The Katangan glaciogenic diamictites and associated over- and
Basin (Wendorff & Key 2009), in the northern Lufilian arc underlying strata contain several economic mineral deposits,
(Robert 1940a), and in the Kibambale area (Dumont & Cahen some of which are currently being exploited. Within the
1977). In the Luapula Beds, the Grand Conglomerat rests non- Mwashya strata below the Grand Conglomerat diamictite, there
conformably on the crystalline granitic basement of the Bang- are two horizons of haematite and jaspilite, some outcrops of
weulu Block, and with an angular unconformity on metavolcanic which have been mined for iron (Jamotte 1947). A mafic pyroclas-
rocks of the Luapula Porphyries and quartzites of the Mporo- tic horizon within the Mwashya Subgroup is the host rock for stra-
koso Group (Thieme 1970, 1971; Andersen & Unrug 1984). tabound copper – cobalt mineralization at the Shituru Mine
The Grand Conglomerat is overlain conformably by carbonate in Katanga (Lefebvre 1974; Cailteux et al. 2007). The Grand Con-
strata of the Kakontwe Limestone Formation or by correlative glomerat diamictite itself is mineralized with disseminated copper
terrigenous ‘fines’ in the regions proximal to the south and sulphides and oxides in a few places in DRC: at Tombolo, NNE of
north margins of the Nguba Rift (Fig. 12.5). Kolwezi; at the Lonshi mine in the Pedicle of Katanga and at
The Petit Conglomerat Formation rests with a disconformity Kamoa, 20 km west of Kolwezi, where a giant high-grade strati-
or low-angle unconformity on upper Nguba Group strata in form orebody was discovered in 2009. The periglacial conglomer-
the DRC (Robert 1940a) and with a pronounced angular unconfor- ates associated with the Grand Conglomerat in the Luapula Beds of
mity on the strongly folded Nguba succession in Zambia northern Zambia contain disseminated copper mineralization in
(Wendorff & Key 2009). In the Luapula Beds, it rests with a dis- several places (Thieme 1970). The Kakontwe carbonate, which
conformity or low-angle unconformity on the Nguba Group and overlies the Grand Conglomerat, is mineralized with epigenetic
also transgressively onlaps onto the crystalline granitic basement, breccia-fill Zn –Pb– Cu –Ge –Ga mineralization at the Kipushi
which it overlies nonconformably. The Petit Conglomerat For- Mine, DRC, as well as in the much smaller Kengere and Lombe
mation is overlain conformably by carbonate of the Lusele For- deposits, which were exploited in the past (see review in Batumike
mation (Dolomie Rose), or by shales where the carbonates are et al. 2007). There is also a stratiform Cu –Co deposit, Tantara,
not developed. consisting primarily of secondary copper carbonates and silicates,
NEOPROTEROZOIC GLACIOGENIC DIAMICTITES OF THE KATANGA SUPERGROUP 181

which is hosted by this stratigraphic unit (Reintjens 1935). In Formation, Mwashya Subgroup. This age provides a maximum
South Katanga, near the Zambian border, there are important age constraint on the overlying diamictites correlated with the
iron deposits hosted by the Kakontwe limestone. The giant Kan- Grand Conglomerat Formation. Key et al. (2002) also obtained a
sanshi Cu –Au deposit in northern Zambia is hosted by flat-lying SHRIMP U –Pb zircon age of 735 + 5 Ma from a volcanic
carbonates and schists of the Kakontwe Formation (Broughton breccia within strongly deformed strata at Mwinilunga (M in
et al. 2002). Ba –Fe mineralization occurs in the Kapumba Fig. 12.2, & Fig. 12.5, Section 7), interpreted as being part of
deposit NW of Likasi, DRC, where barite-specularite veins cut the Nguba beds overlying the Grand Conglomerat. Although its
the Grand Conglomerat and overlying dolomites and shales of stratigraphic position is not entirely certain, the dated volcanic
the Nguba Group (Intiomale & Mbuyi 1997); these appear breccia is spatially associated with carbonate rocks that are not
similar to the barite veins cutting cap carbonates in Mauritania found beneath the Grand Conglomerat in the undeformed succes-
(Shields et al. 2007). The Lusele or Calcaire Rose carbonate sion immediately to the west (LM, Lwaio Mission in Fig. 12.2).
above the Petit Conglomerat diamictite contains copper mineraliz- These carbonate rocks are correlated with the Kakontwe Lime-
ation at the Sokoroshi I and II deposits (Reintjens 1935). The stone Formation (Wendorff & Key 2009). It is therefore concluded
Lusele carbonate in the Luapula Beds contains Pb–Zn mineraliz- that the Grand Conglomerat was most likely deposited prior to
ation (Thieme 1970). The metallic mineralizations associated with 735 + 5 Ma and definitely after the 765 + 5 Ma Mwashya vol-
the Neoproterozoic glacial intervals in the Katangan and elsewhere canism. The youngest detrital zircons from the Grand Conglomerat
have been speculatively linked to global anoxia, and the generation diamictite at Kipushi Mine, DRC, yielded an imprecise age of
of reduced metalliferous fluids from associated reduced strata 729 + 50 Ma (Master et al. 2005), consistent with the above
(Robb et al. 2003, 2004). age constraints.
From the available radiometric data, the age of the Petit Con-
glomerat is not yet well constrained, and is only bracketed
Characteristics of carbonate rocks overlying between 735 + 5 Ma, the age of volcanics in the West Lunga For-
the diamictites mation in the Nguba Group (Key et al. 2002), and c. 620 Ma, the
age of uraninites from veins in thrust zones that affect the Katan-
The Kakontwe dolostone overlying the Grand Conglomerat con- gan stratigraphy to the top of the Kundelungu Group (Cahen
tains several very distinctive features. At its base, where it is 1973a). However, on the basis of regional correlations and their
locally referred to as the ‘serie tigré’, it has a laminated appearance, shared tectonic history as Neoproterozoic rift to passive margin
resembling algal lamination, and at Kipushi it also shows roll-up sequences on the southern edge of the Congo craton, the Petit
structures identical to those which have been described from the Conglomerat and the overlying Calcaire Rose may be equivalent
base of the Rasthof carbonate overlying the Chuos diamictite in to the Ghaub diamictite and Maieberg Formation in the Otavi
Namibia (Hoffman et al. 1998). These basal carbonates also com- Group of the Damara Orogen, Namibia, which are dated at
monly show soft-sediment deformation features, such as convolute 635.5 + 1.2 Ma (Hoffmann et al. 2004).
bedding and isoclinal folds, which have been observed at Kipushi Palynological work on the Nguba Group, including laminated
and at Musoshi (Master, unpublished data). These structures dropstone-bearing shales of the Grand Conglomerat, was carried
resemble similar features seen in cap carbonates overlying Neopro- out by Binda (1972a, b, 1977). Although several microfossil
terozoic glacial diamictites in Brazil (Nogueira et al. 2003) and in forms were identified (including three new varieties of the genus-
Mauritania (Master 2004, unpublished data; Shields et al. 2007). form Fibularix and seven new types of sphaeromorph acritarch),
these had little biostratigraphic usefulness, except in indicating a
Neoproterozoic age. Vavrdova & Utting (1974) described actri-
Palaeolatitude and palaeogeography tarchs from the Luapula beds, which they ascribed to a Lower
Palaeozoic age; however, their age attribution is also considered to
Meert (2003) inferred that at c. 750 Ma, the present-day southern be unreliable, and of little stratigraphic significance (Binda 1977).
part of the Congo Craton was at palaeolatitudes of between 5
and 108S, based on a palaeopole for the Mbozi Complex in
southern Tanzania, for which a SHRIMP U –Pb zircon age of Discussion
748 + 6 Ma has been obtained (Mbede et al. 2004). Wingate
et al. (2010) obtained palaeomagnetic data on the 800-m-thick No striated pavements have ever been observed beneath the two
765 + 5 Ma (SHRIMP U –Pb zircon age) Luakela volcanics of Katangan diamictites thought to be of glacial origin, mainly
the Lwavu Formation, Mwashya Subgroup (Key et al. 2002), because the underlying rocks consisted largely of unconsolidated
based on alternating field (AF) and thermal demagnetization of sediments. On the basis of clast composition, and detrital zircon
65 samples from nine sites. Of three components found, the one ages, the glaciers appear to have fed into the Katangan Basin
component, carried by magnetite and regarded as primary, from the flanking basement regions to the NW (Kibaran Belt)
coincides with the Mbozi palaeopole obtained by Meert et al. and to the NE (Bangweulu Block) (Master et al. 2005). Palaeocur-
(1995). Because the overlying Grand Conglomerat diamictite is rent data from associated periglacial (glaciofluvial) sandstones
bracketed in age between 765 + 5 and 735 + 5 Ma (see below), in the Luapula Beds of north Zambia indicate palaeocurrents
it was most likely deposited in near-equatorial low latitudes. trending towards the SW. Distinctive feldspar-phyric porphyritic
There is abundant palaeogeographic evidence from around the granitoid clasts within the Grand Conglomerat in Kipushi district
Congo Craton, in the form of extensive stromatolitic carbonate are glacial indicators that can be traced to a source in the Luina
platforms, that it was situated in tropical palaeolatitudes during Dome 150 km away, indicating a WSW pre-tectonic transport
much of the Neoproterozoic Era (Verbeek 1970; Cahen 1973b, direction for these glacial erratics (Master et al. 2005). Possible
1982; Alvarez 1993; Evans 2000). Thus although there are no submarine channels within the glaciomarine facies of the Grand
direct palaeomagnetic studies of the Petit Conglomerat, it was Conglomerat in the Itawa area of Zambia trend NNW. The distri-
also likely deposited in tropical palaeolatitudes. bution and thickness variations of facies, and indications of source
areas, show that transport of the glaciers was globally generally
north to south, with many local fluctuations.
Geochronological constraints The size distribution and nature of clasts in the Petit Conglom-
erat indicate a north –south transport direction, corresponding to
Key et al. (2002) obtained a SHRIMP U –Pb zircon crystallization the diminution in size and abundance of extrabasinal clasts
age of 765 + 5 Ma for the Luakela lavas from the Lwavu (derived from the Kibaran Belt and the Bangweulu Block).
182 S. MASTER & M. WENDORFF

However, a facies gradient and clast composition in the syn-glacial Batumike, M. J., Kampunzu, A. B. & Cailteux, J. L. H. 2006. Petrology
fan-delta facies to the west of the Kafue Anticline in Zambia and geochemistry of the Neoproterozoic Nguba and Kundelungu
suggest clast derivation and palaeotransport direction towards Groups, Katangan Supergroup, southeast Congo: implications for
the north, away form the uplifted southern shoulder of the provenance, palaeoweathering and geotectonic setting. Journal of
Roan-Nguba rift basin (Wendorff & Key 2009). African Earth Sciences, 44, 97 – 115.
Batumike, M. J., Cailteux, J. L. H. & Kampunzu, A. B. 2007. Litho-
stratigraphy, basin development, base metal deposits, and regional
Conclusions correlations of the Neoproterozoic Nguba and Kundelungu rock suc-
cessions, central African Copperbelt. Gondwana Research, 11,
The Grand and Petit Conglomerat diamictites of the Katanga 432– 447.
Binda, P. L. 1972a. Preliminary observations on the palynology of the
Supergroup in Central Africa, which were deposited in a variety
Precambrian Katanga Sequence. Geologie en Mijnbouw, 51,
of continental, shallow and deep marine facies, show abundant evi-
315– 319.
dence of a glaciogenic origin. This evidence includes the occur- Binda, P. L. 1972b. Microfossils from the Lower Kundelungu (Late
rence of (i) poorly sorted subangular to subrounded clasts of Precambrian) of Zambia. 24th International Geological Congress,
intra- and extrabasinal origin, ranging in size from millimetric Montreal, Section 1, Precambrian Geology, 179–186.
grains to boulders with a volume of several cubic metres, sitting Binda, P. L. 1977. Microfossils from the Lower Kundelungu (Late
in a fine-grained matrix; (ii) faceted and striated clasts from numer- Precambrian) of Zambia. Precambrian Research, 4, 285–306.
ous localities; (iii) microfabrics and microstructures diagnostic of Binda, P. L. 1994. Stratigraphy of Zambian Copperbelt orebodies.
processes associated with modern tillites; and (iv) associated sedi- Journal of African Earth Sciences, 19, 251– 264.
mentary rocks indicative of glacial or periglacial conditions, Binda, P. L. & Van Eden, J. G. 1972. Sedimentological evidence for the
including laminated facies with dropstones, and sandstones with origin of the Precambrian Great Conglomerate (Kundelungu Tillite),
‘patterned ground’ orthogonal crack networks. Their widespread Zambia. Palaeogeography, Palaeoclimatology, Palaeoecology, 12,
occurrence (over an area of .65 000 km2) and great thickness 151– 168.
(up to 1200 m) make the Katangan diamictites some of the most Bodiselitsch, B. 2004. Geochemical and stable isotope investigations on
important Neoproterozoic glaciogenic units. Despite their rela- ‘Snowball Earth’ samples from the Lufilian tectonic arc, D. R. Congo
tively poor geochronological constraints, the Grand and Petit Con- and Zambia, and on the late Eocene sediment samples from Mas-
glomerat diamictites can be reasonably correlated, respectively, signano, Italy. PhD thesis (unpublished), University of Vienna,
with the Chuos and Ghaub glaciogenic diamictites of the Damara Vienna, Austria.
orogenic belt in Namibia, especially considering similarities in Bodiselitsch, B., Koeberl, C., Master, S. & Reimold, W. U. 2005.
terms of C-isotope chemostratigraphy, textures in overlying car- Estimating duration and intensity of Neoproterozoic snowball glacia-
tions from Ir anomalies. Science, 308, 239–242.
bonate units, and their common tectonic positions on the southern
Broughton, D. W., Hitzman, M. W. & Stephens, A. J. 2002.
rifted margin of the Congo Craton. The presence of the Katangan
Exploration history and geology of the Kansanshi Cu –(Au) deposit,
diamictites within a major metallogenic district, the Central Zambia. Society of Economic Geologists, Special Publication, 9,
African Copperbelt, has resulted in many drill cores being made 141– 153.
available through the diamictites and their over- and underlying Cahen, L. 1954. Géologie du Congo Belge. Vaillant-Carmanne, Liège.
carbonate strata. These cores have facilitated the very detailed Cahen, L. 1963. Glaciations anciennes et dérive des continents. Annales
geochemical studies that resulted in the discovery of Ir anomalies de la Société géologique de Belgique, 86, B19 –B84.
in the carbonate units overlying the diamictites (Bodiselitsch Cahen, L. 1973a. L’uraninite de 620 m.a. post-date tout le Katangien,
et al. 2005), which has contributed significantly to invigorating mise au point. Musée Royale d’Afrique Centrale, Tervuren (Belgi-
the debate on the Snowball Earth model for Neoproterozoic que), Département de Géologie et Minéralogie, Rapport Annuel,
glaciations. 1972, 35– 38.
Cahen, L. 1973b. Corrélations du certains séries du Précambrien supér-
We thank P. Binda, R. Key, P. Hoffman and F. Delpomdor for discussions; ieur du Zaı̈re à la lumière de l’étude des stromatolithes et des données
J. Cailteux, F. Mbuyi and B. Ngoie for guidance in the field; and J. Batumike géochronologie radiométrique. Musée Royale d’Afrique Centrale,
for reprints. We are indebted to R. Hanson, J. Arthurs and G. Halverson for Tervuren (Belgique), Département de Géologie et Minéralogie,
their comprehensive reviews, which have helped considerably in improving the Rapport Annuel, 1972, 38 – 51.
manuscript. We are grateful to G. Halverson for his invitation to produce this Cahen, L. 1978. Les mixtites anté-cambriennes de l’est du Zaı̈re: mise
review, and for his patience and encouragement during the editorial process. au point interimaire. Musée Royale d’Afrique Centrale, Tervuren
This represents a contribution of the IUGS- and UNESCO-funded IGCP (Belgique), Département de Géologie et Minéralogie, Rapport
(International Geoscience Programme) Project #512. Annuel, 1977, 33 –64.
Cahen, L. 1982. Geochronological correlation of the Late Precambrian
sequences on and around the stable zones of Equatorial Africa.
References Precambrian Research, 18, 73 –86.
Cahen, L. & Lepersonne, J. 1979. Upper Proterozoic diamicitites
Abraham, D. 1959. The stratigraphical and structural relationship of the of Shaba (formerly Katanga) and neighbouring regions of
Kundelungu System, Plateau Series and basement rocks in the Mid- Zambia. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s
Luapula valley, Northern Rhodesia. D. Phil. thesis, University of Pre-Pleistocene Glacial Record. Cambridge University Press,
Leeds. Cambridge, 162–166.
Alvarez, Ph. 1993. Un modèle de lagune d’âge Protérozoı̈que supérieur: Cailteux, J., Binda, P. L. et al. 1994. Lithostratİgraphical correlation of
le Schisto-calcaire du Congo. Journal of African Earth Sciences, 17, the Neoproterozoic Roan Supergroup from Shaba (Zaire) and
75 –87. Zambia, in the central African copper-cobalt metallogenic province.
Andersen, L. S. & Unrug, R. 1984. Geodynamic evolution of the Bang- Journal of African Earth Sciences, 19, 265– 278.
weulu Block, northern Zambia. Precambrian Research, 25, 187– 212. Cailteux, J. L. H., Kampunzu, A. B. & Lerouge, C. 2007. The Neopro-
André, L. 1976. Etude aérophotomorphologiques et pétrographique du terozoic Mwashya-Kansuki sedimentary rock succession in the
Katangien de la mosaı̈que contrôlée de Kapulo au Shaba. Mémoire central African Copperbelt, its Cu –Co mineralisation, and regional
de licence, Univ. Libre de Bruxelles, Belgium. correlations. Gondwana Research, 11, 414–431.
Armstrong, R. A., Master, S. & Robb, L. J. 2005. Geochronology of the Cornet, J. 1897. Observations sur le terrains ancien du Katanga faites au
Nchanga Granite, and constraints on the maximum age of the Katanga cours de l’expédition Bia-Francqui (1891– 1893). Annales de la
Supergroup, Zambian Copperbelt. Journal of African Earth Sciences, Société Géologique de Belgique, 24, Mémoires, 1896–1897,
42, 32 –40. 25 – 191.
NEOPROTEROZOIC GLACIOGENIC DIAMICTITES OF THE KATANGA SUPERGROUP 183

Daily, B. & Cooper, M. R. 1976. Clastic wedges and patterned ground in Breakup. Geological Society, London, Special Publications, 206,
the Late Ordovician–Early Silurian tillites of South Africa. Sedimen- 427– 463.
tology, 23, 271– 227. Hennig, E. 1915. Die Glazialerscheinerungen in Äquatorial- und Süd-
Daly, M. C., Chakraborty, S. K. et al. 1984. The Lufilian arc and afrika. Geologische Rundschau, 6, 154–165.
Irumide belt of Zambia: results of a traverse across their intersection. Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P.
Journal of African Earth Sciences, 4, 311– 318. 1998. A Neoproterozoic snowball earth. Science, 281, 1342– 1346.
Delhaye, F. 1920. Les variations de facies du conglomérat inférieur Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
du Système du Kundelungu au Katanga. Annales de la Société 2004. U–Pb zircon dates for the Neoproterozoic Ghaub Formation,
Géologique de Belgique, Publications relatives au Congo Belge et Namibia: constraints on Marinoan glaciation. Geology, 32, 817– 820.
aux régions voisins, 43, 1919–1920, 19– 27. Intiomale, M. M. & Mbuyi, K. 1997. Le gisement Ba –Fe de Kapumba
Delpomdor, F. 2007. Etude des dépôts diamictitiques du Néoprotérozoı̈- (Shaba, Zaı̈re). Musée Royale d’Afrique Centrale, Tervuren (Belgi-
que supérieure en République Démocratique du Congo et au que), Département de Géologie et Minéralogie, Rapport Annuel
sud-ouest du Rwanda. Unpublished report, Musée Royal d’Afrique 1995–1996, 183– 186.
Centrale, Tervuren, Belgium. Jackson, M. P. A., Warin, O. N., Woad, G. M. & Hudec, M. R. 2003.
Delpomdor, F., Tack, L. & Préat, A. 2008. Microstructures in the Neo- Neoproterozoic allochthonous salt tectonics during the Lufilian
proterozoic tillites around the Congo River Basin (CRB), Democratic orogeny in the Katangan Copperbelt, central Africa. Geological
Republic of the Congo (DRC) – comparison with the Karoo tillites Society of America Bulletin, 115, 314– 330.
from the Dekese borehole in the CRB. Extended Abstract, 22nd Jamotte, A. 1947. Esquisse géologique de la région de Kasenga. Bulletin
Colloqium of African Geology, 2 –6 November 2009, Hammamet, de l’Institut Royale Coloniale Belge, 18, 461– 476.
Tunisia. John, T., Schenk, V., Haase, K., Scherer, E. & Tembo, F. 2003. Evi-
De Waele, B., Liégeois, J.-P., Nemchin, A. A. & Tembo, F. 2006. Isoto- dence for a Neoproterozoic ocean in south central Africa from
pic and geochemical evidence of Proterozoic episodic crustal rework- MORB-type geochemical signatures and P –T estimates of
ing within the Irumide belt of south-central Africa, the southern Zambian eclogites. Geology, 31, 243– 246.
metacratonic boundary of an Archaean Bangweulu Craton. Precam- John, T., Schenk, V., Mezger, K. & Tembo, F. 2004. Timing and PT
brian Research, 148, 225–256. evolution of whiteschist metamorphism in the Lufilian arc-Zambezi
Dumont, P. & Cahen, L. 1977. Les complexes conglomeratiques de la belt orogen (Zambia): implications for the assembly of Gondwana.
bordure sud-orientale de la Chaine Kibarienne et leurs relations Journal of Geology, 112, 71– 90.
avec les couches Katangiennes de l’Arc Lufilienne, Rapport Annuelle Kampunzu, A. B. & Cailteux, J. 1999. Tectonic evolution of the Lufilian
1977. Museé royal de l’Afrique centrale, Départment de Géologie et Arc (Central Africa Copperbelt) during the Neoproterozoic
Minéralogie, Tervuren, Belgique, 111–135. Pan-African orogenesis. Gondwana Research, 2, 401– 421.
Dumont, P. & Hanon, M. 1997. Le plateau des Kundelungu, paléograben Kampunzu, A. B., Kanika, M., Kapenda, D. & Tshimanga, K. 1993.
ou aulacogène. In: Charlet, J.-M. (ed.) Proceedings of the Inter- Geochemistry and geotectonic setting of late Proterozoic Katangan
national Cornet Symposium ‘Strata-bound Copper Deposits and basic rocks from Kibambale in Central Shaba (Zaire). Geologische
Associated Mineralizations’ (Mons, 5 – 9 September 1994). Académie Rundschau, 82, 619– 630.
Royale des Sciences d’Outre-Mer, 51– 69. Kampunzu, A. B., Tembo, F., Matheis, G., Kapenda, D. & Huntsman-
Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic Mapila, P. 2000. Geochemistry and tectonic setting of mafic igneous
constraints upon the Neoproterozoic climatic paradox. American units in the Neoproterozoic Katangan basin, Central Africa: impli-
Journal of Science, 300, 347–433. cations for Rodinia breakup. Gondwana Research, 3, 125– 153.
Eyles, N. & Eyles, C. H. 1992. Glacial depositional systems. In: Key, R. M., Liyungu, A. K., Njamu, F. M., Somwe, V., Banda, J.,
Walker, R. G. & James, N. P. (eds) Facies Models: Response to Mosley, P. M. & Armstrong, R. A. 2002. The western end of the
Sea Level Change. Geological Association of Canada, St. John’s, Lufilian arc in NW Zambia and its potential for copper deposits.
73 – 100. Journal of African Earth Sciences, 33, 503–528.
Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for Kirschvink, J. L. 1992. Late Proterozoic low-latitude glaciation: the
Neoproterozoic glaciations during the breakup of Rodinia after snowball Earth. In: Schopf, J. W. & Klein, C. (eds) The Proterozoic
750 Ma. Earth-Science Reviews, 65, 1– 73. Biosphere. Princeton University Press, New York, 51 – 52.
François, A. 1973. L’extremité Occidentale de l’Arc Cuprifère Shabien, Kokonyangi, J. W., Kampunzu, A. B., Armstrong, R., Yoshida, M.,
Etude Géologique. Gécamines, Lubumbashi. Okudaira, T., Arima, M. & Ngulube, D. A. 2006. The Mesoproter-
François, A. & Cailteux, J. 1981. La couverture katangienne entre les ozoic Kibaride belt (Katanga, SE D.R. Congo). Journal of African
socle de Nzilo et de la Kabompo, République du Zaı̈re, région de Earth Sciences, 46, 1– 35.
Kolwezi. Annales de la Musée royal d’Afrique centrale, Sciences Lefebvre, J. J. 1974. Mineralisations cupro-cobaltiferes associées aux
géologiques, 87, 50. horizons pyroclastiques situés dans le faisceau supérieur de la Serie
Grosemans, P. 1935. Contribution à l’etude du conglomérat de base (petit de Roan, à Shituru, Shaba, Zaire. In: Bartholomé, P., de
conglomérat) du Kundelungu supérieur. Annales de Services des Magnée, I, Evrard, P. & Moreau, J. (eds) Gisements stratiformes
Mines, Comité Spécial du Katanga, 5, 38 –57. et provinces cuprifères. Société Géologique de Belgique, Liège,
Grosse, E. 1912. Dwykakonglomerat und Karroosystem im Katanga. 103– 122.
Zeitschrift der Deutsche Geologischer Gesellschaft, Monatsberichte, Master, S. 2007. Neoproterozoic evolution of the Kundelungu Plateau,
64, 320– 321. Katanga Supergroup, Central Africa: from aulacogen to foreland
Grosse, E. 1918. Grundlinien der Geologie und Petrographie des östlichen basin. Extended Abstract, IGCP 485 and IGCP 497 Joint Conference,
Katanga. Neues Jahrbuch für Mineralogie, Geologie und Paläontolo- Chouaı̈b Doukkali University, El Jadida, Morocco, 28 November – 5
gie Beilage, 42, 272–419. December 2007, 62 – 65.
Guillemain, C. 1913. Zur Geologie von Katanga. Zeitschrift der Master, S., Rainaud, C., Armstrong, R. A., Phillips, D. & Robb, L. J.
Deutsche Geologischer Gesellschaft, Monatsberichte, 65, 304– 328. 2005. Provenance ages of the Neoproterozoic Katanga Supergroup
Gysin, M. 1934. Les tillites métamorphiques du Kundelungu de la Haute (Central African Copperbelt), with implications for basin evolution.
Lufira (Congo Belge). Compte Rendu séances, Société de Physique et Journal of African Earth Sciences, 42, 41 –60.
de l’Histoire naturelle, Genève, 51, 218–221. Mbede, E. I., Kampunzu, A. B. & Armstrong, R. A. 2004. Neoproter-
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & ozoic inheritance during Cainozoic rifting in the western and south-
Rice, A. H. N. 2005. Toward a Neoproterozoic composite western branches of the East African Rift System: evidence from
carbon-isotope record. Geological Society of America Bulletin, 117, carbonatite and alkaline intrusions. International Commission on
1181– 1207. Earth Sciences in Africa Conference: The East African Rift, Addis
Hanson, R. E. 2003. Proterozoic geochronology and tectonic evolution of Ababa, Abstracts.
southern Africa. In: Yoshida, M., Windley, B. F. & Dasgupta, S. Meert, J. 2003. A synopsis of events related to the assembly of eastern
(eds) Proterozoic East Gondwana: Supercontinent Assembly, Gondwana. Tectonophysics, 362, 1– 40.
184 S. MASTER & M. WENDORFF

Meert, J. G., van der Voo, R. & Ayub, S. 1995. Paleomagnetic investi- Studt, F. E. 1908. Carte géologique du Katanga (1/500.000e) et notice
gation of the Neoproterozoic Gagwe lavas and Mbozi complex, Tan- explicative. Annales du Musée de Congo belge, sér. 2, 1, 5– 16.
zania and the assembly of Gondwana. Precambrian Research, 74, Studt, F. E. 1913. The geology of Katanga and Northern Rhodesia: an
225– 244. outline of the geology of South Central Africa. Transactions of the
Museu, M. 1987. Considérations sur l’origine du Grand Conglomérat de Geological Society of South Africa, 16, 44– 106.
base du Kundelungu inférieur au Shaba (République du Zaı̈re). Studt, F. E., Cornet, J. & Buttgenbach, H. 1908. Carte géologique du
Musée royale de l’Afrique centrale, Tervuren (Belgique), Départe- Katanga et notes descriptives. Annales Musée du Congo, Bruxelles,
ment de Géologie et Minéralogie, Rapport Annuel 1985–1986, Géologie, Géophysique, Minéralogie & Paléontologie, série 2,
165– 168. Katanga, 1, 94.
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V. & Fair- Stutzer, O. 1911. Über Dwyka Konglomerat im Lande Katanga, Bel-
child, T. R. 2003. Soft-sediment deformation at the base of the Neo- gisch Kongo. Zeitschrift der Deutsche Geologischer Gesellschaft,
proterozoic Puga cap carbonate (southwestern Amazon craton, Monatsberichte, 63, 626– 629.
Brazil): confirmation of rapid icehouse to greenhouse transition in Stutzer, O. 1913a. Über glaziale Konglomerate im Lande Katanga, Bel-
snowball Earth. Geology, 31, 613– 616. gisch Kongo. Zeitschrift der Deutsche Geologischer Gesellschaft,
Porada, H. 1989. Pan-African rifting and orogenesis in southern to equa- Monatsberichte, 65, 114– 117.
torial Africa and eastern Brazil. Precambrian Research, 44, 103– 136. Stutzer, O. 1913b. Über den geologischen Aufbau der südöstlichen
Porada, H. & Berhorst, V. 2000. Towards a new understanding of the Katanga. Jahresbericht der Freiberger Geologischer Gesellschaft,
Neoproterozoic – Early Palaeozoic Lufilian and northern Zambezi VI, 41 – 47.
Belts in Zambia and the Democratic Republic of Congo. Journal of Tembo, F., Kampunzu, A. B. & Porada, H. 1999. Tholeiitic magmatism
African Earth Sciences, 30, 727–771. associated with continental rifting in the Lufilian Fold Belt of Zambia.
Postma, G., Nemec, W. & Kleinspehn, K. L. 1988. Large floating clasts Journal of African Earth Sciences, 28, 403– 425.
in turbidites: a mechanism for their emplacement. Sedimentary Thieme, J. G. 1970. The geology of the Mansa area: Explanation of
Geology, 58, 47– 61. Degree Sheet 1128, parts of NW Quarter and NE Quarter. Report
Rainaud, C., Master, S., Armstrong, R. A. & Robb, L. J. 2005a. Geo- of the Geological Survey of Zambia, Lusaka, 26, 37.
chronology and nature of the Palaeoproterozoic basement in the Thieme, J. G. 1971. The geology of the Musonda Falls area: Explanation
Central African Copperbelt, with regional implications. Journal of of Degree Sheet 1028, SE Quarter. Report of the Geological Survey of
African Earth Sciences, 42, 1 –31. Zambia, Lusaka, 32, 25.
Rainaud, C., Master, S., Armstrong, R. A., Phillips, D. & Robb, L. J. Unrug, R. 1987. Geodynamic evolution of the Lufilian arc and the Kun-
2005b. Monazite U– Pb dating and 40Ar/39Ar thermochronology of delungu aulacogen: Angola, Zambia and Zaire. In: Matheis, G. &
metamorphic events in the Central African Copperbelt during the Schandelmeier, H. (eds) Current Research in African Earth
Pan-African Lufilian orogeny. Journal of African Earth Sciences, Sciences; 14th Colloquium of African Geology, Extended Abstracts.
42, 183– 199. Balkema, Rotterdam, 117–120.
Reintjens, E. 1935. Les gisements cuprifères du Katanga et de la Rhodé- Unrug, R. 1988. Mineralization controls and source of metals in the Lufi-
sie septentrionale. Comité Special du Katanga, Annales des Services lian fold belt, Shaba (Zaire), Zambia and Angola. Economic Geology,
des Mines, 6, 10 – 19. 83, 1247– 1258.
Robb, L. J., Master, S., Armstrong, R. A., Rainaud, C. & Greyling, L. Vanden Brande, P. 1936. Etudes géologiques dans le feuille Lukafu.
2003. Timing of Cu–Co and Pb –Zn mineralisation in the Central Comité Special du Katanga, Annales des Services des Mines, 6, 51–69.
African Copperbelt: a link to Neoproterozoic glaciations? Trans- van Doorninck, N. H. 1928. De Lufilische Plooing in den Boven Katanga
actions of the Institution of Mining and Metallurgy, Section B, (Belgischen Congo). G. Naeff, ’s-Gravenhage.
Applied Earth Sciences, 112, B164 –B166. Vavrdova, M. & Utting, J. 1974. Lower Paleozoic microfossils from the
Robb, L., Master, S., Armstrong, R. A., Greyling, L. & Rainaud, C. Luapula beds of the Mansa area. Records of the Geological Survey
2004. Neoproterozoic glaciations and the link to Cu –Co and Pb –Zn of Zambia, 12, 81– 89.
mineralization in the Central African Copperbelt. Geoscience Africa Verbeek, T. 1970. Géologie et lithologie du Lindien (Précambrien supér-
2004 Conference, Abstracts Volume 2, University of the Witwaters- ieur du Nord de la République Démocratique du Congo). Annales de
rand, Johannesburg, 12 –16 July 2004, 550– 551. la Musée Royale d’Afrique Central, séries In-88, Sciences géologi-
Robert, M. 1912a. La stratigraphie du système du Kundelungu au ques, 66, 311.
Katanga. Annales de la Société Géologique de Belgique, Publications Von Engelin, O. D. 1930. Type form of facetted and striated glacial
relatives au Congo Belge et aux régions voisins, 39, 1911– 1912, fasc. pebbles. American Journal of Science, 19, 9– 16.
I, 5 – 8. Washburn, A. L. 1969. Patterned ground in the Mesters Vig district,
Robert, M. 1912b. Le système du Kundelungu au Katanga. Annales de la northeast Greenland. Biuletyn Peryglacjalny, Łódź, 18, 259–330.
Société Géologique de Belgique, Publications relatives au Congo Wendorff, M. 2003. Stratigraphy of the Fungurume Group– evolving
Belge et aux régions voisins, 40, 1912–1913, 213– 275. foreland basin succession in the Lufilian fold-thrust belt, Neoproter-
Robert, M. 1933. Le système du Kundelungu au Katanga. Bulletin de ozoic –Lower Palaeozoic, Democratic Republic of Congo. South
l’Institut royale colonial belge, 4, 436 –440. African Journal of Geology, 106, 17 –34.
Robert, M. 1940a. La glaciation du Kundelungu au Katanga (Congo Wendorff, M. 2005a. Coarse clastic markers of Rodinia breakup and
Belge). Report of the 17th International Geological Congress, Gondwana assembly in the Lufilian belt, Pan-African orogen of
Moscow, USSR, 1937, 6, 99 –113. Central Africa. In: Wingate, M. T. D. & Pisarevsky, S. A. (eds)
Robert, M. 1940b. Contribution à la géologie du Katanga. Le Système du Supercontinents, Earth Evolution Symposium. Geological Society
Kundelungu et le Système Schisto-dolomitique (1re partie). Mémoire of Australia Abstracts, Perth, 101.
de l’Institut royale colonial belge, Science Naturelle et Médecine, Wendorff, M. 2005b. Evolution of Neoproterozoic – Lower Palaeozoic
Coll. in-4o, 6, 108. Lufilian arc, Central Africa: A new model based on syntectonic con-
Robert, M. 1946. Le Congo Physique, troisième édition, revue et complé- glomerates. Journal of the Geological Society, London, 162, 5– 8.
tée. Vaillant-Carmanne, Liège. Wendorff, M. 2005c. Sedimentary genesis and lithostratigraphy of
Robert, M. 1947. Les traces de glaciation et les périodes climatiques Neoproterozoic megabreccia from Mufulira, Copperbelt of Zambia.
glaciaires au Katanga et en Afrique australe. Bulletin de la Société Journal of African Earth Sciences, 42, 61– 81.
belge de Géologie, Paléontologie et Hydrologie, 56, 62 – 76. Wendorff, M. & Key, R. M. 2009. The relevance of the sedimentary
Selley, D., Broughton, D. et al. 2005. A New Look at the Geology history of the Grand Conglomerat Formation (Central Africa) to the
of the Zambian Copperbelt. Society of Economic Geologists, Inc., interpretation of the climate during a major Cryogenian glacial
Tulsa, 100th Anniversary Volume, 965– 1000. event. Precambrian Research, 172, 127–142.
Shields, G., Deynoux, M., Strauss, H., Paquet, H. & Nahon, D. 2007. Wingate, M. T., Pisarevsky, S. A. & De Waele, B. 2010. Paleomagnet-
Barite-bearing cap dolostones of the Taoudéni Basin, northwest ism of the 765 Ma Luakela volcanics in Northwest Zambia and
Africa: sedimentary and isotopic evidence for methane seepage after implications for Neoproterozoic positions of the Congo Craton.
a Neoproterozoic glaciation. Precambrian Research, 153, 209–235. American Journal of Science, 310, 1333–1344.
Chapter 13

Neoproterozoic sequences of the West Congo and Lindi/Ubangi Supergroups


in the Congo Craton, Central Africa

JENNY TAIT1*, FRANCK DELPOMDOR2, ALAIN PRÉAT2, LUC TACK3, GIJS STRAATHOF1 & VALENTIN KANDA NKULA4
1
School of Geosciences, University of Edinburgh, Edinburgh EH9 3JW, UK
2
Department of Earth Sciences and Environmental Sciences, University of Brussels, 1050 Brussels, Belgium
3
Department of Geology and Mineralogy, Royal Museum for Central Africa (RMCA), Tervuren, Belgium
4
Département des Sciences de la Terre, Faculté des Sciences, Université de Kinshasa (UNIKIN), Democratic Republic of the Congo
*Corresponding author (e-mail: jenny.tait@ed.ac.uk)

Abstract: The focus of this chapter is the West Congo Supergroup in the West Congo Belt (WCB), which extends along the western
margin of the Congo Craton from Gabon in the north to northern Angola in the south, and the Lindi/Ubangi Supergroup of the
Lindian and Fouroumbala – Bakouma Basins exposed on the northern margin of the craton. In both regions, up to two distinct diamictite
horizons have been recognized, the younger of which is often associated with carbonate rocks. Geochronological constraints are gener-
ally rather poor, many of the deposits lack modern sedimentological analysis, and the glacial versus non-glacial genesis of the diamictites
is a matter of debate in the literature. However, recent studies suggest a periglacial influence of diamictite deposition, particularly for the
sequences in the WCB. The stratigraphy of the various basins is described, available geochemical and geochronological information col-
lated, and recent work regarding the periglacial nature of the diamictites discussed. Finally, an updated chronostratigraphic correlation
between the basins is presented. However, much more work is required, particularly in the Neoproterozoic basins on the northern margin
of the Congo Craton, and more accurate geochronological constraints are required before the Neoproterozoic palaeogeography and
depositional environments of the western and northern Congo Craton can be fully understood.

The Congo Craton (Fig. 13.1), which is defined here as the central (Pinna et al. 1996; De Waele et al. 2006, 2008; Noce et al. 2007;
African landmass that amalgamated at the time of Gondwana Delor et al. 2008), and subsequently remained stable and a coher-
assembly (c. 550 Ma; De Waele et al. 2008), is encircled by Neo- ent block throughout Late Palaeoproterozoic and Mesoproterozoic
proterozoic sedimentary basins (Fig. 13.1, see also Hoffman 2011; times (Tack et al. 2006, 2008, 2009).
Master & Wendorff 2011). The Early Neoproterozoic was marked by rifting along the
Delhaye & Sluys (1923– 1924) first identified Precambrian western margin of the Congo Craton related to the break-up
diamictites of assumed glacial origin in the Democratic Republic of Rodinia and opening of the Adamastor Ocean (Tack et al.
of Congo (DRC), which Lepersonne (1951) subsequently inter- 2001), followed by the passive margin-type sedimentation of the
preted in the West Congo Belt (WCB) as representing two distinct West Congolian Group. The western margin of the Congo
diamictite horizons termed the Lower and Upper Diamictite Craton collided with the active São Francisco margin, thus
Formations. The stratigraphy of the WCB was first described in forming the Araçuai – West Congo Orogen (AWCO), including
detail in Gabon by Hudeley (1966), in the Popular Republic of the Brasiliano Araçuai belt now preserved adjacent to the São
Congo by Dadet (1969), in the DRC by Cahen (1978) and in Francisco Craton in Brazil (Pedrosa-Soares et al. 2008), and the
Angola by Stanton et al. (1963), and all available data were WCB in central Africa. Late Neoproterozoic deformation in the
synthesized by Trompette (1994). The glacial origin of the late WCB involved thrusting of Palaeoproterozoic basement rocks
Neoproterozoic WCB diamictites was the centre of much debate (the c. 2.1 Ga Kimezian Supergroup) onto the Neoproterozoic
in the early 1970s (see Kröner & Carreira 1973; Schermerhorn sequences of the West Congo Supergroup, while to the east the
1974; Cahen & Lepersonne 1981). More recently, these deposits external foreland basin sequences of the West Congo Supergroup
have been the subject of facies and microstructural analysis lie unconformably on the Archaean cratonic basement. Opening
(Delpomdor 2007a, b, c; Delpomdor et al. 2008) and a periglacial of the Atlantic in the Cretaceous split the Araçuai – WCB into
influence suggested. two parts, the Brazilian side of which inherited two-thirds of
Geochronological and palaeomagnetic data for the West Congo the AWCO, including all Neoproterozoic ophiolitic slivers, the
Supergroup and Lindi/Ubangi Supergroup are scant, and ages and entire magmatic arc, the suture zone and syn- (c. 585–560 Ma)
palaeolatitudinal constraints are largely derived indirectly by cor- to post- (c. 530–490 Ma) collisional magmatism (Pedrosa-Soares
relation (e.g. Evans 2000). Recent chemostratigraphic studies on et al. 2008).
the Haut-Shiloango and Schisto-Calcaire Subgroups by Frimmel The northern margin of the Congo Craton is marked by the
et al. (2006), Poidevin (2007) and Préat et al. (2011) support Late Neoproterozoic east – west trending Sergipano-Central
general lithological correlation between the Upper Diamictite and African Belt, which stretches from northern Brazil, through
Lower Diamictite Formations with the glaciogenic Chuos and southern Cameroon and the Central African Republic (CAR).
Ghaub Formations in the Kaoko-Damara belt to the south, This south-verging belt marks the boundary between the
and the Petit and Grand Conglomérat Formations in the Zambezi- Central African Mobile Zone to the north, and the northern
Lufilian belt to the SE. margin of the Congo and São-Francisco cratons. This northwes-
ternmost part of the Congo Craton is represented by the Archaean
Ntem complex, which is overlain by Proterozoic metasedimen-
Structural framework and basin setting tary and volcanic rocks (Lower Dja Series), and has been over-
thrust from the north by the Pan African Oubanguide Nappe
The Congo Craton comprises several Archaean nuclei (Fig. 13.1), of the Central African Belt. Nappe emplacement in southern
which welded together during the Eburnean orogeny c. 2.1– 1.8 Ga Cameroon is constrained by 620– 610 Ma ages obtained from

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 185– 194. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.13
186 J. TAIT ET AL.

0°E 10°E lt 20°E 30°E 40°E


an Be
ntra l Afric Fouroumbala-Bakouma
Serg ipe-Ce Basin

CUB LA

(East African Orogen)


Ubangui
N

Mozambique Belt
Lindi Ruw.
0°S Gabon Belt
LV
craton NE
Kibaran
Itombwe Belt
Sao Francisco
CRB

Ru
Mbuji- Tanzania

zis
craton

ian
West
West Congo
Mayi craton

Be
Congo

lt
Araçuai Belt Kibaran LT Ub
Bambui

lt
Belt en

Be
Belt dia
Kimezian Kasai

ran
nB
Belt el

ga
10°S t
Block erb
e l Bangweulu t

a
Us
p p L
Co Block
M
Katanga
B elt
Angola Lufilian m i de
Block Belt Iru h e r n elt
o ut e B Lurio
Legend S d
mi Foreland Fig. 13.1. Geological sketch map of
rI u S.M.
Phanerozoic to Recent cover Zambezi Belt Sub-Saharan Africa (modified after De
Neoproterozoic MB Waele et al. 2008) and showing the
CKB
Mesoproterozoic belts Congo-Uganda Block (CUB) and the
20°S
Palaeoproterozoic belts Zimbabwe Neoproterozoic sedimentary basins in the
Palaeoproterozoic with possible
Archaean basement (unexposed) Damara &
block NE. LA, Lake Albert; LV, Lake Victoria;
Archaean cratons Kaoko Belts km CKB, Choma Kalomo Belt; MB, Magondi
Limpopo 0 500
Belt.

plutonic rocks (Toteu et al. 2006), while a minimum age of Stratigraphy of the West Congolian Group
628 + 12 Ma has been obtained from syn-collisional granites in
the Sergipano Belt of Brazil (Oliveira et al. 2006; Bueno et al. The West Congo Supergroup is exposed along the western margin
2009). Pin & Poidevin (1987) suggest that the Central African of the Congo Craton in the Bas-Congo Basin of the DRC, and the
Belt represents a deeply eroded orogenic belt, whereas Abdelsa- Nyanga Basin of SW Gabon (Fig. 13.2). Up to two Neoproterozoic
lam et al. (2002) interpret it as the result of continent – continent diamictite units, interpreted as being glaciogenic in origin, have
collision between the northern Congo Craton and the Saharan been identified. The stratigraphy of these two sub-basins of the
Metacraton. WCB are described below.

8°E 10°E 12°E 14°E 10°E 12°E 14°E 16°E 18°E 20°E
0° LIBREVILLE 2°N


CONGO

N 2°S

4°S
GABON

2°S
SW Gabon DRC
6°S
Basin N ATLANTIC
YA 8°S
ATLANTIC N OCEAN
BA GA ANGOLA
OCEAN SI -NI
N AR
4°S I Bas-Congo
Basin KINSHASA

LEGEND

POST KAROO 6°S


KAROO SUPERGROUP
INKISI GROUP 100 1 200 300 km
WEST CONGOLIAN GROUP
MAYUMBIAN GROUP
ZADINIAN GROUP
8°S
KIMEZIAN
ARCHAEAN
NOQUI-TYPE GRANITES Fig. 13.2. Geological sketch map of the
LUANDA
West Congo Belt (after Frimmel et al.
2006).
THE WEST CONGO AND LINDI/UBANGI SUPERGROUPS 187

The Bas-Congo Basin, DRC Group, with two diamictite units, is exposed in the Bas-Congo
Basin (Fig. 13.3). Overlying the c. 1000–910 Ma volcanoclastic
The Neoproterozoic West Congo Supergroup of the DRC is divided Zadinian and Mayumbian Groups (Tack et al. 2001; Fig. 13.2),
into the Zadinian, Mayumbian and West Congolian groups. The the West Congolian Group (originally described in detail by
most complete Neoproterozoic sequence of the West Congolian Cahen 1978) corresponds to passive margin siliciclastic and

West-Congolian Group West-Congolian Group Lindi Supergroup Lindi/Ubangi Supergroup


SW Gabon Bas-Congo, DRC Fouroumbala-Bakouma Lindi, DRC
CAR
† 566 Ma

SCHISTO- MPIOKA
Dialanga Formation
GRESEUX SUBGROUP
GROUP BANALIA
Bili Formation GROUP
BAKOUMA
SC4 C5 SERIES
SCHISTO-
SC3
CALCAIRE †† 575 Ma
SC2 GROUP Bakouma Formation
SC1
NIARI TILLITE BONDO TILLITE Alolo Formation
C4 FORMATION
BOUENZA SCHISTO-
FORMATION ARUWIMI
CALCAIRE Mbiana Formation
x x x x x x x
GROUP
x x x x x x
SUBGROUP
x x x x x x x C3 Galamboge Formation
x x x x x x
x x x x x x x KEMBE-
NAKANDO Kole Formation
SERIES Mamungi LOKOMA
C2
Formation GROUP
C1
Bobwamboli Formation
UPPER DIAMICTITE (?)
FORMATION AKWOKWO TILLITE
Asoso Formation
†† 645 Ma (Sh8) †† 730-750 Ma
BOUGBOULOU Lenda ITURI
Formation
SERIES x
x
x
x
x
x
x
x
x
x
x
x
x GROUP
x x x x x x x Penge Formation
x x x x x x
x x x x x x x

† 650 Ma (Sh6) LOWER TILLITE


x x FORMATION
HAUT x
x x
x
x
x
x
x

SHILOANGO x
x
x
x
x
x
x
x
x
x
x
x
x

SUBGROUP

LEGEND

DOLOMITE

v v LOWER LIMESTONE/ ARGILLACEOUS


DIAMICTITE LIMESTONE
v v v
FORMATION v DIAMICTITE/ BASALT-DOLERITE SILL
SHALE
† 923 ± 43 Ma
SANDSTONE-QUARTZITE/
CONGLOMERATE
SANSIKWA x x x x
x x x x
SUBGROUP x x x x PALAEOPROTEROZOIC/ARCHAEAN

v v v
OOLITE, CYANOBACTERIAL MATS,
STROMATOLITE
††† 910 Ma PANAFRICAN TECTONIC
MAYUMBIAN UNCONFORMITY
† Frimmel et al. (2006)
GROUP
†† Poidevin (2007)
††† Tack et al. (2001)

Fig. 13.3. Stratigraphic logs of Neoproterozoic sequences from the western and northern margins of the Congo Craton. Modified from Delpomdor et al. (2011).
188 J. TAIT ET AL.

carbonate platform deposits and is particularly well developed in lacustrine-fluvio-glacial infilling of extensional zones in the
the foreland part of the WCB where it is only gently folded and Central African basement (Alvarez 1995). On the east flank of
unmetamorphosed. The sedimentary rocks have been subdivided the basin, the Niari Group rests on the Bouenza Formation with
into four subgroups: the Sansikwa, Haut-Shiloango, Schisto- a non-angular unconformity.
Calcaire and the Mpioka Subgroups as described below. The The Niari diamictites are overlain by the Schisto-Calcaire Group
uppermost Inkisi ‘Subgroup’ is no longer considered to be part (or Carbonate Subgroup; Prian 2008) which is correlated with the
of the Neoproterozoic sequences (Frimmel et al. 2006 and refer- Schisto-Calcaire Subgroup of the Bas-Congo Basin (DRC) and
ences therein) and considered post-Pan African in age (see discus- thought to be Late Neoproterozoic in age. Again, precise ages
sion further below). from the Gabon sequences are lacking, and the most reliable
The Sansikwa Subgroup forms the base of the West Congolian ages and correlations are summarized in Figure 13.3. The Schisto-
Group and comprises siliciclastic rocks representing a continental Calcaire Group of the SW Gabon Basin is predominantly a carbon-
rift depositional environment. The Lower Diamictite Formation ate sequence composed of four formations consisting of calcareous
marks the top of the Sansikwa Subgroup, and contains up to to dolomitic shales with an uppermost sandy shale-siltstone unit
400 m of diamictite with interbedded sands and shales (Cahen & with interbedded limestones. Carbonate deposition of the upper
Lepersonne 1981). Tholeiitic basalts, sometimes including pillows part was under similar sedimentological environments to those
and hyaloclastic breccia, are interlayered in the diamictite (De of the Bas-Congo (Préat et al. 2010), indicating hypersaline
Paepe et al. 1975; Kampunzu et al. 1991) with accompanying shallow water sub-supratidal conditions, in possible lagoonal
feeder dykes and sills intruded in the underlying Sansikwa Sub- setting with development of lithoherms and cyanobacterial mats.
group. Their emplacement age has yet to be confirmed, as the In SW Gabon, the Schisto-Calcaire Group is overlain by the
40Ar/39Ar resetting age of 566 + 42 Ma is thought to represent Schisto-Gréseux Group (Chevalier et al. 2002), which is correlated
regional greenschist facies metamorphism during Pan African with the Mpioka Subgroup of the Bas-Congo Basin (Tack et al.
orogeny and amalgamation of western Gondwana (Frimmel 2001; Frimmel et al. 2006). The series is thinner in Gabon (up
et al. 2006). to 100 m thick) and of Late Neoproterozoic age (Thiéblemont
The Sansikwa Subgroup is overlain by 700–800-m-thick, pre- et al. 2009). The Schisto-Gréseux Group is divided into two
dominantly siliciclastic rocks of the Haut-Shiloango Subgroup. subgroups: the lower subgroup is predominantly composed of
Conglomerates, quartzarenites, argillites and shales make up the siltstones and argillites with, in places, a 6– 15-m-thick basal con-
lower two-thirds of the Haut-Shiloango Subgroup. The Sekelolo glomerate, and the upper subgroup contains feldspathic sandstones
Limestone forms the upper part and consists of dominantly and argillites.
subtidal, medium- to fine-bedded limestones with a minor clastic
component. The overlying Upper Diamictite Formation, which is
recognized throughout the length of the WCB, from southern Stratigraphy of the Lindi/Ubangi Supergroup
Gabon (the Niari Formation) through to northern Angola, is up
to 200 m thick in the Bas-Congo Basin (Cahen & Lepersonne Neoproterozoic sedimentary rocks exposed on the northern Congo
1981). It is overlain by a sequence of up to 12-m-thick, finely lami- Craton are recognized in the Bangui Basin of southwestern CAR,
nated pink dolomitic carbonates, forming the lowermost part of the Fouroumbala-Bakouma Basin of south central CAR and the
the Schisto-Calcaire Subgroup, which share many lithological Lindian Basin of northeastern DRC (Fig. 13.1). Trompette (1994)
characteristics with basal Ediacarian carbonates elsewhere. The considered all these sub-basins to be part of the Lindian Basin.
remainder of the Schisto-Calcaire Subgroup consists of a range Correlation is problematic due to the limited number of studies
of dominantly shallower-water carbonate shelf facies, including that have been carried out, poor or non-existent age constraints,
oolites, diverse stromatolites, evaporites and cherts (Alvarez poor outcrop conditions and problems concerning accessibility.
1995). The Schisto-Calcaire Subgroup shows a transition from Diamictite units have been recognized in all three basins
east to west between dolomitic intertidal facies with stromatolitic (Fig. 13.4). The best studied is the Lindian Basin of northeastern
biostromes to supratidal facies with evaporitic needles, ooids and DRC, which was described in detail by Verbeek (1970), and the
cross-bedding (Delpomdor 2007c; Préat et al. 2010, 2011). stratigraphy of which is summarized below.
These two facies are separated by a stromatolitic reef barrier
(Trompette 1994).
The Schisto-Calcaire Subgroup is overlain by the Mpioka The Lindian Basin
Subgroup, c. 1000-m-thick siliciclastic succession with conglom-
erates, quartzoarenites and argillites, interpreted as being a The Lindian Basin is an intracratonic basin that developed on
late-orogenic molasse deposit. The Mpioka Subgroup is recog- the northern Congo Craton in the Neoproterozoic. The rocks of
nized throughout the WCB, and the uppermost subgroup has the Lindi Supergroup, which are up to 2500 m thick, rest uncon-
been locally affected by Pan African Late Neoproterozoic/Early formably upon crystalline Archaean and/or Early Proterozoic
Cambrian deformation. In this regard, it is analogous to the basement rocks and have undergone little or no metamorphism.
Mulden Group along the southwestern margin of the Congo To the west of the Lindian Basin, the continuation of this sequence
Craton. in the DRC has been termed the Ubangi Supergroup. As summar-
ized by Verbeek (1970) and Trompette (1994), the Lindi
Supergroup is divided into three groups: the basal Ituri, the
The SW Gabon Basin (or the Nyanga-Niari Basin) Lokoma and the Aruwimi groups, the last of which is subdivided
into the Galamboge, Alolo and Banalia formations. However, in
The West Congo Supergroup of Gabon is exposed in the Nyanga Figure 13.3 the Aruwimi Group as originally defined (Verbeek
synclinorium (Gérard 1958; Dadet 1969), and is subdivided into 1970) is here split into two units, with the uppermost Banalia For-
several informal units. The oldest sedimentary rocks comprise mation ranked as a separate, younger ‘Group’ (see below). Only
the Bouenza Formation (Fig. 13.3), a fluvial sequence up to 100 m one diamictite horizon, occurring at the base of the Lokoma
thick (Prian 2008), which records erosion of the Palaeoprotero- Group and termed the Akwokwo Tillite, is interpreted as being
zoic and Mesoarchaean basement. The sandstones are massive glacial in origin (Fig. 13.3).
with clear cross-bedding and ripple cross-lamination, suggesting The basal clast-supported conglomerate with coarse-grained
palaeocurrents from SW to NE. These are overlain by the arkoses and quartzarenites (30 –50-m-thick Penge Formation,
fluvio-glacial deposits of the Niari Group (or Niari Tillite, Ituri Group) is overlain by thick oolitic and stromatolitic limestone
Fig. 13.3). The Bouenza Formation and the Niari Group record with dolomite layers (the Lenda Formation). Poidevin (2007)
THE WEST CONGO AND LINDI/UBANGI SUPERGROUPS 189

West-Congo West-Congo Lindi Ubangi/Lindi Lindi, DRC


SW Gabon Bas-Congo, DRC Fouroumbala- Bangui, CAR
Bakouma
CAR
POST-PRECAMBRIAN INKISI BANALIA
SUBGROUP GROUP
Panafrican
E Tectonic †566 Ma

D SCHISTO-
I GRESEUX MPIOKA ARUWIMI
A GROUP SUBGROUP GROUP
C BAKOUMA
Post-Marinoan SUBGROUP
A
R ††575 Ma
SCHISTO-
I 600 SCHISTO- BANGUI LOKOMA
CALCAIRE
A Ma CALCAIRE GROUP
GROUP
N SUBGROUP BIMBO

C Marinoan NIARI UPPER BONDO CONGLO. AKWOKWO


R Glaciation TILLITE DIAMICTITE TILLITE TILLITE
Y ††645 Ma KEMBE
O NAKANDO
G SUBGROUP ITURI
E Post-Sturtian †650 Ma GROUP
BOUENZA
N FORMATION HAUT- BOUG- ††730-750 Ma
I SHILOANGO BOULOU ?
700
A SUBGROUP SUBGROUP
N
Sturtian LOWER LOWER
Glaciation DIAMICTITE TILLITE

†923 ± 43 Ma

800

T LEGEND
O Pre-Sturtian
N Diamictites/Mixtites/Tillites
I SANSIKWA
A SUBGROUP
Sandstones/Quartzites
N Cap carbonates/Dolomites
Limestones/Dolomites
900 Shales/Argillites/Pelites
†††910 Ma

† Frimmel et al. (2006)


†† Poidevin (2007) Fig. 13.4. Stratigraphic correlation of
††† Tack et al. (2001) Neoproterozoic sequences of the western and
northern Congo Craton. Correlations based on
Sr isotopic data, radiometric age constraints
and revised lithological relationships.
1000 Modified after Poidevin (2007).

reported the presence of early Neoproterozoic microfaunas in these medium-grained quartzoarenites interbebbed with argillites. The
horizons. The top of the Ituri Group is marked by c. 50 m of mica- quartzoarenites are marked by a rapid transition to the Alolo
ceous shales, carbonates and sandstones of the Asoso Formation. Formation, and comprise 400-m-thick clayey to calcareous argil-
The entire Ituri Group represents lagoonal or shallow marine plat- lites, often stratified, interbedded with dark carbonaceous argillite
form type sedimentary rocks (Daly et al. 1992) and is restricted to layers, fine-grained sandstones, zoned pink and oolitic limestones
the eastern part of the Lindian Basin. It has been tentatively corre- and dolomitic lenses. The uppermost Banalia sequences are
lated with both the Haut-Shiloango (Trompette 1994; Daly et al. composed of .1000 m of massive or finely stratified and fre-
1992) and the Sansikwa Subgroup (Poidevin 2007) of the West quently cross-bedded fine-grained reddish arkoses interbedded
Congolian Group. with reddish micaceous and clayey argillites. The contact between
The Lokoma Group (.500 m thick) overlies the Ituri Group the Banalia and the Alolo Formations is apparently conformable.
with an angular unconformity, and begins with the Akwokwo In terms of thickness and depositional facies, the deltaic Banalia
Tillite, which has only limited exposure in the eastern parts of the Redbeds have been correlated with the similar Inkisi ‘Subgroup’
basin where it sits in erosional contact on the underlying Ituri of the Bas-Congo Basin and interpreted as evidence of post-
Group. It is up to 40 m thick, with diverse and poorly sorted Gondwana amalgamation (550 Ma) and pre-Karoo break-up in
clasts (Verbeek 1970; Cahen 1978) in a grey-green clay matrix. central Africa (Tack et al. 2008, 2009a). For this reason, they
The rest of the Lokoma Group consists of arkoses, conglomerates are ranked here, in both Figures 13.3 and 13.4, as (new) separate
and argillites. groups. On the other hand, new geophysical data in and around
In the original definition of Verbeek (1970), the Aruwimi Group the DRC Cuvette Centrale suggest that not only the Banalia
is up to 1500 m thick and subdivided into three formations. The Redbeds but the complete Aruwimi Group (thus including also
Galamboge Formation (c. 100 m thick) consists of cross-bedded the lowermost Galamboge and overlying Alolo Formations) are
190 J. TAIT ET AL.

post-Pan African in age (Kadima et al. 2011). The precise age of into the fluvio-lacustrine Kembé-Nakando sandstones of the
the Galamboge and Alolo Formations (i.e. upper Neoproterozoic Fouroumbala-Bakouma Basin in eastern CAR, and may be glacial
versus Phanerozoic) clearly needs to be determined using appro- in origin (Alvarez 2000).
priate radiometric dating. The Bimbo sandstones are unconformably overlain by the
Bangui Series (Fig. 13.4), a thick carbonate sequence exposed
mainly along the Ubangui River to the south of Bangui. The
The Fouroumbala-Bakouma Basin Bangui Series carbonates have been affected by Pan-African
related deformation and, therefore, are considered to be Late Neo-
The Fouroumbala-Bakouma Basin is located in the south central proterozoic in age. They have been correlated with the Schisto-
region of the CAR (Fig. 13.1). Three main sedimentary groups Calcaire Subgroup of the West Congolian Supergroup (Alvarez
have been recognized (Fig. 13.3): the Bougboulou, Kembé- 1995). They have been subdivided into the distal Lessé Formation,
Nakando and Bakouma Subgroups (or Series). The stratigraphy the proximal outer shelf Bobassa Formation, the offshore barrier
of the Fouroumbala-Bakouma Basin was described in detail by facies Mboma Formation and the lagoonal Fatima Formation.
Bigotte & Bonifas (1968), Poidevin et al. (1981) and Poidevin These carbonates comprise laminated microbial limestones, rhyth-
(1985) with recognition of two diamictite horizons. The base of mically bedded carbonates with clay and silt-rich interval inter-
the Bougboulou Subgroup rests with an angular unconformity on preted as shallow subtidal or intertidal deposits, more massively
the Archaean Bangui-Ketté basement. A basal diamictite, termed bedded limestones, and are capped by grey microbial limestones
the Lower Tillite, has been identified in the extreme SW of the intercalated with black dolomites indicating deposition in a
basin. The Bougboulou Subgroup comprises predominantly inter- lagoonal environment (Alvarez 1992). Alvarez (1995) interpreted
bedded shales and sandstones, although in the south of the Four- these sedimentary rocks as characterizing an extensive carbonate
oumbala Basin, silicified stromatolitic carbonates (the Kassa ramp that developed along the northern edge of the Central
Formation) have been identified (Mestraud 1952, 1953). The Africa belt. However, it is now thought that they developed in
Bougboulou Subgroup is unconformably overlain by the deltaic local intracratonic basins and they have been correlated with the
to neritic Kembé-Nakando Subgroup comprising sandstones and Bakouma Subgroup in more easterly regions of CAR.
black and red shales of the Mbiana Formation. The base of the
overlying Bakouma Subgroup is marked by the Bondo Tillite,
which is thought to be glacial in origin, containing varves and Glaciogenic character of the WCB diamictites
dropstones. The Bondo Tillite is overlain by the red and grey
Bakouma Formation dolomites, which are similar to other post- Until 1970, the diamictites of the West Congolian Group were gen-
glacial (cap) carbonates of the Congo Craton (Poidevin 2007). erally considered to be glacial in origin, and in 1986 the Upper Dia-
This association, however, is only known from drill cores with mictite of the Congo was described as a marine tillite associated
no surficial exposure. The Bakouma Formation is overlain by the with an ice cap on the Chaillu massif (Trompette & Boudzoumou
Bili and Dialanga Formations (Fig. 13.3). The Bili Formation 1988). In the Congo, however, Cahen & Lepersonne (1981) con-
comprises basal shales overlain by white and grey limestones. sidered the Lower Diamictite Formation to be a tectonically
The clastic Dialanga Formation is the youngest member of the related gravity flow deposit. Subsequently, Trompette (1994) con-
Bakouma Subgroup and comprises black shales and quartzo- sidered the Upper Diamictite Formation to be a gravity-flow
arenites. Poidevin (2007) correlated the Bakouma Group with deposit originating from reworking of glacial material along the
the Aruwimi Group (original definition of Verbeek, 1970) of the margins of the basin by small mountain glaciers developed on per-
Lindi Supergroup. If this is correct, the upper tillites (Bondo Til- ipheral highlands. In Angola, the discussion was more acute. Both
lites) are missing in the Aruwimi Group of the Lindi Supergroup. diamictites were interpreted as being tilloids, that is, non-glacial
and deposited as debris flows (Schermerhorn & Stanton 1963),
although the possibility of reworking of mountain glacier deposits
The Bangui Basin (i.e. Alpine glacial deposits) was suggested for the lower diamic-
tites (Schermerhorn 1974). Kröner & Correia (1973) reinterpreted
The Bangui Basin was described lithostratigraphically by Babet both diamictite levels in Angola as being true glaciomarine dia-
(1935), Legoux & Hourcq (1943), Gerard & Gerard (1952), Bes- mictites with local intercalations of continental tillites. Schermer-
soles & Trompette (1980) and Poidevin (1976, 1979a, 1979b, horn (1981) continued to ascribe the deposits to a deep-water
1985, 2007). Neoproterozoic sedimentary rocks are exposed in submarine turbiditic setting, arguing against a glacial origin. Simi-
SW CAR around the capital Bangui, and extend southwards to larly, Vellutini & Vicat (1983) argued against any glacial influence
the Congo River and overlie the Nola Group quartzopelitic meta- and considered both diamictites of the West Congo Belt as basal
sedimentary rocks, which are thought to be Palaeoproterozoic in conglomerates. In his synthesis of available data and taking all
age. The Nola Group sequences have been overthrust in the the arguments into account, Trompette (1994) surmises that
north by the Pan-African Oubanguides belt (i.e. Central African there was most likely some glacial input, even if they do not rep-
Belt), and intruded by the Nola dolerite dykes. These dykes are resent true glaciomarine deposits.
undated but yield a pan-African resetting Ar/Ar age of 571 Ma More recently, Tack et al. (2006) critically reappraised the
(Moloto-A-Kenguemba et al. 2008). Correlation with 950 Ma origin of Neoproterozoic diamictites from central Afria. As a result,
dolerite dykes in Gabon has been suggested (Vicat et al. 1996) detailed micro-structural and -textural analyses (based on tech-
but remains to be demonstrated. niques applied to recent unconsolidated tills by several authors
The Neoproterozoic sedimentary sequences of the Bangui such as Evans & Benn 2004; Benn & Evans 1996; Menzies
Basin are subdivided into the Kembe conglomerates, the Bimbo 2000a, b, 2006; Phillips 2006; Phillips et al. 2007; Van der Meer
sandstones and the Bangui carbonates (Fig. 13.4). The basal 1993, 1997, 2003) were carried out on samples of the Lower and
Kembe conglomerates (up to 200 m thick and exposed near the Upper Diamictite of the Bas-Congo Basin, the Akwokwo Tillite
village of Bimbo along the Ubangui River), are tectonized and of the Lindian Basin, and the Niari Formation of the Nyanga
comprise fine-grained polymict conglomerates interbedded with Basin (Delpomdor 2007c, Delpomdor et al. 2008, Préat 2008,
mylonitized quartzitic sandstones and schistose shales. The over- Préat et al. 2008). These observations provide important new con-
lying Bimbo sandstones (130 m thick) are a thick fluviatile straints on the various sedimentological and depositional environ-
sequence of massive dark shales interstratified with thick beds ments. The diamictites from the SW Gabon, Bas-Congo and the
of conglomerate, sandstones and quartzarenites with cross- Lindian Basin are principally composed of massive and stratified
stratification, which, according to Alvarez (1995) pass eastwards mud-supported matrix with rare intercalation of varved mudstones
THE WEST CONGO AND LINDI/UBANGI SUPERGROUPS 191

and sandstones. Clasts within the diamictites are only rarely Geochronological constraints
striated, do not display crescent or conchoidal fractures, and no
glacial pavements have been observed. Thus, a directly subglacial Radiometric data from the West Congo and Lindi/Ubangi Super-
environment cannot be identified. The combination of observed groups are sparse and geochronological constraints are based
microstructures demonstrates that all these diamictites were depos- largely on bulk detrital zircon analyses and correlation with bette-
ited under high strain rates with moderate to high stress conditions, dated successions elsewhere on the Congo Craton. The Lower
influenced by water pressure or clays within the beds. The cumu- Diamictite Unit is constrained only to be younger than c. 910 Ma
lative evidence of overprinting microstructures suggests that all based on the youngest U –Pb ages from the Mayumbian Group
these diamictites were deposited in fluvioglacial or glaciomarine volcanic rocks, which underlie the Sansikwa Subgroup (Tack
conditions in proximal subglacial and/or more distal subaqueous et al. 2001). The youngest detrital zircon ages obtained from the
environments. overlying Haut-Shiloango Subgroup yield ages of 547 + 45,
671 + 20 and 709 + 20 Ma (1s errors, 10% discordance)
from which, given the errors, a possible age of 650 Ma is inter-
Chemostratigraphy preted by Frimmel et al. (2006). Clearly, more reliable age
data are required to more accurately constrain maximum and
The only chemostratigraphic data available for the WCB are from minimum depositional ages for both the Lower and Upper Diamic-
a handful of recent studies (Frimmel et al. 2006; Poidevin 2007; tite Formations. The 40Ar/39Ar age of 566 + 42 Ma obtained from
Straathof et al. 2008; Préat et al. 2010, 2011). Strontium isotope a sill intruding the Lower Diamictite Unit and dating greenschist
ratios measured in carbonate rocks of the West Congo Supergroup metamorphism during west Gondwana amalgamation (Frimmel
range from 0.7066 to 0.7109 (Poidevin et al. 2007), although the et al. 2006) provides a loose minimum age constraint for the
highest ratios are surely overprints. The least-altered and most entire sedimentary package.
Sr-rich samples in the Haut Shiloango Subgroup have 87Sr/86Sr
ratios of c. 0.7068 –0.7072, compared to ratios of 0.7074–
0.7075 in the lower Schisto-Calcaire Subgroup (Frimmel et al. Correlation between the West Congolian Group and the
2006; Poidevin 2007). Ten d13Ccarbonate values in the Haut Lindi Supergroup
Shiloango Subgroup range from 3.2 to 8‰. The Schisto-Calcaire
Subgroup defines a larger range in d13Ccarbonate, from – 5.4 to Updated lithostratigraphic correlations are summarized in
9‰. Values in the lowermost part of the subgroup are consistently Figure 13.4 and are based on Sr isotope ratios obtained from
negative (Frimmel et al. 2006; Straathof et al. 2008; Préat et al. carbonates, SHRIMP data (Tack et al. 2001; Frimmel et al.
2011), whereas those in the upper part of the subgroup are 2006; Poidevin 2007) and revised lithological relationships. The
erratic, probably due to temporarily elevated evaporation rates Neoproterozoic West Congolian Group sedimentary rocks are
(Frimmel 2009). The near-shore depositional environment for relatively well constrained in comparison to those of the Lindian
the Bas-Congo carbonates is also demonstrated geochemically Supergroup in the northern part of the Congo Craton. In the Bas-
by Frimmel (2009) who concluded that carbon isotopes obtained Congo Basin, the West Congolian Group starts with siliciclastic
from successions developed in such environments, particularly deposits of the Sansikwa Subgroup. The lower age of the Sansikwa
when elevated evaporation has occurred, should be treated Subgroup is constrained by the underlying 912 + 7 Ma rhyolites
with great care and are not suitable for chemostratigraphic (Tack et al. 2001). The upper age of the Sansikwa Subgroup is
correlation. less well known, but has been inferred by Frimmel et al. (2006)
to be c. 750 Ma based on correlation of the overlying Lower Dia-
mictite unit with the Sturtian glacial event (assuming the Sturtian
Palaeolatitude and palaeogeography to have occurred at 720– 750 Ma, although this age is controver-
sial). No carbonate rocks are observed in the Sansikwa Subgroup;
The Precambrian palaeomagnetic database for the Congo Craton is the oldest carbonate rocks occur at the top of the post-Sturtian
extremely sparse, with very few of the available data meeting Haut-Shiloango Subgroup (Fig. 13.3). However, Poidevin (2007)
modern reliability criteria. Nevertheless, the latest Mesoprotero- proposed that the oldest limestones in the Lindi Supergroup
zoic position of the Congo can be constrained by poles obtained (the Lenda Formation, Ituri Group) are pre-Sturtian and 730–
from the São Francisco Craton (D’Agrella-Filho et al. 1990). 755 Ma in age on the basis 87Sr/86Sr ratios and correlation with
These data place the western margin of the craton (present day the chemostratigraphic compilation of Halverson et al. (2007).
co-ordinates) at intermediate palaeolatitudes (45 –508) for the time However, the isotopic ratios reported by Poidevin (2007) are extre-
period 1.1 –1.0 Ga (De Waele et al. 2008). Neoproterozoic palaeo- mely variable (0.70663 –0.71090), making correlation difficult.
geography is constrained by two poles from the 795 + 7 Ma Given the lack of any carbonate rocks prior to deposition of the
Gagwe lavas of Tanzania (Piper 1972; Meert et al. 1995, Ar/Ar lower diamictite unit (considered to be Sturtian in age in the Bas-
age after Deblond et al. 2001) and the 743 + 30 Ma Mbozi intru- Congo), the limestones of the Lenda Formation (Ituri Group) are
sive rocks (Meert et al. 1995, K –Ar biotite age after Brock 1968). considered to be post-Sturtian in age, as traditionally presented
According to these data, by the early Cryogenian the Congo craton in the literature (Trompette 1994), and correlated with the Haut-
had moved into tropical latitudes, positioning the present-day Shiloango Subgroup of the Bas-Congo Basin and the Bougboulou
western margin close to the equator, and by 755 Ma, had rotated Series in the Fouroumbala-Bakouma Basin (Fig. 13.4). In Gabon,
some 908 while remaining in the tropics. Collisional deformation the Bouenza Group is a correlative of Haut-Shiloango Subgroup
related to consolidation of western Gondwana was initiated c. (Chevallier 2002).
600 Ma in the western Congo and 580 –530 Ma in the south and The Ituri Group is overlain by the Akwokwo Tillite which, as
southeast. Palaeomagnetic data from the Sinyai dolerites of mentioned above, Poidevin (2007) correlated with the Lower Dia-
Kenya (Meert & Van der Voo 1996) and geological evidence indi- mictite Formation of the West Congolian Group. However, recent
cate that proto-Gondwana had assembled by the Early Cambrian. field observations of Neoproterozoic diamictites in SW Gabon
The implications of the available palaeomagnetic data are that and/or in the DRC Bas-Congo and Katanga regions show that car-
the Neoproterozoic sedimentary rocks of the West Congo and bonate rocks comprising massive zoned pink to greyish dolomites
Lindi Supergroups were deposited in tropical latitudes. In terms immediately cap the upper diamictites in these regions, but are not
of more detailed palaeogeography, it is important to point out found in association with the lower diamictites. These ‘cap carbon-
that the paucity of Precambrian data for the Congo Craton ates’ are also recorded in the Fouroumbala-Bakouma Basin (the
means the polarity of the palaeopoles is uncertain. Bakouma Formation capping the Bondo Tillite) (Fig. 13.3). In
192 J. TAIT ET AL.

the Lindian Basin, large lenses of zoned pink dolomites are Babet, V. 1935. Esquisse géologique provisoire de la région comprise
embedded in the Mamungi Formation overlying the Akwokwo entre Bangui et la frontière du Cameroun. Chronique Mineralogique
Tillites. The close association of the Akwokwo Tillite with Col. Paris, 38, 160– 164.
massive zoned pink to greyish dolomites is also confirmed from Benn, D. I. & Evans, D. J. A. 1996. The interpretation and classification
the RMCA (Tervuren) sample collection (Tack, pers. comm.). of subglacially-deformed materials. Quaternary Science Reviews, 15,
Hence, the occurrence in DRC and adjacent CAR of the cap car- 23 – 52.
bonate is thought to be indicative of the Marinoan glacial event Bessoles, B. & Trompette, R. 1980. Géologie de l’Afrique: la chaı̂ne
(Fig. 13.4). No carbonate is observed capping the diamictites of panafricaine ‘zone mobile d’Afrique Centrale (partie sud) et zone
mobile soudanaise’. Mémoire du Bureau de la Recherche Geleolgi-
the Bangui Basin, but the Kembé Formation conglomerates are
que et Minière, 92, 398.
commonly correlated with the Bondo Tillite of the Fouroumbala- Bigotte, G. & Bonifas, G. 1968. Faits nouveaux sur la géologie de la
Bakouma Basin (Poidevin 1976, 1985, 2007; Bessoles & Tromp- région Bakouma. Chronique de Mines et Recherche Minerale, 36,
ette 1980; Cornacchia & Giorgi 1986). Finally, oolitic carbonate 43 – 46.
rocks of the Lindi Supergroup occur in the Alolo Formation Bueno, J. F., Oliveira, E. P., McNaughton, N. J. & Laux, J. H. 2009.
(Aruwimi Group overlying the Lokoma Group; Fig. 13.3). Com- U –Pb dating of granites in the Neoproterozoic Sergipano Belt,
parable oolitic carbonate rocks of post-Marinoan age are also NE-Brazil: implications for the timing and duration of continental
known up sequence in the West Congo and Katanga regions. collision and extrusion tectonics in the Borborema Province. Gond-
In the Bili and Dialinga Formations (upper part of the Bakouma wana Research, 15, 86– 97.
Series) of the Fouroumbala-Bakouma Basin, some metabasalt Cahen, L. 1978. La stratigraphie et la tectonique du Supergroupe Ouest-
layers are intercalated. In the Aruwimi Group of the Lindian Congolien dans les zones médiane et externe de l’orogénèse Ouest-
Basin, Verbeek (1970) considers the Kaparata Breccia to be a con- Congolien (Pan-African) au Bas-Zaı̈re et dans les régions voisines.
glomerate, possibly including some clasts of a volcanic rock. In Annales de la Musée Royale de l’Afrique Centrale, Tervuren, in 88,
fact, the RMCA (Tervuren) Kaparata samples are a hyaloclastic Science Geologique, 83, 150.
breccia, and are reminiscent of rocks observed in the Bas-Congo Cahen, L. & Lepersonne, J. 1981. Proterozoic diamictites of Lower
Basin, suggesting the occurrence in the Lindian Basin of Zaire. In: Hambrey, M. A. & Harland, H. W. (eds) Earth’s
pillow lavas and an up to now radiometrically undated tholeiitic Pre-Pleistocene Glacial Record. Cambridge University Press,
basaltic event. Cambridge, 153–157.
In conclusion, unlike Poidevin (2007), we correlate the Akwokwo Chevallier, L., Makanga, J. F. & Thomas, R. J. 2002. Carte géologique
Tillite (Lindian Basin) with the Bondo Tillite (Fouroumbala- de la république gabonaise, 1:1,000,000. Notice explicative. Council
for Geoscience, South Africa.
Bakouma Basin), the Upper Diamictite (Bas Congo) and Niari
Cornacchia, M. & Giorgi, L. 1986. Les séries Précambriennes d’origine
Tillite (SW Gabon), and conclude a Marinoan age. However, as sédimentaire et volcano-sédimentaire de la République Centrafi-
already suggested by Poidevin (2007), the Upper Bakouma quaine. Annales de la Musée Royale de l’Afrique Centrale, Tervuren,
Series are correlated with the Aruwimi Group, as redefined in in 88, Science Geologique, 93, 51.
this paper (Fig. 13.3). This correlation is tentatively extended to Dadet, P. 1969. Notice explicative de la carte géologique de la république
the Mpioka Subgroup of the WCB (in SW Gabon and DRC Bas- du Congo-Brazzaville au 1:500 000. Mémoire du Bureau de la
Congo). Finally, the Redbeds of the Inkisi ‘Group’ of the WCB, Recherche Geleolgique et Minière, 40, 103.
with a maximum depositional age of c. 550 Ma (i.e. post-Pan D’Agrello-Filho, M. S., Pacca, I. G., Renne, P. R., Onstott, T. R. C.
African) are correlated with the Banalia ‘Group’ of NE DRC (for- & Teixeira, W. A. 1990. Paleomagnetism of middle Proterozoic
merly considered as the uppermost unit of the Neoproterozoic (1.01– 1.08 Ga) mafic dykes in Southeasterrn Bahia State –São Fran-
Lindi Supergroup). cisco Craton, Brazil. Earth and Planetary Science Letters, 101,
Without new geochronological and palaeomagnetic constraints 332– 348.
for these Neoproterozoic basins it will be difficult to refine Daly, M. C., Lawrence, S. R., Diemu-Tshiband, K. & Matouana, B.
these correlations. 1992. Tectonic evolution of the Cuvette Centrale, Zaire. Journal of
the Geological Society, London, 149, 539–546.
In Gabon, the fieldwork was carried out under the terms of the ‘Programme Deblond, A., Punzalan, L. E., Boven, A. & Tack, L. 2001. The Mala-
Sysmin, 8ème Fonds Européen du Développement au groupement BRGM- garazi supergroup of southeast Burundi and its correlative Bukoba
CGS-SANDER-MRAC’. J.T. and G.S. gratefully acknowledge financial supergroup of northwest Tanzania: Neo- and Meso-proterozoic con-
support from the European Commission FP6 Programme (Marie Curie Excel- straints from Ar – Ar ages of mafic intrusive rocks. Journal of African
lence Grant). H. Frimmel and U. Zimmermann are gratefully thanked for Earth Science, 32, 435– 449.
their helpful and constructive reviews. This represents a contribution of the Delhaye, F. & Sluys, M. 1923– 1924 and 1928– 1929. Observations
IUGS- and UNESCO-funded IGCP (International Geoscience Programme) ayant servi a l’elaboration de l’ ‘Esquisse geologique du Congo occi-
Project #512. dental’. Etude du systeme Schisto-Calcaire. 1er– 3e memoire.
Annales de la Société Belge de Géologie, 1923– 1924, C. 50– 91 et
1928– 1929, C. 69– 114.
Delor, C., Theveniaut, H. et al. 2008. New insights into the Precam-
References brian Geology of Angola: basis for an updated lithochronological
framework at 1:2,000,000 scale. 22nd Colloquium African Geology,
Abdelsalam, M. G., Liegeois, J. P. & Stern, R. J. 2002. The Saharan Hammamet, Tunisia, Abstracts volume, 52 –53.
Metacraton. Journal of African Earth Sciences, 34, 119– 136. Delpomdor, F. 2007a. Lithostratigraphie et sédimentologie du faisceau
Alvarez, P. 1992. Répartition de la sédimentation dans le golfe Protéro- de Sekelolo Sh8 (Sous-Groupe du haut-Shiloango), de la Formation
zoı̈que supérieur du Schisto-calcaire au Congo et Gabon. Implications de la Diamictite supérieure et de la Dolomie rose C1 du fiasceau du
en Afrique centrale. Palaeogeography, Palaeoclimatology, Palaeo- Kwilu CI (Sous-Groupe du Schisto-Calcaire), anticlinal de Congo
ecolology, 96, 281– 297. Dia Kati, Bas-Congo, République Démocratique du Congo (RDC),
Alvarez, P. 1995. Evidence for a Neoproterozoic carbonate ramp on the unpublished Internal report of Royal Museum of Central Africa, Ter-
northern edge of the Central African Craton: relations with Late vuren (Belgium).
Proterozoic intracratonic troughs. Geologische Rundschau, 84, Delpomdor, F. 2007b. Etude des dépots diamictitiques du Néoprotérozoı̈-
636– 648. que supérieure en République Démocratique du Congo et au
Alvarez, P. 2000. A quantitative method for the study of non- sud-ouest du Rwanda, unpublished internal report of Royal
fossiliferous clastic formations: pre-Pan-African sandstones from Museum of the Central Africa, Tervuren (Belgium).
central Africa and the northern Democratic Republic of Congo Delpomdor, F. 2007c. Lithostratigraphie et sédimentologie de la chaı̂ne
[ex-Zaire]. Journal of African Earth Sciences, 31, 263– 284. Ouest Congolienne du Néoprotérozoı̈que supérieur (Formation de
THE WEST CONGO AND LINDI/UBANGI SUPERGROUPS 193

la Diamictite supérieure et Sous-groupe du Schisto-Calcaire) Bas- Master, S. & Wendorff, M. 2011. Neoproterozoic glaciogenic diamic-
Congo, République Démocratique du Congo, unpublished MSc tites of the Katanga Supergroup, Central Africa. In: Arnaud, E.,
thesis, Free University of Brussels. Halverson, G. P. & Shields-Zhou, G. A. (eds) The Geological
Delpomdor, F., Tack, L. & Préat, A. 2008. Microstructures in the Neo- Record of Neoproterozoic Glaciations. Geological Society of
proterozoic tillites around the Congo River Basin (CRB), Democratic London, Memoir, 36, 173– 184.
Republic of the Congo (DRC)– Comparison with the Karoo tillites Meert, J. G. & Van der Voo, R. 1996. Paleomagnetic and 40Ar/39Ar
from the Dekese borehole in the CRB. 22nd Colloquium African study of the Sinyai Dolerite, Kenya: implications for Gondwana
Geology, Hammamet, Tunisia, Abstracts volume. assembly. Journal of Geology, 104, 131– 142.
Delpomdor, F., Tait, J., Tack, L. & Préat, A. 2011. Neoproterozoic Meert, J. G., Van der Voo, R. & Ayub, S. 1995. Paleomagnetic investi-
lithostratigraphy in the Democratic Republic of Congo (DRC). 23rd gation of the Neoproterozoic Gagwe lavas and Mbozi complex, Tan-
Colloquiom of African Geology, Johannesburg. zania and the assembly of Gondwana. Precambrian Research, 74,
De Paepe, P., Hertogen, J. & Tack, L. 1975. Mise en évidence de laves 225– 244.
en coussin dans les faciès volcaniques basiques du massif de Kibungu Menzies, J. 2000a. Micromorphological analyses of microfabrics and
(Bas-Zaı̈re) et implications pour le magmatisme ouest-congolien. microstructures, indicative of deformation processes, in glacial sedi-
Annales de la Société Belge de Géologie, 98, 251– 270. ments. In: Maltman, A. J., Hubbard, B. & Hambrey, M. J. (eds),
De Waele, B., Liégeois, J.-P., Nemchin, A. A. & Tembo, F. 2006. Isoto- Deformation of Glacial Materials. Geological Society, London,
pic and geochemical evidence of Proterozoic episodic crustal rework- 176, 245–248.
ing within the Irumide Belt of south-central Africa, the southern Menzies, J. 2000b. Microstructures in diamictites of the lower Gondwana
metacratonic boundary of an Archaean Bangweulu Craton. Precam- Formation (Huronian), near Elliot Lake Ontario: evidence for
brian Research, 148, 225–256. deforming-bed conditions at the grounding line? Journal of Sedimen-
De Waele, B., Johnson, S. P. & Pisarevsky, S. A. 2008. Palaeoproter- tary Research, 70, 210–216.
ozoic to Neoproterozoic growth and evolution of the eastern Congo Menzies, J., van der Meer, J. J. M. & Rose, J. 2006. Till as a glacial ‘tec-
Craton: its role in the Rodinia puzzle. Precambrian Research, 160, tomict’, its internal architecture, and the development of a ‘typing’
127– 141. method for till differentiation. Geomorphology, 75, 172– 200.
Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic Mestraud, J.-L. 1952. Formation du socle en Oubangui-Chari
constraints upon the Neoproterozoic climatic paradox. American central. 19th International Geological Congress (Algiers) A.S.G.A.,
Journal of Science, 300, 347–433. 155– 162.
Evans, D. J. A. & Benn, D. I. 2004. A Practical Guide to the Study of Mestraud, J.-L. 1953. Notice explicative sur la feuille Bangassou Ouest.
Glacial Sediments. Arnold Publishers, London. Carte géologique de reconnaissance au 1/500.000. Direction Mines
Frimmel, H. 2009. Trace element distribution in Neoproterozoic et Géologie, l’Afrique Equatoriale Française, Brazzaville.
carbonates as palaeoenvironmental indicator. Chemical Geology, Moloto-A-Kenguemba, G. R., Trindade, R. I. F., Monie, P., Nedelec,
258, 338– 353. A. & Siqueira, R. 2008. A late Neoproterozoic paleomagnetic pole
Frimmel, H. E., Tack, L., Basei, M. S., Nutman, A. P. & Boven, A. for the Congo Craton: tectonic setting, paleomagnetism and geochro-
2006. Provenance and chemostratigraphy of the Neoproterozoic nology of the Nola dike swarm (Central African Republic). Precam-
West Congolian Group in the Democratic Republic of Congo. brian Research, 164, 214– 226.
Journal of African Earth Sciences, 46, 221– 239. Noce, C. M., Pedrosa-Soares, A. C., da Silva, L. C., Armstrong, R. &
Gérard, G. 1958. Carte géologique de l’Afrique Equatoriale Française Piuzana, D. 2007. Evolution of polycyclic basement complexes in
au 1/2 000 000 avec notice explicative. Brazzaville Directorate des the Araçuai Orogen, based on U–Pb SHRIMP data: implications
Mines et Géologie A.E.F., 198, 4 feuilles. for Brazil– Africa links in Paleoproterozoic time. Precambrian
Gérard, G. & Gérard, J. 1952. Stratigraphie du Précambrien Research, 159, 60– 78.
d’Oubangui-Chari occidental (A.E.F.) et essai de corrélations avec Oliveira, E. P., Toteu, S. F. et al. 2006. Geologic correlation
les territoires voisins. 19th International Geological Congress between the Neoproterozoic Sergipano belt (NE Brazil) and the
(Algiers), A.S.G.A., 145– 153. Yaounde belt (Cameroon, Africa). Journal of African Earth Sciences,
Halverson, G. P., Dudas, F. O., Maloof, A. C. & Bowring, S. A. 44, 470– 478.
2007. Evolution of the 87Sr/86Sr composition of Neoproterozoic sea- Pedrosa-Soares, A. C., Alkmin, F. F., Tack, L., Noce, C. M., Babinski,
water. Palaeogeography Palaeoclimatology Palaeoecology, 256, M., Silva, L. C. & Martins-Neto, M. A. 2008. Similarities and
103– 129. differences between the Brazilian and African counterparts of the
Hoffman, P. F. 2011. Glacigenic and associated strata of the Otavi car- Neoproterozoic Araçuai– West Congo Orogen. In: Pankhurst,
bonate platform and foreslope, northern Namibia: evidence for R. J., Trouw, R. A. J., Brito Neves, B. B. & De Wit, M. J. (eds)
extreme glacioeustatic fluctuation. In: Arnaud, E., Halverson, West Gondwana: Pre-Cenozoic Correlations Across the South Atlan-
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neo- tic Region. Geological Society, London, Special Publications, 294,
proterozoic Glaciations. Geological Society, London, Memoirs, 36, 153– 172.
195– 210. Phillips, E. 2006. Micromorphology of a debris flow deposit: evidence of
Hudeley, H. & Belmonte, Y. 1966. Carte géologique de la République basal shearing, hydrofacturing, liquefaction and rotational defor-
Gabonaise, échelle 1/1 000 000. Bureau de Recherches Géologiques mation during emplacement. Quaternary Science Reviews, 25,
et Minières (B.R.G.M.). 720– 738.
Kadima, K. E., Mwene, N. S. S. & Francis, L. 2011. A Proterozoic-rift Phillips, E., Merrit, J., Auton, C. & Golledge, N. 2007. Microstruc-
origin for the structure and the evolution of the cratonic Congo basin. tures in subglacial and proglacial sediements: understanding faults,
Earth and Planetary Science Letters, 304, 240–250. folds and fabrics, and the influence of water on the style of defor-
Kampunzu, A. B., Kapenda, D. & Manteka, B. 1991. Basic magmatism mation. Quaternary Science Reviews, 26, 1499–1528.
and geotectonic evolution of the Pan African belt in central Africa: Pin, C. & Poidevin, J. L. 1987. U– Pb zircon evidence for a Pan-African
evidence from the Katangan and West Congolian segments. Tectono- granulite facies metamorphism in the Central-African-Republic – a
physics, 190, 363–371. new interpretation of the high-grade series of the northern border of
Kroner, A. & Correia, H. 1973. Further evidence for glaciogenic origin the Congo Craton. Precambrian Research, 36, 303– 312.
of Late Precambrian mixtites in Angola. Nature –Physical Science, Pinna, P., Cocherie, A., Thieblemont, D., Feybesse, J.-L. & Lagny, P.
246, 115– 117. 1996. Evolution géodynamique du Craton Est-Africain et détermi-
Legoux, P. & Hourcq, V. 1943. Esquisse géologique de l’Afrique Occi- nisme gı̂tologique. Geodynamic evolution and metallogenic controls
dental Française, Bulletin du Service des Mines de l’Afrique Occiden- in the East-African Craton (Tanzania, Kenya, Uganda). BRGM,
tale Française. Chronique Recherche Miniere, 525, 33 – 43.
Lepersonne, J. 1951. Données nouvelles sur la stratigraphie des terrains Piper, J. D. A. 1972. Paleomagnetic study of Bukoban System, Tanzania.
anciens du Bas-Congo. Bulletin de la Société Belge de Géologie, 60, Geophysical Journal of the Royal Astronomical Society, 28,
169– 189. 111–127.
194 J. TAIT ET AL.

Poidevin, J.-L. 1976. Les formations du Précambrien supérieur de la (1000 –910 Ma) of the Zadinian and Mayumbian Groups (Bas-
région de Bangui (R.C.A.). Bulletin de la Societé Géologique de Congo), onset of Rodinia rifting at the western edge of the Congo
France, 18, 999– 1003. Craton. Precambrian Research, 110, 277–306.
Poidevin, J.-L. 1979a. Echelle stratigraphie des formations précam- Tack, L., Fernandez-Alonso, M., Kanda Nkula, V., Mpoyi, J.,
briennes de Centrafrique (R.C.A.). Résumé 10ème Colloque de Géolo- Delvaux, D., Trefois, Ph. & Baudet, D. 2006. Neoproterozoic Dia-
gie Africaine, Montpellier. mictites around the Congo River Basin: a critical reappraisal of their
Poidevin, J.-L. 1979b. Les basaltes et dolérites Précambrien supérieur de origin. 21st Colloquium African Geology (CAG21), 3– 5 July 2006,
la région de Bakouma (Empire centraficain). 7ème Réuion Annuelle Maputo, Mozambique, Abstract Book, 152– 153.
des Sciences de la Terre, Lyon, 374. Tack, L., Delvaux, D. et al. 2008. The 1000 m thick Redbeds sequence
Poidevin, J.-L. 1985. Le Protérozoı̈que supérieur de la République centra- of the Congo River Basin (CRB): a generally overlooked testimony in
fricaine. Annales de Muséum Royal d’Afrique Centrale Tervuren, Central Africa of post-Gondwana amalgamation (550 Ma) and pre-
Belgique, Série in 8ème, Sciences Géologiques, 91, 75. Karoo break-up (320 Ma). 22nd Colloquium on African Geology,
Poidevin, J.-L. 2007. Stratigraphie isotopique du strontium et datations Hammamet, Tunisia, Abstracts volume, 86 –88.
des formations carbonatées et glaciogéniques néoprotérozoı̈ques du Tack, L., Delvaux, D. et al. 2009a. The 1000 m thick Redbeds sequence
Nord et de l’Ouest du Craton du Congo. Comptes Rendus Geoscience, of the Congo River Basin (CRB): a generally overlooked testimony in
339, 259– 273. Central Africa of post-Gondwana amalgamation (550 Ma) and pre-
Poidevin, J.-L., Alabert, J. & Miauton, J. D. 1981. Géologie des séries Karoo break-up (320 Ma). Geological Society of London Fermor
du Précambrien supérieur de la région de Bakouma (République Meeting, Edinburgh.
centrafricaine). Bulletin du Bureau de Recherches Geologiques et Tack, L., Wingate, M. T. D. et al. 2009b. The Mesoproterozoic
Minieres, Section IV, 4, 313–320. ‘Kibaran Event’ in Central Africa: a 1375 Ma intracratonic emplace-
Préat, A. 2008. Etude au microscope polarisant de 165 lames minces de ment of a Large Igneous Province (LIP). Geological Society of
roches carbonatées du synclinal de la Nyanga (feuilles à 1:200 000 de London Fermor Meeting, Edinburgh.
Fougamou et de N’Dendé). Rapport Université Libre de Bruxelles, Thiéblemont, D., Castaing, C., Billa, M., Bouton, P. & Préat, A. et
Tervuren, Belgique. collaborateurs. 2009. Notice explicative de la carte géologique et des
Préat, A., Bouton, P., Kolo, K., Prian, J.-P., Simo Ndounze, S. & Ressources minérales de la République gabonaise à 1:1 000 000. Edi-
Thiéblemont, D. 2008. Sédimentologie et isotopes (carbone tions DGMG, Ministère des Mines, du Pétrole, des Hydrocarbures.
et oxygène) des carbonates Précambriens du Gabon: rapport au Libreville.
mode de fonctionnement des bassins Néo- et Paléoprotérozoı̈ques. Toteu, S. F., Fouateu, R. Y. et al. 2006. U– Pb dating of plutonic rocks
Communication Spec. Session Geologica Belgica, Namur, Belgique. involved in the nappe tectonic in southern Cameroon: consequence
Préat, A., Delpomdor, F., Kolo, K., Gillan, D. & Prian, J.-P. 2011. for the Pan-African orogenic evolution of the Central African fold
Stromatolites and cyanobacterial mats in peritidal evaporitive belt. Journal of African Earth Sciences, 44, 479– 493.
environments in the Neoproterozoic of Bas-Congo (Democratic Trompette, R. 1994. Geology of Western Gondwana (2000– 500 Ma).
Republic of Congo) and South of Gabon. In: Tewari, V. C. & Seck- Pan-African-Brasiliano aggregation of South America and Africa.
back, J. (eds) Stromatolites: Interactions of Microbes with Sediment, Balkema, Rotterdam.
Cellular Origin and Life in Extreme Habitats and Astrobiology. Trompette, R. & Boudzoumou, F. 1988. Paleogeographic significance
Springer Verlag, 43 –63. of stromatolitic buildups on Late Proterozoic Platforms– the
Préat, A., Kolo, K., Prian, J.-P. & Delpomdor, F. 2010. A peritidal eva- example of the West Congo Basin. Palaeogeography, Palaeoclima-
porite environment in the Neoproterozoic of South Gabon (Schisto- tology, Palaeoecology, 66, 101–112.
Calcaire Subgroup, Nyanga Basin). Precambrian Research, 177, Van der Meer, J. J. M. 1993. Microscopic evidence of subglacial defor-
253– 265. mation. Quaternary Science Reviews, 12, 553– 587.
Prian, J.-P. 2008. Notice géologique et ressources minérales de la carte Van der Meer, J. J. M. 1997. Particle and aggregate mobility in till:
de N’Dendé à 1/200 000. Document provisoire, carte géologique microscopic evidence of subglacial processes. Quaternary Science
du Gabon. Reviews, 16, 827–831.
Schermerhorn, L. J. G. 1974. No evidence for glacial origin of Late Van der Meer, J. J. M. 1999. Particle and aggregate mobility in till:
Precambrian tilloids in Angola. Nature, 252, 114–115. microscopic evidence of subglacial processes. Quaternary Science
Schermerhorn, L. J. G. 1981. Late Precambrian Tilloids of northwest Reviews, 22, 1659–1685.
Angola. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Van der Meer, J. J. M., Menzies, J. & Rose, J. 2003. Subglacial till:
Pre-Pleistocene Glacial Record. Cambridge University Press, the deforming glacier bed. Quaternary Science Reviews, 22,
Cambridge, 158– 161. 1659– 1685.
Schermerhorn, L. J. G. & Stanton, W. I. 1963. Tilloids in the West Vellutini, P. & Vicat, J. P. 1983. On the origin of basal conglomeratic
Congo geosyncline. Quarterly Journal of the Geological Society of formations of the West-Congolan Geosyncline (Gabon, Congo,
London, 119, 201–234. Zaire, Angola). Precambrian Research, 23, 87– 101.
Stanton, W. I., Schermerhorn, L. J. G. & Korpershoek, H. R. 1963. Verbeek, T. 1970. Géologie et lithostratigraphie du Lindien (Précambrien
The West Congo System. Boletim dos Serviços de Geologia e Minas supérieur du nord de la République Démocratique du Congo). Annale
de Angola, 8, 69 –78. de la Musée Royale de l’Afrique Centreale, série n88, 66, 309.
Straathof, G. B., Tait, J., Cibambula, E., Kanda Nkula, V. & Vicat, J.-P., Léger, J. M., Nsifa, E., Piguet, P., Nzenti, J. P., Tcha-
Zimmermann, U. 2008. Constraints on the glacial events from the meni, R. & Pouclet, A. 1996. Distinction au sein du craton congolais
Neoproterozoic West Congolian Group, 33rd International Geologi- du Sud-Ouest du Cameroun, de deux épisodes doleitiques initiant les
cal Congress, Oslo. cycles orogeniques eburneé en (Paléoprotérozoique) et panafricain
Tack, L., Wingate, M. T. D., Liégeois, J.-P., Fernandez-Alonzo, M. (Néoprotérozoique). Comptes Rendus de l’Académie des Sciences.
& Deblond, A. 2001. Early Neoproterozoic magmatism Série 2. Sciences de la terre et des planètes, 323, 575– 582.
Chapter 14

Glaciogenic and associated strata of the Otavi carbonate platform and foreslope, northern
Namibia: evidence for large base-level and glacioeustatic changes

PAUL F. HOFFMAN1,2
1
Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138, USA
2
School of Earth and Ocean Sciences, University of Victoria, Victoria, BC V8W 2Y2, Canada
(e-mail: hoffman@eps.harvard.edu)

Abstract: Two discrete, mappable, glaciogenic formations occur within the Otavi Group, a 3 + 1-km-thick carbonate-dominated
platform of late Neoproterozoic age, developed on the SW promontory of the Congo craton in northern Namibia and exposed in bordering
late Ediacaran fold belts. Each is overlain abruptly by an expanded postglacial carbonate sequence, the younger of which begins with a
globally-correlative transgressive dolopelarenite. The older Chuos glaciation (,746 Ma) occurred during a time of north-south crustal
stretching. Debris derived from upturned older rocks collected in structural depressions. The younger Ghaub glaciation (635 Ma)
occurred, after stretching ceased, on a thermally-subsiding marine platform and its distally-tapered foreslope. A continuous ice
grounding-zone wedge (GZW) occurs on the distal foreslope, while the upper foreslope and outer platform are devoid of glacial
debris and only small pockets of lodgement facies exist on the inner platform. Debris in the GZW is derived from a distinctive
falling-stand wedge that is unique to the foreslope and from immediately older strata mined preferentially from the inner platform.
The GZW rests on a smooth surface that includes a transverse steep-walled trough presumably cut by an ice-stream, within which is
a towering doubly-crested moraine composed of composite, massive, carbonate diamictite. The surface suggests that the ice-sheet
was grounded on the distal foreslope, implying a large fall in base level at a glacial maximum that predates the GZW. The glacial
record ends with Fe-stained beds, rich in ice-rafted debris, that are notably absent from the moraine, upper foreslope and platform,
which were apparently above sea-level at that time.

The existence of glaciogenic diamictite within the Neoproterozoic discontinuously within it. Condon et al. (2002) argue that this evi-
carbonate succession of the Otavi Mountains in north-central dence for ice-line instability and iceberg-rafting is incompatible
Namibia (Fig. 14.1a) was first recognized by le Roex (1941), with an ice-covered ocean at the time of Ghaub deposition.
who cited as evidence of glacial action the presence of faceted In rejecting a glacial origin altogether, Eyles & Januszczak
and striated clasts, absence of sorting or clast gradation, uniformity (2004a, b) revived the ideas of Frets (1969) and Martin et al.
of gritty fine-grained matrix, and import of granitic and arenaceous (1985). In their opinion, the Ghaub Formation consists ‘entirely
clasts into a conformable carbonate succession. Because its floor is of a wide variety of mass flow facies, . . . derived from faulted
not morphologically exposed, le Roex (1941) forthrightly admitted underlying carbonate platform strata’. This interpretation is diffi-
that the diamictite could have been deposited by grounded ice or cult to reconcile with the regionally consistent stratigraphic pos-
floating ice. Martin (1965a, b) argued that the regional extent of ition of the Ghaub Formation with respect to the overlying
this unit and the existence of similar diamictites of roughly the Keilberg Member (Hoffmann & Prave 1996; Hoffman et al.
same age throughout southern Africa and on other continents 1998a, 2007). It turns a blind eye to prior stratigraphic mapping,
favoured a glacial origin. He interpreted the Otavi Tillite (le sedimentology and chemostratigraphy showing that faulting
Roex 1941) as a glaciomarine unit on account of its stratigraphic ended well before the Ghaub Formation was deposited (Soffer
interposition between marine carbonate formations. Kröner & 1998; Hoffman 1999; Halverson et al. 2002). During Ghaub
Rankama (1972) and Hedberg (1979) also favoured a glaciomarine time, no fault-related structural rotations occurred on the platform
interpretation. Martin (1965b) noted the close association of or on its margin (Hoffman & Halverson 2008).
banded Fe-formation with diamictite in the Otavi Group and
attributed ‘this peculiar combination of sediments’ to ‘oxygen
deficiency in stagnating bottom waters caused by an ice cover’. Structural framework
A glacial origin for the carbonate-clast diamictite (now Ghaub
Formation) on the Fransfontein foreslope (Fig. 14.1b) was rejected The Otavi Group is exposed in an arcuate fold-and-thrust belt
by Frets (1969), who favoured submarine landsliding in response around the junction of the Late Ediacaran – Early Cambrian
to a fall in base level. This view was taken up by Martin et al. Kaoko and Damara orogens (Fig. 14.1a). At the apex of the fold
(1985), who at age 75 (and 20 years removed from Africa) recanted belt is the Kamanjab inlier, a large basement-cored structural cul-
his earlier conclusions and hypothesized that a major eustatic mination. For most of the twentieth century, studies focused on the
event, possibly caused by glaciation in other parts of the globe, Otavi Mountains near the eastern terminus of the fold belt (Miller
had triggered submarine landslides simultaneously throughout 2008; Bechstädt et al. 2009), where major carbonate-hosted Cu –
Namibia. The discovery of large submarine landslides of Quatern- Pb– Zn –V mineralization occurs. Since Namibian independence
ary age on the modern African margin (Dingle 1977) may have in 1990, scientifically motivated research has favoured the better
prompted this reinterpretation. Most present-day workers, exposed sections around the western Kamanjab inlier. Weakly
however, favour the glaciomarine interpretation (Hoffmann & folded Otavi Group strata extend across the Owambo Basin in
Prave 1996; Hoffman et al. 1998a, 2007; Evans 2000; Condon the subsurface, making the total area of the ancient carbonate plat-
et al. 2002; Hoffman 2005, 2011; Domack & Hoffman 2011). form in Namibia .110 000 km2 (Miller 1997).
Although structureless carbonate-clast diamictite (with rare The western margin of the carbonate platform is complicated
basement-derived clasts) is the predominant lithology, thin inter- by a major, sinistral-oblique thrust zone (Sesfontein thrust) bound-
vals of well-stratified detrital carbonate with variable amounts ing the Kaoko Belt (Goscombe et al. 2003). The southern margin
of rafted debris occur at the top and bottom of the formation and is preserved in the east around the Otavi syncline (Smit 1962;

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London, Memoirs,
36, 195– 209. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.14
196 P. F. HOFFMAN

Fig. 14.1. (a) Geological map of the Otavi


Group fold belt and its relation to the
Owambo Basin (Congo craton) and to the
Kaoko and Damara orogenic belts
(Hoffman 2005). Dashed white lines
indicate the southern and western edges of
the Otavi Platform. The black rectangle
shows the area of Figure 14.1b. (b)
Geological map of the Fransfontein Ridge
and its relation to the Otavi Platform of the
Upper Huab outlier and Danubé (after
Hoffman et al. 2007). Also shown is the
division of the foreslope into proximal
and distal zones, the location of the
Duurwater trough and moraine, and
district road C35.

Hedberg 1979) and to the west along the autochthonous southern (Hedberg 1979; SACS 1980), assumed correlative with the
flank of the Kamanjab inlier (Fig. 14.1b). Distal stratigraphic glaciogenic Chuos Formation of the central Damara Belt (Gevers
equivalents of the Otavi Group occur on tectonically stretched and 1931). Hoffmann & Prave (1996) first established the existence
later foreshortened Congo crust within the central and eastern zones of two discrete and extensive glaciogenic formations in the
of the Kaoko Belt (Swart 1992; Stanistreet & Charlesworth 1999; Otavi Group, the older of which (Varianto Formation of SACS
Goscombe & Grey 2007; Paciullo et al. 2007) and in the central 1980) is equivalent to the type Chuos Formation and the younger
and northern zones of the Damara Belt (Hoffmann et al. 2004; of which (the Otavi Tillite of le Roex 1941) they renamed the
Johnson et al. 2006; Clifford 2008; Hoffman & Halverson 2008). Ghaub Formation. This critical distinction has been amply con-
firmed by subsequent studies (Hoffman et al. 1998a, b; Kennedy
et al. 1998; Halverson et al. 2005; Hoffman & Halverson 2008).
Stratigraphy Hoffmann & Prave (1996) divided the Otavi Group into three
subgroups at surfaces corresponding to the floors of the two glacio-
The Otavi Group is a 3 + 1-km-thick, carbonate-dominated suc- genic formations. Thus the Ombombo (om-bomb-boh) Subgroup
cession of Cryogenian and early Ediacaran age, deposited ends below the Chuos Formation, the Abenab (ah-ben-ab) Sub-
between c. 770 and 580 Ma (Halverson et al. 2005). It is conform- group below the Ghaub Formation, and the Tsumeb (tsoo-meb)
ably underlain by fluviatile feldspathic quartzites and conglomer- Subgroup below the Mulden Group. South of the Huab palaeoridge
ates of the Nosib Group (Fig. 14.2), and overlain by marine and (Fig. 14.2), the Ugab (oo-khab) Subgroup occupies a stratigraphic
non-marine, synorogenic clastics of the Mulden Group. The position analogous to the Ombombo Subgroup, but U –Pb zircon
Otavi and Mulden groups are separated by a paraconformity ages show it to be significantly younger (,746 + 2 Ma) than the
with .1.0 km of karstic palaeotopographic relief (Frets 1969; Devede (de-ved-ee) Formation of the Ombombo Subgroup,
Hoffman & Hartz 1999; Hoffman & Halverson 2008). which is .760 + 1 Ma (Halverson et al. 2005).

Subgroups of the Otavi Group Crustal stretching and thermal subsidence in relation
to glaciation
The Otavi Group was originally divided into two subgroups,
Abenab and Tsumeb, at a surface corresponding to the floor of Subsidence accommodating the Otavi Group was driven by north –
what was then thought to be a single glaciogenic formation south crustal stretching, which lasted at least episodically from
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 197

δ13C (per mil)


South 50 Otavi Platform 200 –10 –5 0 5 10
Tsumeb
-0.5 "
Subgroup
Huttenberg
-1.0
Elandshoek
Fore-
slope Keilberg Maieberg
-2.0 Ombaatjie Abenab
Gruis Subgroup
km .
Huab . . Rasthof
Karibib Makalani
. Ridge Ridge . . Ugab Sbgp
Ombombo

Ghaub
. . . Subgroup Fig. 14.2. Stratigraphic restoration of the
Devede Otavi Group in a north–south cross-section
746 ± 2 Ma 759 ± 1 Ma
. . . based on measured sections and mapped
Chuos . . .
Naauw . –10 –5 0 5 10 stratigraphic cutoffs around the western
Ugab
-poort
Palaeoproterozoic metamorphic- . Nosib
. . δ 13C (per mil V-PDB) Kamanjab inlier (after Hoffman et al. 2007).
plutonic complex (Congo craton) Congo
Tanz. Note the existing U –Pb zircon age
Angola
constraints from volcanic lavas and tuffs
Zam.
shallow-water carbonate cannibalized clastics (syn-rift) Mal. (Hoffman et al. 1996; Halverson et al.
Nam. Bots.
Zim.
Moz. 2005), the composite d13C profile
deep-water carbonate bimodal alkaline volcanics Swaz. (Halverson et al. 2005) and the location map

. .
South
Africa showing the Kalahari and Congo cratons
glaciogenic diamictite feldspathic sandstone (palaeoflow)
(shaded).

c. 770 Ma (Halverson et al. 2005) until a rift-sag transition at the tapers from c. 1400 m on the platform to 320 –540 m for isotopi-
base of the Ombaatjie (om-bye-gee) Formation (Fig. 14.2) of cally correlative strata on the distal foreslope (Halverson et al.
the Abenab Subgroup (Halverson et al. 2002). Accordingly, the 2005). The Abenab Subgroup tapers from 500–800 m on the plat-
Chuos glaciation occurred during a prolonged period of active form to 70– 200 m on the distal foreslope (Hoffman 2005). Only
faulting, which is reflected by the diversity of its debris and by the Keilberg cap dolostone exhibits shallow-water bedforms on
a low-angle (1.58) structural unconformity at its base (Fig. 14.2) the distal foreslope, reflecting large-amplitude glacioeustatic fluc-
that truncates .2 km of Ombombo and Nosib, or Ugab and tuation (Hoffman et al. 2007).
Naauwpoort (nauf-port) strata (Hoffman et al. 1998a). No causal The foreslope wedge is best exposed on the 60-km-long Frans-
connection between glaciation and crustal stretching (Eyles & fontein Ridge (Fig. 14.1b), a present-day topographic feature
Januszczak 2004a, b) is indicated: stretching began millions of underlain by a simple, south- to SE-dipping, structural homocline
years before the glaciation and continued for millions of years of Otavi Group carbonates. The strata dip 40–608 on average, with
afterwards. near-vertical cleavage in limestones and fine clastics, and local
In contrast to the Chuos, the Ghaub glaciation occurred at a time north-vergent thrust imbrication of major rheological interfaces
of broad regional subsidence, which is reflected by platform-wide (e.g. dolomite against fine clastics at the Otavi-Mulden group
structural conformity between depositional cycles of the upper contact, and dolostone against marly limestone within the Maie-
Ombaatjie Formation and the Keilberg cap dolostone (Halverson berg Formation). The foreslope is also preserved in structural out-
et al. 2002; Hoffman et al. 2007). The Ombaatjie Formation was liers 60 and 80 km west of Fransfontein (on the farms Bethanis and
regionally flat-lying at the time of the Ghaub glaciation. Toekoms, and Vrede and Opdraend, respectively; see Hoffman &
The Tsumeb Subgroup is a structurally conformable succession Halverson 2008). It also occurs for 250 km east of Fransfontein
on the regional scale and thermal subsidence is the inferred accom- along the southern flank of the Kamanjab inlier and reportedly
modation mechanism for the Ombaatjie through Hüttenberg for- on the southern limb of the Otavi Syncline (Smit 1962; Hedberg
mations (Halverson et al. 2002). 1979). Thus, the foreslope wedge is discontinuously exposed for
330 km along strike.
The foreslope divides naturally on the basis of lithofacies (see
Palaeogeography of the platform and the southern foreslope above) and stratigraphic architecture (Fig. 14.3b) into proximal
and distal zones, projected to lie 0–5 and .5 km seaward of the
The unfaulted southern margin of the platform (Figs 14.1a & 14.2) slope break, respectively (Fig. 14.1b). The western two-thirds of
is well defined lithologically in units above and below the Ghaub the Ridge strike east –west, subparallel to the inferred palaeoslope
Formation (Tsumeb Subgroup and Ombaatjie Formation, respect- contours. It exposes a longitudinal section of the distal slope.
ively). To the north are stacked peritidal parasequences, commonly The eastern third of the ridge strikes NE and exposes a section
bounded by subaerial exposure surfaces marked by tepee struc- that angles obliquely up the proximal slope to the edge of the plat-
tures and tepee breccias (Kendall & Warren 1987). To the south form (or slope break) near its northernmost point on the farm
is a distally tapered foreslope wedge of progressively deeper-water Danubé 59.
carbonate facies (Hoffman & Halverson 2008). More proximal Comparison with recent fine-grained carbonate slopes like the
parts of the foreslope (i.e. 0– 5 km seaward of the slope break) western (leeward) margin of the Great Bahama Bank (Eberli &
are dominated by ribbony lutites and arenites lacking subaerial Ginsburg 1987; Adams & Schlager 2000) suggests that the distal
exposure surfaces. More distally (i.e. .5 km seaward of the foreslope (where the glaciogenic Ghaub Formation occurs) lay at
slope break), parallel-sided rhythmites, rhythmite slump breccias palaeodepths .0.5 km below the rim of the platform (Hoffman
and carbonate turbidites prevail. Overall, the Tsumeb Subgroup 2005, 2011).
198 P. F. HOFFMAN

Glaciogenic and associated strata 1986), but the paucity of bed-scale stratification suggests that
debris flows were of secondary importance in the Chuos Formation
Chuos Formation of the Otavi Group.
The associated strata, in decreasing order of abundance, are con-
The Chuos Formation is highly variable in thickness up to 1.0 km glomerates and pebbly sandstones, laminated siltstones with rafted
and the predominant lithology, especially in the thicker sections, debris, and banded Fe-formation. The coarse clastics are
is structureless to weakly stratified, matrix-supported diamictite. co-derived with the diamictites but are grain-supported,
The diamictite forms composite bodies derived in varying pro- clast-imbricated and crudely stratified or cross-stratified. They
portions from basement rocks (metavolcanic, metaplutonic and were clearly deposited by running water, but whether the flow
metasedimentary rocks, notably orthoquartzite), sub-feldspathic was proglacial or subglacial is not known. The conglomerates
arenites of the Nosib Group, and dolomites, cherts and canniba- and pebbly sandstones generally occur at the top and/or bottom
lized clastics of the Ombombo and Ugab subgroups. Clasts of of the Chuos Formation, and are richer in dolomite clasts and/or
strongly porphyritic, mafic and felsic lava, derived from the felspathic basement debris compared with similar lithofacies in
Naauwpoort Formation (747 + 2 Ma), are locally abundant. The the underlying, non-glaciogenic, Okakuyu (oh-ka-koo-you) For-
matrices of the diamictites are foliated to structureless, mud-rich mation (Ombombo Subgroup). Relatively thin units of finely lami-
wackestones. Some of the foliated fabrics are pre-tectonic. Striated nated siltstone or silty mudstone with rafted lonestones occur
clasts are probably of glacial origin. Fe- and/or Mn-rich cements locally at the base or within the Chuos Formation. Discontinuous
coat the detrital grains, giving the matrix a strong colour – brick beds and lenses of hematitic Fe-formation are intimately interca-
red, olive green or black. Fabric analysis to determine emplace- lated with structureless or stratified diamictites in different areas
ment mechanisms are compromised by tectonic strains, but the (Martin 1965a). Discharges of oxygenated meltwater at the
diamictites appear to represent both ice-contact (foliated) and grounding lines of tidewater glaciers could have triggered the pre-
proximal rain-out deposits. Diamictites in the central zone of the cipitation of Fe-formation precursor in the presence of Fe(II)-
Damara orogen were emplaced as debris flows (Henry et al. laden seawater.

(a) SOUTH Distance inward from the glacial-age slope break (km) NORTH
–20 –10 0 20 40 60 80 100
Depth wrt glacial-age platform rim (m)

400 OUTER PLATFORM subaerial surface INNER PLATFORM

400 m
200 sea-floor
cements iso
che
m
FR AN SF ON TE IN

aeolianite
FO RE SL OP E

0
Trezona anomaly δ13C = 0 o/oo datum
–200
GREAT OTAVI CARBONATE BANK Fig. 14.3. (a) Stratigraphic –
palaeobathymetric cross-section of the
–400 POST- Maieberg Fm: shoalwater dolostone grainstone Otavi carbonate platform and foreslope,
showing showing the Ghaub diamictite
GLACIAL Maieberg Fm: deepwater limestone rhythmite
Ghaub wedge and its relations to the Frannis-aus
–600 prism SYN-DEGLACIAL Keilberg Mb: transgressive 'cap' dolostone falling-stand wedge and the Maieberg
SYN-GLACIAL Ghaub Fm: diamictite prism (ice grounding zone) cap-carbonate sequence. The 0‰
cross-over of the Trezona d13C anomaly
–800 ADVANCE GLACIAL Franni-aus Mb: oolite debris flows (lowstand wedge) serves as the datum for the platform.
Ombaatjie Fm: shelf-slope carbonate (post-Trezona 0 o/oo) Foreslope palaeodepths are estimated
PRE-GLACIAL assuming a profile (below – 200 m) like the
Ombaatjie Fm: shelf-slope carbonate (pre-Trezona 0 o/oo)
(b) western margin of the present Great
Bahama Bank (see text). Ghaub diamictite
Palaeodepth wrt permanent post-glacial flood

20 km inboard of the slope break is the


–0.2 Keilberg cap dolostone
0.0 ‘Otavi Tillite’ (le Roex 1941; Bechstädt
E

et al. 2009). Note parallelism between the


Palaeodepth wrt top of platform (km)

–0.3
OP

–0.1 Ghaub diamictite prism Trezona datum and the top of the Maieberg
SL

sequence, disproving structural rotations


AL

–0.4
Franni-aus Member
M

–0.2 due to faulting associated with the Ghaub


I

low-stand wedge
OX

–0.5 diamictites. (b) Proximal and distal


PR

–0.3 differentiation of the Fransfontein foreslope


–0.6 based on stratigraphic relations between the
–0.4 DISTAL SLOPE Franni-aus falling-stand wedge, the Ghaub
–0.7 diamictite wedge and the Keilberg cap
–0.5 Ombaatjie Fm dolostone. The palaeodepth scale with
–0.8 respect to permanent (i.e. after isostatic
–0.6
Narachaams Member adjustment) post-glacial flood assumes a
–0.9 siltstone of cycle b8 water depth on the outer platform of 140 m
–0.7 for the Maieberg Formation limestone
–1.0
rhythmite, which was deposited below
–0.8 10 9 8 7 6 5 4 3 2 1 0 –1
storm wave and accommodated c. 300 m of
South Horizontal distance from slope break (km) North additional cap-carbonate strata.
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 199

Ghaub Formation incursion in the upper Ombaatjie Formation. Plume fallout forms
laterally continuous drapes of laminated mud and fine silt with
The palaeogeographic distribution of the Ghaub Formation is criti- both normal and reverse grain-size gradation. Plume fallout
cal to its interpretation. As mapped, roughly 99% of its mass is hosts rafted debris of variable size and concentration. Turbidites
contained within a linear wedge that tracks the distal foreslope, and debris flows consist of poorly sorted, sandy to pebbly material
.5 km downslope from the outer edge of the platform in discrete tabular or channelized sedimentation units that are com-
(Fig. 14.1b). It pinches out slightly upslope from the zero-isopach monly grain-size graded towards the top and less commonly
of the Franni-aus Member (Fig. 14.3), a glacioeustatic falling- reverse-graded near the base. Contourites are starved (isolated)
stand wedge composed of carbonate turbidites and debris-flows ripples with foreset laminae consistently indicating current flow
of non-glacial origin (see below). Abundant Franni-aus-derived from ENE to WSW, subparallel to depositional strike. Rafted
debris within the Ghaub Formation indicates that the falling-stand debris are lonestones not carried by debris flows or bedload sedi-
wedge originally extended farther upslope, where it was sub- mentation units. They are most abundant in slowly deposited sedi-
sequently eroded and redeposited downslope during the glaciation. ments (plume fallout) and least abundant in rapidly deposited
In the Summas Mountains, 30 km south of the platform, the Ghaub material (turbidites and debris flows). Subjacent laminae are
Formation is c. 4.0 m thick and its glacial affinity is scarcely recog- pinched or punctured, adjacent ones are laterally ejected, and
nizable (Clifford 2008; Hoffman & Halverson 2008). The geo- superjacent ones onlap or drape the rafted debris.
metry of the Ghaub Formation is a wedge, continuous along The third facies association consists of current-rippled or cross-
strike, pinching-out upslope and severely tapered downslope. bedded, sand- or granule-sized, detrital carbonate with sparse lone-
The Ghaub Formation is totally absent from the proximal fore- stones. Distal facies include fine sand with trains of climbing
slope and the raised outer zone of the platform (Fig. 14.3). There, ripples, implying combined bedload and suspension fallout. Palaeo-
the Keilberg Member (cap dolostone) rests disconformably on the flow directions are southwestward with large apparent dispersion.
upper Ombaatjie Formation (Figs 14.2 & 14.3). Pockets of carbon- Proximal facies involve gravelly sand organized in multi-metric
ate diamictite reappear beneath the Keilberg Member on the inner clinoforms that coarsen up-dip. Grain-supported pebble conglomer-
zone of the platform, which was preferentially eroded by up to ate is the most proximal (and least abundant) component of this
80 m relative to the outer zone during the glaciation (Hoffman facies association. The sandy facies occurs in units up to 5.5 m
et al. 2007). The type section of the Otavi Tillite (le Roex 1941; thick that can rarely be traced laterally for more than a few 100 m.
Bechstädt et al. 2009) is an exceptional 60-m-thick body of They amount to less than 2% of the formation as a whole.
diamictite, situated c. 20 km north of the edge of the platform. West of the Duurwater moraine (Fig. 14.4), the facies asso-
On the whole, however, the platform and proximal foreslope ciations are organized as follows. Each vertical section contains
were areas of erosion during the Ghaub glaciation, and the main 2 to 7 (average 4.6) diamictite units, usually separated by thinner
zone of sedimentation was on the distal foreslope. intervals of well-stratified and/or sandy detritus, but occasionally
In order to document the internal facies architecture and lateral amalgamated diamictite-on-diamictite with erosive contacts. The
variability of the Ghaub Formation, 72 stratigraphic sections were stratified intervals commonly coarsen upwards, and their con-
measured on the Fransfontein Ridge as a whole, including 40 sec- tacts with overlying diamictites are gradational or erosively
tions in a 13-km-long stretch east of the town of Fransfontein truncated. Compositionally, they normally herald the overlying
(Fig. 14.4). The sections examined by Condon et al. (2002) and diamictite. These characteristics are all well illustrated by the
Eyles & Januszczak (2007) are among those measured, as are stratified interval at the base of the formation. The tops of
the sections recently logged in greater detail (Domack & diamictites are commonly abrupt, in some cases associated with
Hoffman 2011). coarse-grained debris flows and in others by a direct transition
into fine-grained, relatively distal sandy facies. In the 5-km
Facies associations and stratal architecture. The Ghaub Formation is stretch east of Fransfontein town, correlation of diamictite bodies
composed of three facies associations. The first consists of weakly from section to section suggests an overall imbricate structure,
stratified to structureless diamictites, individually ranging from younging westward in the line of section (Hoffman 2005). The
1 to 75 m in thickness. Pebble- to boulder-sized, rounded to 2D nature of the homoclinal exposure makes it impossible to deter-
angular clasts of limestone and/or dolostone (rarely basement- mine directly if the imbrication represents lateral accretion, sub-
derived granitoids) ‘float’ in a structureless to faintly laminated parallel to the inferred slope contours, or basinward accretion in
matrix of unsorted detrital carbonate mud, silt, sand and granules. a highly oblique section.
Because of middle greenschist-grade metamorphism, the clasts are The herein-defined Bethanis Member (Fig. 14.4) forms a con-
thoroughly welded to their carbonate matrix, making it impossible tinuous blanket, 5–15 m thick, of well-stratified detrital carbonate
to observe clast surface morphology or micromorphology (e.g. choked with rafted debris of all sizes. It transgresses the imbricate
striae). Isotopic data indicate that the clasts are derived from the structure of the rest of the formation. In contrast to the older well-
upper part of the Ombaatjie Formation (Fig. 14.2) and from the stratified intervals, which coarsen upwards and are followed by
Frannis-aus Member (see below), which contributed the very diamictite, the Bethanis Member becomes finer grained and
coarse-grained oolite clasts and grains conspicuous in many of more distal upwards in all its major constituents (i.e. debrites, tur-
the diamictites. Some diamictites gradationally overlie well- bidites and plume fallout) except for rafted debris. The latter
stratified units, whereas others have erosive bases and internal includes boulders up to 5 m in diameter composed of laterally
reactivation surfaces. Collectively, weakly stratified and structure- linked, hemispheroidal, stromatolitic dolostone. This clast type is
less diamictites make up 68% of the Ghaub Formation west of the not observed in the underlying diamictites or rafted debris.
Duurwater moraine (see below) and 82% overall. Boulder-sized rafted debris continues to the top of the Bethanis
The second facies association, amounting to 17% of the for- Member, where the detrital components give way gradationally,
mation as a whole, consists of well-stratified detritus and com- over a stratigraphic interval of only 2– 5 cm, to peloidal-
prises suspension-plume fallout, turbidites, debrites, contourites micropeloidal dolostone of the Keilberg Member (Hoffmann &
and rafted debris, in varying proportions. All the detritus is carbon- Prave 1996; Hoffman et al. 2007). The isotopic systematics of
ate, except to the west of Fransfontein town (particularly on Betha- the Keilberg Member (Hoffman et al. 2007) preclude a detrital
nis farm), where the basal stratified interval is composed of origin for the dolomite as postulated by Eyles & Januszczak
siliciclastic siltstone with rare outsize quartz granules. This terrige- (2007). The Bethanis Member is dark brown in colour, apparently
nous sediment is likely reworked from the Narachaams (nahr- due to Fe and Mn oxides. Starved ripples generated by westerly
a-khams) Member (Figs 14.3b & 14.5), a fine-grained siliciclastic directed contour currents occur in some sections. The Bethanis
200 P. F. HOFFMAN

Distance in kilometres east of district road C35


7.0 8.0 9.0 10.0 11.0 12.0 13.0
D5 D6 D7 D8 D9 D10

P7010
(a)
20
Keilberg
Bethanis
Bethanis
0

P6510
Ghaub

P7021
P7013

P7024
-20

P7022
P4027

P7026
P7025

P6514
P6508
P7009

P7011
P7012

P7027
P7023
-40 Franni-aus

P7028
P7014
Narachaams Member
Member
P8011

P6506

Distance in kilometres east of district road C35 Datum base Bethanis Mb


0.0 1.0 2.0 3.0 4.0 5.0 6.0
(b) D0 2 4
D1 D D3 D
20 Keilberg Member

Bethanis Member
0

-20

-40
Ghaub Formation
-60
P7004
P2093

P5003
P6537
P6535

-80

P7003
P7002
P7007

P8010

-100
P8007

P6536
P7006
E0801

Franni-aus Member
P7018

P8008
P7008

-120
P6538

Keilberg Member cap dolostone Ghaub Formation


P8012

marly dolostone turbidity flows weakly- and non-stratified limestone-dolomite diamictite


E0803

peloidal dolostone, giant wave ripples weakly- and non-stratified limestone diamictite

-140 dolostone, isopachous sheet-crack cements weakly- and non-stratified dolomite-chert diamictite
P7001

current-rippled carbonate sandstone, minor ice-rafted debris


D2 - drainage number east of C35 stratified diamictite (debris flows, contourites, ice-rafted debris)

Fig. 14.4. Facies associations of the Ghaub Formation on the distal foreslope of the Fransfontein Ridge: (a) between the western edge of Duurwater trough (at 7.5 km
east) and the western edge of Duurwater moraine (at 13.0 km east) and (b) west of the Duurwater trough.

Member is lithologically distinct from earlier stratified units of the 0.1 km deep, as defined with respect to the underlying stratigraphy
Ghaub Formation and represents the terminal collapse of the ice (Fig. 14.5). On the floor of the trough, carbonates of the falling-
sheet on the Otavi platform. stand wedge (Franni-aus Member) and terrigenous siltstone of
the underlying Narachaams Member are missing, and the Ghaub
Duurwater trough and moraine. In the central part of the Fransfon- Formation rests directly on older units of the Abenab Subgroup.
tein Ridge (Fig. 14.1b), the erosion surface on which the Ghaub The trough is roughly coaxial with a major complex of deepwater
Formation rests cuts out a steep-walled trough, 18 km wide and carbonate (oolitic) sand bodies (submarine channel and levee
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 201

0 WEST Duurwater Trough EAST


–100 DT Platform
N Otavi crystal fans
–200 above cap
dolostone
Palaeodepth in metres below top of platform

–300
DM Belt
Damara
–400
Karibib Formation
–500 (includes Keilberg cap dolostone) Fig. 14.5. Stratigraphic relations in the
Abenab Subgroup and Ghaub Formation on
Ghaub Formation moraine
–600 the Fransfontein Ridge, reconstructed to
falling-stand o o o o o ridge show the inferred palaeobathymetry of the
o o o wedge
o o o
–700 o Duurwater trough and moraine (see
Narachaams Member Fig. 14.1b for location) at the end of the
–800 Ghaub glaciation (simplified after Hoffman
2005). Inset shows a postulated
grainstone
–900 palaeogeography (see text) in which the
prism
crystalline Duurwater trough (DT) marks a tributary
–1000 crystalline basement Chuos basement ice streamlet to a trunk ice stream flowing
diamictite westward along the northern zone of the
Abenab Subgroup Damara Belt. Duurwater moraine (DM) is
–1100 slope carbonates interpreted as one of a set of terminal
moraines (black dots) formed by the
–1200
–5 0 5 10 15 20 km Abenab retreating trunk ice stream,
Subgroup contemporaneous with deposition of the
–1300 Distance east of district road C35 Bethanis Member.

deposits) in the lower Abenab Subgroup. It straddles a palaeoval- Quaternary ice streams are mostly lateral moraines (e.g. Isfjorden
ley on the crystalline erosion surface lined by basement-derived Svalbard, Ottesen et al. 2005) or terminal moraines oriented at
boulder diamictite and minor iron-formation of the Chuos For- high angles to the direction of ice-stream flow (Ottesen et al.
mation (Fig. 14.5). Hoffman (2005) attributes the Ghaub-age 2005; McMullen et al. 2007).
trough to a transverse, southward-flowing ice stream. It occupied
a preexisting drainage system developed long before the Ghaub
glaciation. Compared with Quaternary ice-stream troughs, the Duur- Boundary relations with overlying and underlying
water trough is quite modest in size, similar to the Isfjorden trough non-glacial units
on the European continental shelf west of Svalbard (Ottesen
et al. 2005). Like the Ghaub Formation itself, what makes the Chuos Formation
Duurwater trough unusual is its location on the distal slope, well
below the depth of any known Quaternary ice-stream trough. The Chuos Formation truncates the basement complex and all
The Ghaub wedge thins on both sides of the Duurwater trough to older Neoproterozoic strata at a low-angle (,1.58) unconformity
about half its average thickness (Fig. 14.5), without change in (Fig. 14.2). Carbonates directly beneath the unconformity are
facies. In the middle of the trough, it swells to a pair of thickness typically shattered and silicified. In contrast, the top of the Chuos
maxima of 288 m (east) and 602 m (west). The maxima are 3 km Formation is a sharp but conformable contact lacking evidence
apart and separated by a 182-m-thick saddle. As the wedge thick- for substantial hiatus and only locally of significant reworking
ens, it loses its well-stratified facies and at both crests consists (Hoffman & Halverson 2008). The top of the Chuos Formation
exclusively of amalgamated, weakly stratified to structureless, is strongly enriched in hematite, but concentrations along tectonic
boulder diamictites. cleavage planes suggests secondary Fe migration. The basal
For want of a datum in the upper Abenab Subgroup beneath Rasthof facies is generally a dark grey, parallel-laminated dolo-
the trough (see above), the true palaeo-elevation of the morainal micrite with variable proportions of calcitic, allodapic, centimetric
crest above the rims of the trough is uncertain. That it was a density-flows (Yoshioka et al. 2003; Pruss et al. 2010). It appears
major bathymetric feature is reflected by the anomalous nature to be the deepest-water facies within the Rasthof Formation,
of the directly overlying units. The Keilberg cap dolostone is meaning that the Rasthof lacks a transgressive facies tract.
less than half its thickness off the moraine, and the normal fine- Hoffman & Schrag (2002) suggested that seawater failed to reach
grained micropeloids are winnowed away, leaving lags of unu- critical oversaturation with respect to CaCO3 or [CaMg]CO3
sually coarse-grained peloids (Hoffman et al. 2007, fig. 5b). The until after the glacioeustatic rise accompanying Chuos deglacia-
directly overlying Maieberg limestone hosts spectacular crystal tion. This failure could be related to insufficient availability of
fans (Hoffman 2005, fig. 11a), pseudomorphous after aragonitic exposed carbonate globally, as only carbonate weathering could
sea-floor cements. These are elsewhere absent from the distal fore- have supplied sufficient alkalinity on the timescale of glacio-
slope, but are modestly well-developed on the proximal foreslope eustatic recovery (Higgins & Schrag 2003).
(e.g. Hoffman et al. 2007, section S5).
The orthogonal aspect ratio of 0.08 for the morainal buildup
(7.5 km wide at the base) in the line of exposure implies a trans- Ghaub Formation
verse orientation assuming a ridge-like form. It is termed a
medial moraine (Hoffman 2005) with respect to its location Relation to the Franni-aus Member. The Ghaub Formation paracon-
within the Duurwater trough, and its double crest could reflect formably overlies the Franni-aus Member (Fig. 14.3) on the distal
the fact that true medial moraines consist of paired lateral moraines foreslope (Hoffman 2005). The contact is a smooth, planar, erosion
back-to-back. On the other hand, morainal ridges associated with surface despite the coarse fragmental nature of the underlying
202 P. F. HOFFMAN

debris flows. The Franni-aus Member (Hoffman & Halverson Carbon isotopes
2008) is a coarsening-upward stack of limestone and dolomite
turbidites and debris flows, characterized in the upper part by Pre-glacial negative anomalies. A composite d13Ccarb curve for the
variably disaggregated slabs and grains of very coarse-grained Otavi platform is given in Figure 14.2. The major features of the
(,3 mm) oolite, typically heavily silicified. The contact with the curve have been replicated in numerous sections and compare
Ghaub Formation is marked by an abrupt decrease in grain size, favourably with correlative sections outside Namibia (Halverson
a change in colour from pale grey to medium brown, and the et al. 2005; Halverson 2006; Nogueira et al. 2007). Both glacia-
gradual appearance of increasingly numerous and sizeable drop- tions were preceded by precipitous declines in d13C, following
stones, which grade upwards into the first, non-stratified, rain-out persistent (.500 m) intervals of strong 13C enrichment
diamictite. The ooids are inferred to have originally formed in (d13C ¼ 4 –8‰ PDB). The decline before the Ghaub glaciation
highly agitated waters at the surf zone and were redeposited grav- is deeper, from þ7.5‰ to –5.0‰ PDB; it represents the
itationally before complete lithification. As there is no known ‘Trezona anomaly’ (Halverson et al. 2005; Halverson 2006)
source for these ooids in the Ombaatjie or any other formation observed below the Elatina, Stelfox (Ice Brook) and Smalfjord gla-
on the platform, it is hypothesized that they originally formed ciations in Australia, Canada and Norway, respectively. The older
on the proximal foreslope during glacioeustatic regression decline is observed at the top of the Ugab Subgroup, south of the
(Hoffman 2005; Hoffman et al. 2007). Continued base-level fall Huab palaeoridge, and its nadir of only 0‰ is probably truncated
hastened their redeposition more distally as debris flows. The by sub-Chuos erosion. It possibly represents the here-named ‘Islay
only strata of equivalent age on the platform are ,30 m of anomaly’, observed to precede the Port Askaig, Petrovbreen and
aeolian(?) dolarenite sandwiched between the youngest marine Julius River glaciations of Scotland, East Svalbard and Tasmania,
cycle of the Ombaatjie Formation and the Keilberg Member respectively (Calver 1998; Brasier & Shields 2000; Halverson
close to the outer edge of the platform in the Upper Huab area et al. 2005; Halverson 2006; McCay et al. 2006; Prave et al. 2009).
(Fig. 14.1b) (Soffer 1998; Halverson et al. 2002). The Franni-aus
Member is cut out by the Duurwater trough, where the Ghaub For- Post-glacial negative anomalies. Negative d13C anomalies are
mation rests directly on siliciclastic siltstone of the Narachaams associated with the post-glacial Rasthof and Maieberg formations
Member or underlying slope carbonates of the lower Abenab (Fig. 14.2), and differences between them are common to the pairs
Subgroup (Fig. 14.5). Reworked siltstone at the base of the of Cryogenian glaciations in Australia and Canada (Kennedy et al.
Ghaub Formation outside the trough is most likely derived from 1998; Hoffman & Schrag 2002). The anomaly in the Rasthof
the Narachaams Member, implying that the Duurwater trough or Formation has been documented in fine detail by Yoshioka et al.
others like it had been eroded before the Ghaub Formation was (2003) and their findings are valid over most of the Otavi platform
deposited (Hoffman 2005). If the Duurwater trough was cut by (Pruss et al. 2010). The negative anomaly is limited to the lower
an ice stream, glacial maximum conditions are not represented in (abiotic rhythmite) member, which is 2–70 m thick regionally
the Ghaub Formation. and 14 m in its detailed section. Its top and bottom are dolomitic
but the middle 10 m is calcitic. The basal dolomite rhythmite
Relation to the Ombaatjie Formation on the platform. Lenticles of rises asymptotically in d13C from –4.0‰ towards –2.0‰, the
Ghaub Formation overlie a continuous paraconformity on the plat- middle calcite rhythmite rises gradually by an additional 1‰,
form having ,80 m of stratigraphic relief with respect to the and the upper dolomite rhythmite rises from –1.0‰ asymptoti-
Ombaatjie Formation (Halverson et al. 2002; Hoffman et al. cally to þ2‰. The middle (microbialaminite) and upper (grain-
2007). The erosion surface must be the combined product of kar- stone) members of the Rasthof Formation have d13C values
stification during the initial glacioeustatic fall and subsequent between þ4 and þ6‰.
erosion by the ice sheet after it developed. It can be reconstructed The negative d13C anomaly following the Ghaub glaciation
in detail using the 0-per-mil intersection of the Trezona extends through the full ,400-m thickness of the Maieberg For-
carbon-isotope anomaly as a datum (Halverson et al. 2002; mation (Fig. 14.2). Isotopic profiles of the Keilberg cap dolostone
Hoffman et al. 2007). At the end of the glaciation, the platform differ significantly according to palaeogeography because of dia-
had a raised rim and a bowl-shaped interior with local highs. chroneity associated with glacioeustatic transgression: shallow-
The Keilberg cap dolostone consistently swells over positive-relief water facies in distal slope sections are entirely older than the
features on the platform (Hoffman et al. 2007) and sea-floor same facies on the platform (Hoffman et al. 2007). Distal slope
cements (pseudomorphosed aragonite crystal fans) are localized sections begin at –0.5‰ and fall asymptotically towards
above them in the overlying Maieberg limestone rhythmite. –3.0‰; platform sections begin near –3.0‰ and rise impercept-
ibly to –2.5‰ over most of their thickness, before falling steeply
to –4.5‰ at the dolostone-limestone transition. Proximal slope
Relation to the Keilberg Member. The Ghaub Formation is every-
sections occupy the middle part of the overall sigmoidal trend.
where overlain conformably and without significant hiatus by
If the timescale of the global meltdown was on the order of 104
the Keilberg Member (Hoffman et al. 2007). Well-stratified pro-
years (Peltier et al. 2004), severe constraints are placed on the
glacial deposits choked with dropstones (Bethanis Member)
origin of the secular d13C change of 4‰ (Hoffman et al. 2007).
precede the Keilberg except on the Duurwater moraine, where
Independent of the secular changes, the absolute d13C values
the Bethanis Member is absent and the Keilberg Member rests
vary systematically across the platform. Sections at the edge of
directly on non-stratified diamictite. The base of the Keilberg is
the platform and on the proximal slope are 1.0‰ and 1.5‰
generally a laminated peloidal-micropeloidal dolostone with low-
enriched in 13C, respectively, compared with sections in the plat-
angle cross-stratification. In distal-slope sections, however, the
form interior. It is implausible, due to rapid mixing, that this
cross-stratified facies is commonly preceded by ,1.0 m of deeper-
lateral isotopic gradient could be maintained in ocean surface
water dolomitic turbidites (Hoffman et al. 2007; Hoffman &
waters but, if the dissolved inorganic carbon (DIC) pool was domi-
Macdonald 2010).
nated by CO2 (Snowball Earth hypothesis), the gradient in d13Ccarb
could reflect a lateral temperature gradient resulting from strong
surface warming over the platform and upwelling of colder
Chemostratigraphy waters at the slope (Hoffman et al. 2007). The isotopic gradient
would result from temperature-dependent isotopic fractionation
The Otavi Group is ideally suited for chemostratigraphy on account of CO23 – ions with respect to aqueous and atmospheric CO2.
of its carbonate-dominated lithology, ocean margin setting and Thus, the lateral gradient in d13C reflects spatial variation in
estimated duration of c. 190 Ma (Halverson et al. 2005). surface temperature, and the secular changes in d13C in all areas
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 203

reflect progressive warming over time as global ice cover recedes. isotopic change straddles the dolostone – limestone transition,
Importantly, this interpretation is contingent upon low-pH (,7.2) which also coincides with the disappearance of wave-generated
conditions, unique to a snowball aftermath, in which CO2 is the bedforms and is interpreted as the incursion due to flooding of
dominant carbon-bearing species. anoxic sulphidic waters onto the platform, accompanied by the
The observed consistent spatial and temporal variations in d13C oxidation of sulphide to sulphate at the oxic – anoxic interface
(Hoffman et al. 2007) demonstrate that the Keilberg cap dolostone (Hurtgen et al. 2006).
does not consist of detrital dolomite (Eyles & Januszczak 2007).
Early diagenetic breccias and cements like those in presumed cor-
relative cap dolostones in South China and West Africa (Jiang Reactive Fe and Mn concentrations
et al. 2003, 2006; Shields et al. 2007) are largely absent from
the Keilberg Member (Hoffman et al. 2007). Hurtgen et al. (2006) report reactive Fe and Mn concentrations for
The lower limestone rhythmites of the Maieberg Formation three, widely spaced, closely sampled sections of the Maieberg
have rather uniform d13C values of –6.0 to –6.5‰ (Fig. 14.2). cap-carbonate sequence. Fe and Mn concentrations from acid-
The drop of c. 0.75‰ (and 2– 3‰ in d18O) at the dolostone – lime- soluble (mainly Fe2þ and Mn2þ carbonates) and diothionite-
stone transition is a step function apparently related to mineralogy. extractable (mainly oxides and oxyhydroxides) sources were
The combined d13C and d18O changes are consistent with measured, but not sulphide-bound Fe. There is excellent agreement
low-temperature equilibrium fractionation between coexisting between the sections. All concentrations rise from very low values
dolomite and calcite (Friedman & O’Neil 1977; Vasconcelos in the lower half of the Keilberg cap dolostone towards high values
et al. 2005). Reverse steps occur at the limestone-dolostone tran- in the middle Maieberg limestone member, before falling in the
sition in the upper Maieberg Formation. The Maieberg is the upper Maieberg dolomitic grainstone. A sharp peak in acid-soluble
only formation in the Otavi Group (including the Rasthof For- Fe (and Ba) concentration occurs at the cap dolostone – limestone
mation) exhibiting a systematic isotope fractionation between stra- transition. Combined with d13C, d18O, d34Spyrite and d34SCAS data
tigraphically associated calcite and dolomite (Hoffman & from the same samples, Hurtgen et al. (2006) propose a model
Halverson 2008). involving progressive mixing between cold, euxinic (i.e. anoxic-
The rise in d13C with stratigraphic height through the middle and sulphidic), deep water and a lid of warm, oxic, brackish water,
upper Maieberg Formation varies laterally in its trajectory, rising during and after deglaciation (see also Shields 2005).
less near the edge of the platform and more in the platform interior.
This may reflect faster sedimentation rates on the outer platform
and diachronous shoaling from the edge of the platform toward Oxygen isotopes
the interior. The ‘isochems’ in Figure 14.3a denote schematically
the surfaces of uniform d13C composition. Oxygen isotopes in carbonate rocks are susceptible to alteration by
fluid flows under diagenetic and metamorphic conditions. d18O
values for relatively unaltered Otavi Group carbonates fall
Strontium isotopes mostly in the range of – 4 to –6‰ VPDB (Hurtgen et al. 2006,
unpublished data). Of the three formations containing both lime-
87
Sr/86Sr ratios have been measured in limestones of the Otavi stone and dolostone, including fabric-retentive dolostone, one
Group deposited before, between and after the Cryogenian glacia- (Maieberg Formation) displays consistent isotopic fractionations
tions. Because Sr is a mobile trace element in carbonate rocks, of both d18O and d13C between the two phases, whereas the
samples with low Sr concentrations (,800 ppm), high Rb –Sr others (Rasthof and Ombaatjie formations) show no pattern what-
ratios, low d18O values and high 87Sr/86Sr ratios are screened out. soever on a cross-plot of d13C v. d18O (Hoffman & Halverson
No data from dolostones survive screening because of their low 2008). In the Maieberg Formation (including its Keilberg
Sr concentrations. According to these criteria, the best limestones Member), limestone is isotopically depleted by c. 1.7‰ in d13C
from the Ombombo Subgroup have 87Sr/86Sr ratios c. 0.7065 and c. 3.8‰ in d18O, relative to dolostone. This fractionation
(P. F. Hoffman & A. J. Kaufman, unpublished data), the Rasthof remains locked despite the large secular change (5‰ in d13C)
cap carbonate c. 0.7068 (Yoshioka et al. 2003), the lower Ombaat- the formation records. This analysis includes the upper dolostone
jie Formation c. 0.7073 (Halverson et al. 2007) and the Maieberg grainstone member as well as the Keilberg Member. The Maieberg
cap carbonate c. 0.7072 (Halverson et al. 2007). These data are data are consistent with low-temperature (c. 50 8C) isotopic equi-
broadly consistent with results from both western and eastern librium fractionation between dolomite and calcite (Friedman &
Laurentia (Brasier & Shields 2000; Halverson et al. 2007) and O’Neil 1977), and by implication that the dolomite formed in
support the correlation of the pairs of glaciogenic formations in contact with seawater.
each region. If one ‘corrects’ the limestone values to bring them in line with
adjacent dolomites (e.g. add 3.8‰ in d18O), a well-defined nega-
tive d18O anomaly of 1 to 3‰ is evident in the first 130 m of the
Sulphur isotopes (CAS) Maieberg Formation (limestone from 13 to 330 m). This is poss-
ibly an isotopic record of the meltwater plume attending deglacia-
Secular variations in the concentration and sulphur isotopic com- tion, before it was mixed with salty glacial deepwater (Hoffman
position (d34S) of carbonate-associated sulphate (CAS) have 1999; Shields 2005).
been profiled in sections of the Abenab Subgroup and the Maieberg
cap-carbonate sequence (Hurtgen et al. 2002, 2006). CAS concen-
trations are mostly ,200 ppm, lower than Phanerozoic carbonates, Boron and calcium isotopes
and this combined with large (.20‰) and apparently rapid isoto-
pic fluctuations suggests that seawater sulphate concentrations Kasemann et al. (2005, 2010) report d11B and d44Ca measurements
were lower than in Phanerozoic oceans (Hurtgen et al. 2002). for carbonates of the Ombaatjie and Maieberg formations as
Large, complex, but broadly similar d34S anomalies are associated proxies for pre- and post-glacial seawater. There is a large (c.
with the Rasthof and Maieberg cap-carbonate sequences. Their 20‰) isotopic fractionation between borate ion (which is incorpor-
lower and upper parts are isotopically enriched to 40 and 50‰ ated into carbonate) and boric acid in seawater, and the relative
(CDT), respectively, while their middle part, corresponding proportions of these species change from nearly 0 to almost 1.0
closely to the deepest-water facies on the platform, are isotopically over the pH range of 7.5– 9.5. Changes in seawater d11B over geo-
depleted to c. 15‰. In the Maieberg Formation, the steepest logical time (i.e. millions of years) may also reflect changes in
204 P. F. HOFFMAN

ocean inputs and outputs of boron. Because continental weathering No constraints on the Ghaub glaciation have been obtained
tends to lower seawater d11B and mid-ocean ridge hydrothermal from the Otavi platform, but a tuff layer within carbonates of the
exchange to raise it, we may suppose that seawater should central Damara Belt carrying ice-rafted debris, believed correla-
become relatively more enriched if continental weathering was tive with the Ghaub Formation, gives an age of 635.5 + 0.5 Ma
held down for millions of years under snowball conditions. (Hoffmann et al. 2004; Condon et al. 2005). This is indistinguish-
Instead, post-glacial d11B values are more depleted, reaching a able from the age of 635.2 + 0.6 Ma for a tuff lying upon the post-
nadir at the top of the Keilberg cap dolostone and in the directly glacial (Nantuo diamictite) cap dolostone in South China (Condon
overlying Maieberg limestones. Using their d44Ca data to correct et al. 2005). The Ghaub glaciation is therefore inferred to have
for changes in Ca (and indirectly boron) ocean input and output, ended in 635 Ma.
Kasemann et al. (2005) calculate changes in the maximum levels
of atmospheric pCO2 over time. Their results allow for a rise in
pCO2 of up to 0.1 atm (100 000 ppm) in the glacial aftermath, Discussion
broadly consistent with the Snowball Earth hypothesis.
Tectonic/palaeogeographical setting and origin of the
Ghaub Formation
Palaeolatitudes and palaeogeography
Eyles & Januszczak (2007; see also Eyles 2004; Eyles & Januszc-
Neoproterozoic sedimentary rocks in Namibia were extensively zak 2004a, b) examined the Ghaub Formation in two sections, near
and thoroughly remagnetized during the orogenic amalgamation the western terminus of the Fransfontein Ridge (Narachaams) and
of Gondwanaland in the Cambrian. Palaeopoles from mafic at the Fransfontein drainage gap (Fig. 14.1b). They conclude that it
members of the post-tectonic Mbozi igneous complex in Tanzania consists of subaqueous mass-flow deposits of non-glacial origin,
(Meert et al. 1995), felsic components of which give K –Ar cooling coincident with rifting and breakup of the Congo craton. Accord-
ages of 755 + 25 Ma (recalculated from Cahen & Snelling 1966; ingly, the Ghaub Formation has no bearing on the Neoproterozoic
Evans 2000), suggest a palaeolatitude of 10 + 58 for the Otavi glacial record and provides no support for the Snowball Earth
platform c. 40 Ma before the Chuos glaciation, assuming a rigid hypothesis (Eyles & Januszczak 2007).
Congo craton (Evans 2000). Dual-polarity palaeopoles from the Their argument boils down to four interconnected points. First,
Nola dyke swarm (Central African Republic) dating from its the Ghaub Formation lacks key glacial indicators such as faceted,
metamorphism and remagnetization at 571 + 9 Ma (Moloto-A- striated and bullet-shaped clasts. Second, the pre-, syn- and post-
Kenguemba et al. 2008) place the Otavi platform near the palaeoe- Ghaub facies are indistinguishable, meaning that the identification
quator (04 + 58) c. 65 Ma after the Ghaub glaciation. of a glaciogenic interval is entirely arbitrary. Third, the Ghaub dia-
Preliminary palaeomagnetic data from the basal part of the mictites occur strictly in base-of-slope palaeoenvironments, con-
Keilberg cap dolostone place the Otavi platform in the (southern) sistent with their origin as subaqueous mass-flows. Fourth, they
subtropics with its present southern margin facing the palaeoequa- were deposited contemporaneously with rift faulting and continen-
tor (R.I.P. Trindade, pers. comm.). These fragmentary results are tal break-up at the edge of the Congo craton, suggesting that active
at least consistent with more robust data sets indicating that faulting was the trigger for mass wasting. These conclusions are so
other Neoproterozoic carbonate-dominated successions formed at variance with those favoured by other authors (e.g. Condon et al.
at palaeolatitudes ,358 (e.g. Trindade et al. 2003; Macouin 2002; Domack & Hoffman 2011) that a point-by-point discussion
et al. 2004; Kilner et al. 2005; Maloof et al. 2006), and therefore is warranted.
that the meridional climatic gradient was not reversed as proposed
in the large orbital obliquity hypothesis for low-latitude glaciation Diagnostic glacial indicators. In the type area of the Ghaub diamic-
(Williams 1993; see also Evans 2006). tite, le Roex (1941) reported that ‘a high percentage of the pebbles
show typical glacial faceting, particularly the hard quartzitic
types’. Some also show striations, ‘which are much better pre-
Geochronological constraints served on the quartzitic than the calcareous types’ (le Roex
1941). Quartzitic pebbles do not occur in the Ghaub Formation
Zircons were extracted from tuffaceous layers within the Chuos on the Fransfontein foreslope, only carbonate clasts (along with
Formation, from strata directly preceding the Ghaub glaciation rare granitic detritus at Bethanis). The problem is that the carbon-
on the platform and slope, and from both cap carbonates. No ate clasts and the carbonate matrix were thoroughly welded
primary volcanogenic zircons were recovered. The glacial history together during middle greenschist facies metamorphism. Facets
of the Otavi platform is currently constrained by just three U –Pb and striae might exist on every clast, or on none: it would be
zircon ages. impossible to tell the difference.
The first provides a maximum constraint for the Chuos For- In the absence of clast morphology, dropstones are the most
mation on the platform. An age of 760 + 1 Ma (Halverson et al. diagnostic glacial feature (along with the overall facies associ-
2005) was obtained for a tuff near the top of the Devede Formation ations and stratal architecture). Classic attributes of dropstones
of the Ombombo Subgroup (Fig. 14.2), nearly 350 m stratigraphi- are well-developed in the Ghaub Formation, including pinched and
cally below the Chuos at that locality. The age is close to punctured substrata, thickened and ejected side-strata, draped or
756 + 2 Ma for the Oas quartz-syenite pluton (Hoffman et al. onlapping superstrata, and independence from bottom sediment
1996), representing a suite of peralkaline igneous rocks associated transport events. Ejection folds are directly analogous to the ‘over-
with crustal stretching. turned flap’ on the rim of Meteor Crater, Arizona (Shoemaker
The younger Naauwpoort bimodal (basalt-rhyolite) volcanics 1963). Eyles & Januszczak (2004a, 2007) emphasize that drop-
provide a tighter maximum constraint on Chuos glaciation in the stones also occur in non-glacial settings, but the frequency of
Summas Mountains area (Fig. 14.1a). There, rhyolite lava that is their occurrence in the Ghaub Formation is unmatched by any non-
unconformably overlain by (tectonically overturned) Chuos dia- glacial deposit. With a conservative average of one dropstone per
mictite has an age of 747 + 2 Ma (Hoffman et al. 1996). An 1000 ppm3, the Bethanis Member alone held 50 trillion (5  1013)
ash-flow tuff with an indistinguishable age of 746 + 2 Ma in the dropstones given its original dimensions of 0.1  5  100 km3.
same area is separated from the Chuos by 720 m of shallow-water
mixed carbonates and clastics of the Ugab Subgroup (Hoffman Pre-, syn- and post-glacial facies. The lower and upper contacts of
et al. 1996; Hoffman & Halverson 2008). Accordingly, the onset of the Ghaub Formation can be unambiguously mapped on the
the Chuos glaciation must be substantially younger than 746 Ma. basis of lithology with an accuracy of centimetres along the
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 205

length of the Fransfontein foreslope. The same is true of the out- and granitic basement in the Otavi Mountains by hundreds of
liers to the west (Bethanis-Toekoms and Vrede-Opdraend). The metres of shallow-water (stromatolitic and oolitic) carbonate
carbonate turbidite-hosted debris flows of the underlying strata forming a conformable stratigraphic succession. He con-
Franni-aus Member and overlying Maieberg Formation are oligo- cluded that the clasts must have been transported from distant
mictic and intraformational; those of the Ghaub Formation are sources, where quartzite and granitic basement were exposed in
polymictic and extraformational. Weakly stratified to structureless Ghaub time.
diamictites make up 82% of the Ghaub Formation (based on 72 Halverson et al. (2002) greatly strengthened le Roex’s (1941)
measured sections) but are uncommon or absent in the adjacent argument with detailed isotopic chemostratigraphy of the upper
formations. Similarly, dropstones are profligate in the Ghaub Ombaatjie Formation. They documented a steep decline in d13C
but rare or non-existent in the Franni-aus and Maieberg. In short, from þ5‰ or higher down to – 5‰ in the final pair of depositional
the Ghaub is a true formation, lithologically mappable on the cycles (parasequences) at 17 locations distributed from the edge of
regional scale. the platform far into its interior. This isotopic shift has been corre-
How could Eyles & Januszczak (2007) conclude that the same lated globally and is named the Trezona anomaly after the for-
essential features occur within and below the Ghaub Formation? mation in South Australia where it was first encountered
The answer is the following. At Narachaams (see their figs 4 and (Halverson et al. 2002, 2005). Plotting all the sections using the
5), they place the base of the Ghaub Formation at the base of the interpolated 0‰ intersection as the datum provides a quantitative
Bethanis Member, which constitutes only the top 2 m of the basis for reconstructing the palaeotopography of the erosion
54-m-thick Ghaub Formation in that section. At Fransfontein surface on the platform beneath the Ghaub diamictite, or the Keil-
(their figs 3 and 5), everything they call Ghaub Formation lies berg Member where the Ghaub is absent (Halverson et al. 2002;
below the Bethanis Member. In other words, their ‘Ghaub Hoffman et al. 2007). There is up to 80 m of local relief on the
Formation’ at Fransfontein is stratigraphically equivalent to erosion surface but no evidence of uplift or back-rotation of the
their ‘Abenab Subgroup’ at Narachaams. Small wonder their outer platform. If a basinward-dipping normal fault was active at
‘pre-Ghaub’ and ‘syn-Ghaub’ facies are similar; they are one the edge of the platform (Eyles & Januszczak 2007), uplift and
and the same. back-rotation would occur as an isostatic response to tectonic
unloading of the footwall. The resulting unconformities as well
Palaeoenvironmental setting of the Ghaub Formation. Eyles & as cannibalistic clastic deposits on the back-rotated dip-slopes
Januszczak (2007) concur with the interpretation of the distal are precisely the criteria successfully used to determine the
Fransfontein foreslope (Fig. 14.1b) as a deep-water marine location and timing of rift faulting in the lower Otavi Group,
environment (Henry et al. 1990; Halverson et al. 2002, 2005; prior to the Ombaatjie Formation (Soffer 1998; Hoffman 1999;
Hoffman 2005). Their interpretation of the Ghaub Formation as Halverson et al. 2002; Hoffman & Halverson 2008). However,
a stack of submarine mass flows is consistent with this setting, there is no evidence of such activity at the time of the Ghaub
but it does not account for sedimentary structures (low-angle cross- Formation. The edge of the platform preserves more upper
lamination and giant wave ripples) in the overlying Keilberg cap Ombaatjie strata than does the interior, not less, and over the plat-
dolostone, which demonstrate that the distal foreslope was above form as a whole the upper Ombaatjie cycles and the Keilberg
storm wave-base at the glacial – deglacial transition (Hoffman Member are parallel. This is a very sensitive test because even a
et al. 2007). The diachronous nature and extent of the cap dolo- small angular rotation of a rift ‘shoulder’ of modest dimensions
stone (same-depth deposits on the platform and distal foreslope) would result in significant stratigraphic truncation in a shallow
support a large-amplitude glacioeustatic fluctuation, as a result marine environment. None is observed. Before Eyles & Januszc-
of which the ice grounding-line advanced to the distal foreslope zak (2004a, 2007) began their study, the history of rift faulting
before retreating at the glacial termination (Hoffman et al. 2007). on the Otavi platform was known to have ended before the
The facies associations and stratal architecture of the Ghaub Ombaatjie Formation was deposited, millions of years before the
Formation have much in common with ice grounding-line wedges Ghaub glaciation (Hoffman et al. 1998a; Halverson et al. 2002;
of Quaternary age on the continental shelves and upper slopes of Hoffman 2002).
polar seas (e.g. Alley et al. 1989; Boulton 1990; Powell 1990;
King et al. 1991; King 1993; Powell & Domack 1995). The well-
stratified facies association represents proglacial marine deposits. Significance of ice-rafted dropstones for maximum
The sandy facies association resembles ice grounding-line fans ice-shelf extent
(Powell 1990). Weakly-stratified to structureless diamictites rep-
resent ice-proximal and ice-contact deposits. Diamictites that The presence of ice-rafted dropstones in the Ghaub Formation on
grade downwards into stratified proglacial deposits probably accu- the Fransfontein Ridge was taken as evidence for the existence of
mulated as rain-out deposits immediately seaward of the ice free-floating icebergs (Condon et al. 2002), which would be incon-
grounding line. Ice-contact deposits include diamictites with sistent with a globe-encircling ice shelf as predicted by the Snow-
erosive bases, internal reactivation surfaces and discontinuous ball Earth hypothesis (Hoffman & Schrag 2002; Warren et al.
stringers of finely laminated siltstone. The silt-stringers are inter- 2002; Goodman & Pierrehumbert 2003). Two issues are raised
preted as ice-bed separation cracks beneath stagnant grounded by this interpretation. Were the dropstones carried by icebergs or
ice. They were susceptible to deformation associated with recur- were they released by basal melting of ice-shelf ice carrying eng-
rent ice flowage. lacial debris across the grounding line? Free-floating icebergs
Eyles & Januszczak (2007) fail to mention the widespread relics transport debris farther from the grounding line than ice-shelf
of Ghaub diamictite on the Otavi platform, including the 60 m ice. Iceberg transport is favoured for the Bethanis Member, but
of diamictite with granitic and quartzitic clasts in the Otavi the high proportion (88%) of weakly stratified and structureless
Mountains type section (le Roex 1941; Hoffmann & Prave diamicitites in the rest of the Ghaub Formation supports deposition
1996). As these diamictites are underlain and overlain by shallow- close to the grounding line. Thus, in the absence of iceberg
water carbonates (Ombaatjie Formation and Keilberg Member, ‘dumps’ (Condon et al. 2002), the origin of dropstones below
respectively), it is improbable that they originated as submarine the Bethanis Member is ambiguous with respect to the continuity
mass-flows. of ocean ice cover.
The second issue is whether the Ghaub Formation represents the
Role of rift faulting. Le Roex (1941) argued that the quartzitic and maximum glacial conditions, or only the waning stages of glacia-
granitic clasts in his diamictite could not be locally derived (e.g. tion. If the smooth, planar and laterally continuous surface at the
through faulting) because the diamictite is separated from quartzite base of the Ghaub Formation on the Fransfontein Ridge represents
206 P. F. HOFFMAN

a subglacial erosion surface, then the overlying glacial deposits do sediments forming today only in latitudes ,358 (Köppen &
not represent the first, or even necessarily the maximum, glacial Wegener 1924, p. 19; Rodgers 1957). The zonal distribution of car-
stage. Likewise the Duurwater trough, if it was cut by an ice bonate sediments did not change between the warm (non-glacial)
stream, must represent a glacial maximum that is unrepresented and cool (glacial – interglacial) periods of the Phanerozoic (Black-
by deposits in the Ghaub Formation itself (Hoffman 2005). ett 1961; Briden & Irving 1964; Briden 1970; Ziegler et al. 1984;
Accordingly, any evidence for open water within the Ghaub For- Opdyke et al. 1990; Witzke 1990; Kiessling 2001). This is because
mation on the Fransfontein Ridge is ambiguous with regards to carbonate production and preservation reflect relative, not absolute
the maximum extent of the ice. A complete glacial record will temperatures. Moreover, it implies that the presence or absence
only be found seaward of the maximum ice grounding line. of polar ice sheets caused no perceptible change in the flux of
alkalinity entering the ocean. It is therefore a reasonable inference
that the Otavi carbonate platform occupied one of the warmer parts
Significance of grounding-line oscillations of the global ocean before, between and after the Cryogenian
glaciations.
Condon et al. (2002) also suggested that the evidence in the If one accepts that an ice sheet on the Otavi platform had a con-
Ghaub Formation for grounding-line oscillations was inconsistent tinuous tidewater grounding line in one of the warmer parts of the
with an ice-covered ocean. The logic behind this assertion is Cryogenian ocean, then contemporaneous ice sheets must have
unclear: changes in the mass-balance of an ice sheet (driven, existed at higher elevations and palaeolatitudes, wherever there
for example, by orbital forcing) or, alternatively, episodicity was net meteoric precipitation or glacial influx.
in ice-sheet dynamics (e.g. the binge – purge cycles of
MacAyeal 1993) might just as well occur if the hydrological Research on the Otavi Group was initially supported by the Geological Survey
cycle was driven by the sublimation of sea ice as by the evapor- of Namibia and the Canadian National Science and Engineering Research
ation of seawater. Relative to grounding-line oscillations on Qua- Council (NSERC). Since 1994, support was generously provided by the United
ternary continental shelves, however, the oscillations recorded in States National Science Foundation (NSF), NASA Astrobiology Institute,
the Ghaub Formation appear to have been more limited in scale, Harvard University Center for the Environment (HUCE), and the Canadian
possibly because of the inclination of the Otavi foreslope Institute for Advanced Research (CIFAR). Among the many who have contribu-
(Pollard & deConto 2007; but see also Alley et al. 2007). ted to this work, I am particularly indebted to S. A. Bowring, E. W. Domack,
G. P. Halverson, J. A. Higgins, M. T. Hurtgen, A. J. Kaufman, F. A. Macdonald,
A. C. Maloof, D. P. Schrag, G. Soffer and R. Trindade. Roy McG. Miller gave
Magnitude of base-level and glacioeustatic changes valuable advice on stratigraphic names. G. Jiang is thanked for his perceptive
comments on the manuscript. This represents a contribution of the IUGS- and
The existence of grounded ice on the distal foreslope at an UNESCO-funded IGCP (International Geoscience Programme) Project #512.
inferred palaeodepth of c. 0.5 km below the rim of the platform,
combined with shallow-water structures (low-angle cross-strata
and giant wave ripples) in the directly overlying cap dolostone
and post-glacial flooding of the platform below storm wave-base References
in the succeeding limestone (Fig. 14.3a) imply a base-level
Adams, E. W. & Schlager, W. 2000. Basic types of submarine slope
change of .0.5 km at the glacial termination. The glacioeustatic curvature. Journal of Sedimentary Research, 70, 814– 828.
change must have been far greater than this because of the iso- Alley, R. B., Blankenship, D. D., Rooney, S. T. & Bentley, C. R.
static rebound of the platform due to the removal of the ice 1989. Sedimentation beneath ice shelves– the view from ice stream
sheet, the hydroisostatic sinking of the sea floor (and therefore B. Marine Geology, 85, 101– 120.
of sea level) due to the addition of glacial meltwater, and the Alley, R. B., Anandakrishnan, S., Dupont, T. K., Parizek, B. R. &
additional lowering of sea level in the vicinity of a retreating Pollard, D. 2007. Effect of sedimentation on ice-sheet grounding-
ice sheet due to the weakening of its gravitational attraction on line stability. Science, 315, 1838–1841.
the adjacent ocean (Clark 1976; Farrell & Clark 1976). The Bechstädt, T., Jäger, H., Spence, G. & Werner, G. 2009. Late Cryo-
time scale over which the platform was flooded, given the thick- genian (Neoproterozoic) glacial and post-glacial successions at the
ness of the deepwater deposits (Fig. 14.3a), must have far southern margin of the Congo Craton, northern Namibia: facies,
exceeded the timescale of the isostatic adjustments. If the thick- paleogeography and hydrocarbon perspective. In: Craig, J.,
ness of the ice sheet near the edge of the Otavi platform was Thurow, J., Thusu, B., Whitham, A. & Abutarruma, (eds)
c. 1.0 km, comparable to the edge of the East Antarctic ice Global Neoproterozoic Petroleum Systems: The Emerging Potential
sheet today (Lythe et al. 2001), then the isostatic rebound in North Africa. Geological Society, London, Special Publications,
would have been c. 0.27 km (rice/rmantle). If the global average 326, 255– 287.
thickness of ice sheets on all continents was .2.5 km, as Blackett, P. M. S. 1961. Comparison of ancient climates with the ancient
suggested by climate models of a Snowball Earth (Donnadieu latitudes deduced from rock magnetic measurements. Proceedings of
et al. 2003; Pollard & Kasting 2004), then .1.25 km of meltwater the Royal Society of London, Series A, 263, 1 –30.
Boulton, G. S. 1990. Sedimentary and sea-level changes during
would have been added to the ocean because its area is roughly
glacial cycles and their control on glacimarine facies architecture.
twice that of the continents. The resulting hydroisostatic sinking
In: Dowdeswell, J. A. & Scourse, J. D. (eds) Glacimarine Environ-
of the seafloor would have been .0.38 km (1.25/rmantle). The ments: Processes, Sediments. Geological Society of London, Special
additional lowering of sea level in the vicinity of the former ice Publications, 53, 15 –52.
sheet could be as great as c. 0.1 km (Farrell & Clark 1976). Brasier, M. D. & Shields, G. 2000. Neoproterozoic chemostratigraphy
Accordingly, a glacioeustatic change of .1.25 km is fully compa- and correlation of the Port Askaig glaciation, Dalradian Supergroup
tible with the inferred base-level change of .0.5 km (i.e. of Scotland. Journal of the Geological Society, London, 157, 909–914.
1.25–0.27–0.38 –0.1 ¼ 0.5 km). Briden, J. C. 1970. Palaeolatitude distribution of precipitated sediments.
In: Runcorn, S. K. (ed.) Palaeogeophysics. Academic Press,
London, 437– 444.
Significance of the associated carbonates Briden, J. C. & Irving, A. 1964. Palaeolatitude spectra of sedimentary
palaeoclimatic indicators. In: Nairn, A. E. M. (ed.) Problems in
The ‘reverse’ solubility of calcium carbonate accounts for the Palaeoclimatology. John Wiley & Sons, New York, 199–224.
preferential accumulation of carbonate sediments in the warmer Cahen, L. & Snelling, N. J. 1966. The Geochronology of Equatorial
parts of the surface ocean. Carbonate dominates shallow-water Africa. North-Holland Publishing Co., Amsterdam.
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 207

Calver, C. R. 1998. Isotope stratigraphy of the Neoproterozoic Togari Halverson, G. P. 2006. A Neoproterozoic chronology. In: Xiao, S. &
Group, Tasmania. Australian Journal of Earth Sciences, 45, Kaufman, A. J. (eds) Neoproterozoic Geobiology and Paleobiology.
865– 874. Springer, Dordrecht, 231– 271.
Clark, J. A. 1976. Greenland’s rapid postglacial emergence: A result of Halverson, G. P., Hoffman, P. F., Schrag, D. P. & Kaufman, J. A.
ice-water gravitational attraction. Geology, 4, 310– 312. 2002. A major perturbation of the carbon cycle before the Ghaub
Clifford, T. N. 2008. The geology of the Neoproterozoic Swakob-Otavi glaciation (Neoproterozoic) in Namibia: prelude to snowball
transition zone in the Outjo District, northern Damara Orogen, Earth? Geophysics, Geochemistry, Geosystems, 3, doi: 10.1029/
Namibia. South African Journal of Geology, 111, 117– 140, 3 maps. 2001GC000244.
Condon, D. J., Prave, A. R. & Benn, D. I. 2002 Neoproterozoic Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
glacial rainout intervals: observations and implications. Geology, Rice, A. H. N. 2005. Toward a Neoproterozoic composite
30, 35 – 38. carbon-isotope record. Geological Society of America Bulletin, 117,
Condon, D., Zhu, M., Bowring, S. A., Wang, W., Yang, A. & Jin, Y. 1181–1207, doi: 10.1130/B25630.1
2005. U– Pb ages from the Neoproterozoic Doushantuo Formation, Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A.
China. Science, 308, 95– 98. 2007. Evolution of the 87Sr/86Sr composition of Neoproterozoic
Dingle, R. V. 1977. The anatomy of a large submarine slump on a sheared seawater. Palaeogeography, Palaeoclimatology, Palaeoecology,
continental margin (SE Africa). Journal of the Geological Society 256, 103–129.
(London), 134, 293 –310. Hedberg, R. M. 1979. Stratigraphy of the Owamboland Basin, South
Domack, E. W. & Hoffman, P. F. 2011. An ice grounding-line wedge West Africa. Precambrian Research Unit Bulletin, 24, 325, 6 maps,
from the Ghaub glaciation (635 Ma) on the distal foreslope of University of Cape Town, South Africa.
the Otavi platform, Namibia, and its bearing on the Snowball Henry, G., Stanistreet, I. G. & Maiden, K. J. 1986. Preliminary results
Earth hypothesis. Geological Society of America Bulletin, 123, of a sedimentological study of the Chuos Formation in the central
1448– 1477. zone of the Damara Orogen: evidence for mass flow processes and
Donnadieu, Y., Fluteau, F., Ramstein, G., Ritz, C. & Besse, J. 2003. glacial activity. Communications of the Geological Survey of South
Is there a conflict between the Neoproterozoic glacial deposits and the West Africa/Namibia, 2, 75 –92.
snowball Earth interpretation: an improved understanding with Henry, G., Clendenin, C. W., Stanistreet, I. G. & Malden, K. J.
numerical modeling. Earth and Planetary Science Letters, 208, 1990. A multiple detachment model for the early rifting stage of
101– 112. the Late Proterozoic Damara Orogen in Namibia. Geology, 18,
Eberli, G. P. & Ginsburg, R. N. 1987. Segmentation and coalescence of 67– 71.
Cenozoic carbonate platforms, northwestern Great Bahama Bank. Higgins, J. A. & Schrag, D. P. 2003. Aftermath of a snowball Earth. Geo-
Geology, 15, 75 –79. physics, Geochemistry, Geosystems, 4, doi: 10.1029/2002GC000403.
Evans, D. A. D. 2000. Stratigraphic, geochronological, and paleomagnetic Hoffman, P. F. 1999. The break-up of Rodinia, birth of Gondwana, true
constraints upon the Neoproterozoic climatic paradox. American polar wander, and the snowball Earth. Journal of African Earth
Journal of Science, 300, 347–433. Sciences, 28, 17 –33.
Evans, D. A. D. 2006. Proterozoic low orbital obliquity and axial-dipolar Hoffman, P. F. 2002. Carbonates bounding glacial deposits: evidence
geomagnetic field from evaporite palaeolatitudes. Nature, 444, for Snowball Earth episodes and greenhouse aftermaths in the Neo-
51 – 55. proterozoic Otavi Group of northern Namibia. Excursion Guide,
Eyles, N. 2004. Frozen in time: concepts of ‘global glaciation’ from 1837 16th International Sedimentological Conference, Auckland Park,
(die Eiszeit) to 1998 (the Snowball Earth). Geoscience Canada, 31, South Africa, 39.
157– 166. Hoffman, P. F. 2005. 28th DeBeers Alex Du Toit Memorial Lecture: On
Eyles, N. & Januszczak, N. 2004a. ‘Zipper-rift’: a tectonic model for Cryogenian (Neoproterozoic) ice-sheet dynamics and the limitations
Neoproterozoic glaciations during the breakup of Rodinia after of the glacial sedimentary record. South African Journal of Geology,
750 Ma. Earth-Science Reviews, 65, 1– 73. 108, 557–576.
Eyles, N. & Januszczak, N. 2004b. Interpreting the Neoproterozoic Hoffman, P. F. 2011. Strange bedfellows: glacial diamictite and cap
glacial record: the importance of tectonics. In: Jenkins, G. S., carbonate from the Marinoan (635 Ma) glaciation in Namibia. Sedi-
McMenamin, M. A. S., McKey, C. P. & Sohl, L. (eds) The mentology, 58, 57– 119.
Extreme Proterozoic: Geology, Geochemistry, and Climate. Geophy- Hoffman, P. F. & Halverson, G. P. 2008. The Otavi Group of the North-
sical Monograph 146. American Geophysical Union, Washington, ern Platform and the Northern Margin Zone. In: Miller, R. McG.
DC, 125–144. (ed.) The Geology of Namibia, vol. 2. Geological Survey of
Eyles, N. & Januszczak, N. 2007. Syntectonic subaqueous mass flows Namibia, Windhoek, 13.69– 13.136.
of the Neoproterozoic Otavi Group, Namibia: where is the evidence Hoffman, P. F. & Hartz, E. H. 1999. Large, coherent, submarine
of global glaciation? Basin Research, 19, 179–198, doi: 10.1111/ landslide associated with Pan-African foreland flexure. Geology,
j.1365-2117.00319.x 27, 687– 690.
Farrell, W. E. & Clark, J. A. 1976. On postglacial sea level. Geophysi- Hoffman, P. F. & Macdonald, F. A. 2010. Sheet-crack cements and
cal Journal of the Royal Astronomical Society, 46, 647–667. early regression in Marinoan (635 Ma) cap dolostones: regional
Frets, D. C. 1969. Geology and structure of the Huab-Welwitschia area, benchmarks of vanishing ice-sheets? Earth and Planetary Science
South West Africa. Precambrian Research Unit Bulletin, 5, 235, 2 Letters, 300, 374– 384.
maps, University of Cape Town, South Africa. Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
Friedman, I. & O’Neil, J. R. 1977. Compilation of stable-isotope frac- testing the limits of global change. Terra Nova, 14, 129– 155.
tionation factors of geochemical interest. In: Fleischer, M. (ed.) Hoffman, P. F., Hawkins, D. P., Isachsen, C. E. & Bowring, S. A. 1996.
Data of Geochemistry, 6th edn. United States Geological Survey, Precise U– Pb zircon ages for early Damaran magmatism in the
Washington, DC, Professional Paper 44-KK. Summas Mountains and Welwitschia Inlier, northern Damara belt,
Gevers, T. W. 1931. An ancient tillite in South-West Africa. Transactions Namibia. Communications of the Geological Survey of Namibia,
of the Geological Society of South Africa, 34, 1 –17. 11, 47– 52.
Goodman, J. & Pierrehumbert, R. T. 2003. Glacial flow of floating Hoffman, P. F., Kaufman, J. A. & Halverson, G. P. 1998a. Comings
marine ice in ‘Snowball Earth’. Journal of Geophysical Research, and goings of global glaciations on a Neoproterozoic carbonate plat-
108, doi: 10.1029/2002JC001471. form in Namibia. GSA Today, 8, 1 – 9.
Goscombe, B. & Gray, D. R. 2007. The Coastal Terrane of the Kaoko Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P.
Belt, Namibia: outboard arc-terrane and tectonic significance. Pre- 1998b. A Neoproterozoic snowball Earth. Science, 281, 1342– 1346.
cambrian Research, 155, 139–158. Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
Goscombe, B., Hand, M. & Gray, D. 2003. Structure of the Kaoko Belt, Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635 Ma)
Namibia: progressive evolution of a classic transpressional orogen. post-glacial ‘cap dolostones’ diachronous? Earth and Planetary
Journal of Structural Geology, 25, 1049– 1081. Science Letters, 258, 114–131.
208 P. F. HOFFMAN

Hoffmann, K.-H. & Prave, A. R. 1996. A preliminary note on a revised Society of America Bulletin, 118, 1099–1124, doi: 10.1130/
subdivision and regional correlation of the Otavi Group based on gla- B25892.1.
ciogenic diamictites and associated cap dolomites. Communications Martin, H. 1965a. The Precambrian geology of South West Africa and
of the Geological Survey of Namibia, 11, 77 –82. Namaqualand. Precambrian Research Unit Bulletin, 1, 159, 11
Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L. plates, 2 maps, University of Cape Town, South Africa.
2004. U– Pb zircon date from the Neoproterozoic Ghaub Martin, H. 1965b. Beobachtungen zum Problem der jung-präkam-
Formation, Namibia: constraints on Marinoan glaciation. Geology, brischen Glazialen Ablagerungen in Südwestafrika (Observations
32, 817– 820. concerning the problem of the late Precambrian glacial deposits in
Hurtgen, M. T., Arthur, M. A., Suits, N. S. & Kaufman, A. J. 2002. South West Africa). Geologische Rundschau, 54, 115– 127.
The sulfur isotopic composition of Neoproterozoic seawater sulfate: Martin, H., Porada, H. & Walliser, O. H. 1985. Mixtite deposits of the
implications for a snowball Earth? Earth and Planetary Science Damara Sequence, Namibia, problems of interpretation. Palaeogeo-
Letters, 203, 413–429. graphy, Palaeoclimatology, Palaeoecology, 51, 159–196.
Hurtgen, M. T., Halverson, G. P., Arthur, M. A. & Hoffman, P. F. McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial
2006. Sulfur cycling in the aftermath of a 635 Ma snowball glacia- trinity: Neoproterozoic Earth history within the British-Irish Caledo-
tion: evidence for a syn-glacial sulfidic deep ocean. Earth and Plane- nides. Geology, 34, 909–912, doi: 10.1130/G22694A.1
tary Science Letters, 245, 551– 570. McMullen, K., Domack, E., Leventer, A., Olson, C., Dunbar, R. &
Jiang, G., Kennedy, M. J. & Christie-Blick, N. 2003. Stable isotopic Brachfeld, S. 2006. Glacial morphology and sediment formation in
evidence for methane seeps in Neoproterozoic postglacial cap car- the Metz Trough, East Antarctica. Palaeogeography, Palaeoclima-
bonates. Nature, 426, 822– 826. tology, Palaeoecology, 231, 169–180.
Jiang, G., Kennedy, M. J., Christie-Blick, N., Wu, H. & Zhang, S. Meert, J. G., Van der Voo, R. & Ayub, S. 1995. Paleomagnetic investi-
2006. Stratigraphy, sedimentary structures, and textures of the late gation of the Neoproterozoic Gagwe lavas and Mbozi complex,
Neoproterozoic Doushantuo cap carbonate in South China. Journal Tanzania, and the assembly of Gondwana. Precambrian Research,
of Sedimentary Research, 76, 978–995. 74, 225– 244.
Johnson, S. D., Poujol, M. & Kisters, A. F. M. 2006. Constraining the Miller, R. McG. 1997. The Owambo Basin of northern Namibia. In:
timing and migration of collisional tectonics in the Damara Belt, Selley, R. C. (ed.) African Basins. Sedimentary Basins of the
Namibia: U– Pb zircon ages for the syntectonic Salem-type Stinkbank World, 3. Elsevier, Amsterdam, 237– 268.
granite. South African Journal of Geology, 109, 611– 624. Miller, R. McG. 2008. Otavi Group of the Otavi Mountainland (OML),
Kasemann, S. A., Hawkesworth, C. J., Prave, A. R., Fallick, A. E. & the region west of the OML, and the eastern Kaokoveld. In: Miller,
Pearson, P. N. 2005. Boron and calcium isotope composition in Neo- R. McG. (ed.) The Geology of Namibia, 2. Geological Survey of
proterozoic carbonate rocks from Namibia: evidence for extreme Namibia, Windhoek, 13.52– 13.69.
environmental change. Earth and Planetary Science Letters, 231, Moloto-A-Kenguemba, G., Trindade, R. I. F., Monié, P., Nédélec, A.
73 –86. & Siqueira, R. 2008. A late Neoproterozoic paleomagnetic pole for
Kasemann, S. A., Prave, A. R., Fallick, A. E., Hawkesworth, C. J. & the Congo craton: tectonic setting, paleomagnetism and geochronol-
Hoffmann, K.-H. 2010. Neoproterozoic ice ages, boron isotopes, ogy of the Nola dike swarm (Central African Republic). Precambrian
and ocean acidification: implications for a snowball Earth. Geology, Research, 164, 214– 226.
38, 775– 778. Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V., Trin-
Kendall, C. G. St. C. & Warren, J. 1987. A review of the origin and dade, R. I. F. & Fairchild, T. R. 2007. Carbon and strontium isotope
setting of tepees and their associated facies. Sedimentology, 34, fluctuations and paleoceanographic changes in the late Neoprotero-
1007– 1028. zoic Araras carbonate platform, southern Amazon craton, Brazil.
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K.-H. & Chemical Geology, 237, 168– 190.
Arthur, M. A. 1998. Two or four Neoproterozoic glaciations? Opdyke, B. N. & Wilkinson, B. H. 1990. Palaeolatitude distribution of
Geology, 26, 1059–1063. Phanerozoic marine ooids and cements. Palaeogeography, Palaeocli-
Kiessling, W. 2001. Paleoclimatic significance of Phanerozoic reefs. matology, Palaeoecology, 78, 135–148.
Geology, 29, 751–754. Ottesen, D., Dowdeswell, J. A. & Rise, L. 2005. Submarine landforms
Kilner, B., Mac Niocaill, C. & Brasier, M. 2005. Low-latitude glacia- and the reconstruction of fast-flowing ice streams within a large
tion in the Neoproterozoic of Oman. Geology, 33, 413– 416. Quaternary ice sheet: the 2500-km-long Norwegian– Svalbard
King, L. H. 1993. Till in the marine environment. Journal of Quaternary margin (578 – 808N). Geological Society of America Bulletin, 117,
Science, 8, 347–358. 1033– 1050.
King, L. H., Kokoengen, K., Fader, G. B. J. & Gunleiksrud, T. 1991. Paciullo, F. V. P., Ribeiro, A., Trouw, R. A. J. & Passchier, C. W.
Till-tongue stratigraphy. Geological Society of America Bulletin, 103, 2007. Facies and facies association of the siliciclastic Brak River
637– 659. and carbonate Gemsbok formations in the Lower Ugab River
Köppen, W. & Wegener, A. 1924. Die Klimate der geologischen Vorzeit. valley, Namibia, W. Africa. Journal of African Earth Sciences, 47,
Gebrüder Borntraeger, Berlin. 121– 134.
Kröner, A. & Rankama, K. 1972. Late Precambrian glaciogenic sedi- Peltier, W. R., Tarasov, L., Vettoretti, G. & Solheim, L. P. 2004.
mentary rocks in southern Africa: a compilation with definitions Climate dynamics in deep time: modeling the ‘snowball bifurcation’
and correlations. Precambrian Research Unit Bulletin, 11, 37, and assessing the plausibility of its occurrence. In: Jenkins, G. S.,
University of Cape Town, South Africa. McMenamin, M. A. S., McKey, C. P. & Sohl, L. (eds) The
Lythe, M. B. & Vaughan, D. G.the BEDMAP Consortium. 2001. Extreme Proterozoic: Geology, Geochemistry, and Climate. Geophy-
BEDMAP: a new ice thickness and subglacial topographic model sical Monograph 146, American Geophysical Union, Washington,
of Antarctica. Journal of Geophysical Research, 106, 11335–11351. DC, 107–124.
MacAyeal, D. R. 1993. Binge-purge oscillations of the Laurentide ice Pollard, D. & DeConto, R. M. 2007. A coupled ice-sheet/ice-shelf/
sheet as a cause of the North Atlantic’s Heinrich events. Paleoceano- sediment model applied to a marine margin flowline: forced and
graphy, 8, 775–784. unforced variations. In: Hambrey, M. J., Christoffersen, P.,
Macouin, M., Besse, J., Ader, M., Gilder, S., Yang, Z., Sun, Z. & Glasser, N. F. & Hubbard, B. (eds) Glacial Marine Processes and
Agrinier, P. 2004. Combined paleomagnetic and isotopic data Products. International Association of Sedimentologists Special
from the Doushantuo carbonates, South China: implications for the Publication No. 39, Blackwell Publishing, Malden, MA, 37– 52.
‘snowball Earth’ hypothesis. Earth and Planetary Science Letters, Pollard, D. & Kasting, J. F. 2004. Climate-ice sheet simulations of Neo-
224, 387– 398. proterozoic glaciation before and after collapse to Snowball Earth.
Maloof, A. C., Halverson, G. P., Kirschvink, J. L., Schrag, D. P., In: Jenkins, G. S., McMenamin, M. A. S., McKey, C. P. & Sohl,
Weiss, B. P. & Hoffman, P. F. 2006. Combined paleomagnetic, L. (eds) The Extreme Proterozoic: Geology, Geochemistry, and
isotopic and stratigraphic evidence for true polar wander from the Climate. Geophysical Monograph 146, American Geophysical
Neoproterozoic Akademikerbreen Group, Svalbard. Geological Union, Washington, DC, 91 –105.
THE OTAVI CARBONATE PLATFORM AND FORESLOPE 209

Powell, R. D. 1990. Glacimarine processes at grounding-line fans and Smit, J. M. 1962. Stratigraphy and metamorphism of the Otavi Series
their growth to ice-contact deltas. In: Dowdeswell, J. A. & southeast of Otavi, South West Africa. Transactions of the Geological
Scourse, J. D. (eds) Glacimarine Environments: Processes and Sedi- Society of South Africa, 65, 63 –78.
ments. Geological Society Special Publication No. 53, London, Soffer, G. 1998. Evolution of a Neoproterozoic continental margin
53 – 73. subject to tropical glaciation. BA thesis, Harvard College,
Powell, R. D. & Domack, E. 1995. Modern glacimarine environments. Cambridge, USA.
In: Menzies, J. (ed.) Glacial Environments, vol. 1: Modern Glacial Stanistreet, I. G. & Charlesworth, E. G. 1999. Damaran basement-
Environments: Processes, Dynamics and Sediments. Butterworth- cored fold nappes incorporating pre-collisional basins, Kaoko Belt,
Heinemann, Oxford, 445– 486. Namibia, and controls on Mesozoic supercontinental breakup. Econ-
Prave, A. R., Strachan, R. A. & Fallick, A. E. 2009. Global C cycle omic Geology Research Unit, Information Circular, 332, University
perturbations recorded in marbles: a record of Neoproterozoic of the Witwatersrand, Johannesburg, 14.
Earth history within the Dalradian succession of the Shetland Swart, R. 1992. Facies analysis of late Proterozoic carbonate turbidites in
Islands, Scotland. Journal of the Geological Society, London, 166, the Zerrissene Basin, Damara Orogen, Namibia. Journal of African
129– 135. Earth Sciences, 14, 283–294.
Pruss, S. B., Bosak, T., Macdonald, F. A., McLane, M. & Hoffman, Trindade, R. I. F., Font, E., D’Agrella-Filho, M. S., Nogueira,
P. F. 2010. Microbial facies in a Sturtian cap carbonate, the Rasthof A. C. R. & Riccomini, C. 2003. Low-latitude and multiple geomag-
Formation, Otavi Group, northern Namibia. Precambrian Research, netic reversals in the Neoproterozoic Puga cap carbonate, Amazon
181, 187– 108. craton. Terra Nova, 15, 441– 446, doi: 10.1046/j.1365-3121.2003.
Rodgers, J. 1957. The distribution of marine carbonate sediments: a 00510.x.
review. In: Le Blanc, R. J. & Breeding, J. G. (eds) Regional Vasconcelos, C., McKenzie, J. A., Warthmann, R. & Bernasconi, S.
Aspects of Carbonate Deposition. Society of Economic Paleontolo- M. 2005. Calibration of the d18O paleothermometer for dolomite
gists and Mineralogists (SEPM) Special Publication No. 5, Tulsa, precipitated in microbial cultures and natural environments.
Oklahoma, 2– 14. Geology, 33, 317– 320. doi: 10.1130/G20992.1.
le Roex, H. D. 1941. A tillite in the Otavi Mountains, S.W.A. Trans- Warren, S. G., Brandt, R. E., Grenfell, T. C. & McKay, C. P. 2002.
actions of the Geological Society of South Africa, 44, 207–218, 2 Snowball Earth: ice thickness on the tropical ocean. Journal of Geo-
plates, 1 map. physical Research, 107, doi: 10.1029/2001JC001123.
SACS (South African Committee for Stratigraphy) 1980. Stratigraphy of Williams, G. E. 1993. History of the Earth’s obliquity. Earth-Science
South Africa. Part 1 (comp. L.E. Kent): Lithostratigraphy of the Reviews, 34, 1 –45.
Republic of South Africa, South West Africa/Namibia and the repub- Witzke, B. J. 1990. Palaeoclimatic constraints for Palaeozoic palaeo-
lics of Bophuthatswana, Transkei and Venda. Handbook, Geological latitudes of Laurentia and Euramerica. In: McKerrow, W. S. &
Survey of South Africa, 8, 415– 437. Scotese, C. R. (eds) Palaeozoic Palaeogeography and Biogeo-
Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal graphy. Geological Society of London Memoir, 12, 57 –73.
of existing models and the plumeworld hypothesis. Terra Nova, 17, Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003.
299– 310. Systematic variations in C, O, and Sr isotopes and elemental concen-
Shields, G. A., Deynoux, M., Strauss, H., Paquet, H. & Nahon, D. trations in Neoproterozoic carbonates in Namibia: implications for
2007. Barite-bearing cap dolostone of the Taoudéni Basin, northwest a glacial to interglacial transition. Precambrian Research, 124,
Africa: sedimentary and isotopic evidence for methane seepage after 69– 85.
a Neoproterozoic glaciation. Precambrian Research, 154, 209– 235. Ziegler, A. M., Hulver, M. L., Lottes, A. L. & Schmachtenberg,
Shoemaker, E. M. 1963. Impact mechanics at Meteor Crater, Arizona. W. F. 1984. Uniformitarianism and paleoclimates: inferences from
In: Middlehurst, B. M. & Kuiper, G. P. (eds.) The Moon Meteorites the distribution of carbonate rocks. In: Brenchley, P. J. (ed.)
Comets 4. University of Chicago Press, Chicago, 301– 336. Fossils and Climate. John Wiley & Sons, New York, 3 –25.
Chapter 15

The Witvlei Group of East-Central Namibia

A. R. PRAVE1*, K.-H. HOFFMANN2, W. HEGENBERGER2 & A. E. FALLICK3


1
Department of Earth Sciences, University of St Andrews, St Andrews KY16 9AL, UK
2
Geological Survey of Namibia, 1 Aviation Road, Windhoek, Namibia
3
Scottish Universities Environmental Research Centre, East Kilbride G75 0QF, UK
*Corresponding author (e-mail: ap13@st-andrews.ac.uk)

Abstract: The Witvlei Group is preserved in two regional synclinoria in the Gobabis-Witvlei area of east-central Namibia and as isolated
outcrops 90 km SW of Rehoboth, itself some 200 km south of that area. It consists of mixed, coarse- to fine-grained siliciclastic and
carbonate strata deposited in deep- to shallow-marine, and locally non-marine, settings along the post-rift continental margin of the
Kalahari Craton prior to the onset of foreland basin sedimentation recorded by the overlying terminal Neoproterozoic–Cambrian
Nama Group. No direct age constrains exist for the Witvlei Group, but it post-dates c. 800 Ma rift-related rocks and pre-dates the
c. 548 Ma base of the Nama Group, thereby placing it as Cryogenian to Ediacaran in age.
The Witvlei Group consists of three main units, from oldest to youngest, the Blaubeker, Court and Buschmannsklippe Formations. The
Blaubeker Formation is highly variable in thickness and can be as much as 1000 m thick. It consists mostly of massive, polymict diamictite
and, in the area of the type locality, contains conglomerate and pebbly sandstone beds. The diamictic strata combined with the presence of
numerous faceted and striated clasts provide the evidence for glaciogenic influences on sedimentation. The highly variable thickness
pattern likely reflects the infill of palaeo-valleys formed by the deep erosion and scouring of bedrock by ice, and the conglomerates
and pebbly sandstones record glacial outwash processes. The Tahiti Formation is a locally developed, fine-grained sandstone above
the Blaubeker Formation. It is poorly exposed and its exact stratigraphic relationship to the Blaubeker rocks and overlying Court Formation
remains to be determined.
The Blaubeker rocks are overlain sharply by the basal unit of the Court Formation, the Gobabis Member. This Member is from
20 to 60 m thick and consists mostly of dark and light grey laminated dolostones that display a d13Ccarbonate profile that rises
from values of 24‰ in the lowermost beds to values of 5‰ in the topmost. The Gobabis Member is conformably overlain by the
shales, marls and thin limestones of the Constance Member followed by quartzites of the uppermost unit of the Court Formation, the
Simmenau Member.
The basal unit of the Buschmannsklippe Formation is the light to tan and pink grey dolostone of the Bildah Member. Its basal contact
is sharp everywhere, and it is gradationally overlain by a coarsening (shoaling) upward succession from shales, thin limestones (some
exhibiting formerly aragonitic fans) and fine sandstones of the La Fraque Member, to interbedded quartzites and stromatolitic and
cherty dolostones of the Okambara Member. The d13Ccarbonate profile for the Buschmannsklippe rocks shows that the basal beds of the
Bildah Member begin at –4‰, followed by a decline to –6‰ in the lower La Fraque limestones and then a rise to –3‰ in the dolostones
of the Okambara Member before being truncated by the base of the regionally unconformably overlying basal Weissberg Quartzite
Member of the lower Nama Group. Although no glacial sediments have been recognized below the Bildah Member, its lithofacies char-
acter, stratigraphic position and C-isotopic profile are compatible with and strikingly similar to younger Cryogenian cap carbonates. Thus,
the Witvlei Group arguably contains both the older and younger cap carbonates of Neoproterozoic time, but only the older Cryogenian
glacial deposit.

Glaciogenic rocks of the Witvlei Group form part of a mixed carbonates’ are present (Fig. 15.2). The older of the two is the
siliciclastic and carbonate sedimentary succession of Late Neopro- Gobabis Member of the Court Formation; it caps the Blaubeker
terozoic (Cryogenian – early Ediacaran) age that is preserved along diamictite. The younger one is the basal unit of the Buschmannsk-
the NW margin (present-day coordinates) of the Kalahari Craton lippe Formation, the Bildah Member; it sits sharply on and trans-
(Fig. 15.1). It averages 200 –400 m in thickness (but has a compo- gresses a variable substrate of the Court Formation and, where
site thickness of more than 1000 m) and rests unconformably on a absent, older Kamtsas and Eskadron-Doornpoort Formation
varied substrate comprised of thick sequences of pre-glacial, early – rocks, but no glaciogenic rocks have been recognized.
middle Neoproterozoic (Tonian – early Cryogenian) rift-related
arkosic sandstones, conglomerates and argillites of the Eskadron
and Doornpoort Formations (Tsumis Group) and the Kamtsas For- Structural framework
mation (Nosib Group) and, locally, older Palaeo/Mesoproterozoic
metasedimentary and igneous rocks. The Witvlei Group records The Witvlei Group is preserved in two NE– SW trending regional
sedimentation in a range of shallow- to deep-marine and locally synclinoria located within the southern foreland zone of the
non-marine settings that developed following rifting and antece- Damara Belt in east-central Namibia (Fig. 15.1; Hegenberger
dent to the onset of foreland basin development represented by 1993). The western one, the Witvlei Synclinorium, averages
the unconformably overlying terminal Neoproterozoic (late Edia- about 20 km in width and is 180 km in length. It comprises
caran) – Cambrian Nama Group. Hegenberger (1993) provides the several open to tight second- and numerous third-order folds,
most complete description and synthesis of the Witvlei Group and exposures are good along many of the moderately to steeply
rocks, while Hoffmann (1989) presents lithostratigraphic corre- dipping limbs. The eastern structure is termed the Gobabis Syncli-
lations of the Witvlei Group succession within a regional strati- norium and covers an area of c. 80 km by 100 km; dips are rela-
graphic framework encompassing the Naukluft Nappe Complex, tively shallow and exposure is poor. Minor faulting disrupts the
the southern and northern Damara Belt (including the Otavi stratigraphy in both synclinoria, but displacements are small and
Group) and the Gariep Belt of southern Namibia. typically no more than a few tens of metres. Several larger steep
Only one glaciogenic unit has been recognized in the Witvlei faults are associated with the Witvlei Synclinorium but they do
Group, the diamictic Blaubeker Formation, but two ‘cap not compromise the stratigraphic coherence of the succession.

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 211– 216. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.15
212 A. R. PRAVE ET AL.

Fig. 15.1. Generalized geological map of


the southern Damara foreland showing the
distribution of Witvlei Group rocks in the
Witvlei-Gobabis area (after Hegenberger
1993). Inset map shows location of main
figure and the major tectonic regions and
sedimentary basins. B, Blaubeker; GB,
Gariep Belt; NNC, Naukluft Nappe
Complex; R, Rehoboth.

Folding was associated with very low-grade metamorphism, Damaran structural front, defined by thrust faults involving pre-
decreasing from NW to SE, and is attributed to the late Precam- Witvlei Group rocks including crystalline basement. To the east,
brian –Cambrian Damaran orogeny (Miller 1983). south and SW, the Witvlei Group rocks are buried beneath Phaner-
The original extent of the Witvlei Group is not known but was ozoic strata (Karoo and Kalahari Groups) and how far they extend
likely much greater than the currently preserved area of outcrop. under that cover sequence is not known, although drill holes some
To the NW, the Witvlei Group outcrop belt is truncated by the 40 km south of the southwestern tip of the Witvlei Synclinorium
have intersected Witvlei Group rocks (Hegenberger 1993).
Further SW, within the western part of the southern Damara fore-
land, glacial sediments of the Blaubeker Formation are preserved
as local erosional remnants on Blaubeker farm, the designated
type locality (Schalk 1970) some 90 km SW of Rehoboth, with
thick strata of the Nama Group of the northern Nama (Zaris)
basin resting directly on the older Neoproterozoic Tsumis and
Nosib Group rift-related sediments, and, still further west,
Palaeo- and Mesoproterozoic basement. Absence of rocks equival-
ent to the Court and Buschmannsklippe Formation is either
due to non-deposition or erosional removal prior to deposition of
the Nama.
In the Gobabis-Witvlei area, the lower units of the Witvlei
Group, the Blaubeker and Court Formation, display lateral thick-
ness and facies variations and a major intraformational unconfor-
mity at the base of the Simmenau Member that has been
interpreted as evidence for influence by synsedimentary tectonism
(Hegenberger 1993). No unequivocal basin-bounding faults and/
or growth structures have been identified, but the Nina Anticline
(which separates the Gobabis and Witvlei Synclinoria) was
actively up-warped during this time, resulting in erosion of
lower units of the Court Formation across the crest of that struc-
ture. The upper unit of the Witvlei Group, the Buschmannsklippe
Formation, exhibits systematic thickness variations and similar
facies characteristics throughout the region, and displays an
overall westward thickening (Hegenberger 1993). This suggests
that sedimentation occurred during a phase of relatively uniform
subsidence.

Fig. 15.2. Stratigraphy of the Witvlei Group (after Hoffmann 1989; Stratigraphy
Hegenberger 1993). Glaciogenic rocks (Blaubeker Formation) and cap
carbonates (Gobabis and Bildah Members) are highlighted in bold and with The rocks that make up the Witvlei Group, including the locally
shaded ornament. See text for discussion. preserved diamictite beds in the Gobabis-Witvlei area, were first
THE WITVLEI GROUP 213

described collectively as the Buschmannsklippe Formation by (1993) subdivided the Okambara into an upper and lower
Martin (1965) and regarded for many years as a local, northeastern dolostone-dominated unit separated by a quartzite-dominated unit.
facies of the Nama Group (e.g. Hegenberger & Seeger 1980). The top of the Witvlei Group is marked by the regional uncon-
Since then, lithostratigraphic subdivision and nomenclature of formity at the base of the Weissberg Member (Dabis Formation) of
the succession has undergone several revisions, the most important the lower Nama Group.
of which is the recognition that the lower and middle parts of the
original Buschmannsklippe Formation are older than the Nama
Group, which led to the establishment of the Witvlei Group as a Glaciogenic deposits and associated strata
major new division for these units (Hoffmann 1989; see Hegenber-
ger 1993 for a review). Here we adopt the subdivision by Hegen- The Blaubeker Formation
berger (1993), but include the Blaubeker Formation at the base
as recommended by Hoffmann (1989). The Witvlei Group thus No detailed modern sedimentological investigations of the Blaube-
consists of three main units, the Blaubeker, Court and Busch- ker Formation have been carried out, and the following summary is
mannsklippe Formations (Fig. 15.2). The Blaubeker Formation is based on descriptions by Martin (1965), Schalk (1970), Kröner &
patchily preserved, the Court Formation is most complete in the Rankama (1972), Hegenberger & Seeger (1980) and field obser-
Gobabis Synclinorium, and the Buschmannsklippe Formation is vation by the authors. In the type locality and adjacent areas
best developed and exposed in the Witvlei Synclinorium. west of Rehoboth, exposures are patchy, but show that the Blaube-
The Blaubeker Formation is characterized by polymict diamic- ker Formation consists of a lower unit of pebbly quartzite with rare
tite. At the type locality on Blaubeker farm and surrounding areas, shale interbeds. This unit is several tens of metres thick and is
it overlies the Kamtsas Formation with an angular unconformity sharply overlain by massive polymict diamictite between 200
and consists of a basal unit of interbedded conglomerate and and 300 m thick. In the Gobabis-Witvlei area, the Blaubeker con-
quartzite followed by massive polymict diamictite, which is sists entirely of massive, non-graded, non-stratified, polymict dia-
unconformably overlain by the Nama Group; both the Court and mictite; the best exposures are at Farm Tahiti in the Black Nossob
Buschmannsklippe Formations are absent. Total thickness varies River valley. In both synclinoria, clasts are dispersed in a grey-
between 250 and 300 m (Schalk 1970). In the Gobabais-Witvlei green to reddish-grey, fine siliciclastic matrix; the matrix is
area, the Blaubeker Formation is present in two main outcrop massive and no sandstone lenticules or laminae have been
areas where it unconformably overlies the Kamtsas Formation. reported. Clasts typically range between 0.05 and 0.5 m in the
The thickest and best-developed exposures are in the south-central longest dimension, but can be as large as 1 m. They consist
portions of the Gobabis Synclinorium (particularly in the valleys of mainly of quartzite, lesser felsic and mafic volcanics, granitoids,
the White and Black Nossob Rivers) and it is also present along the gneiss and schist. Shapes vary from angular to sub-rounded and
easternmost edge of the south-central flanks of the Witvlei Syncli- many of the clasts display variably orientated striations and facet-
norium (Hegenberger 1993). In these areas the Blaubeker For- ing. This is the primary observational evidence to infer a glacially
mation is composed entirely of massive polymict diamictite and influenced origin for the Blaubeker diamictite. At Gobabis, a 0.1 –
is highly variable in thickness, ranging from as little as 10 m to 0.5-m-thick, crumbly weathering, poorly sorted, massive brown
as much as 1000 m (Hegenberger & Seeger 1980). sandstone containing dispersed quartzite clasts and cobbles
The Court Formation consists of four units (Hegenberger occurs between the underlying Kamtsas Formation and the sharp
1993). The oldest is the Tahiti Member, a 1– 10-m-thick fine- base of the Gobabis Member. This sandstone may be a facies of
grained feldspathic arenite. This member is poorly exposed and the Blaubeker Formation, but this remains to be proven.
few sedimentological details exist. It is restricted in its occurrence
to a few outcrops along the extreme east-southeastern edge of
the Gobabis Synclinorium (and may well be a facies of the The Doornpoort/Eskadron and Kamtsas Formation
Blaubeker Formation, but this remains to be proven). The first, well-
developed and laterally extensive unit of the Court Formation is These rocks are pre-glacial in age but are briefly described here
the Gobabis Member. It mostly consists of grey laminated because many of the clasts in the Blaubeker diamictites were
dolostones and local stromatolitic limestone, and varies between reworked from these older units. All three formations can reach
20 and 60 m in thickness. The youngest units of the Court Formation thicknesses measured in many thousands of metres and are domi-
are the Constance Member, a recessive, poorly exposed 100– nated by feldspathic quartzite, pebbly quartzite and arenite and
200-m-thick succession of pale-coloured shale, marl and thin- local conglomerate. The conglomeratic portions contain clasts
bedded limestone, and the Simmenau Member, a 100 –150-m- ranging in size from pebbles to boulders that consist of quartz,
thick fine- to medium-grained quartzite to feldspathic arenite. quartzite, felsic and mafic volcanic rocks, granites and granitic
The Buschmannsklippe Formation is divided into three units, gneiss (Hegenberger & Seeger 1980).
from oldest to youngest, the Bildah, La Fraque and Okambara
Members (Hegenberger 1993). The Formation as a whole defines
a west-SW-thickening wedge, from about 50 m thick in the east- The Court Formation
ernmost sections of the Gobabis Synclinorium to more than
300 m thick in the Witvlei Synclinorium. The Bildah Member is Resting sharply on the Blaubeker Formation is the Gobabis
typically 40– 60 m thick and consists of light grey to pale tan Member of the Court Formation (locally, the Tahiti Formation
and pink dolostone. Its base defines a regional unconformity that sandstones occur on top of the Blaubeker rocks). In the vicinity
oversteps the underlying Court Formation to rest unconformably of the type locality, the Gobabis Member is 50–60 m thick and
on older Neoproterozoic rocks. Overlying the Bildah is the La consists of two main units. The lower unit is of variable thickness
Fraque Member, a succession of interbedded shale, thin-bedded (10 –30 m) and characterized by thin-bedded, planar- to wavy-
limestone and fine sandstone. It thickens to the west and reaches parallel laminated dolomicrite. The centimetre-thick laminae alter-
a maximum thickness of c. 150 m in the central portions of the nate between dark grey and lighter grey in colour, giving the
Witvlei Synclinorium. The Okambara Member, like the La Gobabis Member a distinctive striped pattern. In places, syn-
Fraque, similarly thickens westward attaining a maximum thick- sedimentary (or early lithification) disruption of bedding has
ness of about 125 m. It is characterized by cream to pink-grey resulted in locally formed intraformational breccias; these,
coloured commonly stromatolitic dolostone and cherty dolostone however, are rare. Much of the remainder of the Member is com-
and subordinate fine, white quartzite. In the thicker sections in posed of similarly dark –light coloured dolomicrite, but the lami-
the central portions of the Witlvlei Synclinorium, Hegenberger nation is crinkly-wrinkly in character and exhibits microbial
214 A. R. PRAVE ET AL.

roll-up structures. Columnar stromatolites are developed locally, have expanded upon the C-isotopic work and the following is a
and some of these are in limestone rather than dolostone. preliminary description of new, unpublished, data from the
carbonate-bearing units of the Witvlei Group by the authors. A
total of 193 stable isotopic analyses have been carried out, but
The Buschmannsklippe Formation additional work is required before more robust trends can be con-
structed, especially for the less well exposed interglacial units
At its type locality, the Bildah Member rests with sharp contact on The C-isotopic profile of the older cap carbonate, the Gobabis
quartzites of the Kamtsas Formation and consists of four units, Member, displays a progressively rising trend: the basal layer
from the base upward these are: (i) light grey to tan coloured, has d13Ccarbonate values of –4‰ to – 3‰ and these rise to –2‰
flat to wavy-parallel laminated dolomicrite containing locally to –1‰ in the overlying thin-bedded and laminated dolomicrites.
developed soft-sediment slumping and irregularly shaped The contact between the lower planar- to wavy-laminated unit and
calcite-spar-filled vugs; (ii) microbially laminated to stromatolitic the upper crinkly-wrinkly laminated unit with roll-up structures is
light grey dolomicrite with quartz- and calcite-spar-filled vugs and everywhere covered and/or mantled with calcrete, but the lowest
tubes (developed in variable densities); (iii) fine-grained dolo- exposed beds of the upper unit begin with d13Ccarbonate values of
stone, light grey in colour, exhibiting cross-bedding and flat lami- 5‰, and these stay between 3‰ and 5‰ to the top of the
nation; and (iv) thin-bedded to laminated, fine-grained, grey Gobabis Member. Because the contact is covered, we are uncertain
dolostone with mudstone partings. The Bildah cap dolostone if the rise to positive values is gradational or abrupt.
varies in thickness from 60 to 100 m, and much of this variability The inter-glacial succession consists of the Constance and
is due to the lesser or greater development of the stromatolitic unit. Simmenau Members of the Court Formation. The former is
Gradationally overlying the Bildah dolostones is the reddish- poorly exposed and the latter does not contain carbonate rocks,
brown shale and thin-bedded limestone of the lower part of the so we have few data on these units and more work needs to be
La Fraque Member; some of the limestones exhibit centimetre- carried out before anything substantial can be stated regarding
scale aragonite fan structures. These beds pass transitionally their chemostratigraphy.
upward into tan-brown coloured, cross-bedded and hummocky The basal laminated unit of the Bildah cap dolostone consistently
cross-stratified, commonly calcareous-cemented, sandstone marking has d13Ccarbonate values around –4‰; these rise slightly to values
the upper part of the La Fraque Member. Gradationally overlying the around –3‰ in the middle and upper units of the cap dolostone.
La Fraque rocks are the light-coloured quartzites and cream-coloured At the contact with, and through the transition into, the La Fraque
dolostones of the Okambara Member. The Okambara rocks contain a Member, C-isotopic values decline to –5‰ to –6‰ through the
variety of sedimentary features including crinkly-wrinkly laminated limestone rhythmite and thin-bedded limestone units (it is note-
dolomicrite, columnar, branching stromatolites, as well as laterally worthy that at this stratigraphic level, fans considered to be formerly
linked heads, imbricated (edgewise)-, wave-rosette- and flat-pebble aragonitic in composition are present). Above this and through
intraclastic limestones and dolostones, cross-bedded dolograin- the Okambara Member, the d13Ccarbonate profile rises to values
stones, rare oolitic dolostone, chicken-wire-fabric quartz nodules, between –3‰ and –2‰ before being truncated by the basal uncon-
cauliflower chert and enterolithic bedding. The tops of some beds formity of the overlying Weissberg Member of the Nama Group.
exhibit desiccation cracks.

Other characteristics (e.g. economic deposits, biomarkers)


Boundary relations with overlying and underlying
non-glacial units No biomarker data have been obtained on the Witvlei Group rocks,
nor are these rocks known to contain any major deposits or miner-
The Blaubeker Formation everywhere rests unconformably on the alizaton of economic importance. The Bildah Member has been
Kamtsas Formation. The highly varied thickness pattern of the quarried locally for dimension stone.
Blaubeker (from 10 m or less to more than 1000 m) is evidence
that its base must be deeply erosive. In the area of the eponymous
type locality, strata of the lower Nama Group (Dabis Formation) Palaeolatitude and palaeogeography
rest unconformably on the Blaubeker rocks. In the Gobabis-
Witvlei area, with the exception of where the areally restricted Kröner et al. (1980) obtained palaeomagnetic data from three sites
Tahiti Member is present, the diamictic strata of the Blaubeker of the Blaubeker Formation at its type locality. The data indicated
Member are everywhere sharply capped by the basal laminated that the magnetization post-dated folding and was attributed to
dolostone of the Gobabis Member. The Gobabis Member is over- thermal events associated with Damaran orogenesis. No other
lain by the shales and thin limestones of the Constance Member, palaeomagnetic data exist on the Witvlei Group rocks, but it is gen-
but the contact between these two Members is poorly exposed erally assumed that they, as part of the Kalahari craton, were
and commonly obscured by calcrete, so we are uncertain if it is deposited in low-latitudinal positions during most of the Cryogen-
sharp or gradational. ian (e.g. Li et al. 2008).
The base of the Bildah Member rests unconformably on a variable
substrate consisting of the Court and/or Kamtsas or Eskadron-
Doortpoort Formations and is everywhere sharp. The contact Geochronological constraints
between the Bildah Member and the overlying La Fraque Member
is gradational over a few metres; the topmost several metres of the No direct geochronological data exist on the Witvlei Group rocks.
Bildah Member exhibit an upward increase in the proportion and The Witvlei Group rests unconformably on rocks that are correla-
thickness of mudstone partings concomitant with a thinning tive with those elsewhere in central and southern Namibia con-
upward and proportional decrease in dolostone beds. sidered as c. 1000– 800 Ma in age; it is overlain unconformably
by the late Ediacaran – Cambrian Nama Group (Hegenberger
1993). Age constraints are good for the lower part of the Nama
Chemostratigraphy Group and an ash bed within the Zaris Formation in the main
Nama basin to the SW has yielded a U – Pb zircon age of 548 Ma
Kaufman et al. (1991, 1997) Saylor et al. (1998) and Kennedy et al. (Grotzinger et al. 1995). These broad constraints show that the
(1998) did reconnaissance C-isotopic studies on the Witvlei Group Witvlei Group is entirely Neoproterozoic (Cryogenian to Edia-
rocks and Gorjan et al. (2003) performed S-isotopic analyses. We caran) in age.
THE WITVLEI GROUP 215

The lithological and chemostratigraphic similarity of the Bildah the most extensively and consistently developed (e.g. Hoffman
Member to the Keilberg cap dolostone (Hoffmann 1989; Kennedy & Schrag 2002; Fairchild & Kennedy 2007). Since then, the C-
et al. 1998) can be used to infer an age of 635 Ma, that is, the isotopic profile and overall facies of the Bildah Member has
timing of meltback associated with the Marinoan-equivalent been shown to be identical to those known to cap the younger
glacial and cap-carbonate rocks in northern Namibia and southern Cryogenian glaciation elsewhere and is now interpreted as being
China (Hoffmann et al. 2004; Condon et al. 2005). The post- equivalent to the Keilberg cap dolostone of the Otavi Group in
Marinoan d13C crossover to positive values (e.g. Halverson et al. northern Namibia (Hoffmann & Prave 1996; Kennedy et al.
2005) is not recorded in the Buschmannsklippe C-isotopic data, 1998; Hoffman & Schrag 2002). The overlying La Fraque –
most likely due to erosional truncation along the unconformity Okambara succession records the post-glacial transgression
surface defining the base of the Nama Group. If so, then the time through shoaling sequence: from the deeper-marine settings of
gap between the base of the Nama Group and top of the Witvlei the lower La Fraque (interbedded shale and originally aragonite-
Group could be considerable and on the order of many tens of fan-bearing limestone rhythmites), to storm-dominated inner
millions of years. shelf settings (hummocky cross-stratified sandstones) of the
Direct age constraints are also lacking for the base of the Witvlei upper La Fraque, to shallow subtidal and intertidal (quartz-
Group, but can be inferred from the proposed correlation of the replacement evaporite fabrics and desiccated microbial laminites)
Gobabis Member and Blaubeker Formation with the Rasthof cap depositional settings of the Okambara rocks (Hegenberger 1993).
carbonate and Chuos Formation of northern Namibia (Hoffmann
K.-H. H. and W. H. thank the Geological Survey of Namibia for support. A. R. P.
1989; Kennedy et al. 1998), which overlies felsic volcanic rocks and A. E. F. acknowledge NERC and The Carnegie Trust for The Universities of
with a U –Pb zircon age of 746 Ma (Hoffman et al. 1996). A still Scotland for support. This represents a contribution of the IUGS- and UNESCO-
younger age has been suggested as c. 710 Ma (Halverson et al. funded IGCP (International Geoscience Programme) Project #512.
2005). It is reasonable, then, to infer a similar age for the
Blaubeker-Gobabis glacial –cap couplet. If this is correct, then
the Witvlei Group likely spans some 100 Ma in time, from the References
Cryogenian through to at least the early –middle Ediacaran.
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005.
U–Pb ages from the Neoproterozoic Doushantuo Formation, China.
Discussion Science, 308, 95– 98.
Fairchild, I. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the
The Blaubeker Formation is the only unit in the Witvlei Group suc- Earth System. Journal of the Geological Society, London, 164,
cession that can be attributed to recording glacial influences on 895– 921.
sedimentation in the form of diamictic rocks containing striated Gorjan, P., Walter, M. R. & Swart, R. 2003. Global Neoproterozoic
and faceted clasts. It occurs in narrow outcrop belts displaying (Sturtian) post-glacial sulfide – sulfur isotope anomaly recognized in
highly variable thicknesses, and these features most likely are Namibia. Journal of African Earth Sciences, 36, 89 –98.
the result of glacial scouring and carving of valleys and channels. Grotzinger, J. P., Bowring, S., Saylor, B. Z. & Kaufman, A. J. 1995.
The pebbly sandstones and conglomerates associated with the Biostratigraphic and geochronologic constraints on early animal
diamictites represent glacial outwash. evolution. Science, 270, 598– 604.
The Blaubeker is sharply capped by the Gobabis Member dolo- Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
Rice, A. H. N. 2005. Towards a Neoproterozoic composite carbon
stones. The overall facies development of this Member is identical
isotope record. Geological Society of America Bulletin, 117,
to that documented and described from the Rasthof Formation, the
1181–1207.
older of the two cap carbonates in the Otavi succession of northern Hegenberger, W. 1993. Stratigraphy and sedimentology of the Late
Namibia (Hoffmann 1989; Hoffmann & Prave 1996; Kennedy Precambrian Witvlei and Nama Groups, East of Windhoek.
et al. 1998; Hoffman & Schrag 2002). This has been used in com- Geological Survey of Namibia, Memoir 17, 82.
bination with C-isotopic data to correlate the Gobabis dolostones Hegenberger, W. & Seeger, K. G. 1980. The geology of the Gobabis
and the Blaubeker diamictite with the Rasthof Formation and area. Explanation of Sheet 2218, scale 1:250,000. Geological
Chuos Formation diamictite (Hoffmann 1989; Kennedy et al. Survey of South West Africa/Namibia Windhoek, 11.
1998; Hoffman & Schrag 2002). The overlying Constance and Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
Simmenau Members form the inter-glacial package and are testing the limits of global change. Terra Nova, 14, 129– 155.
typified by fine-grain sizes, flat lamination, low-angle cross- Hoffman, P. F., Hawkins, D. P., Isachsen, C. E. & Bowring, S. A. 1996.
stratification and current ripples. Rare desiccation cracks are Precise U– Pb zircon ages for early Damaran magmatism in the
preserved in the Constance Member and the Simmenau Member Summas Mountains and Welwitschia Inlier, northern Damara belt,
contains mud-chip conglomerates in the base of broad, shallow- Namibia. Communications of the Geological Survey of Namibia,
channelled beds. These features have been used to infer that the 11, 47– 52.
Constance Member was deposited in a low-energy intermittently Hoffmann, K. H. 1989. New aspects of lithostratigraphic subdivision and
exposed, shallow-water setting and that the Simmenau Member correlation of late Proterozoic to early Canbrianrocks of the southern
records a distal, fine-grained braidplain (Hegenberger 1993). Com- Damara Belt and their correlation with the central and northern
bined, the Court Formation records a post-glacial transgression Damara Belt and the Gariep Belt. Communications of the Geological
from deeper marine (Gobabis Member) through shoaling into shal- Survey of Namibia, 5, 59 –67.
Hoffmann, K. H. & Prave, A. R. 1996. A preliminary note on a revised
lower marine (Constance Member) settings.
subdivision and regional correlation of the Otavi Group based on gla-
Nowhere along the outcrop belt of the Witvlei Group have rocks
ciogenic diamictites and associated cap dolostones. Communications
been recognized as glacial in origin beneath the Bildah Member. of the Geological Society of Namibia, 11, 81 –86.
Based on striking similarities in lithology and facies, it was pro- Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
posed to be equivalent to the characteristic pink-grey dolostones 2004. A U– Pb zircon date from the Neoproterozoic Ghaub For-
that cap diamictites in the Naukluft Nappe Complex and the mation, Namibia: constraints on Marinoan glaciation. Geology, 32,
Gariep Belt (Hoffmann 1989). This established, for the first 817– 820.
time, that the Bildah, even given the absence of an underlying dia- Kaufman, A. J., Hayes, J. M., Knoll, A. H. & Germs, G. J. B. 1991. Iso-
mictite, was the same (cap) carbonate unit overlying a regional topic compositions of carbonates and organic carbon from upper Pro-
glacial unconformity surface. This is similar to what is now recog- terozoic successions in Namibia: stratigraphic variation and the
nized from many places worldwide where glacial rocks are only effects of diagenesis and metamorphism. Precambrian Research,
patchily preserved and it is the overlying cap carbonates that are 49, 301– 327.
216 A. R. PRAVE ET AL.

Kaufman, A. J., Hayes, J. M., Knoll, A. H. & Narbonne, G. M. Li, Z. X., Bogdanova, S. V. et al. 2008. Assembly, configuration, and
1997. Isotopes, ice ages, and terminal Proterozoic earth break-up history of Rodinia: a synthesis. Precambrian Research,
history. Proceedings of the National Academy of Science, 94, 160, 179– 210.
6600– 6605. Martin, H. 1965. The Precambrian Geology of South West Africa and
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K. H. & Namaqualand. Precambrian Research Unit, University of Cape
Arthur, M. 1998. Two or four Neoproterozoic glaciations? Town, Cape Town.
Geology, 26, 1059–1063. Miller, R. McG. 1983. The pan-African Damara Orogen of South West
Kröner, A., McWilliams, M. O., Germs, G. J. B., Reid, A. B. & Schalk, Africa/Namibia. In: Miller, R. McG. (ed.) Evolution of the Damara
K. E. L. 1980. Paleomagnetism of late Precambrian to early Paleozoic Orogen of Sout West Africa/Namibia. Special Publication of the Geo-
mixtite-bearing formations in Namibia (South West Africa): the logical Survey of South Africa, 11, 431– 515.
Nama Group and Blaubeker Formation. American Journal of Saylor, B. Z., Kaufman, A. J., Grotzinger, J. P. & Urban, F. 1998.
Science, 280, 942–968. A composite reference section for terminal Proterozoic strata of
Kröner, A. & Rankama, K. 1972. Late Precambrain glaciogenic sedi- southern Namibia. Journal of Sedimentary Research, 68, 1223– 1235.
mentary rocks in southern Africa: a compilation with definitions Schalk, K. E. L. 1970. Some late Precambrian formations in central South
and correlations. Bulletin of the Precambrian Research Unit, Uni- West Africa. Annals of the Geological Survey of South Africa, 8,
versity of Cape Town, 11, 37. 29 – 47.
Chapter 16

The Chameis Gate Member, Chameis Group, Marmora Terrane, Namibia

HARTWIG E. FRIMMEL1,2
1
Department of Geological Sciences, University of Cape Town, Rondebosch 7701, South Africa
2
Present address: Geodynamics & Geomaterials Research Division, Institute of Geography and Geology, University of Wuerzburg,
Am Hubland, D-97074 Wuerzburg, Germany
(e-mail: hartwig.frimmel@uni-wuerzburg.de)

Abstract: The Chameis Gate Member is a poorly exposed and poorly investigated diamictite in the Chameis Subterrane of the Marmora
Terrane, which forms the western, completely allochthonous part of the Pan-African Gariep Belt (southwestern Namibia). Its significance
lies in its position in an entirely oceanic unit, the Dernburg Formation, which is dominated by mafic volcanic rocks. The diamictite con-
tains exotic dropstones in a mafic volcaniclastic matrix, thus providing evidence for transport by ice away from the continental margin
into an oceanic environment. No direct age data are available and stratigraphic relationships are obscured by limited outcrop and intense
syn-orogogenic deformation. Preliminary chemostratigraphic data obtained on carbonate rocks below and above the diamictite, impre-
cise Pb–Pb age data on the largely volcaniclastic silicate fraction within associated stromatolitic reef carbonates, and imprecise Ar–Ar
data on early hornblende related to sea-floor metamorphism of the associated volcanic rocks all point to an age loosely constrained
between 640 and 580 Ma. Based on a comparison between the tectono-stratigraphic units of the Marmora Terrane with the continental
Port Nolloth Group on the one side and the Rocha Group of the Punta del Este Terrane in Uruguay on the other side of the terrane, it is
suggested that the diamictite was deposited in a back-arc basin that developed in response to the 640 –590 Ma volcanic arc of the Dom
Feliciano Belt in southeastern Brazil and eastern Uruguay.

The Neoproterozoic Marmora Terrane is an internal, entirely onto the para-autochthonous continental Port Nolloth Zone along
allochthonous tectonic unit in the western part of the Gariep Belt the Schakalsberge Thrust (Fig. 16.1). The Marmora Terrane has
(Fig. 16.1). Although the Marmora Terrane stretches along the been subdivided into three tectono-stratigraphic units, the Scha-
Atlantic coast of Namibia from 27.468S southwards into South kalsberge, Oranjemund and Chameis complexes (Hartnady &
Africa to 28.768S, just south of the Orange River mouth, exposures Von Veh 1990) or subterranes (Frimmel 2000b). The tectonically
of the Chameis Gate Member are restricted to a narrow coastal lowest of these units, the Schakalsberge Subterrane, consists
strip in the so-called Sperrgebiet (‘restricted area’) of southwestern mainly of mafic metavolcanic rocks of the Grootderm Formation
Namibia. As all outcrops of this member are effectively located with a dolomitic, in places stromatolitic to oolitic carbonate on
within the mining lease area of the alluvial diamond operations top, interpreted as a former guyot that evolved from an oceanic
of Namdeb Pty Ltd. along the coast NW of Oranjemund, access seamount (Frimmel et al. 1996). The tectonically intervening
to these outcrops is restricted. Moreover, the outcrops are in a pre- Oranjemund Subterrane consists of predominantly siliciclastic,
dominantly sandy desert devoid of infrastructure, and much of the commonly turbiditic, metasedimentary rocks, with only very minor
rocks of the Chameis Gate Member are covered by shifting sand carbonate and ferruginous chert, originally referred to as the Oran-
dunes. As a consequence, this unit is only very poorly investigated. jemund Formation (Frimmel 2000b) but since raised to the rank of
The area was first mapped by Kaiser (1926), but he did not dis- group (Basei et al. 2005). The Chameis Subterrane, which litho-
tinguish the diamictite that characterizes the Chameis Gate stratigraphically comprises the Chameis Group and contains the
Member. The largest exposures of the member can be found on Chameis Gate Member, is the tectonically highest unit. It is inten-
wind-blown surfaces to the west and south of a security gate, sely deformed, resembling a tectonic melange zone, which makes
named Chameis Gate, at 27.8378S, 15.7278E. Minor outcrops a stratigraphic subdivision difficult. The lower part consists of pre-
can be found at Bakers Bay at 27.6558S, 15.5368E and further dominantly mafic metavolcanic rocks, similar to those of the
NW at 27.4538S, 15.4258E (Fig. 16.2). The Chameis Gate Grootderm Formation, whereas the upper part consists of carbon-
Member is so far the only glaciogenic unit that has been recognized ates and siliciclastic metasedimentary rocks that resemble those of
in that subterrane, and it was first described by Frimmel (2000b). the Holgat Formation in the Port Nolloth Zone and those of the
Preliminary lithogeochemical and chemostratigraphic data for upper Oranjemund Group.
the bounding strata have been presented by Frimmel & Jiang No basement has been identified in the Chameis Subterrane.
(2001). No further follow-up studies have been carried out, and Available geochemical and petrological data for the mafic rocks
much of its stratigraphic interpretation hinges on circumstantial (both extrusive and intrusive) across the Marmora Terrane are
evidence. most compatible with an oceanic intra-plate setting (Frimmel
The main significance of the Chameis Gate Member lies in its et al. 1996). For some metagabbros in the Chameis Subterrane,
palaeodepositional environment. In contrast to most, if not all, geochemical characteristics of mid-ocean ridge basalt are indi-
other diamictite units described so far, this diamictite occurs cated. Although exact ages for the time of magmatism are lacking,
associated with oceanic rocks with very little continental influence some crude constraints on the timing of sea-floor metamorphism
except for pre-Gariep basement-derived clasts and dropstones in (Frimmel & Frank 1998) and of oceanic volcanism (Frimmel &
the diamictite (Frimmel & Jiang 2001). Fölling 2004), detailed below, suggest an interval sometime
between 640 and 580 Ma for the time of oceanic crust formation.
Based on zircon provenance data from the Oranjemund Group
Structural framework and tectonic evolution siliciclastic rocks and comparable rocks from the Rocha Group
in the Punta del Este Terrane in eastern Uruguay, a back-arc
The Chameis Gate Member forms part of the largely oceanic basin to the 640–590 Ma volcanic arc of the Dom Feliciano Belt
Marmora Terrane, which represents an allochthonous nappe in southeastern Brazil and eastern Uruguay has been suggested
complex that was thrust in an easterly to southeasterly direction as the most likely depositional environment for the lower

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 217– 221. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.16
218 H. E. FRIMMEL

Fig. 16.1. Tectonic sub-division of the Marmora Terrane in the western Gariep
Belt (modified after Hartnady & von Veh 1990).

stratigraphic units of the Marmora Terrane (Basei et al. 2005).


During amalgamation of SW Gondwana, the basin was closed,
resulting in syn-orogenic sedimentation in a foredeep in front
and on top of already stacked thrust sheets of the Marmora
Terrane. This model explains the remarkable similarity in the
younger sedimentary deposits across tectonic boundaries in both
the Port Nolloth Zone and the Marmora Terrane (Frimmel &
Fölling 2004).
A penetrative foliation (s1) dips to the west and is axial –planar
relative to tight to isoclinal, east-vergent F1 folds. Locally, s1 was Fig. 16.2. Map showing the distribution of the Chameis Gate Member within
refoliated around open to tight, NE-trending F2 kink bands. The the Chameis Subterrane, Marmora Terrane (Gariep Belt), in southwestern
contact with the underlying Port Nolloth Zone is a major thrust Namibia.
fault with top-to-SE transport (Schakalsberge Thrust). The inten-
sity of folding increases towards the Schakalsberge Thrust. At
the northern boundary of the Marmora Terrane, F1 folds within
the Chameis Subterrane are truncated by the Schakalsberge with NE-, east- and SE-directed transport along the northeastern,
Thrust, highlighting its syn-D2 timing. The contact between the eastern and southeastern basin margin, respectively (Frimmel
Oranjemund and the Schakalsberge Subterranes is a west- to 2000a). This was followed by the second stage, sinistral transpres-
NW-dipping fault plane. The contact between the Chameis and sion with overall top-to-SE transport. The resulting folds and
Oranjemund Subterranes is a zone of intense F2 folding with thrusts are east-verging along the northeastern lateral ramps and
top-to-SE transport and subsequent back-folding that resulted in SE-verging along the southern frontal ramps (in South Africa)
locally steeply SE-dipping crenulation. within the outer Port Nolloth Zone (Gresse 1994).
The preferred kinematic model for the evolution of the main The Marmora Terrane differs from the Port Nolloth Zone not
orogenic structures involves two stages. The first is eastward only with regard to stratigraphy and deformation, but also in its
transport of the oceanic rocks of the Marmora Terrane, internal metamorphic history. In the Port Nolloth Zone, only one stage
stacking of these, and eventually obduction onto the continental of regional metamorphism is discerned, the peak of which
margin of the Port Nolloth Zone. The variable angle between (c. 520 8C and 3.0 –3.5 kbar) was reached after D2 but prior
principal direction of compression and the pre-existing arcuate to D3; it is ascribed to crustal thickening following overthrusting
north- to NW-trending continental margin led to strain partitioning of the Marmora Terrane onto the Port Nolloth Zone. The
THE CHAMEIS GATE MEMBER 219

metamorphic evolution of the Marmora Terrane is more complex. Glaciogenic deposits and associated strata
There, the mafic and ultramafic rocks experienced polyphase
metamorphism that is not recorded by their sedimentary cover. The Chameis Gate Member in the Dernburg Formation
As many as three metamorphic amphibole generations therein
reflect different metamorphic conditions of formation (Frimmel The Dernburg Formation, which contains in its upper part the
& Hartnady 1992) and yield distinctly different Ar –Ar age glaciogenic unit of interest here, consists predominantly of thinly
spectra (Frimmel & Frank 1998). They reflect three metamorphic laminated greenschist (metatuff), mafic hyaloclastites, metabasalt
stages: (i) sea floor metamorphism, (ii) a relatively high-pressure and serpentinized metapicrite. Compositional layering on a
metamorphic event, probably related to first accretion of mafic millimetre- to centimetre-scale reflects variations in magma
crust and (iii) regional metamorphism as a result of continental composition from mafic to intermediate. Acidic layers are rare.
collision. The metabasalt displays pillow structures in strain-protected
domains. In places, mica-rich tuffitic to pelitic beds are inter-
calated. Locally, the mafic rocks show evidence of intense meta-
Stratigraphy somatism, leading to epidosite, albitite, hornblendite and
ophicarbonate.
The Chameis Gate Member forms part of the Chameis Group, the The Chameis Gate Member is distinguished from the remainder
rocks of which constitute the Chameis Subterrane of the Marmora of the formation by its diamictite lithology. No stratification is dis-
Terrane. The stratigraphic subdivision of this group given here cernable. The matrix of the diamictite is similar to the laminated
(Fig. 16.3) follows a preliminary subdivision presented by greenschist (metatuff and -tuffite) and intercalated metapelite
Frimmel (2000b). It should be noted that the intensity of syn- above, but the clast composition is distinguishing. It contains
orogenic tectonic overprint of the entire subterrane makes it vir- exotic lonestones and dropstones, up to 1.5 m in diameter. They
tually impossible to provide reliable constraints on stratigraphic comprise various granitoids, gneisses, quartzite and dolomite,
thickness and, in places, even on stratigraphic boundaries. and display a variable degree of rounding. In less intensely
A thick pile of mafic metavolcanic rocks at the base of the deformed places, the dark grey pelitic or green chloritic beds are
succession (Dernburg Formation) with minor, local intercalations deflected only on the lower side of a given larger clast, thus attest-
of carbonates (Sholtzberg Member) and diamictite (Chameis Gate ing to its emplacement as a dropstone. Apart from the exotic clast
Member) is overlain by a carbonate sequence (Dreimaster Mem- content, no other glacial indicators, such as glacial striations, have
ber) and a siliciclastic, mainly arenitic succession with turbiditic been observed.
affinities. The latter two units constitute the Bogenfels Formation. Associated with the metavolcanic and volcaniclastic rocks of the
The Dernburg Formation was intruded by a series of tectonically Dernburg Formation are also distinct, in places Fe-rich, dolostone
dismembered metagabbro bodies (Bakers Bay Suite). outcrops, distinguished as Sholtzberg Member (Frimmel 2000b).

Fig. 16.3. Generalized lithostratigraphy of the Marmora Terrane and correlation with that of the Port Nolloth Zone, all of which constitutes the Gariep Supergroup
(modified after Frimmel & Fölling 2004).
220 H. E. FRIMMEL

Their exact stratigraphic position is unclear because of tectonic with very low Rb –Sr ratios (,0.0005). The presence of perfectly
contacts, but they appear in the upper part of the formation, preserved dolomitic ooids in the succession and the lack of geo-
close to the diamictite of the Chameis Gate Member. At one chemical evidence of diagenetic alteration suggest that the above
locality (27.6078S, 15.0408E), relics of circular stromatolite Sr isotope ratios are very close to primary. That dolostone has posi-
mounds are preserved. Elsewhere, the dolostone is associated tive d13CCarb ratios (as much as 2.82‰ relative to the PDB stan-
with magnesian metapelite that contains layers or lenses of tour- dard) and d18OCarb as high as –4.3‰ (relative to PDB), whereas
malinite and albitite. Sodic metasomatism is widespread, resulting the limestone of the younger Dreimaster Member is characterized
in the crystallization of abundant magnesioriebeckite in all litho- by consistently negative d13CCarb ratios slightly below zero, with
types in and around the Sholtzberg Member. An evaporitic the lowest ratios ( –2.6‰) determined within the first few metres
origin of this member is indicated by mineralogical and fluid above the bottom contact. A further distinctive feature of this lime-
inclusion data as well as B isotopic evidence (Frimmel & Jiang stone is its high Sr content (as much as 1412 ppm).
2001). For the lack of any continental influence in these deposits
(Rb, Zr, Nb, Th and U concentrations are an order of magnitude
lower than in normal marine carbonates) an atoll setting on an Other characteristics
oceanic seamount with temporary evaporation of seawater in the
central basin is envisaged for the Sholtzberg Member depositional No economic ore deposits are known from the Chameis Subter-
environment (Frimmel 2000b). rane, but base metal anomalies have been noted, particularly in
association with Na-metasomatism that affected most of the area
(H. E. Frimmel, unpublished data). The fluids involved were
The Bogenfels Formation highly oxidizing, which is evident from the widespread occurrence
of hematite veins. The metasedimentary rocks of the Marmora
The term ‘Bogenfels Formation’ was introduced by Martin (1965) Terrane still await a micropalaeontological investigation, which
for those rocks that had previously been described as ‘Folded might help in further constraining their age.
Nama’ (Kaiser 1926). At that time, no distinction was made
between the Marmora Terrane and the Port Nolloth Zone. To
conform to the current tectonic nomenclature, it has been sug- Palaeolatitude and palaeogeography
gested that the Bogenfels Formation be redefined as the unit of car-
bonate and siliciclastic rocks above the Dernburg Formation No reliable palaeomagnetic data are available for the entire
within the Chameis Subterrane (Frimmel 2000b). It thus represents Gariep Belt and thus also for the Marmora Terrane. The low-grade
the upper part of the Chameis Group. Its base is a regionally exten- regional metamorphic overprint most likely led to the remagneti-
sive, laterally continuous, laminated, medium-grey limestone, zation of the rocks during the Gariepian orogeny. Thus, any
followed by a massive, thick-bedded, fine-grained, light creamy- assumption about palaeolatitude and palaeogeography is based
white to grey dolostone and finally a dolomitic breccia in which only on circumstantial evidence and comparison with continen-
the immediately underlying carbonates appear re-worked. This tal-scale models (e.g. Collins & Pisarevsky 2005). The presence
carbonate succession of the Dreimaster Member is overlain by a of former evaporite deposits in the Marmora Terrane speak
siliciclastic succession of alternating feldspathic and quartz against a high-latitude position at the time of sedimentation.
arenite, chlorite schist and calcpelite. The feldspathic arenite, a
greywacke to arkose, displays rare graded bedding of turbiditic
appearance. Geochronological constraints
In contrast to the Dernburg Formation, which is entirely differ-
ent to any unit in the Port Nolloth Zone, the Bogenfels Formation The minimum age is provided by the age of metamorphism related
is lithologically comparable to the Holgat Formation of the Port to the accretion of the Marmora Terrane, which has been con-
Nolloth Zone (Fig. 16.3). strained at 575 + 2 Ma by Ar –Ar data on a second, relatively
high-pressure metamorphic amphibole generation in metagabbro
of the Bakers Bay Suite (Frimmel & Frank 1998). Slightly older
Boundary relations with overlying and underlying Ar –Ar ages of 600–610 Ma, obtained in the same study on an
non-glacial units early, very low-pressure metamorphic amphibole generation
related to sea floor metamorphism, might set a minimum constraint
Upper and lower boundaries of the Chameis Gate Member are on the timing of oceanic crust formation, but are subject to a large
either tectonic or covered by sand. The spatial distribution of the uncertainty. An attempt to date the dolostone of the Sholtzberg
member is confined to areas made up of metavolcanic and Member by the double-spike Pb– Pb technique did not yield a
-volcaniclastic rocks of the Dernburg Formation. The lower con- reliable result for the carbonate fraction. Its silicate fraction, con-
tact of that formation is not exposed. The upper contact of the trolled by minute, presumably volcanogenic zircon grains, gave an
Dernburg Formation with the Dreimaster Member (Bogenfels imprecise age of 598 + 60 Ma (Frimmel & Fölling 2004), which is
Formation) is sharp and conformable, defined by the first appear- in agreement with the mentioned Ar –Ar data.
ance of limestone. The stratigraphic relationship of the Chameis
Gate Member with the Sholtzberg Member within the Dernburg
Formation is unclear and only inferred because of a lack of Discussion
primary contacts.
The presence of exotic dropstones in marine, volcaniclastic rocks,
the massive, unbedded nature of the diamictite and lack of graded
Chemostratigraphy bedding support a glaciomarine origin of the Chameis Gate
Member, but no definitive glaciogenic features have been
Limited trace element concentrations as well as C, O and Sr isotope observed. Deposition distal from a glaciated continental margin
data have been reported for carbonates of the Sholtzberg and Drei- by processes of settling from suspension and ice rafting is
master Members in the stratigraphic vicinity of the glaciogenic suggested. The Dreimaster Member carbonates are interpreted as
Chameis Gate Member (Frimmel 2000b; Frimmel & Jiang 2001; a cap-carbonate sequence based on the stratigraphic position and
Frimmel & Fölling 2004). 87Sr/86Sr ratios of 0.7075– 0.7078 C isotopic compositions equivalent to those of the post-Numees
were obtained on evaporitic dolostone of the Sholtzberg Member Bloeddrif Member in the Port Nolloth Zone.
THE CHAMEIS GATE MEMBER 221

The lack of precise radiometric age control for the Chameis Frimmel, H. E. 2000a. The Pan-African Gariep Belt in southwestern
Gate Member precludes a reliable stratigraphic correlation with Namibia and western South Africa. Communications of the Geologi-
glaciogenic deposits elsewhere. Although the available data base cal Survey of Namibia, 12, 197–209.
is rather limited, tentative correlation with other glaciogenic Frimmel, H. E. 2000b. The stratigraphy of the Chameis Sub-terrane in the
deposits in the region can be attempted on the basis of Sr and Gariep Belt in southwestern Namibia. Communications of the Geo-
C isotope data on over- and underlying carbonates and the logical Survey of Namibia, 12, 179–186.
limited radiometric age data. The C isotopic evidence for the Frimmel, H. E. & Fölling, P. G. 2004. Late Vendian closure of the
Sholtzberg Member is ambiguous. The noted enrichment in 13C Adamastor Ocean: timing of tectonic inversion and syn-orogenic
sedimentation in the Gariep Basin. Gondwana Research, 7,
therein could be merely a reflection of elevated evaporation
685– 699.
rates. However, its Sr isotopic composition is lower than in late Frimmel, H. E. & Frank, W. 1998. Neoproterozoic tectono-thermal evol-
Ediacaran carbonates, but compares well with that reported region- ution of the Gariep Belt and its basement, Namibia/South Africa.
ally for the Pickelhaube Formation (Fölling & Frimmel 2002) Precambrian Research, 90, 1 – 28.
and globally for early to middle Ediacaran carbonates (Halverson Frimmel, H. E. & Hartnady, C. J. H. 1992. Blue amphiboles and
et al. 2007). their significance for the metamorphic history of the Pan-African
Gariep belt, Namibia. Journal of Metamorphic Geology, 10,
Fieldwork in the Sperrgebiet would have been impossible without the logistic 651– 669.
support and permissions to enter restricted areas granted by Namdeb Pty Ltd. Frimmel, H. E. & Jiang, S.-Y. 2001. Marine evaporites from an oceanic
Funding by the South African National Research Foundation is gratefully island in the Neoproterozoic Adamastor ocean. Precambrian
acknowledged. Comments from G. Halverson in particular helped in improving Research, 105, 57– 71.
the manuscript. This work represents a contribution of the IUGS- and Frimmel, H. E., Hartnady, C. J. H. & Koller, F. 1996. Geochemistry
UNESCO-funded IGCP (International Geoscience Programme) Projects #512 and tectonic setting of magmatic units in the Pan-African Gariep
and #478. Belt, Namibia. Chemical Geology, 130, 101–121.
Gresse, P. G. 1994. Strain partitioning in the southern Gariep Arc as
reflected by sheath folds and stretching lineations. South African
References Journal of Geology, 97, 52– 61.
Halverson, G. P., Dudas, F. Ö., Maloof, A. C. & Bowring, S. A.
Basei, M. A. S., Frimmel, H. E., Nutman, A. P., Preciozzi, F. & Jacob, 2007. Evolution of the 87Sr/86Sr composition of Neoproterozoic
J. 2005. A connection between the Neoproterozoic Dom Feliciano seawater. Palaeogeography, Palaeoclimate, Palaeoecology, 256,
(Brazil/Uruguay) and Gariep (Namibia/South Africa) orogenic 103– 129.
belts — evidence from a reconnaissance provenance study. Precam- Hartnady, C. J. H. & von Veh, M. W. 1990. Tectonostratigraphic and
brian Research, 139, 195–221. structural history of the Late Proterozoic –Early Palaeozoic Gariep
Collins, A. S. & Pisarevsky, S. A. 2005. Amalgamating eastern Gond- Belt, Cape Province, South Africa. Geological Society of South
wana: the evolution of the circum-Indian orogens. Earth-Science Africa, Cape Town.
Reviews, 71, 229–270. Kaiser, E. 1926. Die Diamanten Wüste. Dietrich Reimer, Berlin.
Fölling, P. G. & Frimmel, H. E. 2002. Chemostratigraphic correlation of Martin, H. 1965. The Precambrian geology of South West Africa and
carbonate successions in the Gariep and Saldania Belts, Namibia and Namaqualand. Precambrian Research Unit, University of Cape
South Africa. Basin Research, 14, 69– 88. Town, Bulletin, 4, 1– 177.
Chapter 17

The Kaigas and Numees formations, Port Nolloth Group, in South Africa and Namibia

HARTWIG E. FRIMMEL1,2
1
Department of Geological Sciences, University of Cape Town, Rondebosch 7701, South Africa
2
Present address: Geodynamics & Geomaterials Research Division, Institute of Geography and Geology, University of Wuerzburg,
Am Hubland, D-97074 Wuerzburg, Germany
(e-mail: hartwig.frimmel@uni-wuerzburg.de)

Abstract: The Port Nolloth Group makes up the eastern, external part of the Pan-African Gariep Belt (Port Nolloth Zone) in southern
Namibia and western South Africa. It contains two glaciogenic diamictite units, the older Kaigas Formation and the younger Numees
Formation, with intercalated and overlying carbonate-dominated units. Available chemostratigraphic information include O, C and Sr
isotope data. Micropalaeontological and geochronological data point to an early Cryogenian age (c. 750 Ma) of the Kaigas Formation
and possibly a middle Ediacaran age (c. 580 Ma) for the Numees Formation. The former was deposited in an evolving, but eventually
failed, continental rift on the western flank of the Kalahari Craton, probably at low latitude. The Numees Formation is a laterally con-
tinuous, up to 600-m-thick glaciomarine deposit for which a passive continental margin setting has been suggested. Alternatively, based
on more recent data, the depositional setting might have been a back-arc basin. The eroded remnants of the corresponding arc are present
in the Dom Feliciano Belt.

The Kaigas and Numees Formations represent the older and at the so-called Numees Prospect (28.29168S, 16.97018E) were
younger glaciogenic diamictite-dominated units, respectively, initially regarded as part of the Numees Formation (De Villiers
within the para-authochthonous Port Nolloth Zone of the Pan- & Söhnge 1959) but since have been recognized to be older and
African Gariep Belt (Fig. 17.1). Owing to intense transpressional part of the Kaigas Formation (Kröner 1974; von Veh 1993).
overprinting of large parts of this zone, the distinction between Although the name Numees Formation is derived from a place
the two formations has been unclear for decades. This has led to (Numees Prospect) that is now assigned to the Kaigas Formation,
considerable confusion in the stratigraphic terminology used in the formation name is retained because the main occurrence of the
the older literature, with some workers even having argued for Numees Formation is only about 5 km to the west and the name is
the existence of only a single diamictite unit in the entire belt firmly entrenched in the literature. The type area for the Numees
(Hälbich & Alchin 1995; Jasper et al. 2000). Regional mapping Formation is south of the Orange River at 28.18 –28.338S,
(von Veh 1993; Frimmel 2008), as well as the distinctly different 16.87 –16.938E.
litho-, chemo-, chrono- and biostratigraphic characteristics in the Following the initial description of the South African portion of
respective overlying carbonate-rich successions (Fölling et al. the Port Nolloth Zone by Rogers (1916), the area was remapped in
2000; Fölling & Frimmel 2002; Frimmel & Fölling 2004; greater detail by De Villiers & Söhnge (1959). Subsequent struc-
Gaucher et al. 2005), however, have demonstrated the existence tural and lithostratigraphic studies by Kröner (1974) and von
of the two separate diamictite units at different stratigraphic pos- Veh (1993) laid the foundation for the current lithostratigraphic
itions. Here, the stratigraphic nomenclature as accepted by the subdivision of the entire zone (Frimmel 2003b). The Namibian
South African Commission on Stratigraphy (Frimmel 2003a) is sector was mapped in greater detail by McMillan (1968). Davies
followed but with the understanding that from a sequence strati- & Coward (1982) provided a first interpretation of the Gariep
graphic perspective, it is in need of revision. Belt’s structural evolution. This structural work was followed up
Rocks of both the Kaigas and Numees Formations are exposed at in the South African sector of the Port Nolloth Zone by Gresse
several sites in southernmost Namibia and in the adjoining part of (1994) and in the Namibian sector by Hälbich & Alchin (1995)
South Africa (Fig. 17.2). The spatial distribution of outcrops and Jasper et al. (2000). An attempt to resolve the sedimentology
follows that of the outer zone of the Gariep Belt, which occurs in of parts of the stratigraphic succession was made by Jasper et al.
an arcuate, south-trending, 20–60-km-wide zone stretching from (1992 –1993). Metamorphism (Frimmel 1995), geochemistry and
Lüderitz in Namibia, across the Orange River to the western Rich- tectonic setting of the various igneous bodies (Frimmel et al.
tersveld region of the Northern Cape Province in South Africa, and 1996a), timing of tectonic events and the thermal evolution have
striking out to sea south of Port Nolloth (Fig. 17.1). The Numees also been studied (Reid et al. 1991; Grotzinger et al. 1995;
Formation, including outcrops that are now assigned to the Frimmel et al. 1996b; Frimmel & Frank 1998).
Kaigas Formation, was the first Neoproterozoic unit in southern Attempts to place the rock record of the Port Nolloth Zone,
Africa to have been recognized as being of glacial origin (Rogers including the Kaigas and Numees Formations, into a regional
1916; Beetz 1926), an interpretation that has never been seriously tectonic framework and to correlate units on a subcontinental
challenged by subsequent workers (Martin 1965a). scale have been made by Frimmel et al. (2002). Our current
The original Kaigas Series of Rogers (1916) included most of understanding of the tectonic settings for the deposition of the
the strata now assigned to the Hilda Subgroup. The Kaigas For- glaciogenic deposits is based on work by Basei et al. (2005), and
mation as redefined by De Villiers & Söhnge (1959) was limited the lithostratigraphic correlation has been updated by Gaucher
to beds along the Kaigas (also known as Gaigas) River, where et al. (2005).
excellent, almost continuous exposures along a ridge to the east
of the river bed at 28.6318S, 17.0988E (Fig. 17.2) constitute the
holostratotype (Fig. 17.3). Diamictite in the type area for the Structural framework and tectonic evolution
Numees Series of Rogers (1916) was later included by von Veh
(1993) in the Kaigas Formation because of lithological similarities. In the past, the entire Port Nolloth Group was interpreted as
Likewise, the diamictite exposures west and NW of Numees Peak having been deposited in a basin (Gariep Basin) that continuously

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 223– 231. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.17
224 H. W. FRIMMEL

Fig. 17.1. Position of the Port Nolloth Zone within the Gariep Belt (modified
after Hartnady & von Veh 1990).

evolved from a continental rift to a passive margin and eventually


became a foredeep prior to continental collision at c. 545 Ma
(Frimmel et al. 1996a; Frimmel & Frank 1998). An alternative
to the traditional Wilson Cycle model for the Gariep Belt has
emerged from a comparison with the tectono-thermal evolution
of the South American counterparts in southeastern Brazil and
eastern Uruguay (Basei et al. 2005; Frimmel & Basei 2006).
According to these more recent studies, the Gariep Basin might
have been a successor basin in which sediments had been deposited Fig. 17.2. Map showing the distribution of the Kaigas and Numees Formations
in two different basins and were subsequently stacked on top of within the Port Nolloth Zone (made up of rocks of the Port Nolloth Group),
each other. Gariep Belt, in southern Namibia and westernmost South Africa (western
The Kaigas Formation belongs to an older sequence that reflects Richtersveld). Also shown are the localities of the type sections for the Kaigas
progressive continental rifting. Predominantly siliciclastic sedi- and Numees Formations.
ments of the Stinkfontein Subgroup were first deposited into a
widening rift graben, bounded by listric faults. The presence of
debris flow and turbidity fan deposits as well as large slump Pinah Graben in the east, with a number of smaller, separate
masses indicates widespread syn-depositional deformation related basins, from a half-graben to the west. At an advanced rifting
to continental rifting. Large, but only locally developed dolomitic stage, eustatic sea-level drop was accompanied by the entry of gla-
olistostromes near the northeastern basin margin are explained by ciers into the basin along its eastern margin to the mainland. Debris
seismic activity along active, basin-bounding growth faults. These released from those glaciers makes up the bulk of the Kaigas
normal faults were subsequently inverted during the Gariepian Formation. This stage was accompanied by bimodal, predomi-
orogeny, but the position of some of the major syn-rift faults is nantly felsic, rift volcanism (Rosh Pinah Formation) on a growth
still recognizable. A major growth fault must have existed along fault along the eastern basin margin. Although still in an overall
today’s eastern to northeastern margin of the belt, separating base- extensional regime, the eustatic drop in sea level led to the star-
ment rocks to the east and NE from the rift graben fill. This fault is vation of the Rosh Pinah Basin and consequently to anoxic
now present as an inverted NW –SE trending thrust fault system, bottom waters, which provided a suitable environment for sedi-
with the Rosh Pinah Fault being a good example. Along those mentary – exhalative base-metal mineralization (Rosh Pinah
faults, an elongated, roughly north –south trending basement mine and protore of the secondary Skorpion deposits). Renewed
horst (Aurus Horst) developed, separating the so-called Rosh transgression led to the drowning of parts of the basement horst
KAIGAS AND NUMEES FORMATIONS 225

the emergence of the Dom Feliciano volcanic arc towards the


(north)west. Consequently, the upper Hilda Subgroup and the
overlying Numees Formation would have been deposited in a
back-arc basin behind the Dom Feliciano arc.
Inversion from extension to compression took place during
post-Numees sedimentation. A foredeep developed in which the
youngest sediments of the Port Nolloth Group (Holgat Formation)
were deposited (Frimmel & Fölling 2004). At the same time, first
stacking and accretion of the advancing oceanic rock pile in the
shrinking basin took place. This was followed by continental
collision in a SE-directed transpressional regime, which was
accompanied by regional low-grade metamorphism, east-vergent
folding, thrusting and back-thrusting. Many of the original
normal faults were reactivated at that stage (von Veh 1993;
Hälbich & Alchin 1995). In places, syn-tectonic foliation occurred
at a high angle to bedding, but primary sedimentary contacts
remained well preserved in strain-protected domains.

Stratigraphy

Both the Kaigas and Numees Formations form part of the Port
Nolloth Group (Fig. 17.4), which, in turn, is the part of the
Gariep Supergroup that makes up the bulk of the Port Nolloth
Zone. The Port Nolloth Group unconformably overlies rocks of
the Mesoproterozoic Namaqua Metamorphic Province and, in
places, alkaline intrusive rocks of the 833– 771 Ma Richtersveld
Suite (Frimmel et al. 2001). It is, in turn, unconformably overlain
by the Kuibis Subgroup of the Nama Group in the east, and towards
the west it is overlain tectonically by the Marmora Terrane, the
allochthonous internal sector of the Gariep Belt.
The stratigraphy of the Port Nolloth Group begins with a pre-
dominantly siliciclastic, minor volcanic to volcaniclastic continen-
tal rift graben fill, the Stinkfontein Subgroup. This is followed by
the stand-alone Kaigas Formation, which, in turn, is overlain by the
Hilda Subgroup. Within the Hilda Subgroup, four formations are
distinguished, some of which are time-equivalents of each other.

Fig. 17.3. Holostratotype of the Kaigas Formation (for location see Fig. 17.2;
modified from Frimmel et al. 2003a).

as well as the Kaigas Formation and the deposition of a varied,


carbonate-dominated succession (Pickelhaube Formation).
One of the great enigmas of the Gariepian stratigraphy has been
the lack of sediment in the post-rift successions. If a successful rift
is assumed for the tectonic evolution of the Gariep Basin (Adamas-
tor Ocean), as postulated by most workers in the past (Hartnady
et al. 1985; Germs 1995; Frimmel et al. 1996a), the post-Kaigas
period should have been marked by further transgression and the
development of a thick sedimentary passive margin sequence.
First regional drowning of the rift shoulders is indeed indicated
by the laterally continuous post-Kaigas carbonates, and the
various carbonate-dominated facies of the Hilda Subgroup
provide good evidence of the development of carbonate platforms.
Yet the total thickness of the Hilda Subgroup (except for the vol-
canically controlled Rosh Pinah area) is not more than 650 m.
This lack of sediment is explained by a major hiatus and/or
slope failure as the Hilda Subgroup is largely composed of turbi-
dites, olistostromes and potentially cannibalizing debris flows.
Fieldwork by the author has shown that the erosive surface
beneath the siliciclastic Wallekraal Formation within the Hilda
Subgroup extends across the entire Port Nolloth Zone and is a
far more widespread feature than previously recognized. It is a
first-order sequence boundary.
The temporal extent of the pre-Wallekraal hiatus remains un-
resolved. If the above back-arc model applies, the original basin- Fig. 17.4. Generalized lithostratigraphy of the Port Nolloth Group (Gariep
bounding faults would have become re-activated in response to Supergroup); unit thickness in metres given in parentheses.
226 H. W. FRIMMEL

In the volcanically influenced basin margin, the Kaigas Formation siliciclastic successions higher up in the stratigraphy, lateral
diamictite is followed by the volcanic to volcaniclastic Rosh Pinah facies changes that make long-distance correlation with the type
Formation, whereas elsewhere it is followed by the carbonate- sections in South Africa problematic, and the commonly tectonic
dominated Pickelhaube Formation. The former comprises rhyolitic nature of its upper and lower contacts. Occurrences of the sub-
lava flows, subaerial and submarine pyroclastic deposits of highly group are confined to the northeastern basin margin. Lithologi-
variable grain size (from agglomerate to ash tuff and ignimbrite), cally, the Stinkfontein Subgroup in Namibia is similar to the
whereas the latter consists predominantly of carbonate rocks Vredefontein Formation in the Richtersveld, whereas the relatively
(laterally discontinuous, massive dolomitic marble and laminated clean quartz arenite, typical of the Lekkersing Formation in the
to medium-bedded siliceous and graphite-bearing dolomitic, lower part of the Stinkfontein Subgroup, is only poorly developed
partly allodapic limestone, with slump breccias) and subordinate in Namibia.
feldspathic arenite and graphitic metapelite alternating locally The overall higher feldspar content in the siliciclastic rift depos-
with calcpelite. These are overlain paraconformably, or with an its in Namibia might reflect an overall more proximal position of
erosional unconformity, by coarse-grained clastic sedimentary the depositional environments relative to the area further south
rocks of the Wallekraal Formation in which parts of the older Pick- in South Africa. This finds support from the observation that
elhaube carbonate platform appear re-worked. The Wallekraal along the NW –SE trending branch of the basin north of the
Formation consists predominantly of very coarse-grained felds- Aurus Horst, the entire Stinkfontein Formation is missing and
pathic arenite and less abundant well-rounded polymict conglom- rocks of the next younger stratigraphic units, the Kaigas For-
erate, phyllitic to calcareous, locally graphitic, metapelite, isolated mation, where present, or the Hilda Subgroup rest directly on
lenses of dolostone, typically re-worked into olistostromes, and the basement.
thin-bedded allodapic limestone. The youngest unit of the Hilda
Subgroup is the Dabie River Formation, which consists essentially
of stromatolitic reef facies, including dolomite breccia and oolite. The Kaigas Formation
The Hilda Subgroup is followed by diamicite and minor banded
iron-formation of the Numees Formation, which, in turn, is over- The Kaigas Formation is only locally developed along the eastern
lain by the Holgat Formation. The latter commences with a carbon- and northeastern margin of the basin. It is characterized by its
ate sequence above the Numees Formation diamictite (Bloeddrif marked variations in thickness along strike, with individual occur-
Member), and in places a thick package of turbiditic, siliciclastic rences pinching out over only several hundred metres. At its type
metasedimentary rocks. locality (Fig. 17.2) in the southern Richtersveld (South Africa), it
Because glaciogenic deposits are the emphasis of this paper, the reaches a thickness of 115 m. The formation consists predomi-
stratigraphic description of the two glaciogenic units and those that nantly of medium- to thick-bedded diamictite and subordinate
are associated with them follow in more detail below. massive, locally cross-bedded or graded bedded, feldspathic
arenite and argillite. Ripple cross laminations, indicating palaeo-
current directions from the east, rip-up of mudstone clasts, load
Glaciogenic deposits and associated strata and flute casts, and sinuous ripple marks are locally present. The
diamictite beds display sharp upper and lower contacts and lack
Descriptions of the two principal glaciogenic units in the Port internal structures except for crudely graded bedding. The clast
Nolloth Group, the Kaigas and Numees Formations, have been size distribution is strongly bimodal with generally larger clasts
provided by von Veh (1993), Frimmel (2000, 2008) and present in thicker diamictite beds. In places, outsized erratic
Frimmel et al. 2003a, b). To also obtain some information on blocks can also be found in pelitic beds. An overall decrease in
the rock record that provides possible information on the events clast size away from the former basin margin is noted. Individual
that led up to the glacial intervals and those that followed, the clasts and blocks commonly appear sandwiched (cold deformation
bounding stratigraphic units below and above the glaciogenic by ice). Faceted clasts can be found in less deformed regions.
units will be also described briefly. Unequivocal glacial striations are difficult to identify because of
the tectonic overprint under greenschist-facies temperatures,
which modified most clast surfaces.
The Vredefontein Formation Locally, dolostone is a common clast lithology in the southern
part of the Port Nolloth Zone, but elsewhere the Kaigas Formation
Strata underlying the Kaigas Formation are known as the Vrede- diamictite is notably free of carbonate clasts. The clast population
fontein Formation (Stinkfontein Subgroup). The Vredefontein is generally dominated by basement-derived rock types, such as
Formation is up to 300 m thick. Internal sedimentary structures, granite, gneisses, amphibolite, schists and re-worked arenite
accentuated by heavy mineral concentrations along laminations, from the underlying Stinkfontein Subgroup. The general lack of
are well preserved in the dominantly medium-bedded feldspathic carbonate clasts (with the exception of the southern Port Nolloth
arenite. These include various forms of cross-bedding, asymmetri- Zone) is a major feature by which the Kaigas Formation differs
cal ripple marks indicating palaeocurrent directions from the east from the younger Numees Formation. Further features that make
or SE, as well as various soft sediment deformation features possible a distinction of the Kaigas from the Numees Formation
(von Veh 1993). are the absence of a near-basal banded iron-formation or ferrugi-
In the upper part of the formation, minor felsic, intermediate and nous arenite zone and the relative abundance of interbedded
mafic metavolcanic rocks are intercalated (Middlemost 1963). arenite and pelite.
These rocks are only known from the Richtersveld in South In Namibia, the formation reaches its maximum thickness in an
Africa, but they provide an important reference for the genesis area east of the Aurus Mountains around Borehole 5 at 16.5278E,
of more abundant volcanic rocks at a higher stratigraphic position 27.6798S (Fig. 17.2), where it rests directly on basement. From
in Namibia (Rosh Pinah Formation). Sandy to gritty breccia that there it pinches out rapidly towards the SE, but re-emerges in
occurs towards the top of the Vredefontein Formation contains several small occurrences along the basement edge to the north
fragments of angular to sub-rounded vein quartz, granitoid, micro- and SE of Rosh Pinah (Alchin et al. 2005; Frimmel 2008). On
granite, gneiss, orthoquartzite, calcareous grit and felsic volcanic the farm Spitzkop 111, the Kaigas Formation diamictite grades lat-
rocks in a massive or indistinctly cross-bedded feldspathic erally into a coarse-grained, proximal pyroclastic deposit (at
grit matrix. 16.7068E, 27.8278S) that is assigned to the Rosh Pinah Formation,
The distribution of the Stinkfontein Subgroup in Namibia indicating sedimentation that overlapped in time with volcanism.
remains enigmatic, because of the occurrence of similar Elsewhere, the background sedimentation towards the top of the
KAIGAS AND NUMEES FORMATIONS 227

Kaigas Formation gradually changes from siliciclastic to calcar- comprise intercalated ripple-marked quartzite and dolostone,
eous. An example of that can be found east of the Aurus Mountain which is, in places, highly ferruginous due to fumarolic activity
at 16.4868E, 27.6518S, where the matrix of the diamictite and (Frimmel & Lane 2005).
intercalated arenite becomes progressively dolomitic towards the
top of the formation, thus heralding a fundamental change in the
style of sedimentation that led to the next stratigraphic unit. The Numees Formation

In places where the stromatolitic reef facies of the Dabie River


The Pickelhaube and Rosh Pinah Formations Formation is not developed, the siliciclastic Wallekraal Forma-
tion grades conformably into the Numees Formation. The strato-
Where the Kaigas Formation is developed, the Pickelhaube For- type section of the formation as accepted by the South African
mation rests on it conformably. Elsewhere, the Pickelhaube For- Committee for Stratigraphy (Frimmel et al. 2003b) is located in
mation occurs with an erosional unconformity on basement. It the Richtersveld region in westernmost South Africa (Fig. 17.5).
starts with laminated, variably dolomitized, medium to dark grey That area suffers, however, from considerable tectonic compli-
limestone, followed by a varied sequence of predominantly argil- cation and less tectonized sections of the formation occur in the
lite and marl with minor feldspathic quartz arenite in which Namibian part of the belt. The total thickness of the Numees For-
small-scale channels and asymmetric ripple marks indicate a mation is estimated to be not more than 600 m, but decreases
palaeocurrent direction from the SE. Intercalated are a number towards the east where the formation onlaps directly onto pre-
of limestone and dolostone beds, which become more prominent Gariep basement. The dominant lithology is a generally massive,
towards the top of the formation. The absence of conglomerate, very thick bedded (bedding poorly defined) diamictite with small
scarcity of cross-bedding, presence of thin bedding, and the to very large (2 mm –10 m) subrounded clasts. The lithology of
increasing carbonate content of the Pickelhaube Formation are- the clasts comprises pegmatite, leucogranite, granitic-gneiss,
nites indicate quieter, submerged conditions compared to the quartzite, schist, dolostone, minor limestone of the immediately
underlying Stinkfontein Subgroup and Kaigas Formation, as can underlying Dabie River Formation and re-worked diamictite.
be expected in the distal parts of a fan complex. Some of the Carbonate clasts are particularly abundant in the lower parts of
clastic beds may represent turbidites and minor debris flow the formation but become rare in the upper parts, where basement-
deposits. The original limestone deposits are largely dolomitized derived clasts dominate. Clast and matrix grain sizes decrease
in the more proximal positions but retained their syn-sedimentary upwards. Intercalated minor ferrugineous, feldspathic arenite is
mineralogy farther away from the basin margin. Laminated alloda- coarse-grained, medium-bedded and forms the matrix of those
pic limestone (calcarenite) is a ubiquitous feature of this formation, parts of diamictite that are dark grey to blue-green in colour. The
which reaches a maximum thickness of about 280 m. latter becomes dark brown on weathered surfaces. In addition,
Where the formation onlaps directly onto basement, basement- thin-bedded, varve-like siltstone beds with intercalated coarser-
derived clasts are embedded in allodapic limestone of the lower grained, graded sandstone beds occur. Some of these beds
Pickelhaube Formation. The carbonate beds in the middle of the contain dropstones with soft-sediment deformation around their
formation show extensive syn-sedimentary brecciation that is bottom contact.
reminiscent of tempestite. This is in contrast to slump breccias A highly ferruginous unit, consisting of thinly laminated chlorite
and carbonate debris flow deposits, some of them containing extra- schist, iron-formation and diamictite, is distinguished as the
basinal clasts, which are common in the thicker carbonate beds. Jakkalsberg Member near the base of the formation (Frimmel
The Rosh Pinah Formation (Kapok Formation of Martin 1965b) et al. 2003b). This member is best developed in the southern
is distinguished from the Pickelhaube Formation by the presence portion of the belt in South Africa (Fig. 17.5), but is largely
of felsic volcanic and volcaniclastic rocks in the former. The great- missing in Namibia (except for a few outcrops just north of the
est accumulation of felsic volcanic rocks, also referred to as Spitz- Orange River). In places, exotic, mainly basement-derived drop-
kop Formation by some authors (Siegfried & Moore 1990), occurs stones are found within the iron-formation.
some 15 –20 km north of Rosh Pinah (on farm Spitzkop 111) and is
considered to represent the main volcanic centre. The Rosh Pinah
Formation displays repetitive sedimentary cycles that reflect quies- The Bloeddrif Member (Holgat Formation)
cent periods intervening with rapid deposition related to reactiva-
tion on basin bounding faults with subsequent thermal and On top of the Numees Formation follows a distinct carbonate
mechanical subsidence (Alchin et al. 2005). Although deposition unit that is distinguished as the Bloeddrif Member within the other-
of the volcanosedimentary succession was sustained in the more wise siliciclastic Holgat Formation. The true thickness of the for-
active, proximal parts of the basin, distal background sedimen- mation is difficult to constrain because of intense folding and thrust
tation ranged from deposition of planar laminated allodapic lime- duplication. It appears, however, that stratigraphic onlap led to a
stone to mudstone, followed by platform carbonates (Pickelhaube decrease in thickness from several hundred metres in the west to
Formation). In proximal positions, tilting of the rift shoulders led to less than 100 m in the east. The formation is best developed in
partial erosion of the Rosh Pinah Formation and thus to intraforma- the western Richtersveld in South Africa, where the type sections
tional breccias and olistostromes. are located.
Volcanic lithotypes in the Rosh Pinah Formation comprise In less deformed regions near the basin margin, such as some
massive to flow banded quartz-alkali feldspar rhyolite to rhyoda- 17 km to the NE of Rosh Pinah and on farm Nord Witputs 22 as
cite, hyaloclastic breccias, often locally reworked (lapilli tuff brec- well as on farm Namuskluft east of Rosh Pinah, internal, vertical
cias), and a variety of volcaniclastic units, which reflect different tube-like structures of infilled micritic sediment and cement
energy and distance from the eruptive centre. They range from form a conspicuous feature of the Bloeddrif Member. They
proximal agglomerate, to lapillistone, and distal coarse- and fine- resemble structures that have been described from many other
grained tuffs. Inversely graded lapillistone/tuff units are present post-glacial carbonates (Cloud 1974), and have been variably
and indicate water-lain pumice ash flow deposits. A special interpreted as microbial in origin (Hegenberger 1993; Hoffman
variety is an alkali feldspar-rich crystal tuff to tuffite. A bimodal et al. 1998; Corsetti & Grotzinger 2005) or by gas escape following
character of the Rosh Pinah volcanism is indicated by the local the destabilization of gas hydrate during warming of terrestrial
presence of metabasalt and metagabbro and the lack of igneous permafrost (Kennedy 2001). Some microbial activity in the Bloed-
rocks of intermediate composition (Frimmel et al. 1996a). drif Member is clearly indicated by the local presence of stromat-
Sedimentary rocks directly associated with the volcanic facies olites. More distal, relatively pure limestone deposits that are
228 H. W. FRIMMEL

Fig. 17.5. Location of the holostratotypes


of the Numees Formation (a) and the
Jakkalsberg Member (b). The type section
of the Numees Formation is 380 m thick,
rests in fault contact on feldspathic arenite
of the Wallekraal Formation, and begins
with 30 m of thickly bedded diamictite.
Banded iron-formation, thinly laminated
Fe-rich metapelite and quartzite of the
Jakkalsberg Member, displaying graded
bedding, follow conformably and attain a
thickness of 5 m. They are overlain by 95 m
of Fe-rich, green, thickly bedded diamictite
and feldspathic arenite intercalations,
locally displaying horizontal lamination.
The diamictite is rich in dolomite clasts.
This is followed by 250 m of thickly
bedded, yellow-brown diamictite with only
few dolomite clasts but predominantly
basement-derived granite and gneiss clasts.
The upper contact of the formation is
tectonic, with a repetition of the
Jakkalsberg Member (from Frimmel et al.
2003b).

characteristically devoid of continental detritus can be distin- towards the top of the formation. The upper contact with the
guished from more proximal deposits that experienced more Pickelhaube Formation is in many places conformable, but
intense, early diagenetic dolomitization and have intercalations locally (mostly in South Africa) unconformable.
of thin arenite beds (Frimmel & Fölling 2004). A particularly The lower contact of the Numees Formation to the underlying
good example of a very shallow, proximal facies of this member Hilda Subgroup is locally conformable (Kröner 1974), although
occurs to the north of Aurus Mountain at 16.2168E, 27.4538S, in most places it is tectonic. Wherever the Dabie River Formation
where a thinly laminated dark grey to cream-coloured limestone carbonates are not developed, the siliciclastic Wallekraal
with numerous erosion surfaces and mud cracks grades upwards Formation grades conformably into the Numees Formation. The
into a chaotic dolomitic breccia that is unconformably overlain upper boundary of the Numees Formation with the carbonates of
by white, locally cross-bedded dolomitic arenite that became dis- the Bloeddrif Member is typically sharp, conformable or paracon-
rupted and filled with sand dykes (Frimmel 2008). In most formable. The stratigraphic relationships of the Jakkalsberg
places, the laminated limestone is overlain by a massive, dark Member within the Numees Formation with the bounding units
grey dolostone that is followed by a thickly bedded, distinctly are quite clear from a regional perspective. The iron-formations
pink dolostone. It should be noted, however, that the distal sections are always, without a single exception, near the bottom of the
through the Holgat Formation appear continuous without breaks in Numees Formation diamictite, and with conformable contacts,
sedimentation. either sharp or gradational. To take those out and place them
into a different stratigraphic position altogether is not supported
by the field evidence.
Boundary relations with overlying and underlying
non-glacial units
Chemostratigraphy
The boundary relations of the Kaigas Formation with the under-
lying strata are marked by the laterally discontinuous distribution The first stable isotopic compositional data for carbonates from the
of the formation. At the type locality in the southern part of the Port Port Nolloth Zone were reported by Kaufman et al. (1991), but
Nolloth Zone, the contact of the Kaigas Formation with the under- those data suffered from poor stratigraphic control. Since then, a
lying feldspathic arenite of the Vredefontein Formation is con- number of chemostratigraphic profiles with isotope ratios carefully
formable and gradational, and is defined by the first appearance screened for the effects of post-depositional alteration, through
of diamictite. Further north, the contact with the underlying Stink- various sections of carbonate-dominated successions in the Port
fontein Subgroup is, in many places, an angular discordance, or the Nolloth Zone (Fölling & Frimmel 2002), have helped in dis-
Kaigas Formation rests directly on the pre-Gariep basement, but a tinguishing between post-Kaigas and post-Numees carbonates
conformable contact exists NE of Rosh Pinah (Fig. 17.4). On the because of distinctly different isotopic trends. The former are
farm Spitzkop 111 north of Rosh Pinah, the Kaigas Formation dia- characterized by an increase in d13CCarb from negative ratios
mictite grades laterally into a coarse-grained, proximal pyroclastic (,4‰ relative to PDB standard) above the Kaigas Formation dia-
deposit of the Rosh Pinah Formation, indicating sedimentation that mictite to values as high as þ8.7‰ further up-section. This trend is
overlapped in time with volcanism. Elsewhere, the background accompanied by a decrease in 87Sr/86SrCarb from 0.7076 to 0.7071.
sedimentation towards the top of the Kaigas Formation gradually The Sr isotope ratios are considered near primary because of
changes from siliciclastic to calcareous; the matrix of the diamic- high Sr (1828 –2916 ppm) and generally low Rb concentrations
tite and intercalated arenite becomes progressively dolomitic (,13 ppm, Rb/Sr 0.007, Fölling & Frimmel 2002). In contrast,
KAIGAS AND NUMEES FORMATIONS 229

the post-Numees carbonates of the Bloeddrif Member have signifi- provenance of the entire Port Nolloth Group on the Kalahari
cantly higher, but consistent 87Sr/86SrCarb ratios of 0.7085 in the Craton. This is supported by detrital zircon age spectra that
limestone immediately above the diamictite and 0.7082 in the conform to the ages known from the immediate basement on the
remainder of the member. Considering Rb/Sr ratios of much Kalahari Craton but differ from those typical for the Rio de la
less than 0.001 (Sr and Rb concentrations of 1078 –2483 and Plata Craton (Basei et al. 2005). Both the Kaigas and the
,5.3 ppm, respectively), the measured 87Sr/86SrCarb ratios are Numees Formation diamictites would have been deposited in a
again regarded as effectively primary (Fölling & Frimmel 2002). marginal cratonic position for which, in most palaeogeographic
The bottom of the Bloeddrif Member limestone is markedly reconstructions, a low palaeolatitude is assumed, though on cir-
depleted in 13C (d13CCarb ¼ 24.6‰), whereas the rest has cumstantial evidence, throughout the Neoproterozoic time span
d13CCarb consistently clustering within the narrow range between of interest here (Collins & Pisarevsky 2005).
21.0 and þ1.0‰. In a follow-up study (Frimmel & Fölling
2004) it could be shown, however, that the increase in d13CCarb
is strongly dependent on palaeoenvironment, with much higher Geochronological constraints
isotope ratios (þ4.0‰) achieved in the most proximal positions
for the same stratigraphic interval. Alternatively, deposition of A maximum age constraint for the onset of sedimentation in the
the cap carbonate could be diachronous from the slope up to the Gariep Basin is given by the youngest age obtained on pre-Gariep
shelf (e.g. Hoffman et al. 2007). basement, that is, 771 + 6 Ma for the Lekkersing Granite
Although O isotopes are more susceptible to diagenetic altera- (Frimmel et al. 2001). The Kaigas Formation must be younger
tion, a distinct negative d18OCarb anomaly (decrease by as much than that age but cannot be younger than the volcanism evident
as 4‰) has been noted in the first 25 m of the post-Numees cap in the Rosh Pinah Formation. Felsic volcanic rocks above the
carbonates in various outcrops and has been interpreted as reflect- Kaigas Formation yielded Pb–Pb and U – Pb single zircon ages
ing freshwater mixing with seawater during large-scale melting of 741 + 6 Ma and 752 + 6 Ma, respectively (Frimmel et al.
of glaciers, in agreement with the slightly higher 87Sr/86SrCarb at 1996b; Borg et al. 2003). This is supported by a double-spike
the same position (Fölling & Frimmel 2002). Pb– Pb carbonate age of 728 + 32 Ma obtained for the overlying
By and large, the post-Numees carbonates are more enriched in cap carbonates of the Pickelhaube Formation (Fölling et al.
Sr than the post-Kaigas carbonates. Strontium concentrations of 2000), which has been interpreted as dating the time of diagenesis.
several thousand ppm have been reported for the distal portions More problematic is the absolute age of the Numees Formation,
(characterized by a lack of Rb, Y and Zr) of the Bloeddrif as no radiometric data are available for this unit. Minimum age
Member limestone, whereas the near-shore equivalents contain constraints are given by a double-spike Pb–Pb carbonate age of
an order of magnitude less Sr (Frimmel & Fölling 2004). 555 + 28 Ma for the overlying limestone of the Bloeddrif
Member of the Holgat Formation (Fölling et al. 2000) and the
age of regional, syn-collisional metamorphism that is constrained
Other characteristics at 545 + 2 Ma by Ar –Ar data on syn-tectonic hornblende and
micas (Frimmel & Frank 1998).
The Rosh Pinah Formation is economically significant for its
base-metal enrichment. Stratabound and largely stratiform sulphi-
dic Pb and Zn ore bodies are being mined at Rosh Pinah (past Discussion
production and resources are c. 50 Mt at an average grade of
7 wt% Zn, 2 wt% Pb, 0.1 wt% Cu and 11 ppm Ag). The ore is The sedimentary features of the Kaigas Formation diamictite, such
hosted by a variably silicified carbonaceous mudstone that as sharp upper and lower bedding contacts, lack of internal struc-
reaches as much as 50 m in thickness, with intercalated carbonate tures (except for crudely graded bedding), correlation between
lenses, within an overall arkosic succession of the middle Rosh clast size and bedding thickness, rapid decrease in clast size
Pinah Formation (Alchin et al. 2005). A protore in a similar strati- towards the basin centre, and the limited geographical distribution
graphic position is regarded as the source for the secondary oxi- suggest rapid deposition as debris flows. Transport by turbidity
dized Zn deposit at Skorpion (25 Mt, average grade of 10.6 wt% currents in a subaqueous fan environment is indicated by com-
Zn; Borg et al. 2003), located c. 22 km NW of Rosh Pinah. mon upwards fining beds in the intercalated feldspathic arenite.
In spite of a low-grade metamorphic overprint, it has been The presence of westward-directed gravity flow deposits stepping
possible to recover a number of organic-walled microfossils that back onto the basement foreland suggests a phase of marine trans-
help in setting some biostratigraphic constraints for units above gression along an active fault scarp. This is also indicated by local
and below the Numees Formation (Gaucher et al. 2005). Poorly dolostone clasts in the diamictite, which implies the presence of
preserved, highly carbonized acritarchs characterize units shallow-water, pre-Kaigas carbonate platform deposits, at least
beneath the Numees Formation. Of biostratigraphic significance in the south.
is the occurrence of Bavlinella faveolata in the Pickelhaube and A component of melt-out glacial debris in the Kaigas Formation
upper Wallekraal Formations. Acritarchs in the Bloeddrif Member, diamictite, such as dropstones, bimodal clast distribution, sand-
Holgat Formation, are characterized by a low-diversity assemblage wiched and faceted clasts, and outsized erratic blocks in shaly
consisting of Soldadophycus, Myxococcoides, Coniunctiophycus beds, suggest a glaciomarine or fluvioglacial origin of parts of
and Leiosphaeridia, with a dominance of Soldadophycus bossii, the Kaigas Formation. Uplifted horsts along the margin of the
and the absence of acanthomorphs and large sphaeromorphs. basin may have been the sites of mountain glaciers that contributed
Associated with this acritarch assemblage is the agglutinated some glacial material.
foraminifer Titanotheca sp. Although there is little doubt of a glacial influence during
Kaigas times, most of the Kaigas Formation reflects sediment
gravity flows in consequence of subaqueous slumping, probably
Palaeolatitude and palaeogeography triggered by tectonic activity along the growing rift shoulders.
In contrast, the Numees Formation represents a truly glacial
No reliable palaeomagnetic data are available for the Gariep Belt deposit. Deposition adjacent to a glaciated continental margin by
or for the location of the Kalahari Craton in the Neoproterozoic processes of settling from suspension and ice rafting is indicated
in general. Today, the Gariep Belt is positioned between the Kala- by the extensive lateral distribution of the diamictite, its textural
hari and the Rio de la Plata Cratons. Both Pb and Nd isotopic evi- homogeneity and largely unbedded nature, and the presence of
dence (Frimmel et al. 2004; Frimmel & Basei 2006) point to a out-sized extrabasinal lonestones. The associated sediments, in
230 H. W. FRIMMEL

particular the banded iron-formation of the Jakkalsberg Member, Borg, G., Kärner, K., Buxton, M., Armstrong, R. & Van der Merwe,
record marine rather than glaciofluvial depositional conditions. S. W. 2003. Geology of the Skorpion zinc deposit, southern Namibia.
Huge blocks, more than 10 m in length, of completely different Economic Geology, 98, 749–771.
lithology (dolomite, gneiss) occurring next to each other within Cloud, P. E. 1974. Giant stromatolites and associated vertical tubes from
the same diamictite bed (commonly found near the Orange the Upper Proterozoic Noonday Dolomite, Death Valley region,
River) can only be explained reasonably by melt-out from drifting eastern California. Geological Society of America, Bulletin, 85,
icebergs. The Bloeddrif Member carbonates are interpreted as a 1869– 1882.
typical cap carbonate sequence. Collins, A. S. & Pisarevsky, S. A. 2005. Amalgamating eastern Gond-
wana: the evolution of the circum-Indian orogens. Earth-Science
Although the Kaigas and Numees Formations have been tenta-
Reviews, 71, 229–270.
tively correlated with the global Sturtian and Marinoan glaciations Corsetti, F. A. & Grotzinger, J. P. 2005. Origin of tube structures in
(Frimmel et al. 2002), this interpretation has been questioned. The Neoproterozoic post-glacial cap carbonates: example from
age of the Kaigas Formation is c. 750 Ma, older than the Sturtian Noonday Dolomite, Death Valley, United Sates. Palaios, 20,
glaciation and, purely based on evidence from the Gariep Belt, 348– 362.
there is insufficient evidence from the Gariep Belt to postulate a Davies, C. & Coward, M. P. 1982. The structural evolution of the
global glacial event for the Kaigas Formation. On a regional Gariep Arc in southern Namibia. Precambrian Research, 17,
scale, the Kaigas glaciogenic deposits may be correlated with the 173– 198.
Chuos Formation of northern Namibia, based on similar radio- De Villiers, J. & Söhnge, P. G. 1959. The geology of the Richtersveld.
metric ages for associated felsic volcanic rocks there (Hoffman Memoirs of the Geological Survey of South Africa, 48, 1– 295.
et al. 1996; Hoffman & Prave 1996). As evidence for glacial Fölling, P. G. & Frimmel, H. E. 2002. Chemostratigraphic correlation of
conditions at that time exists also from other continents (Vieira carbonate successions in the Gariep and Saldania Belts, Namibia and
et al. 2007; Zheng et al. 2007), it may well represent a global South Africa. Basin Research, 14, 69– 88.
glacial event. Fölling, P. G., Zartman, R. E. & Frimmel, H. E. 2000. A novel
The lack of precise radiometric age control for the Numees approach to double-spike Pb –Pb dating of carbonate rocks: examples
Formation precludes a reliable stratigraphic correlation with gla- from Neoproterozoic sequences in southern Africa. Chemical
ciogenic deposits elsewhere. A precise age of 636 + 1 Ma has Geology, 171, 97 –122.
been obtained on an ash bed in metamorphosed glaciomarine dia- Frimmel, H. E. 1995. Metamorphic evolution of the Gariep Belt. South
mictite from the Central Zone of the Damara Belt (Hoffmann et al. African Journal of Geology, 98, 176–190.
2004). These authors consider the dated diamictite unit as equival- Frimmel, H. E. 2000. The Pan-African Gariep Belt in southwestern
Namibia and western South Africa. Communications of the Geologi-
ent to the Ghaub Formation in the northern platform and conse-
cal Survey of Namibia, 12, 197–209.
quently postulate a correlation with the Marinoan glaciation. A Frimmel, H. E. 2003a. Gariep Supergroup. In: Johnson, M. R. (ed.)
correlation of the Numees Formation with the Ghaub Formation Catalogue of South African Lithostratigraphic Units. Council for
appears intuitively probable (as suggested by Frimmel et al. Geoscience, Pretoria, 5 –7.
2002), but is not supported by more recent biostratigraphic evi- Frimmel, H. E. 2003b. Port Nolloth Group. In: Johnson, M. R. (ed.)
dence (Gaucher et al. 2005). The microfossil assemblage in the Catalogue of South African Lithostratigraphic Units. Council for
post-Numees rocks is markedly different from those in the pre- Geoscience, Pretoria, 35 – 37.
Numees strata, and the former compares well with Ediacaran Frimmel, H. E. 2008. The Gariep Belt. In: Miller, R. M. (ed.) The
assemblages elsewhere. An Ediacaran age of the Holgat Formation Geology of Namibia. Geological Survey of Namibia, Windhoek, 2,
is also supported by the Pb–Pb age of 555 + 28 Ma obtained for 14-1 –14-39.
the Bloeddrif cap carbonate (Fölling et al. 2000) and the compara- Frimmel, H. E. & Basei, M. A. S. 2006. Tracking down the Neoprotero-
tively high initial 87Sr/86SrCarb (Fölling & Frimmel 2002). A cor- zoic connection between southern Africa and South America – a
relation of the Numees Formation with the c. 580 Ma Gaskiers revised geodynamic model for SW-Gondwana amalgamation. In:
glaciation is therefore preferred. Thus, if the successor basin Gaucher, C. & Bossi, J. (eds) V. South American Symposium on
model is accepted for the Gariep Basin, the apparent absence of Isotope Geology. 24 –27 August, Punta del Este, 94– 97.
any syn-Marinoan glacial deposits in the Port Nolloth Zone may Frimmel, H. E. & Fölling, P. G. 2004. Late Vendian closure of
be explained by the hiatus between the continental rift stage the Adamastor Ocean: timing of tectonic inversion and
(770– 740 Ma) and the back-arc basin stage (c. 610– 580 Ma) in syn-orogenic sedimentation in the Gariep Basin. Gondwana
the overall basin evolution. Research, 7, 685– 699.
Frimmel, H. E. & Frank, W. 1998. Neoproterozoic tectono-thermal
The South African National Research Foundation is thanked for funding the
evolution of the Gariep Belt and its basement, Namibia/South
research on the Gariep Belt for many years. Reviews by F. MacDonald, R.
Africa. Precambrian Research, 90, 1 –28.
McMiller and G. Halverson significantly improved the original manuscript.
Frimmel, H. E. & Lane, K. 2005. Geochemistry of carbonate beds in the
This work represents a contribution of the IUGS- and UNESCO-funded IGCP
Neoproterozoic Rosh Pinah Formation, Namibia: implications on
(International Geoscience Programme) Projects #512 and #478.
depositional setting and hydrothermal ore formation. South African
Journal of Geology, 108, 5 – 18.
Frimmel, H. E., Hartnady, C. J. H. & Koller, F. 1996a. Geochemistry
and tectonic setting of magmatic units in the Pan-African Gariep Belt,
References Namibia. Chemical Geology, 130, 101– 121.
Frimmel, H. E., Klötzli, U. & Siegfried, P. 1996b. New Pb– Pb single
Alchin, D. J., Frimmel, H. E. & Jacobs, L. E. 2005. Stratigraphic setting zircon age constraints on the timing of Neoproterozoic glaciation and
of the metalliferous Rosh Pinah Formation and the Spitzkop and continental break-up in Namibia. The Journal of Geology, 104,
Koivib Suites in the Pan-African Gariep Belt, southwestern 459– 469.
Namibia. South African Journal of Geology, 108, 19 –34. Frimmel, H. E., Zartman, R. E. & Späth, A. 2001. Dating Neoprotero-
Basei, M. A. S., Frimmel, H. E., Nutman, A. P., Preciozzi, F. & Jacob, zoic continental break-up in the Richtersveld Igneous Complex,
J. 2005. A connection between the Neoproterozoic Dom Feliciano South Africa. The Journal of Geology, 109, 493– 508.
(Brazil/Uruguay) and Gariep (Namibia/South Africa) orogenic Frimmel, H. E., Fölling, P. G. & Eriksson, P. 2002. Neoproterozoic
belts – evidence from a reconnaissance provenance study. Precam- tectonic and climatic evolution recorded in the Gariep Belt,
brian Research, 139, 195–221. Namibia and South Africa. Basin Research, 14, 55 –67.
Beetz, W. 1926. Über Glazialschichten an der Basis der Nama- und Frimmel, H. E., von Veh, M. W. & Fölling, P. G. 2003a. Kaigas For-
Konkipformation in der Namib Südwestafrikas. Neues Jahrbuch f. mation. In: Johnson, M. R. (ed.) Catalogue of South African Litho-
Mineralogie, Geologie und Paläontologie, Abteilung B, 56, 437– 481. stratigraphic Units. Council for Geoscience, Pretoria, 17– 19.
KAIGAS AND NUMEES FORMATIONS 231

Frimmel, H. E., von Veh, M. W. & Fölling, P. G. 2003b. Numees sedimentological evolution of the late Proterozoic/early Palaeozoic
Formation. In: Johnson, M. R. (ed.) Catalogue of South African Gariep Belt, southern Namibia. Communications of the Geological
Lithostratigraphic Units. Council for Geoscience, Pretoria, 25– 28. Survey of Namibia, 8, 99 –118.
Frimmel, H. E., Jonasson, I. & Mubita, P. 2004. An Eburnean base metal Jasper, M. J. U., Stanistreet, I. G. & Charlesworth, E. G. 2000.
source for sediment-hosted zinc-lead deposits in Neoproterozoic units Neoproterozoic inversion tectonics, half-graben depositories and
of Namibia: lead isotopic and geochemical evidence. Mineralium glacial controversies, Gariep fold-thrust belt, southern Namibia.
Deposita, 39, 328–343. In: Miller, R. M. (ed.) Henno Martin Commemorative Volume.
Gaucher, C., Frimmel, H. E. & Germs, G. J. B. 2005. Organic-walled Communications of the Geological Survey of Namibia, Windhoek,
microfossils and biostratigraphy of the upper Port Nolloth Group 1887–1896.
(Namibia): implications for the latest Neoproterozoic glaciations. Kaufman, A. J., Hayes, J. M., Knoll, A. H. & Germs, G. J. B. 1991.
Geological Magazine, 142, 539– 559. Isotopic composition of carbonates and organic carbon from upper
Germs, G. J. B. 1995. The Neoproterozoic of southwestern Africa, with Proterozoic successions in Namibia: stratigraphic variation and the
emphasis on platform stratigraphy and paleontology. Precambrian effects of diagenesis and metamorphism. Precambrian Research,
Research, 73, 137– 151. 49, 301– 327.
Gresse, P. G. 1994. Strain partitioning in the southern Gariep Arc as Kennedy, M. J. 2001. Are Proterozoic cap carbonates and isotopic excur-
reflected by sheath folds and stretching lineations. South African sions a record of gas hydrate destabilization following Earth’s coldest
Journal of Geology, 97, 52 – 61. intervals? Geology, 29, 443–446.
Grotzinger, J. P., Bowring, S. A., Saylor, B. Z. & Kaufman, A. J. Kröner, A. 1974. The Gariep Group, Part I: Late Precambrian forma-
1995. Biostratigraphic and geochronologic constraints on early tions in the western Richtersveld, northern Cape Province. Pre-
animal evolution. Science, 270, 598–604. cambrian Research Unit, University of Cape Town, Bulletin, 13,
Hälbich, I. W. & Alchin, D. J. 1995. The Gariep belt: stratigraphic- 1– 115.
structural evidence for obliquely transformed grabens and back- Martin, H. 1965a. Beobachtungen zum Problem der jung-präkam-
folded thrust stacks in a combined thick-skin thin-skin structural brischen glazialen Ablagerungen in Südwestafrika. Geologische
setting. Journal of African Earth Sciences, 21, 9 –33. Rundschau, 54, 115– 127.
Hartnady, C., Joubert, P. & Stowe, C. 1985. Proterozoic crustal evol- Martin, H. 1965b. The Precambrian geology of South West Africa and
ution in southwestern Africa. Episodes, 8, 236– 244. Namaqualand. Precambrian Research Unit, University of Cape
Hartnady, C. J. H. & von Veh, M. W. 1990. Tectonostratigraphic and Town, Bulletin, 4, 1– 177.
structural history of the Late Proterozoic – Early Palaeozoic Gariep McMillan, M. D. 1968. The geology of the Witputs-Sendelingsdrif area.
Belt, Cape Province, South Africa. Geological Society of South Precambrian Research Unit, University of Cape Town, Bulletin, 4,
Africa, Cape Town. 1– 177.
Hegenberger, W. 1993. Stratigraphy and sedimentology of the Late Middlemost, A. E. K. 1963. Geology of the southeastern Richtersveld.
Precambrian Witvlei and Nama Groups, east of Windhoek. Unpublished PhD thesis, University of Cape Town, Cape Town.
Memoirs of the Geological Survey of Namibia, 17, 1– 82. Reid, D. L., Ransome, I. G. D., Onstott, T. C. & Adams, C. J. 1991.
Hoffman, P. F., Hawkins, D. P., Isachsen, C. E. & Bowring, S. A. 1996. Time of emplacement and metamorphism of Late Precambrian
Precise U–Pb zircon ages for early Damaran magmatism in the mafic dykes associated with the Pan-African Gariep orogeny,
Summas Mountains and Welwitschia inlier, northern Damara belt, Southern Africa: implications for the age of the Nama Group.
Namibia. Communications of the Geological Survey of Namibia, Journal of African Earth Sciences, 13, 531–541.
11, 47 – 52. Rogers, A. W. 1916. The geology of part of Namaqualand. Transactions
Hoffman, P. F., Kaufman, A. J. & Halverson, G. P. 1998. Comings and of the Geological Society of South Africa, 18, 72 – 101.
goings of global glaciations on a Neoproterozoic tropical platform in Siegfried, P. R. & Moore, J. M. 1990. The Rosh Pinah Zn– Pb– Cu– Ag
Namibia. GSA Today, 8, 1– 9. massive sulphide deposit–a product of early rift-related volcanism?
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., In: Geocongress 90. Geological Society of South Africa, Cape
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran Town, 512– 513.
(635 Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Pla- Vieira, L. C., Trindade, R. I. F., Nogueira, A. C. R. & Ader, M. 2007.
netary Science Letters, 258, 114–131. Identification of a Sturtian cap carbonate in the Neoproterozoic Sete
Hoffmann, K.-H. & Prave, A. R. 1996. A preliminary note on a revised Lagoas carbonate platform, Bambuı́ Group, Brazil. Comptes Rendus
subdivision and regional correlation of the Otavi Group based on gla- Geoscience, 339, 240–258.
ciogenic diamictites and associated cap dolostones. Communications von Veh, M. W. 1993. The stratigraphy and structural evolution of the
of the Geological Survey of Namibia, 11, 77 – 82. Late Proterozoic Gariep Belt in the Sendelingsdrif-Annisfontein
Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L. area, northwestern Cape Province. Precambrian Research Unit,
2004. A U– Pb zircon date from the Neoproterozoic Ghaub For- University of Cape Town, Bulletin, 38, 1– 174.
mation, Namibia: constraints on Marinoan glaciation. Geology, 32, Zheng, Y.-F., Wu, Y.-B., Gong, B., Chen, R.-X., Tang, J. & Zhao,
817– 820. Z.-F. 2007. Tectonic driving of Neoproterozoic glaciations: evi-
Jasper, M. U., Stanistreet, I. G. & Charlesworth, E. G. 1992– 1993. dence from extreme oxygen isotope signature of meteoric water
Report: preliminary results of a study of the structural and in granite. Earth and Planetary Science Letters, 256, 196– 210.
Chapter 18

The Karoetjes Kop and Bloupoort formations, Gifberg Group, South Africa

HARTWIG E. FRIMMEL1,2
1
Department of Geological Sciences, University of Cape Town, Rondebosch 7701, South Africa
2
Present address: Geodynamics & Geomaterials Research Division, Institute of Geography and Geology, University of Wuerzburg,
Am Hubland, D-97074 Wuerzburg, Germany
(e-mail: hartwig.frimmel@uni-wuerzburg.de)

Abstract: The Vredendal Outlier near the South African west coast lies in an intermediate position between the late Neoproterozoic
Gariep Belt further north and the Cambrian Saldania Belt further south. It consists of a sedimentary succession of siliciclastic and car-
bonate rocks, unified as the Gifberg Group, which contains two diamictite-bearing units, the Karoetjes Kop Formation at the base and the
Bloupoort Formation near the top of the group. The diamictite of the Karoetjes Kop Formation represents mainly debris flow deposits in a
continental rift setting, with some contribution from retreating glaciers. In contrast, the younger diamticite in the Bloupoort Formation is
glaciomarine, is associated with banded iron-formation, is underlain by stromatolitic reef carbonates and is overlain by carbonates. Most
of the Gifberg Group is poorly exposed and poorly investigated. In the absence of radiometric age data, stratigraphic interpretation and
correlation is based on lithological and chemostratigraphic evidence. The entire Gifberg Group is considered to be equivalent to the much
better investigated Port Nolloth Group in the Gariep Belt (sensu stricto). Whereas the Karoetjes Kop Formation is correlated with the
c. 750 Ma Kaigas Formation of the Gariep Belt, the diamictite and associated banded iron-formation of the Bloupoort Formation are
regarded as correlatives of the Numees Formation of the Gariep Belt. The entire Gifberg Group was subjected to transpressional defor-
mation and accompanying low-grade metamorphism during continental collision between the Rio de la Plata and Kalahari plates at the
end of the Neoproterozoic and again during the Cambrian accretionary orogeny along the southwestern margin of Gondwana, which led
to the development of the Saldania Belt further south.

The Karoetjes Kop and Bloupoort Formations both contain dia- between the two diamictite-bearing formations of the Gifberg
mictite. They occur in a succession of metasedimentary Neoproter- Group (Frimmel 2008) helped in better constraining the strati-
ozoic rocks that are relatively poorly exposed along the South graphic position of that group.
African west coast between latitudes 308S and 328S and in the
plains that stretch from the coast up to 80 km inland in the area
around Vanrhynsdorp (Figs 18.1 and 18.2). This area was first Structural framework and tectonic evolution
mapped by Rogers (Rogers & Schwarz 1904; Rogers 1911). Sub-
sequent studies (Brink 1950; Jansen 1960; Kröner 1968; Lamont Although a polyphase deformational history is evident from field
1947) led to rather contrasting stratigraphic interpretations. The observations and structural data (Gresse 1992), the overall struc-
basis for the current stratigraphic subdivision is the work of Gresse ture appears to be relatively simple with a gentle dip of primary
(1992), who re-mapped large parts of the area and studied the sedi- stratigraphic contacts towards the SW. The local distribution of
mentology, structure and metamorphism. Originally, Gresse rock types can be largely explained by the interference between
included this metasedimentary succession as the Gifberg Subgroup two generations of folds (F2 and F3). Earlier F1 folds are rarely pre-
in the Vanrhynsdorp Group. The younger part of that group has served. As all fold axes follow a similar trend that is subparallel to
since been recognized as a correlative of the lower to middle the orogenic belt and colinear with the stretching direction, a trans-
Nama Group in Namibia and represents the fill of a southern sub- pressive regime is inferred. The style of deformation is identical to
basin of the larger Nama foreland basin (Gresse & Germs 1993). that in the Gariep Belt further north, where it has been explained
The Gifberg rocks occur unconformably below these foreland by sinistral transpression (Von Veh 1993; Gresse 1994). Similarly,
basin deposits and bear strong lithological similarities to sedimen- the deformation in the Vredendal Outlier can be explained by sinis-
tary successions of the Neoproterozoic Port Nolloth Group in the tral transpression that reflects compression towards the east with
Gariep Belt further north. Subsequently, they were, therefore, SSE-verging folds and thrusts (F1), west-verging back-folds and
described as a separate group, the Gifberg Group (de Beer et al. back-thrusts (F2), ENE-verging F3 folds and conjugate kinks.
2002). Their area of distribution is referred to as the Vredendal By analogy with findings from the Gariep Belt further north
Outlier, which occupies an intermediate position between the (Frimmel & Frank 1998), an age of c. 545 Ma is assigned to F1
Gariep Belt in the north and the Saldania Belt in the south. and F2. The F3 fold- and thrust-event also affected the siliciclastic
A glacial influence during deposition of the Karoetjes Kop sedimentary rocks of the Vanrhynsdorp Group (Gresse 1992) and
Formation was indirectly inferred by Gresse (1992), who was the is thus younger. Ar – Ar age data on syn-F3 micas are in the range
first to compare the diamicite of that formation with the glacio- 500– 480 Ma (Gresse et al. 1988; Frimmel & Frank 1998).
genic Kaigas Formation diamictite in the Port Nolloth Zone. Syn-orogenic deformation was accompanied by regional meta-
Less ambiguous is the glacial origin in the case of the diamictite morphism. The grade of metamorphism increases from the fore-
within the younger Bloupoort Formation. Although it was des- land in the east towards the west. Although the Vanrhynsdorp
cribed originally as agglomerate (Rogers & Schwarz 1904), Group rocks in the east have experienced only sub-greenschist
general agreement exists now on its glaciogenic character follow- facies conditions, those near the thrust contact with the rocks of
ing a more detailed description of the rock association by Buehr- the Gifberg Group reached greenschist-facies conditions. Conse-
mann (1981). The correlation of individual formations between quently, this phase of dynamic metamorphism must have taken
the Port Nolloth and Gifberg Groups has remained speculative, place after the deposition of the Vanrhynsdorp foreland basin fill
and uncertainty exists particularly with regard to the age of the and it was syn-kinematic with respect to F3. An increase in the
Bloupoort Formation diamictite (de Beer et al. 2002). Recent litho- metamorphic grade from lowest greenschist facies near Vanrhyns-
geochemical and isotopic data on carbonates that are positioned dorp to lowermost amphibolite facies at the Atlantic coast is

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 233– 237. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.18
234 H. E. FRIMMEL

starts with a white to blue-grey basal conglomerate in which the


basement lithotypes appear reworked. This is followed by a grey,
fine-grained quartzitic greywacke in which thin, well-rounded
quartz (or quartzite) pebble beds are intercalated. In places, this
greywacke rests directly on the basement. A massive, poorly
bedded, laterally discontinuous diamictite follows gradationally
above the quartzitic greywacke. It reaches a maximum thickness
of a few tens of metres and contains poorly sorted pebbles,
cobbles and boulders of basement-derived gneiss, quartzite, vein
quartz and biotite schist, all of which are set in a metamorphosed
biotite-rich matrix. In places, the diamictite rests directly on base-
ment and grades laterally into the quartzitic greywacke.
Above the diamictite follows a reddish weathering, laterally
continuous feldspathic quartzite with intercalated metapelite (ser-
icite schist). Large trough cross-bedded sets are common, reaching
several metres in thickness. This lithotype grades upwards into
thick-bedded, coarse-grained, white to light blue-grey quartzite.
No complete section through the entire formation is exposed.
Consequently, no information on the true stratigraphic thickness
is available.

The Widouw Formation

Carbonate intersections of more than 200 m were recorded in bore-


holes through the central part of the Vredendal Outlier (Gresse
1992), but the true thickness may be less due to tectonic thickening
by isoclinal folding. The carbonate beds thin out towards the north
where a basement high probably existed throughout Neoprotero-
zoic to early Cambrian times (Kamieskroon Ridge, Fig. 18.1).
Variably coloured (from white, pink, ochre to black), flat lami-
nated, medium- to thick-bedded limestone dominates the lower
part of the formation. In several places, individual beds are later-
ally discontinuous, irregular and reflect former dissolution surfaces
(Frimmel 2008). Although the limestone unit is generally very
uniform, intra-formational syn-sedimentary limestone breccia
beds as well as stratiform black chert boudins occur in the upper
parts. Contorted dolomitized beds reach 20–30 cm in thickness
and contain white blades of calcite, several centimetres long,
which have been compared with herringbone calcite and inter-
Fig. 18.1. Map showing the location of the Vredendal Outlier within the
preted as pseudomorphs after enterolithic gypsum (Frimmel
network of Pan-African orogenic belts of southwestern Africa.
2008). Impure limestone contains as much as 40 vol% coarse-
grained, angular quartz clasts, which indicate an overall proximal
setting for the depositional environment of these carbonates.
evident from metapelites of the Gifberg Group, with kyanite- Marked differences in grain size reflect different degrees of
staurolite-garnet-micaschist developed in the extreme NW of the metamorphic recrystallization. Several types of dolostone, ranging
Vredendal Outlier. in relative age from syn-sedimentary/early diagenetic to meta-
morphic, can be distinguished. The former includes stromatolitic
dolostone (Gresse 1992).
Stratigraphy

The oldest unit of the Gifberg Group (Fig. 18.3) is the mixed sili- The Aties Formation
ciclastic, diamictite-bearing Karoetjes Kop Formation. Variably
recrystallized limestone and dolostone, as well as minor meta- Although the lower contact of the Aties Formation with the under-
greywacke, quartzite and phyllite of the Widouw Formation lying carbonates of the Widouw Formation is typically sheared, an
follow above the Karoetjes Kop Formation. The overlying Aties originally conformable contact is inferred. The Aties Formation is
Formation is dominated by carbonaceous schist, with subordinate made up of predominantly black carbonaceous, pyritic schist with
phyllite to biotite-schist (in the westernmost exposures kyanite- interbedded phyllite to biotite schist, meta-arenite and thin, impure
bearing), meta-arenite, and impure, thin carbonate beds. The limestone and dolostone beds. The arenitic beds are quartz-
youngest formation is the Bloupoort Formation, which comprises sandstone and greywacke in composition, but some of them are
a variety of carbonate rocks, calcarenite, diamictite, pyritic and highly feldspathic (arkose), coarse-grained and grade in places
graphitic phyllite, ferruginous chert and iron-formation. into intraformational feldspathic conglomerate beds. Many of the
well-bedded greywacke beds display normal grading and trough
cross-bedding on a metre scale. Locally, monomictic quartz
Glaciogenic deposits and associated strata pebble conglomerate occurs within white quartzite. The presence
of, inter alia, conglomerate and diamictite, associated with lime-
The Karoetjes Kop Formation stone and dolostone has been reported as ‘interbedded’ with phyl-
lite of the Aties Formation (de Beer et al. 2002). It is suspected,
This basal formation overlies high-grade metamorphic gneisses however, that this diamictite-bearing succession is part of the
of the Mesoproterozoic Namaqualand Metamorphic Complex. It overlying Bloupoort Formation.
THE KAROETJES KOP AND BLOUPOORT FORMATIONS 235

Fig. 18.2. Areal distribution of the


individual formations that constitute the
Gifberg Group.

The Bloupoort Formation Boundary relations with overlying and underlying


non-glacial units
The distinction of this formation from the underlying Aties Forma-
tin has been problematic and prompted Gresse (1992) to include all The lower contact of the diamictite in the Karoetjes Kop Forma-
the Bloupoort rocks as a separate member within the Aties For- tion with the underlying quartzitic greywacke is gradational.
mation. Subsequently, it was recognized that they are younger Where the diamictite rests directly on basement gneiss, the
and represent a relatively complex unit, and their stratigraphic contact is an erosional unconformity. The upper contact of this
rank was raised to that of formation (de Beer et al. 2002). diamictite with the overlying quartzite is sharp and conformable.
The main distribution of this formation is to the SE of Vanrhyns- The contact between the Karoetjes Kop Formation and the
dorp (Fig. 18.2), where a mixed carbonate unit grades into a Widouw Formation is not exposed, and the relative age relation-
diamictite-dominated unit. The former is almost entirely dolomitic ship between the two formations is only inferred from their
and comprises laminated, stromatolitic (Conophyton-like), oolitic, spatial distribution.
intraclastic, conglomeratic, and brecciated and siliceous dolo- The lower contact of the diamictite-hosting Bloupoort For-
stone. Sandy to gritty dolostone is rich in variably rounded mation is typically tectonic. The upper contact of the formation
quartz clasts and contains up to 15 vol% feldspar (Gresse 1992). is either a thrust-fault, or it is covered by younger sedimentary
Oolites, pelletoids, pellets and lumps are common and were depos- rocks of the Cape Supergroup or by aeolian sand. Within the for-
ited in carbonate mud (micrite). Dolomitic breccia consists of mation, the lower boundary of the diamictite with the underlying
micritic fragments that are cemented by sparry dolomite or micrite. dolostone is gradational, with an intervening dolomitic arenite,
This dolomitic unit grades upwards into dolomitic arenite and whereas the upper boundary against the overlying dolostone is
ferruginous diamictite, which constitute the Swartleikrans Bed sharp and appears conformable. The entire succession is,
of Buehrmann (1981). The diamictite consists of very poorly however, strongly deformed. A fivefold rhythmical repetition of
sorted, angular fragments and subrounded pebbles, cobbles and an ‘interbedded sequence of dolomite, ironstone and ferruginous
boulders of basement-derived orthogneiss and granulite, as well mixtite’, each about 5 –15 m in thickness, described by Gresse
as basin-derived dolostone and arenite, and minor ferruginous (1992), could well reflect tectonic repetition.
chert, jasper, vein quartz, limestone, oolitic dolostone and feldspar
grains, all of which are set in a ferruginous, hematite-rich argillitic
matrix. Bedding is only very poorly developed in the overall Chemostratigraphy
massive diamictite, but in places very crude graded bedding
from diamictite into ferruginous sandstone and siltstone to argillite Limited trace element analyses as well as C, O and Sr isotope data
can be observed on a metre scale. Overall, the proportion of inter- have been reported for carbonates of the Widouw Formation
bedded ferruginous argillite increases towards the top of the unit. (Frimmel 2008). Limestone of this formation is very rich in Sr
Laterally, the diamictite grades into banded iron-formation with (as much as 4728 ppm) and has extremely low Rb/Sr ratios.
isolated dropstones and interbedded chert and jasper (Rooihoogte Their 87Sr/86Sr ratios can be taken as near-primary and they
Bed of Buehrmann 1981). range consistently throughout the succession between 0.70844
Locally, the diamictite is overlain by thinly laminated pink and 0.70847. The C and O isotope ratios in the limestone cover
dolostone and dolomitic calcarenite. Intense thrust-faulting and a wide range of d13C between – 4.2 and þ4.8‰ (relative to
folding in all of these units makes it virtually impossible to con- V-PDB) and d18O between –9.4 and –3.1 (relative to V-PDB).
strain their true sedimentary thickness. Syn-sedimentary or early diagenetic dolostone is enriched in 18O
236 H. E. FRIMMEL

(e.g. Collins & Pisarevsky 2005). The depositional basin for the
Vanrhynsdorp Group, undifferentiated Gifberg Group was most likely the southern continuation of the
(c. 550–530 Ma) Gariep Basin, specifically the basins in which the sediments of
the Port Nolloth Group were deposited. Today the Gariep Belt is
positioned between the Kalahari and the Rio de la Plata Cratons.
Both Pb and Nd isotopic evidence (Frimmel et al. 2004; Frimmel
& Basei 2006) point to a provenance of the entire Gariep Super-
group, including the Gifberg Group, on the Kalahari Craton. For
Bloupoort
parts of the Widouw Formation, an evaporitic influence has been
Formation suggested based on petrographic observations, trace element distri-
bution and isotope data (Frimmel 2008), which speaks against a
high-latitude position at the time of sedimentation.
Gifberg Group

Geochronological constraints
Aties
Formation No radiometric age data that might help in constraining the absol-
ute age of the diamictite units in the Gifberg Group are available.
By analogy with the Port Nolloth Group in the eastern Gariep Belt,
Widouw the maximum age of sedimentation is given by the youngest age
Formation obtained on the pre-Gariep basement. This is a U –Pb single
zircon age of 771 + 6 Ma for the Lekkersing granite (Frimmel
et al. 2001). The minimum age for sedimentation is given by the
age of metamorphism. Ar – Ar age data obtained on syn-tectonic
muscovite and biotite from coastal outcrops in the west of the
Karoetjes Kop Vredendal Outlier (Frimmel & Frank 1998) range from 513
Formation to 491 Ma. This range includes a Ar – Ar whole-rock age of
496 + 2 Ma obtained on a phyllite from the Vanrhynsdorp
Group (Gresse et al. 1988). These ages have been interpreted to
date a late Pan-African tectonic pulse that led to NE-directed
thrusting of both the Gifberg Group and the younger Vanrhynsdorp
Basement (1.0 Ga) Group. It is therefore likely that the first regional metamorphism
that affected the Gifberg Group rocks took place earlier during
Limestone, the continental collision phase of the Gariepian orogeny. By
Arenite, rudite analogy with results obtained on the Port Nolloth Group (Frimmel
dolostone
& Frank 1998) and the Nama Group (Grotzinger et al. 1995)
Iron Formation Argillite further north, an age of c. 545 Ma is assumed for that stage of meta-
morphism and tectonism that caused the unconformity between the
Diamictite Gneiss, schist Gifberg and Vanrhynsdorp Groups.

Fig. 18.3. Generalized lithostratigraphy of the Gifberg Group (modified from


de Beer et al. 2002 and Frimmel & Fölling 2004). Discussion

No evidence exists for any marine influence in the Karoetjes Kop


relative to the limestone and it shows a similar, but narrower vari- Formation for which a narrow continental rift, dominated by allu-
ation in d13C. vial fans, plains and fan deltas, is indicated. The depositional
environment envisaged for this formation is therefore similar to
that inferred for the Stinkfontein Subgroup and the Kaigas For-
Other characteristics mation in the Port Nolloth Zone (Chapter 17), except for a total
lack of any evidence for marine influence in the Karoetjes Kop
Calcitic marble of the Widouw Formation is mined at several quar- Formation. A number of very small outcrops of the Karoetjes
ries in the vicinity of Vredendal. The pyritic schists of the Aties Kop Formation also exist along the coast further to the NW of
Formation are strongly weathered with numerous quartz veins the Vredendal Outlier, thus suggesting a continuation of the
that carry Cu, Fe and Mn staining. Numerous surface Fe-oxide former basin margin from the Port Nolloth Zone southwards into
deposits (‘gossans’) are derived from the palaeoweathering of the area around Vredendal. Consequently, the current eastern
the pyrite-rich beds on a Paleocene palaeosurface, the so-called boundary of these Neoproterozoic rocks would be a good approxi-
African Surface (Partridge & Maud 1987). mation of the original eastern basin margin. As with the Kaigas
No biostratigraphic data have been published so far for rocks Formation, the diamictite in the Karoetjes Kop Formation is
of the Gifberg Group. interpreted as mainly a debris-flow deposit because of its lack of
internal structure and very limited geographical distribution
(lateral discontinuity) along the basin edge. By analogy with the
Palaeolatitude and palaeogeography Kaigas Formation, a fluvio-glacial origin for parts of the formation
is suggested.
No reliable palaeomagnetic data are available for the Vredendal The isotopic record of the Widouw Formation carbonates has
Outlier. The low-grade regional metamorphic overprint most been interpreted as reflecting an overall 13C-depleted sea, with
likely led to the remagnetization of the rocks during the the higher d13C ratios resulting from elevated evaporation rates
Pan-African/Gariepian orogeny. Thus, any assumption about rather than changes in the global seawater composition (Frimmel
palaeolatitude and palaeogeography is based only on circumstan- 2008). These shallow-water marine deposits are therefore corre-
tial evidence and comparison with continental-scale models lated with the post-Kaigas transgression.
THE KAROETJES KOP AND BLOUPOORT FORMATIONS 237

The regional and global correlation of the diamictite in the Blou- Frimmel, H. E. & Frank, W. 1998. Neoproterozoic tectono-thermal
poort Formation remains problematic. Little doubt exists about the evolution of the Gariep Belt and its basement, Namibia/South
glacial nature of the diamictite. The association with, and lateral Africa. Precambrian Research, 90, 1 – 28.
transition into, banded iron-formation with occasional lonestones Frimmel, H. E., Jonasson, I. & Mubita, P. 2004. An Eburnean base metal
and dropstones clearly indicate a glaciomarine origin. The appear- source for sediment-hosted zinc-lead deposits in Neoproterozoic units
ance of the underlying carbonates, which bear a remarkable resem- of Namibia: lead isotopic and geochemical evidence. Mineralium
blance to those of the Dabie River Formation in the Port Nolloth Deposita, 39, 328–343.
Zone, the ferruginous matrix of the diamictite, its massive struc- Frimmel, H. E., Zartman, R. E. & Späth, A. 2001. Dating Neoprotero-
zoic continental break-up in the Richtersveld Igneous Complex,
ture and its association with iron-formation are all features that
South Africa. The Journal of Geology, 109, 493 – 508.
are identical to the Numees Formation in the Port Nolloth Zone. Gresse, P. G. 1992. The tectono-sedimentary history of the Vanrhyns-
In the absence of radiometric and biostratigraphic evidence, the dorp Group. Memoirs of the Geological Survey of South Africa, 79,
lower dolomitic unit of the Bloupoort Formation is correlated 1– 163.
with the Dabie River Formation, the diamictite and iron-formation Gresse, P. G. 1994. Strain partitioning in the southern Gariep Arc as
with the Numees Formation, and the upper carbonates and silici- reflected by sheath folds and stretching lineations. South African
clastic rocks with the Holgat Formation. Journal of Geology, 97, 52– 61.
Gresse, P. G., Fitch, F. J. & Miller, J. A. 1988. 40Ar/39Ar dating of the
Fieldwork in the Vredendal area was supported by Kumba Resources. Funding by Cambro-Ordovician Vanrhynsdorp tectonite in southern Namaqua-
the South African National Research Foundation is gratefully acknowledged. land. South African Journal of Geology, 91, 257–263.
F. MacDonald and G. Halverson are thanked for helpful comments on the original Gresse, P. G. & Germs, G. J. B. 1993. The Nama foreland basin:
manuscript. This work represents a contribution of the IUGS- and UNESCO- sedimentation, major unconformity bounded sequences and
funded IGCP (International Geoscience Programme) Projects #512 and #478. multisided active margin advance. Precambrian Research, 63,
247– 272.
Grotzinger, J. P., Bowring, S. A., Saylor, B. Z. & Kaufman, A. J.
References 1995. Biostratigraphic and geochronologic constraints on early
animal evolution. Science, 270, 598–604.
Brink, W. C. 1950. The geology, structure and petrology of the Nuwerus Jansen, H. 1960. The geology of the area around Nieuwoudtville. In: Von
area. Annals of the University of Stellenbosch, 16, 97 –221. Backström, J. W. (ed.) Explanation Sheet 241 (Nieuwoudthville).
Buehrmann, H. T. 1981. The geology of the Lower Kobe Valley, Geological Survey of South Africa, Pretoria, 49.
Vanrhynsdorp District. Annals of the University of Stellenbosch, 3, Kröner, A. 1968. The gneiss-sediment relationship northwest of
Series A1, 67 –144. Vanrhynsdorp, Cape Province. Bulletin, Precambrian Research
Collins, A. S. & Pisarevsky, S. A. 2005. Amalgamating eastern Gond- Unit, University of Cape Town, 3, 1– 233.
wana: the evolution of the circum-Indian orogens. Earth-Science Lamont, G. T. 1947. The geology of part of the Van Rhynsdorp Division,
Reviews, 71, 229–270. Cape Province. Unpublished PhD thesis, University of Cape Town,
de Beer, C. H., Gresse, P. G., Theron, J. N. & Almond, J. E. 2002. The Rondebosch.
Geology of the Calvinia Area, Explanation Sheet 3118 Calvinia Partridge, T. C. & Maud, R. R. 1987. Geomorphic evolution of southern
1:250000 scale. Council for Geoscience, Pretoria. Africa since the Mesozoic. South African Journal of Geology, 90,
Frimmel, H. E. 2008. An evaporitic facies in Neoproterozoic post-glacial 179– 208.
carbonates: the Gifberg Group, South Africa. Gondwana Research, Rogers, A. W. 1911. Report on the geological survey of parts of the div-
13, 453– 468. isions of Van Rhyn’s Dorp, and Namaqualand. Annual Report of the
Frimmel, H. E. & Basei, M. A. S. 2006. Tracking down the Neoprotero- Geological Commission, Cape of Good Hope, 1911, 7– 84.
zoic connection between southern Africa and South America – a Rogers, A. W. & Schwarz, E. H. L. 1904. Geological survey of the north-
revised geodynamic model for SW-Gondwana amalgamation. In: western part of Van Rhyn’s Dorp. Annual Report of the Geological
Gaucher, C. & Bossi, J. (eds) V. South American Symposium on Commission, Cape of Good Hope, 1904, 9– 46.
Isotope Geology. 24 –27 August, Punta del Este, 94 –97. Von Veh, M. W. 1993. The stratigraphy and structural evolution of the
Frimmel, H. E. & Fölling, P. G. 2004. Late Vendian closure of the Late Proterozoic Gariep Belt in the Sendelingsdrif-Annisfontein
Adamastor Ocean: timing of tectonic inversion and syn-orogenic area, northwestern Cape Province. Precambrian Research Unit,
sedimentation in the Gariep Basin. Gondwana Research, 7, 685– 699. University of Cape Town, Bulletin, 38, 1– 174.
Chapter 19

The Ayn Formation of the Mirbat Group, Dhofar, Oman

PHILIP A. ALLEN1*, RUBEN RIEU2, JAMES L. ETIENNE3, ALBERT MATTER4 & ANDREA COZZI5
1
Department of Earth Science and Engineering, Imperial College London, South Kensington Campus, London SW7 2AZ, UK
2
TOTAL E&P Nederland, Bordewijklaan 18, Den Haag, The Netherlands
3
Neftex Petroleum Consultants Ltd, 97 Milton Park, Abingdon, Oxfordshire OX14 4RY, UK
4
Institute of Geological Sciences, University of Bern, Baltzerstrasse 1, CH-3012 Bern, Switzerland
5
ENI Angola, Rua Nicola Gomes Spencer, 140, PO Box 1289, Luanda, Angola
*Corresponding author (e-mail: philip.allen@imperial.ac.uk)

Abstract: Glacial deposits are found in the Ayn Formation and Shareef Formation of the Mirbat Group close to Mirbat in Dhofar,
southern Oman. The Mirbat Group is most likely a correlative of the Abu Mahara Group of the Huqf Supergroup of northern Oman.
The Ayn Formation, the main subject of this chapter, comprises ,400 m of mainly coarse-grained glaciogenic deposits, ponded in 2-
to .8-km-wide N- to NW-oriented palaeovalleys eroded into crystalline basement, with few or no deposits preserved on intervening
palaeohighs. The Shareef Formation occurs as thin, lenticular, erosional remnants beneath the unconformably overlying Cretaceous.
The Ayn Formation is overlain by a thin (,3 m), discontinuous cap carbonate that passes from carbonate-cemented talus on the
basin margin to stromatolitic carbonate on palaeohighs and resedimented gravity flows on palaeovalley flanks.
The Ayn Formation is younger than its youngest detrital zircons and the youngest late plutons in crystalline basement, constraining it
to ,c. 720 Ma, but its exact age is unknown. The detrital zircon population comprises exclusively Neoproterozoic sources, suggesting
derivation from the juvenile Neoproterozoic crust of the Arabian area.
The composition of fine-grained matrix in glaciogenic diamictite units and of non-glacial mudstones, plotted using the chemical index
of alteration (CIA), suggests strong variations in the intensity of palaeoweathering on contemporary land surfaces between the mechan-
ical weathering-dominated Ayn Formation, and the chemical weathering-dominated overlying Arkahawl Formation, which supports the
notion of major glaciation followed by rapid climatic transit as basin margins were flooded and buried with sediment during post-glacial
transgression. The carbon isotopic ratio (d13C) of the post-glacial carbonate is strongly variable from 23.5‰ to þ5.8‰, whereas
carbonate fissures in the underlying basement range between þ4.1‰ and þ5.7‰.
Two independent palaeomagnetic studies have yielded low palaeomagnetic latitudes for the Mirbat Group.

Precambrian sedimentary rocks have been known in Oman since lacustrine deposits. Rieu et al. (2006) focused on a number of pre-
early surveys of the Dhofar area of southern Oman (Carter 1852; viously undescribed outcrops and interpreted the Ayn Formation in
Lees 1928) when the Mirbat Sandstone was recognized. Subse- the context of the whole outcrop area north and west of Mirbat.
quently, a fourfold division of the sedimentary rocks in the Huqf Many Neoproterozoic sedimentary successions are preserved in
area of east-central Oman was devised (Henson & Elliot 1958; marine basins strongly influenced by subaqueous gravity flows,
Morton 1959; Beydoun 1960, 1964), and the Neoproterozoic – leading to an active debate on the glacial v. non-glacial origin of
Early Cambrian succession referred to as the Huqf Group (Glennie the poorly sorted deposits labelled diamictites (Schermerhorn
1977; Gorin et al. 1982). The Huqf Group was later raised in status 1974; Hambrey & Harland 1981; Young & Gostin 1991; Young
to a Supergroup, comprising Abu Mahara, Nafun and Ara Groups 1992; Ross et al. 1995; Arnaud & Eyles 2002b; Allen et al.
(Glennie et al. 1974; Gorin et al. 1982; Hughes-Clarke 1988; 2004; Eyles & Januszczak 2004, 2007). In contrast, the Ayn For-
Wright et al. 1990), the former containing glaciomarine deposits mation records a terrestrial to marginal marine succession, and
in the Jabal Akhdar region of northern Oman. Rieu et al. (2006, therefore provides a valuable example of a glaciated basin margin
2007a) and Allen (2007) recommended the use of the term where glaciofluvial and glaciodeltaic deposits are well developed.
Mirbat Group, to include the Ayn, Arkahawl, Marsham and The Ayn Formation is overlain by a discontinuous cap carbonate
Shareef Formations, in Dhofar. The Mirbat Group is unfossilifer- that exhibits large and abrupt lateral changes in sedimentary facies.
ous and comprises ,400-m-thick glacially influenced deposits The presence of glaciogenic diamictites in thin (,40 m) iso-
of the Ayn Formation, overlain by c. 1 km of non-glacial marine lated remnants beneath the erosional sub-Cretaceous unconfor-
deposits of the Arkahawl and Marsham Formations, followed by mity in Jabal Samhan was recognized by Kellerhals (1993) and
a ,40-m-thick unit of diamictite of the Shareef Formation. confirmed by Rieu (2006), who termed them the Shareef
The precise correlation of the Mirbat Group with the Huqf Formation. Their presence demonstrates a second glacial epoch
Supergroup of east-central Oman and the Jabal Akhdar of northern preserved in the Mirbat Group, but the paucity of outcrop limits
Oman has been a subject of ongoing debate, largely because of the further studies.
inadequacy of the geochronological database. Recent publications
(Rieu et al. 2006, 2007a; Allen 2007) correlate the Mirbat Group of
Dhofar with the Abu Mahara Group of northern Oman, whereas Structural framework and basement geology
Kilner et al. (2005) correlated it with the Nafun Group.
The glaciogenic origin of the Ayn Formation was first recog- The Mirbat Group crops out superbly along a 20 km NE –SW strik-
nized by Qidwai et al. (1988) and the central part of its outcrop ing escarpment near the town of Mirbat (Fig. 19.1). It comprises up
area was studied sedimentologically by Kellerhals (1993) and to 1.5 km of weakly deformed, almost entirely siliciclastic rocks
Kellerhals & Matter (2003), who reported a wide range of facies that generally dip c. 108 to the NW. In the northeastern part of
including subglacial, glaciomarine, fluvial, deltaic, turbiditic and the outcrop belt, a sub-vertical fault juxtaposes the lower part of

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 239– 249. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.19
240 P. A. ALLEN ET AL.

Fig. 19.1. (a) Outcrop areas of Neoproterozoic basement and sedimentary rocks in Oman. (b) Geological map of the Mirbat area showing the location of palaeovalleys
discussed in the text. (c) Summary stratigraphic column of the Mirbat Group. Based on Rieu et al. (2006) and Rieu et al. (2007a, b).

the Mirbat Group and basement with the sub-Cretaceous unconfor- sediment transport direction in the Mirbat Group derived from
mity (see map in fig. 3 of Rieu & Allen 2008), and in the southern palaeocurrent data was to the NW (Kellerhals 1993; Kellerhals
extremity of the outcrop belt small thrust faults cutting the Mirbat & Matter 2003; Rieu et al. 2006, 2007b; Rieu & Allen 2008).
granodiorite fold the overstepping carbonate and terminate This direction would therefore correspond to a transverse supply
upwards in the Arkahawl Formation. In general, however, the from a margin situated to the SE.
Mirbat area escaped the deformation associated with the The Mirbat Group unconformably overlies a complex crys-
‘Western Deformation Front’ to the NW (Allen 2007). talline and metamorphic basement (Platel et al. 1992a, b; Hauser
It has been suggested that the Mirbat Group was formed on the & Zurbriggen 1992, 1994; Würsten 1994; Worthing 2005; Mer-
southeastern flank of a c. 1000-km-long NE –SW striking rift basin colli et al. 2006). Four major units are recognized in the basement
imaged from the subsurface (Loosveld et al. 1996). The general at Mirbat (Mercolli et al. 2006; Bowring et al. 2007): (i) the Juffa
THE AYN FORMATION 241

Group, composed of metasedimentary mica-gneisses and amphi- (ii) fluviodeltaic, (iii) proximal glaciomarine and (iv) distal glacio-
bolites with a .1300 Ma (Mesoproterozoic) siliciclastic sedimen- marine. Details can be found in Rieu et al. (2006).
tary protolith; (ii) the Sadh Group, comprising highly deformed
and metamorphosed banded gneisses intruded by slightly Subaerial facies association. Units of the subaerial facies associ-
deformed diorites and tonalites (Mahall Complex); (iii) the Tona- ation, generally ,1 m thick, unconformably overlie crystalline
lite Group, composed of three large calc-alkaline plutons dated and metamorphic basement at many localities, and comprise
780– 800 Ma; and (iv) the Granite Group, comprising different in situ brecciated bedrock interpreted as regolith, and poorly
types of dykes and small intrusive bodies including the Pegmatite sorted, clast-supported breccia and crudely stratified diamictite
Complex (770 –750 Ma), and the Mirbat Granodiorite thought to represent slope deposits. Clast type commonly reflects
(706 + 40 Ma), Leger Granite (726 + 0.4 Ma) and Shaat Dyke the directly underlying bedrock lithology. There is no evidence
Swarm (c. 750 –700 Ma). A phase of accretion of juvenile Neopro- such as striated clasts or an underlying striated pavement that sup-
terozoic crust (.900 Ma) was followed by extensive calc-alkaline ports a glacial origin of the diamictites in this facies association.
magmatism, amphibolite facies metamorphism and intense defor-
mation culminating at c. 800 Ma. Worthing (2005) believed the Fluviodeltaic facies association. The fluviodeltaic facies association
NW –SE orientated calc-alkaline Shaat Dyke Swarm to originate comprises mainly coarse-grained sandstones and conglomerates
from subduction-modified lithospheric mantle, mobilized by that make up the lower part of the stratigraphy of palaeovalleys
crustal extension. The dykes appear to be similar in orientation 1, 2 and 3. In palaeovalley 3, this facies association is also found
and geochemistry to a dyke swarm (c. 700 Ma) in the Al Bayda higher up in the stratigraphy in two 20–30-m-thick units, alter-
island-arc terrane of Yemen (Windley et al. 1996). Late stocks nated with glaciomarine deposits. Characteristic architectural
such as the Mirbat Granodiorite (and probably the Leger elements include ,25-m-wide and ,2-m-thick channel fills,
Granite) and the Shaat dykes were most likely exhumed by the and 10– 15-m-thick, stacked tabular foreset beds with a c. 258
time of deposition of the oldest sedimentary rocks of the Mirbat apparent dip.
Group (Qidwai et al. 1988; Platel et al. 1992a, b; Kellerhals & Massive, graded and cross-stratified sandstones and conglomer-
Matter 2003; Rieu et al. 2006, 2007a). ates are thought to represent low-sinuosity, bedload-dominated
fluvial deposits, whereas the stacked tabular units are Gilbert-type
deltas that identify the delta front. Interbedded siltstones were
Stratigraphy deposited by hemipelagic fallout in pro-delta bottomsets, and
dropstone-bearing silty laminites demonstrate rainout from float-
The ,1.5-km-thick Mirbat Group comprises a lowermost glacial ing ice (Ovenshine 1970; Smith 2000) in a background glaciomar-
interval (,400-m-thick Ayn Formation) occupying deep palaeo- ine or glaciolacustrine setting characterized by dilute underflows
valleys incised in crystalline basement (Fig. 19.1). The overlying or interflows (e.g. Mackiewicz et al. 1984; Powell & Molnia 1989).
Arkahawl Formation records at its base a major post-glacial trans- In the central outcrop belt (palaeovalley 4), Kellerhals & Matter
gression over the previous basin margin, followed by a 300– (2003) describe glacial striations preserved within the fluviodeltaic
400-m-thick turbidite complex consisting of 1–5-km-wide lobes facies association, demonstrating dynamic glacier advance over
embedded in fine-grained distal fan deposits (Rieu & Allen 2008). fluvioglacial environments.
The turbidite complex gradationally passes up into c. 500 m of
distal marine mudstone and siltstone. The overlying c. 100- Proximal glaciomarine facies association. Coarse-grained proximal
m-thick Marsham Formation records the progradation of shallow glaciomarine deposits constitute the bulk of the upper Ayn strati-
marine and fluviatile deposits, before a second glaciation rep- graphy throughout the area. Deposits of this facies association
resented by the Shareef Formation, barely preserved as thin rem- overstep palaeovalley margins and may comprise single units
nants beneath the sub-Cretaceous unconformity. that reach up to c. 150 m in thickness in palaeovalley 4 (Kellerhals
& Matter 2003). Massive and weakly stratified diamictites with
abundant striated clasts are the most common lithology in this
facies association. Stratified diamictites contain common outsized
Glaciogenic deposits and associated strata clasts (,1 m) deflecting underlying lamination, and many striated
and/or faceted clasts. Interbedded sandstones and conglomerates
In the Jabal Akhdar of the Oman Mountains, Neoproterozoic gla- locally fill channels or occur as pebble lags.
ciomarine deposits are found in the Abu Mahara Group (Ghubrah The presence of many large outsized, striated and faceted clasts
and Fiq Formations), occupying the tectonically deformed core is suggestive of rainout from floating ice (e.g. Ovenshine 1970;
of a major east –west oriented anticline, with no basement rocks Syvitski et al. 1996; Smith 2000) of debris carried by subglacial
exposed (Allen et al. 2011). In contrast, the Abu Mahara- transport prior to deposition (Boulton 1978; Benn & Evans
equivalent glaciogenic sedimentary rocks of the Ayn and Shareef 1998). Conglomerates within stratified diamictite units represent
Formations of the Mirbat Group of Dhofar are relatively unde- either local concentrations of clasts derived from iceberg overturn
formed and occur above an irregular crystalline basement (Keller- (Ovenshine 1970), lag deposits resulting from intensive wave or
hals & Matter 2003; Rieu et al. 2006). Rieu et al. (2007a) proposed current reworking of ice-rafted diamictite in relatively shallow
that the Arkahawl and Marsham Formations represent the missing marine conditions (Eyles et al. 1985; Eyles 1988; Moncrieff &
stratigraphy between the Ghubrah and Fiq Formations in the Jabal Hambrey 1990) or debris flows of remobilized sediment (Wright
Akhdar. & Anderson 1982). Channel bodies comprising graded and cross-
stratified sandstones ‘embedded’ in diamictites may have been
deposited by currents issuing from meltwater tunnel outlets
Ayn Formation (Rust & Romanelli 1975; Banerjee & MacDonald 1975; Powell
1981). The entire facies association suggests a proximal glacio-
The Ayn Formation comprises ,400 m of mainly coarse-grained marine environment, close to the grounded ice-margin (Eyles
glaciogenic deposits, ponded in 2- to .8-km-wide N- to NW- et al. 1985; Moncrieff & Hambrey 1990; Syvitski et al. 1996),
oriented palaeovalleys eroded into basement, with few or no possibly in a grounding-line fan (Powell 1990; Lønne 1995).
deposits preserved on intervening palaeohighs. Sedimento-
logical analysis of near-continuous sections through the three Distal glaciomarine facies association. The distal glaciomarine
westernmost palaeovalley-fills of the Ayn Formation (Fig. 19.2) facies association consists mainly of laminated mudstones with
allow four facies associations to be distinguished: (i) subaerial, subordinate sandstone laminae. Outsized, commonly striated, clasts
242
P. A. ALLEN ET AL.
Fig. 19.2. Summary sedimentological logs of the Ayn Formation from palaeovalleys 1 to 4 and interpretive correlation panel (after Rieu et al. 2006). (a) Logged sections through palaeovalleys and possible correlation (section
PV-4 modified after Kellerhals & Matter, 2003). Logs; m, mudstone; s, sandstone; g, gravel/conglomerate. Facies notation follows Rieu et al. (2006). (b) Lateral correlation of palaeovalley fills and elevation of basement, based
on logged sections and field mapping. Field map in Rieu (2006) and Rieu & Allen (2008).
THE AYN FORMATION 243

deflecting laminae occur throughout this facies association, but are several populations of cross-cutting, millimetre-to-centimetre-
much less common (c. 1%) than in the stratified diamictites of the wide and ,10 m deep, fissures filled with detrital carbonate
proximal glaciomarine facies association. Distal glaciomarine (dolomite) mixed with minor amounts of sand-sized siliciclas-
facies also include graded sandstones and rare gravel and thin tic particles, and void-filling cements. The fills of the fissures
pebble conglomerates. The predominance of fine-grained lami- originate from the overlying carbonate (dolomite), which is
nated deposits points to a relatively quiet depositional environment further supported by their similarity in C-isotopic values (see
dominated by hemipelagic sedimentation. Sediment gravity flow below). In one location a small-scale (,2 cm high) set of fore-
deposits are more sheet-like and thinner than those found in the sets of a ‘mini delta’ developed over a centimetre-high negative
proximal glaciomarine facies association, which together with the step in the wall of the fracture indicates pumping of sediment
low abundance of outsized clasts, interpreted as ice-rafted debris, and fluids upwards through a connected fracture system (Rieu
supports a relatively distal glaciomarine depositional environment et al. 2006). Fissures were evidently opened up at the same
(e.g. Mackiewicz et al. 1984; Eyles et al. 1985; Syvitski et al. 1996). time as carbonate deposition during post-glacial transgression.
In the central outcrop area (palaeovalley 4), Kellerhals & Matter
(2003, p. 60) describe graded, turbiditic sandstones with drop-
stones at the tops of beds, interleaved with diamictite units, demon- Boundary relations with overlying and underlying
strating contemporaneous sediment gravity flow processes and
glacial strata
rainout from melting icebergs.
The Mirbat Group is sandwiched between a basal unconformity
represented by a series of deep palaeovalleys cut into crystalline
Post-glacial carbonate above Ayn Formation basement, exposed in the bevelled coastal plain close to Mirbat
town, and a regional sub-Cretaceous planar unconformity
The post-glacial Arkahawl Formation, in contrast to the Ayn For- exposed high in the cliffs below Jabal Samhan. A map of the
mation, is laterally extensive and covers all remaining basement Group and its boundary relations is found in Rieu & Allen
relief left after deposition of the Ayn Formation (Rieu 2006). At (2008, fig. 3).
the base of the Arkahawl Formation is a ,3-m-thick transgressive
cap carbonate that sharply overlies the underlying glaciomarine or
fluviodeltaic deposits of the Ayn Formation or, where these are Chemostratigraphy
absent, directly overlies crystalline basement (Rieu et al. 2006).
This carbonate is locally overlain by marine sandstones but more Carbon-isotopic composition of the post-glacial carbonate
commonly by a 30 –40-m-thick succession of distal marine
shales, indicating rapid deepening of water depths. Rieu et al. (2006) measured the carbon isotopic composition of
inorganic carbonate (d13C) in the carbonate at five different sites
Carbonate facies. The carbonate is laterally discontinuous and across the area and in the carbonate-filled fissures in basement
varies considerably in thickness and sedimentary facies. Carbon- (Fig. 19.3). Tests for diagenetic effects are discussed in Rieu
ates are locally well developed on palaeohighs, whereas they are et al. (2006). The most negative d13C values of the carbonate are
absent or represented by discontinuous, mixed carbonate – silici- found in the central part of the field area where they increase
clastic mass flow deposits in the deeper parts of the basin (i.e. in from 23.5‰ at the base to þ5.1‰ at the top in laminated and stro-
palaeovalleys). The carbonate comprises shallow-water, mass- matolitic carbonate (sections C4 and C5). Further to the west, lami-
flow and fissure-filling facies. nated carbonate overlying basement shows an increasing trend from
þ0.9‰ at the base to þ2.7‰ at the top (section C3) and nearby car-
† Shallow-water facies association. In the western part of the bonate associated with conglomerate shows upward-increasing
outcrop area, onlapping onto crystalline basement or thin values from þ3.5‰ to þ4.3‰ (section C3). The most positive
breccia, the carbonate is mainly a pink, recrystallized micritic values are found in the westernmost part of the area in carbonate-
dolomite, passing upward into wave-rippled calcareous sand- filled fissures in crystalline basement (between þ4.1‰ and
stones and then a .1.5-m-thick, generally massive dolomite. þ5.7‰) and in the overlying laminated and wave-rippled carbon-
In the centre of the outcrop area a locally well-developed car- ate, which shows an upward decreasing trend from þ5.8‰ at the
bonate with a basal conglomeratic lag overlies fluviodeltaic base to þ2.0‰ at the top (sections C1 and C2).
deposits. The carbonate is a dark grey, locally organic-rich, The lateral variation in d13C in the carbonate is interpreted to
recrystallized micritic limestone showing mainly millimetre- represent a secular variation of the C-isotopic composition of
thin, undulating laminations, small-scale symmetrical wave ocean water. Owing to progressive flooding of the palaeotopogra-
ripples (eastern part of palaeovalley 3) and ,1-m-high, phically highest parts of the basin (i.e. in the west), carbonate depo-
elongate domal stromatolites (centre of palaeovalley 3), sition in these locations would have been delayed with respect to
which together indicate shallow-water deposition. the basin centre. The resulting d13C signal of the carbonate on
† Carbonate mass-flow facies association. In palaeovalleys 1, 2 the basement high is therefore ‘base-truncated’. This interpretation
and 4, the carbonate is either absent or represented by allows the construction of a composite isotope profile through the
carbonate-rich mass flow deposits containing a significant carbonate, showing an increasing trend from 23‰ to þ5.8‰ and
(,c. 75%) amount of siliciclastic material. Mass-flow deposits then decreasing to þ2‰ at the top (Fig. 19.3). In the central area
include centimetre-thick graded beds, interpreted as turbidites, (sections C4, C5; palaeovalley 3) values reach a maximum of
decimetre-thick beds with soft-sediment deformation, inter- þ5.1‰, but the decreasing arm is missing (‘top-truncated’),
preted as slump deposits, and poorly sorted conglomerates inter- which is in accord with the observation that carbonate deposition
preted as debris flows. A ,15-m-thick erosive unit of boulder was restricted mainly to shallower environments.
and cobble conglomerates overlies basement and glacial depos-
its in palaeovalley 1, containing a matrix of siliciclastic sand and
detrital carbonate. The presence of bedding-parallel veneers Neoproterozoic weathering: Chemical Index of Alteration
of carbonate beneath and above this unit indicates that the con-
glomerates, which contain rare carbonate clasts, were deposited Rieu et al. (2007b) studied the compositional and mineralogical
while carbonate was being produced in the basin. changes in the Mirbat Group in order to test the possible effect
† Carbonate-filled fissures. In the western part of the area, a of severe climate swings in the Neoproterozoic on the intensity
slightly foliated granodiorite (Mirbat Granodiorite) is host to of chemical weathering in source areas (Nesbitt & Young 1982;
244 P. A. ALLEN ET AL.

Fig. 19.3. Summary logs, d13C data and major element data of the post-glacial carbonate (after Rieu et al. 2006). See Figure 19.2 for location of C1 –C5. (a) Isotopic
profiles and summary logs are shown in their relative vertical positions, high in the SW where the cap carbonate onlaps basement, and low in the centre of the outcrop area
where the cap overlies fluvial deposits. (b) Composite isotopic profile, with sections C1 and C4 compressed in order to account for different sedimentation rates and to
match sections C2 and C5, respectively. Dashed arrows indicate the relation between cap carbonate and carbonate fissures. (c) Element concentrations for selected
samples from composite section.
THE AYN FORMATION 245

McLennan et al. 1993; Fedo et al. 1995; Nesbitt et al. 1996; Schef- trends were best explained by variations in the intensity of palaeo-
fler et al. 2003; Bahlburg & Dobrzinski 2011). A similar study of weathering of contemporary land surfaces.
Neoproterozoic sedimentary rocks using the chemical index of
alteration (CIA) is found in Dobrzinski et al. (2004). To ensure a
well-mixed provenance and to minimize the effects of hydrodyn- Palaeolatitude and palaeogeography
amic sorting, Rieu et al.’s (2007b, c) study was limited to mud-
stone beds and the mudstone matrices of diamictites. Kilner et al. (2005) reported palaeomagnetic data that passed fold
Throughout the Mirbat Group there are significant compo- and reversal tests from Oman, which gave the Muscat region a
sitional and mineralogical variations (Fig. 19.4). When plotted in palaeolatitude of 138 in the late Neoproterozoic. They recognized
A –CN –K and Qtz – Pl– Kfs space, the data define trends parallel a remnant magnetization that was acquired between deposition and
to A – CN and Qtz –Pl boundaries, thus suggesting that compo- a period of deformation in the Cambrian – Ordovician (Blendinger
sitional variations result from variability in the extent of chemical et al. 1990; Gass et al. 1990), and argued (p. 415) that the magne-
weathering in the regolith in source areas. Glacial deposits of the tization was acquired at deposition since a similar pattern of rever-
Ayn Formation (AY2 –AY8) plot close to the feldspar join and sals occurs at the same stratigraphic level in widely separated
close to the composition of unaltered granodiorite, indicating locations in the north, centre and south of Oman. Subsequent
that their composition has been little affected by chemical altera- work, however, points strongly to different ages for these strati-
tion, as reflected by their low average CIA value of 54. Pre-glacial graphic sections (Allen & Leather 2006; Rieu et al. 2006; Rieu
deposits of the Ayn and Marsham Formation, and non-glacial et al. 2007a). The difference between the Oman palaeopoles and
deposits of the Arkahawl Formation in particular, are enriched in those of Gondwanan terranes (Meert 2003) suggested to Kilner
Al2O3. This enrichment suggests the incorporation of weathered et al. (2005) that Oman did not accrete with the ‘African’ terranes
material, in agreement with the relatively high abundance of comprising greater Gondwana until after c. 600 Ma, the approxi-
clay minerals, and is reflected by average CIA values of 60 and mate age of the youngest stratigraphy sampled for palaeomagnet-
74 for the pre-glacial and non-glacial intervals, respectively. In ism. Kempf et al. (2000) recovered a Precambrian pole position
Qtz – Pl– Kfs space, a similar trend is revealed, with samples from the south of Oman (implying a palaeolatitude of 98) but
from the glacial, pre-glacial and non-glacial intervals, respect- noted a similarity with published poles from Gondwana (Congo,
ively, being progressively enriched in quartz with respect to India, Australia) at c. 550 Ma, which was thought to be the deposi-
plagioclase, resulting in progressively higher average MIA (miner- tional age. The rocks in south Oman providing the low palaeolati-
alogical index of alteration ¼ Qtz/(Qtz þ Kfs þ Pl)) values (18, tude are c. 700 Ma in age, not 550 Ma (Rieu et al. 2006, 2007a),
36 and 46, respectively). Rieu et al. (2007b) believed that these which suggests that either the south of Oman was in low

CIA MIA
unconformity 50 60 70 0 20 40 60 80

glacial epoch II
Mhm Shf

decreasing
1200 temperatures

Non-glacial interlude
or ‘interglacial’
Arkahawl Fm

800

cap carbonate rapidly


400 increasing
temperatures
Ayn Fm

Fig. 19.4. Variations in weathering indices


glacial epoch I (CIA and MIA) in the Mirbat Group,
displayed as a function of their stratigraphic
height (from Rieu et al. 2007b). Low CIA
and MIA values are associated with glacial
and pre-glacial conditions inferred from
0m
sedimentological facies, and relatively high
CIA and MIA values are associated with
50 60 70 0 20 40 60 80
non-glacial marine deposits. Errors on CIA
CIA MIA are ,2%.
246 P. A. ALLEN ET AL.

palaeolatitudes at c. 700 Ma and experienced similar palaeolati-


tudes at 550 Ma, or that the magnetization was acquired during a
later event close to 550 Ma, when Oman and the adjacent parts
of Gondwana were indeed situated in the tropics on the basis of
their common evaporitic sedimentary facies (Allen 2007). Geo- 870
logical evidence summarized in Allen (2007) indicates an earlier
accretion of Oman into the megasuture forming the eastern
border (in present-day coordinates) of Gondwana. Currently,
therefore, the palaeomagnetic data described by Kempf et al.
(2000) and Kilner et al. (2005) lack a rigorous stratigraphic and
geochronological context. 810
Shareef Fm
(glaciation)
Geochronological constraints
n = 177 (69)
The age of the Mirbat Group is uncertain. The maximum age must
be younger than the age of the youngest intrusions in the basement
complex truncated by the basal unconformity. The carbonate
stratigraphically above the Ayn Formation is therefore younger
than the age of the underlying Mirbat Granodiorite (706 +
40 Ma, Rb –Sr date, Gass et al. 1990). The Leger Granite
(726 + 0.4 Ma U –Pb zircon date, Bowring et al. 2007) cannot 810
be proven to pre-date the Mirbat Group because it is truncated 840

Mirbat Group
by the sub-Cretaceous unconformity to the NE of the outcrop
belt of the Mirbat Group, but it is nevertheless highly likely (see
below). These late intrusions were most likely exhumed by the
time of deposition of the oldest sedimentary rocks of the Mirbat
Group (Qidwai et al. 1988; Platel et al. 1992a, b; Kellerhals & Arkahawl +
Matter 2003; Rieu et al. 2006, 2007a). The dykes at Mirbat do Marsham Fms
not cut the unconformably overlying sedimentary Mirbat Group,
as recognized by Qidwai et al. (1988), so the Mirbat Group sedi-
mentary rocks must be younger than the 696 –744 Ma age of the
dykes derived from Sm– Nd and Rb – Sr dating (Worthing 2005). n = 92 (59)
In order to further constrain the age of the Mirbat Group and the
contributing source areas, use has been made of detrital zircon geo-
chronology. A total of 1057 new U– Pb detrital zircon ages were
produced by laser ablation ICP-MS from the Huqf Supergroup
(Rieu et al. 2007a; Allen 2007). The sampled stratigraphic inter- 840
vals in the Mirbat Group include the Ayn, Arkahawl, Marsham
810
and Shareef Formations. Samples of the Ayn Formation come
from three different diamictite units (D1, D2 and D3 of Kellerhals
& Matter 2003) in Wadi Autunt. The sample of the Arkahawl For-
mation was obtained from turbiditic sandstones in Wadi Hinuna,
c. 75 m above the top of the Ayn Formation. From the Marsham
Formation two samples were collected near Jabal Shareef from
shoreface sandstones. Diamictites of the Shareef Formation were Ayn Fm
sampled in the same section, as well as in a section 3 km to the NE. (glaciation)
The stratigraphic units of the Mirbat Group contain remarkably 722
similar detrital zircon populations, comprising zircons almost n = 180 (111)
exclusively of Neoproterozoic age (Rieu et al. 2007a) (Fig. 19.5).
When only the most concordant data and those with the lowest
analytical errors are considered, main peaks in the Ayn Formation
are revealed at c. 722, 810 and 840 Ma. In the Arkahawl and 0.5 1.0 1.5
Marsham formations, only the c. 810 Ma and c. 840 Ma peaks
are present, whereas in the Shareef Formation the zircon popu-
lation is dominated by c. 870 Ma ages with only a small ‘shoulder’ Age (x109 years)
at c. 810 Ma. The maximum age constraint for the depositional age
Fig. 19.5. Frequency distribution of detrital zircons from the Mirbat Group
of the Ayn Formation and the overlying part of the Mirbat Group is (modified from Rieu et al. 2007a). Grey areas are data produced by filter 1
provided by the c. 722 Ma subpopulation in the Ayn Formation, (,25% discordant; 2s errors ,20% for 235U– 207Pb and ,10% for
represented by four grains with a mean age of 722 +19 Ma. 238
U – 206Pb). Black lines are data produced by filter 2 (,10% discordant; 2s
This age is in agreement with the constraints provided by the errors ,10% for 235U – 207Pb and 238U– 206Pb). Number of analyses (n)
700–750 Ma age of the underlying basement based on Rb – Sr, produced with filters 1 and 2 are displayed outside and inside parentheses
K –Ar and Sm– Nd whole-rock and mineral analyses (Gass et al. respectively. Numbers for peaks are in Ma.
1990; Worthing 2005; Mercolli et al. 2006), but is considered to
be more robust.
The detrital zircon populations are dominated by Neoprotero- the Leger Granite (726 + 0.4 Ma), the 722 +19 Ma subpopu-
zoic ages that are in good agreement with the age range (0.7 – lation of zircons in the Ayn Formation may have been derived
1.0 Ga) of local basement rocks in the Mirbat area. Although from this or associated intrusions, suggesting that the Leger
there is no stratigraphic contact between the Mirbat Group and Granite intruded prior to deposition of the Ayn Formation and
THE AYN FORMATION 247

was exhumed shortly after. This subpopulation is similar in age to Formation is a second, little-known glaciogenic unit preserved
the zircons found in tuffaceous ashes of the Ghubrah Formation in patchily beneath the sub-Cretaceous unconformity.
northern Oman (Brasier et al. 2000). The sedimentary evolution of the Mirbat Group is paralleled
by variations in the CIA derived from mudstones and the fine-
grained matrices of diamictites. The CIA suggests conditions
Discussion: the nature of Neoproterozoic glaciation with minimal chemical weathering during the deposition of the
Ayn Formation, followed by intensive weathering in the overlying
The Ayn stratigraphy suggests that in basin-marginal settings, Arkahawl Formation above the cap carbonate. This geochemical
periods of glaciomarine deposition alternated with periods of trend strongly supports the Ayn Formation as glaciogenic, fol-
fluvial and deltaic deposition during glacial recession. Thus, terres- lowed by climatic transit as basin margins were flooded during
trial sediment routing systems functioned strongly during these post-glacial transgression.
periods of recession, by which plentiful sediment could be trans- The age of the Ayn Formation is constrained by (i) the age
ported by high-energy fluvial systems and delivered to more of the youngest detrital zircons (four zircons with a mean of
distal, hemipelagic-dominated areas. CIA analysis suggests a 722 + 19 Ma), (ii) the age of the Shaat dyke swarm cutting
dominance of mechanical weathering at this time. Although the basement and which does not penetrate the Ayn Formation
simplest interpretation is that the entire glacial epoch was charac- (696 –744 Ma), (iii) the age of late plutons such as the Mirbat
terized by dynamic glacier advance and retreat cycles, it is possible Granodiorite (706 + 40 Ma), which is definitely erosionally over-
that preservation of stratigraphy took place preferentially during a lain by cap carbonate and the Leger Granite (726 + 0.4 Ma), which
long, pulsed deglacial phase. The stratigraphy of the Ayn For- is most likely but not definitely older than the Ayn Formation. It is
mation is most easily reconcilable with ‘soft snowball’ models therefore highly likely that the Ayn Formation is younger than
allowing for ice margin fluctuations and the presence of equatorial c. 720 Ma. The age of the Shareef Formation is unknown.
oceans (Hyde et al. 2000; Crowley et al. 2001; Peltier et al. 2007). The Mirbat Group contains detrital zircons of exclusively Neo-
A central tenet of the Snowball Earth hypothesis is the occur- proterozoic age, closely matching the range of ages of meta-
rence of laterally extensive cap carbonates, which sharply morphic and igneous rocks in the underlying basement, which
overlie most Neoproterozoic glacial deposits, with typically nega- formed during a period of major crustal growth associated with
tive C-isotopic ratios that have been used for global correlation subduction at the southeastern periphery of the Arabian-Nubian
(Kaufman & Knoll 1995; McKirdy et al. 2001; Halverson et al. Shield.
2005). Cap carbonates have long been considered paradoxical
because they suggest an abrupt change from glacial to tropical con- The senior author is grateful for the financial and logistical support of Petroleum
ditions (Hoffman & Schrag 2002, and references therein). Models Development Oman (PDO) and Enterprise Ireland, supplemented in later years by
requiring less extreme environmental changes have been proposed the Research Committee of ETH-Zürich. We are grateful for the constructive
to explain the presence of cap carbonates (Grotzinger & Knoll input on field visits of J. Grotzinger, S. Bowring, P. Hoffman, G. Halverson,
1995; Kennedy 1996; Hoffman et al. 1998; Kennedy et al. 2001; PDO staff (particularly J. Amthor) and the participants of the field trip associated
Jiang et al. 2003). Eyles & Januszczak (2004) doubted that cap with the IAS Regional Meeting in Muscat in 2005. I am grateful for the reviews
carbonates had any precise palaeoenvironmental significance, of N. Dobrinski, I. Fairchild and E. Arnaud. This represents a contribution of
the IUGS- and UNESCO-funded IGCP (International Geoscience Programme)
and claimed that many of them are made up of detrital carbonate
Project #512.
derived by erosion rather than representing a distinct oceano-
graphic condition.
It is important to stress first that the Mirbat cap carbonate is
References
clearly associated with a significant palaeoclimatic change,
because it marks the termination of a glacial epoch. Second, a Allen, P. A. 2007. The Huqf Supergroup of Oman: basin development
detrital origin is unlikely, because it occurs within a .1-km-thick and context for Neoproterozoic glaciation. Earth Science Reviews,
succession devoid of any other primary or detrital carbonate. 84, 139– 185.
However, the cap carbonate described here contrasts with many Allen, P. A. & Leather, J. 2006. Siliciclastic marine sedimentation
other cap carbonates worldwide, as it is discontinuous and shows in the aftermath of a Marinoan glacial epoch: the Masirah Bay
considerable variations in facies, thickness and C-isotopic signal Formation, Huqf Supergroup, of Oman. Precambrian Research,
over very short lateral distances (hundreds of metres). Rieu 144, 167–198, doi: 10.1016/j.precamres.2005.10.006.
et al.’s (2006) data suggest that carbonate was produced in the shal- Allen, P. A., Leather, J. & Brasier, M. D. 2004. The Neoproterozoic
lower parts of the basin and redistributed downslope from these Fiq glaciation and its aftermath, Huqf Supergroup of Oman. Basin
palaeohighs, rather than as a blanket of carbonate precipitated Research, 16, 507– 534, doi: 10.1111/j.1365-2117.2004.00249.x.
from an oversaturated ocean (Hoffman et al. 1998; Hoffman & Allen, P. A., Leather, J. et al. 2011. The Abu Mahara Group (Ghubrah
Schrag 2002; Shields 2005). Carbonate production therefore and Fiq Formations), Jabal Akhdar, Oman. In: Arnaud, E.,
appears to have been determined by the same depth-dependent Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record
effects as we see in the photosynthesis-mediated processes of of Neoproterozoic Glaciations. Geological Society, London,
Phanerozoic carbonate production. If carbonate precipitation took Memoirs, 36, 251–262.
place beyond the shallow waters of the palaeohighs, steep submarine Arnaud, E. & Eyles, C. H. 2002. Glacial influence on Neoproterozoic
relief in the palaeovalleys may have prevented the primary depos- sedimentation: the Smalfjord Formation, northern Norway. Sedimen-
tology, 49, 765– 788.
ited carbonate from being preserved in situ.
Bahlburg, H. & Dobrzinski, N. 2011. A review of the Chemical Index
of Alteration (CIA) and its application to the study of Neoproterozoic
glacial deposits and climate transitions. In: Arnaud, E., Halverson,
Conclusions G. P. & Shields-Zhou, G. (eds) The Geological Record of Neo-
proterozoic Glaciations. Geological Society, London, Memoirs, 36,
The Ayn Formation of the Mirbat Group is a c. 400-m-thick suc- 81– 92.
cession of glaciofluvial, glaciodeltaic and glaciomarine sediments Banerjee, I. & MacDonald, B. C. 1975. Nature of esker sedimentation.
deposited in several deeply incised palaeovalleys cutting the meta- In: Jopling, A. V. & McDonald, B. C. (eds) Glaciofluvial and Gla-
morphic and igneous basement in the Mirbat area of Dhofar. The ciolacustrine Sedimentation. Society of Economic Paleontologists
Formation is overlain by a discontinuous, thin (,3 m) and dia- and Mineralogists, Special Publication, 23, 132– 154.
chronous cap carbonate that marks an abrupt overstepping of the Benn, D. I. & Evans, D. J. A. 1998. Glaciers and Glaciation. Arnold,
basin margin and deepening of palaeoenvironments. The Shareef London.
248 P. A. ALLEN ET AL.

Beydoun, Z. R. 1960. Synopsis of the geology of East Aden Protectorate. Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
Report of the 21st International Geological Congress, Copenhagen, Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
Part 21, 131– 149. isotope record. Geological Society of America Bulletin, 117,
Beydoun, Z. R . 1964. The stratigraphy and structure of the East Aden 1181– 1207.
Protectorate. Overseas geology and mineral resources supplement Hambrey, M. J. & Harland, W. B. 1981. Earth’s Pre-Pleistocene
series. Bulletin supplement, 5, 1– 107. Glacial Record. Cambridge University Press, Cambridge.
Blendinger, W., van Vliet, A. & Hughes Clarke, M. W. 1990. Hauser, A. & Zurbriggen, R. 1992. Geology of the Crystalline Basement
Updoming, rifting and continental margin development during the of the Hadbin area (Salalah area, Dhofar, Sultanate of Oman).
Late Palaeozoic in northern Oman. In: Robertson, A. H. F., Unpublished MSc thesis, University of Berne, Switzerland, 6– 221.
Searle, M. P. & Ries, A. C. (eds) The Geology and Tectonics of Hauser, A. & Zurbriggen, R. 1994. Geology of the crystalline basement
the Oman Region. Geological Society, London, Special Publication, of the Hadbin area (Salalah area, Dhofar, Sultanate of Oman).
49, 27 –37. Schweizerische Mineralogisch Petrographische Mitteilungen, 74,
Boulton, G. S. 1978. Boulder shapes and grain-size distribution of 213– 226.
debris as indicators of transport paths through a glacier and till Henson, F. R. S. & Elliot, G. F. 1958. Exhibition of specimens of
genesis. Sedimentology, 25, 773– 799. ‘Collenia’ from pre-Permian sediments in south Arabia. Proceedings
Bowring, S. A., Grotzinger, J. P., Condon, D. J., Ramezani, J., Geological Society London, 1561, 89– 90.
Newall, M. J. & Allen, P. A. 2007. Geochronologic constraints Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
on the chronostratigraphic framework of the Neoproterozoic Huqf testing the limits of global change. Terra Nova, 14, 129–155, doi:
Supergroup, Sultanate of Oman. American Journal Science, 307, 10.1046/j.1365-3121.2002.00408.x.
1097– 1145. Hoffman, P. F., Kaufmann, A. J., Halverson, G. P. & Schrag, D. P.
Brasier, M. D., McCarron, G., Tucker, R., Leather, J., Allen, P. A. 1998. A Neoproterozoic snowball Earth. Science, 281, 1342–1346,
& Shields, G. 2000. New U– Pb zircon dates for the Neopro- doi: 10.1126/science.281.5381.1342.
terozoic Ghubrah glaciation and for the age of the top of the Huqf Hughes Clarke, M. W. 1988. Stratigraphy and rock nomenclature
Supergroup, Oman. Geology, 28, 175–178, doi: 10.1130/0091- in the oil-producing area of interior Oman. Journal Petroleum
7613(2000)028,0175:NUPZDF.2.2.CO;2. Geology, 11, 5 –60.
Carter, H. J. 1852. Memoir on the geology of the south east coast of Arabia. Hyde, W. T, Crowley, T. J., Baum, S. K. & Peltier, W. R. 2000. Neo-
Journal of the Royal Asiatic Society, Bombay Br., 4, 21–96. proterozoic ‘snowball Earth’ simulations with a coupled climate/
Crowley, T. H., Hyde, W. T. & Peltier, W. R. 2001. CO2 levels required ice-sheet model. Nature, 405, 425– 429, doi: 10.1038/35013005.
for glaciation of a ‘near-snowball’ Earth. Geophysical Research Jiang, G., Kennedy, M. J. & Christie-Blick, N. 2003. Stable isotopic
Letters, 28, 283–286, doi: 10.1029/2000GL011836. evidence for methane seeps in Neoproterozoic post-glacial cap car-
Dobrzinski, N., Bahlburg, H., Strauss, H. & Zhang, Q. 2004. bonates. Nature, 426, 822– 826.
Geochemical climate proxies applied to the Neoproterozoic glacial Kaufman, A. J. & Knoll, A. H. 1995. Neoproterozoic variations in the
succession on the Yangtze Platform, South China. In: Jenkins, C-isotopic composition of seawater: stratigraphic and biogeochem-
G. S., McMenamin, M. A. S., McKay, C. P. & Sohl, L. (eds) The ical implications. Precambrian Research, 73, 27– 49.
Extreme Proterozoic: Geology, Geochemistry and Climate. Kellerhals, P. 1993. Igneous Petrology of the Mirbat Complex and
American Geophysical Union Monograph, 146, 13– 32. Facies Analysis of the Mirbat Sandstone Formation (Middle and
Eyles, C. H. 1988. Glacially and tidally influenced shallow marine sedi- Upper Member). Unpublished MSc thesis, University of Berne,
mentation of the Late Precambrian Port Askaig Formation. Palaeo- Switzerland.
geography, Palaeoclimatology, Palaeoecology, 68, 1 –25. Kellerhals, P. & Matter, A. 2003. Facies analysis of a glaciomarine
Eyles, N. & Januszczak, N. 2004. Zipper-rift: a tectonic model for sequence, the Neoproterozoic Mirbat Sandstone Formation, Sultanate
Neoproterozoic glaciations during the break-up of Rodinia after of Oman. Eclogae Geologicae Helvetiae, 96, 49 – 70.
750 Ma. Earth Science Reviews, 65, 1 – 73, doi: 10.1016/ Kennedy, M. J. 1996. Stratigraphy, sedimentology and isotope geo-
S0012-8252(03)00080-1. chemistry of Australian Neoproterozoic post-glacial cap dolostones:
Eyles, N. & Januszczak, N. 2007. Syntectonic subaqueous mass flows deglaciation, d13C excursions and carbonate precipitation. Journal
of the Neoproterozoic Otavi Group, Namibia: where is the evidence Sedimentary Research, 66, 1050– 1064.
for global glaciation? Basin Research, 19, 179–198, doi: 10.1111/ Kennedy, M. J., Christie-Blick, N. & Sohl, L. E. 2001. Are Proterozoic
j.1365-2117.2007.00319.x. cap carbonates and isotopic excursions a record of gas hydrate desta-
Eyles, C. H., Eyles, N. & Miall, A. D. 1985. Models of glacimarine bilization following Earth’s coldest intervals? Geology, 29, 443– 446.
sedimentation and their application to the interpretation of ancient Kempf, O., Kellerhals, P., Lowrie, W. & Matter, A. 2000. Paleomag-
glacial sequences. Palaeogeography, Palaeoclimatology, Palaeo- netic directions in late Precambrian glaciomarine sediments of the
ecology, 51, 15 –84. Mirbat Sandstone Formation, Oman. Earth & Planetary Science
Fedo, C. M., Nesbitt, H. W. & Young, G. M. 1995. Unravelling Letters, 175, 181– 190, doi: 10.1016/S0012-821X(99)00307-6.
the effects of potassium metasomatism in sedimentary rocks and Kilner, B., MacNiocaill, C. & Brasier, M. D. 2005. Low-latitude
paleosols, with implications for paleoweathering conditions glaciation in the Neoproterozoic of Oman. Geology, 33, 413– 416,
and provenance. Geology, 23, 921– 924, doi: 10.1130/0091- doi: 10.1130/G21227.1.
7613(1995)023,0921:UTEOPM.2.3.CO;2. Lees, G. M. 1928. The geology and tectonics of Oman and parts of
Gass, I. G., Ries, A. C., Shackleton, R. M. & Smewing, J. D. 1990. southeastern Arabia. Quarterly Journal Geological Society London,
Tectonics, geochronology and geochemistry of the Precambrian 84, 585– 670.
rocks of Oman. In: Robertson, A. H. F., Searle, M. P. & Ries, Lønne, I. 1995. Sedimentary facies and depositional architecture of
A. C. (eds) The Geology and Tectonics of the Oman Region. Geologi- ice-contact glaciomarine systems. Sedimentary Geology, 98, 13– 43.
cal Society, London, Special Publications, 49, 585– 599. Loosveld, R., Bell, A. & Terken, J. 1996. The tectonic evolution of
Glennie, K. W. 1977. Outline of the geology of Oman. Mémoires de la interior Oman. GeoArabia, 1, 28 –50.
Societé Géologique de France, hors série, 8, 25– 31. Mackiewicz, N. E., Powell, R. D., Carlson, P. R. & Molnia, B. F.
Glennie, K. W., Boeuf, M. G. A., Hughes Clark, M. W., Moody- 1984. Interlaminated ice-proximal glacimarine sediments in Muir
Stuart, M., Pilaar, W. F. H. & Reinhardt, B. M. 1974. Geology Inlet, Alaska. Marine Geology, 57, 113–147.
of the Oman Mountains. KSEPL, Rijswijk, The Netherlands. McKirdy, D. M., Burgess, J. M. et al. 2001. A chemostratigraphic over-
Gorin, G. E., Raacz, L. G. & Walter, M. R. 1982. Late Precambrian – view of the late Cryogenian interglacial sequence in the Adelaide
Cambrian sediments of Huqf Group, Sultanate of Oman. American Fold-Thrust Belt, South Australia. Precambrian Research, 106,
Association Petroleum Geologists Bulletin, 66, 2609– 2627. 149– 186.
Grotzinger, J. P. & Knoll, A. H. 1995. Anomalous carbonate pre- McLennan, S. M., Hemming, S., McDaniel, D. K. & Hanson, G. N.
cipitates: is the Precambrian the key to the Permian? Palaios, 10, 1993. Geochemical approaches to sedimentation, provenance and
578– 596. tectonics. In: Johnsson, M. J. & Basu, A. (eds) Processes
THE AYN FORMATION 249

Controlling the Composition of Clastic Sediment. Geological Society of the Mirbat area, southern Oman. Journal of the Geological
of America Special Paper, 284, 21 –40. Society of London, 164, 997– 1009.
Meert, J. G. 2003. A synopsis of events related to the assembly of eastern Rieu, R., Allen, P. A., Plötze, M. & Pettke, T. 2007b. Compositional
Gondwana. Tectonophysics, 362, 1 –40. and mineralogical variations in a Neoproterozoic glacially influenced
Mercolli, I., Briner, A. P., Frei, R., Schönberg, R., Nagler, T. F., succession, Mirbat area, south Oman: implications for paleoweather-
Kramers, J. & Peters, T. 2006. Lithostratigraphy and geochronol- ing conditions. Precambrian Research, 154, 248– 265, doi: 10.1016/
ogy of the Neoproterozoic crystalline basement of Salalah, Dhofar, j.precamres.2007.01.003.
Sultanate of Oman. Precambrian Research, 145, 182– 206. Rieu, R., Allen, P. A., Plötze, M. & Pettke, T. 2007c. Climatic cycles
Moncrieff, A. C. M. & Hambrey, M. J. 1990. Marginal marine glacial during a Neoproterozoic ‘snowball’ glacial epoch. Geology, 35,
sedimentation in the late Precambrian succession of east Greenland. 299– 302, doi: 10.1130/G23400A.1.
In: Dowdeswell, J. A. & Scourse, J. D. (eds) Glacimarine Environ- Ross, G. M., Bloch, J. D. & Krouse, H. R. 1995. Neoproterozoic strata
ments: Processes and Sediments. Geological Society, London, of the southern Canadian Cordillera and the isotopic evolution of
Special Publications, 53, 387– 410. seawater sulfate. Precambrian Research, 73, 71 –99.
Morton, D. M. 1959. The Geology of Oman. 5th World Petroleum Rust, B. R. & Romanelli, R. 1975. Late Quaternary subaqueous outwash
Congress, 1959, Section 1, paper 14. deposits near Ottawa, Canada. In: Joplin, A. V. & McDonald, B. C.
Nesbitt, H. W. & Young, G. M. 1982. Early Proterozoic climates and (eds) Glaciofluvial and Glaciolacustrine Sedimentation. Society of
plate motions inferred from major element chemistry of lutites. Economic Paleontologists and Mineralogists, Special Publication,
Nature, 299, 715– 717, doi: 10.1038/299715a0. 25, 177– 192.
Nesbitt, H. W., Young, G. M., McLennan, S. M. & Keays, R. R. 1996. Scheffler, K., Hoernes, S. & Schwark, L. 2003. Global changes
Effects of chemical weathering and sorting on the petrogenesis of during Carboniferous-Permian glaciation of Gondwana: linking
siliciclastic sediments, with implications for provenance studies. polar and equatorial climate evolution by geochemical proxies.
Journal of Geology, 104, 525– 542. Geology, 31, 605–608, doi: 10.1130/0091-7613(2003)031,
Ovenshine, A. T. 1970. Observations of iceberg rafting in Glacier 0605:GCDCGO.2.0.CO;2.
Bay, Alaska, and the identification of ancient ice-rafted deposits. Schermerhorn, L. J. G. 1974. Late Precambrian mixtites: glacial and/or
Geological Society America Bulletin, 81, 891– 894. nonglacial? American Journal Science, 274, 673– 824.
Peltier, W. R., Liu, Y. & Crowley, J. W. 2007. Snowball Earth preven- Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal
tion by dissolved organic carbon remineralization. Nature, 450, of existing models and the plumeworld hypothesis. Terra Nova, 4,
813– 818. 299– 310.
Platel, J. P., Roger, J., Peters, T., Mercolli, I., Kramers, J. D. & Le Smith, I. 2000. Diamictic sediments within high Arctic lake sediment
Métour, J. 1992a. Geological Map of Salalah, Sheet NE 40-09, 1: cores: evidence for lake ice rafting along the lateral glacial margin.
250 000, with explanatory notes. Directorate General of Minerals, Sedimentology, 47, 1157– 1179.
Oman Ministry of Petroleum and Minerals. Syvitski, J. P. M., Andrews, J. T. & Dowdeswell, J. A. 1996. Sediment
Platel, J. P., Le Métour, J., Berthiaux, A., Buerrier, M. & Roger, J. deposition in an iceberg-dominated glacimarine environment, East
1992b. Geological map of Juzor Al Halaaniyat, sheet NE 40-10, scale Greenland: basin-fill implications. In: Solheim, A., Riis, F., Elver-
1:250,000. Directorate General of Minerals, Oman Ministry of hoi, A., Faleide, J. I., Jensen, L. N. & Cloetingh, S. (eds) Impact
Petroleum and Minerals. of Glaciations on Basin Evolution: Data and Models from the
Powell, R. D. 1981. A model for sedimentation by tidewater glaciers. Norwegian Margin and Adjacent Areas. Global Planetary Change,
Annals Glaciology, 2, 129–134. 12, 251– 270.
Powell, R. D. 1990. Glacimarine processes at grounding-line fans and Windley, B. F., Whitehouse, M. J. & Ba-Bttat, M. A. O. 1996. Early
their growth to ice-contact deltas. In: Dowdeswell, J. A. & Precambrian gneiss terranes and Pan-African island arcs in Yemen:
Scourse, J. D. (eds) Glacimarine Environments: Processes and crustal accretion of the eastern Arabian Shield. Geology, 24, 131–134.
Sediments. Geological Society, London, Special Publications, 53, Worthing, M. A. 2005. Petrology and geochronology of a Neoprotero-
53 – 73. zoic dyke swarm from Marbat, south Oman. Journal African Earth
Powell, R. D. & Molnia, B. F. 1989. Glacimarine sedimentary pro- Sciences, 41, 248–265.
cesses, facies and morphology of the south-southeast Alaska Shelf Wright, R. & Anderson, J. B. 1982. The importance of sediment gravity
and fjords. Marine Geology, 85, 359– 390. flow sediment transport and sorting in a glacial marine environment:
Qidwai, H. A., Khalifa, M. I. & Ba-Mkhalif, K. A. 1988. Evidence of eastern Weddell Sea, Antarctica. Geological Society America
Permo-Carboniferous glaciation in the basal Murbat Sandstone Bulletin, 93, 951–963.
Formation, Southern Region, Sultanate of Oman. Journal Petroleum Wright, V. P., Ries, A. C. & Munn, S. G. 1990. Intraplatformal basin-fill
Geology, 11, 81 –88. from the Infracambrian Huqf Group, east-central Oman. In:
Rieu, R. 2006. Sedimentology, Stratigraphy and Geochemistry of the Robertson, A. H. F., Seale, M. P. & Ries, A. C. (eds) The
Glacially Influenced Neoproterozoic Mirbat Group, Oman. PhD Geology and Tectonics of the Oman Region. Geological Society,
thesis, ETH-Zürich. London, Special Publications, 49, 601–616.
Rieu, R. & Allen, P. A. 2008. Siliciclastic sedimentation in the inter- Würsten, F. 1994. The Precambrian crystalline basement of Salalah.
lude between two Neoproterozoic glaciations, Mirbat area, southern Unpublished PhD thesis, University of Berne, Switzerland.
Oman: a missing link in the Huqf Supergroup. GeoArabia, 13, Young, G. M. 1992. Neoproterozoic glaciation in the Broken Hill area,
45 – 72. New South Wales, Australia. Geological Society America Bulletin,
Rieu, R., Allen, P. A., Etienne, J. L., Cozzi, A. & Weichert, U. 2006. 104, 840–850.
A Neoproterozoic glacially influenced basin margin succession and Young, G. M. & Gostin, V. A. 1991. Late Proterozoic (Sturtian) succes-
‘atypical’ cap carbonate associated with bedrock paleovalleys, sion of the North Flinders Basin, South Australia: an example of
Mirbat area, southern Oman. Basin Research, 18, 471– 496. temperate glaciation in an active rift setting. In: Anderson, J. B. &
Rieu, R., Allen, P. A., Cozzi, A., Kosler, J. & Bussy, F. 2007a. A Ashley, G. M. (eds) Glacial Marine Sedimentation: Paleoclimatic
composite stratigraphy for the Neoproterozoic Huqf Supergroup of Significance. Geological Society America Special Paper, 261,
Oman: integrating new litho-, chemo- and chronostratigraphic data 207– 223.
Chapter 20

The Abu Mahara Group (Ghubrah and Fiq formations), Jabal Akhdar, Oman

PHILIP A. ALLEN1*, JONATHAN LEATHER2, MARTIN D. BRASIER3, RUBEN RIEU4, MARGARET MCCARRON3,
ERWAN LE GUERROUÉ5, JAMES L. ETIENNE6 & ANDREA COZZI7
1
Department of Earth Science and Engineering, Imperial College London, South Kensington Campus, London SW7 2AZ, UK
2
Tullow Oil, 5th Floor, Block C, Central Park Leopardstown, Dublin 18, Ireland
3
Department of Earth Sciences, Oxford University, Parks Road, Oxford OX1 3PR, UK
4
TOTAL E&P Nederland, Bordewijklaan 18, Den Haag, The Netherlands
5
Beicip-Franlab, 232, Avenue Napoléon Bonaparte, 92502 Rueil-Malmaison, PO Box 213, France
6
Neftex Petroleum Consultants Ltd, 97 Milton Park, Abingdon, Oxfordshire OX14 4RY, UK
7
ENI Angola, Rua Nicola Gomes Spencer, 140, PO Box 1289, Luanda, Angola
*Corresponding author (e-mail: philip.allen@imperial.ac.uk)

Abstract: The Abu Mahara Group (c. 725–,645 Ma) of the Huqf Supergroup in the Jabal Akhdar of northern Oman hosts two glacial
successions in the Ghubrah and Fiq formations, separated by the ,50-m-thick volcanogenic Saqlah Member. The .400-m-thick
Ghubrah Formation is dominated by distal glaciogenic rainout diamictites, laminites and turbiditic siltstones, whereas the ,1.5-km-
thick Fiq Formation exhibits a cyclical stratigraphy of proximal and distal marine glaciogenic facies, and non-glacial sediment
gravity flow and shallow marine facies. The Fiq Formation is overlain by a transgressive, isotopically light carbonate known as the
Hadash Formation. A tuffaceous ash interbedded with glacial diamictites of the Ghubrah Formation in Wadi Mistal has yielded a
U –Pb zircon age of 713.7 + 0.5 Ma. The Fiq Formation contains detrital zircons as young as 645 Ma. The use of the CIA (Chemical
Index of Alteration) shows the Fiq Formation to be climatically cyclic, with alternations of high and low chemical weathering of con-
temporary land surfaces driven by phases of glaciation and deglaciation. The transgression into the post-glacial Masirah Bay Formation is
marked by a major increase in chemical weathering.

The Huqf Supergroup crops out in northern Oman (Jabal Akhdar), Masirah Bay Formation assigned to the Nafun Group. The strati-
east-central Oman (Huqf area) and southern Oman (Mirbat area of graphic nomenclature currently in use divides the Abu Mahara
Dhofar) (Fig. 20.1) (Allen et al. 2011) and is penetrated by many Group into a lower Ghubrah Formation and an upper Fiq For-
boreholes in the salt basins of the Oman interior. The Huqf Group mation, separated locally by a volcanogenic unit. This unit is tran-
recognized by Glennie (1977) in the Huqf area of east-central sitional upwards into Fiq Formation, so is included in the Fiq as the
Oman was subdivided into the Abu Mahara, Khufai, Shuram and Saqlah Member.
Buah Formations by Gorin et al. (1982), but no glaciogenic depos- Neoproterozoic glacial deposits are found in the Abu Mahara
its are found in this region. In the Jabal Akhdar of northern Oman, Group in the Jabal Akhdar of the Oman Mountains, occupying
Kapp & Llewellyn (1965) defined the Mistal, Hajir, Mu’aydin and the core of a major east – west oriented anticline (Fig. 20.2).
Kharus Formations, which were correlated (Tschopp 1967) with Partly coeval rocks are found in Dhofar (southern Oman) (Allen
the Huqf succession of east-central Oman. Kapp & Llewellyn et al. 2011), but the Abu Mahara Group is absent from the Huqf
(1965), Glennie et al. (1974), and geologists of the Bureau de area of east-central Oman. The Group has been penetrated by a
Recherches Géologiques et Minières (Beurrier et al. 1986; Rabu limited number of boreholes in the salt basins of Oman, and has
et al. 1986) recognized and mapped the Mistal Formation (syn. been imaged on seismic reflection profiles, which show it to be
Mistal Conglomerate Formation) as the oldest exposed Neoproter- confined to large north – south oriented, c. 50-km-wide basins
ozoic sedimentary rocks in the Jabal Akhdar. Rabu et al. (1986) that have been interpreted as rifts (Loosveld et al. 1996). This
divided the Mistal Formation into four members, in ascending rift interpretation is supported by sedimentological and strati-
order, the Ghubrah, Saqla, Fiq and Amq Members. In an attempt graphic evidence derived from field studies (Rabu 1988; Rabu
to streamline stratigraphic nomenclature across the surface out- et al. 1993). In the Jabal Akhdar, the sedimentology of the glacio-
crops and subsurface well penetrations of the salt basins, PDO genic Fiq Formation indicates sediment derivation from two
geoscientists (PDO internal reports, Loosveld et al. 1996) rede- opposing margins (Leather et al. 2002; Allen et al. 2004). The
fined the Abu Mahara Group as comprising a lower diamictite-rich intervening basin has the same scale and orientation as the structu-
Ghadir Manqil Formation and an upper Masirah Bay Formation, rally confined basins imaged on seismic profiles.
and raised the status of the succession to the Huqf Supergroup, The phase of sedimentation responsible for the Fiq Formation
comprising Abu Mahara, Nafun and Ara Groups (Glennie et al. was preceded by a period of extrusive basaltic volcanism,
1974; Gorin et al. 1982; Hughes-Clarke 1988; Wright et al. represented by the Saqlah Member (Rabu 1988; Rabu et al.
1990). McCarron (2000) referred to the distinctive carbonate 1993; Le Guerroué et al. 2005). Although the Saqlah volcanic
capping the diamictites of the Fiq Member as the Hadash For- and volcaniclastic rocks and the Fiq glaciogenic and non-glacial
mation, based on the name of the village close to its type locality sedimentary rocks can be attributed to a continental rift basin,
above Wadi Mistal in the Jabal Akhdar. When the basin evolution the Ghubrah stratigraphy below the Saqlah Member is far less
of the Huqf Supergroup was better understood, the term Abu well constrained.
Mahara Group was restricted to the stratigraphy below the trans- A transgressive, 13C-depleted (down to –5‰) carbonate
gressive Hadash Formation carbonate (McCarron 2000; Leather (Hadash Formation) oversteps basement-cored Abu Mahara
et al. 2002; Allen et al. 2004; Allen & Leather 2006), and the basin margins, as seen in the Huqf area, where the carbonate

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 251– 262. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.20
252 P. A. ALLEN ET AL.

Thickness (m)
(a) (b) Oman
composite
Nomenclature
stratigraphy

CAMBRIAN
Dhahaban
IRAN A6
Al Noor

ARA GROUP
Musandam

Subsurface
A5

Ara cycles
0
26 N
Athel A4
ARABIAN GULF A3

Birba
Fara A2
Makran Fold Belt 4000 A1
A0
Buah

Jabal Akhdar and Huqf


GULF OF OMAN

NAFUN GROUP
EDIACARAN
Shuram
Qatar Arch

UNITED ARAB EMIRATES


Muscat
Jabal Akhdar
Khufai
Fig. 2 A Saih Hatat 23O N
Oman Mountains Masirah Bay 3200
Qalhat
Hadash
Southern Gulf F7
Salt Basin Jebel Ja'alan
Fahud
Basin

xis
2800

fa
q
Hu
gh

Ghaba F6
Hi

SAUDI ARABIA Basin


rem
ka

Al Jobah

Jabal Akhdar
Ma

F5

HUQF SUPERGROUP
Western Rub Al-Khali B
Salt Basin Huqf area Fiq/Shareef
F3
Masirah Island
O
20 N
2000
South Oman

CRYOGENIAN

ABU MAHARA GROUP


Salt Basin
OMAN n k
si n
gh dB
a
Fla
Hi a r n
F1

sfa
h Wa ste 1600
a Ea Saqlah
Kh
u n-
ud
Gh ARABIAN SEA
Marsham
?
sin
Ba

1200
ld
ha
ag

Ain Sarit Mirbat


W

YEMEN

Mirbat
C 100 km
Al Hota 17o N
Arkahawl 800
52o E 55oE 58o E
Major transgressive
Evaporites of Ara Sandstones and conglomerates surface

Ayn/Ghubrah 400
Carbonate rocks Shales and siltstones
northern Huqf
Mirbat and

Basement
Extrusives in basin-fill/
Glacial diamictites crystalline basement/late plutons

Fig. 20.1. (a) Neoproterozoic outcrops (black) and salt basins (grey) of Oman. Main outcrop areas are Jabal Akhdar of northern Oman (A), Huqf area of east-central
Oman (B) and Mirbat area of Dhofar (C), as well as the metamorphic window of Saih Hatat. (After Allen et al. 2004.) (b) Composite stratigraphy of the Huqf Supergroup
in Oman. (After Allen 2007 and Rieu et al. 2007a.) Arrows indicate main transgressive surfaces, and F1 to F7 are Fiq units of Leather (2001).

rests via an unconformity lined with silcrete upon felsic ignim- of the Oman Mountains, and in the intensely deformed and meta-
brites and tuffs dated 802 Ma and granitic basement dated 822 morphosed domal culmination of the Saih Hatat (Glennie et al.
and 825 Ma (Allen & Leather 2006; Allen 2007; Bowring et al. 1974; Le Métour et al. 1986; Rabu et al. 1993; Gray et al.
2007). Over basement highs, therefore, the unconformity 2005). The Saih Hatat structural deformation was generated
beneath the Nafun Group stretches from c. 800 Ma to the onset during a top-to-the-NE shear, producing a major, refolded,
of deposition of the Nafun Group, whereas in the Abu Mahara NE-facing, recumbent anticlinal fold-nappe within pre-Ordovician
basins, as in the Jabal Akhdar, the stratigraphic succession is sedimentary rocks, including lateral equivalents of the Huqf
more complete and conformable. Supergroup. No detailed stratigraphic or sedimentological work
has been carried out on account of this deformation and meta-
morphism, although the stromatolitic Khufai Formation of the
Structural framework Jabal Akhdar has been correlated lithologically with the Hiyam
Formation of the Saih Hatat (Rabu et al. 1986).
The Huqf Supergroup is exposed in the core of a 700-km-long, The Huqf Supergroup of the Jabal Akhdar is less intensely
30–130-km-wide, east –west-trending antiform in the Jabal Akhdar deformed than in the Saih Hatat and is exposed in a series of
THE ABU MAHARA GROUP 253

Fig. 20.2. Geological map of the Jabal


Akhdar showing log localities, main towns
and villages. Log localities: WS1 and WS2,
Wadi Sahtan; WH1 and WH2, Wadi Hajir;
WBK1, Wadi Bani Kharus; WM1, WM2,
WM3 and WM4, Wadi Mistal; MU1 Wadi
Mu’aydin; WBJ1, Wadi Bani Jabir. Arrows
indicate main entrances to wadis. (Modified
from Allen et al. 2004.)

erosional windows or ‘bowls’, chiefly (from west to east) the laterally discontinuous volcaniclastic and volcanic (pillowed and
Sahtan, Kharus and Mistal bowls. Huqf Supergroup sedimentary vesicular basalts) Saqlah Member (Le Guerroué et al. 2005). The
rocks were tilted (block-faulting and large-amplitude folding Fiq Formation is overlain by a well-developed carbonate named
most likely associated with the Hercynian orogeny, Rabu et al. the Hadash Formation (McCarron 2000), which is the lowermost
1986) prior to the deep erosion represented by the pre-Permian unit of the Nafun Group.
unconformity, and underwent further deformation during an
‘Alpine’ (c. 90 Ma) event associated with the emplacement of
the Samail ophiolite and thrust sheets involving Mesozoic Glaciogenic deposits and associated strata
Tethyan carbonate-dominated rocks (Mann & Hanna 1990).
Alpine deformation imparted a strong cleavage, particularly in Glaciogenic deposits are found in the Abu Mahara Group in the
the Ghubrah Formation, where schistosity produces a near- Jabal Akhdar at two different stratigraphic levels. The younger
horizontal pencil cleavage and a roughly north –south stretching Cryogenian glaciogenic sedimentary rocks are found in the cycli-
lineation measured from the long axes of clasts in diamictite cal Fiq Formation (Leather et al. 2002; Allen et al. 2004). The
units. The gradient in intensity of deformation and metamorphism older Cryogenian glaciogenic deposits are found in the much
towards the east and NE of the Oman Mountains (Le Métour et al. less well-understood Ghubrah Formation (Rabu et al. 1986;
1986), reaching a peak in the Saih Hatat, indicates that the same Brasier et al. 2000; Leather 2001; Le Guerroué et al. 2005).
tectonic events affected both the Jabal Akhdar and Saih Hatat
regions. The fact that the penetrative cleavage that affects fine-
grained units of the Huqf Supergroup extends into the overlying The Ghubrah Formation
Permian –Mesozoic succession implies that it is Alpine in age,
with an increasing shear deformation downwards. The structural Very little is known about the Ghubrah Formation. It was initially
geology of the Jabal Akhdar illustrated in the work and map of recognized by Kapp & Llewellyn (1965) as the lower part of the
Rabu et al. (1986), Rabu (1988) and Rabu et al. (1993) and the Mistal Conglomerate, and Rabu et al. (1986) referred to it as the
separation of the pre-Permian from Alpine deformation, awaits a Ghubrah Member of the Mistal Formation. The base is not seen.
further detailed analysis. The Ghubrah Formation crops out in Wadi Mistal (.200 m
thick) and Wadi Mu’aydin in the east, and in Wadi Sahtan
(.400 m thick) in the west of the Jabal Akhdar. Rabu et al.
Stratigraphy (1986) contrasted it with the Fiq unit as comprising large thick-
nesses of poorly stratified diamictic conglomerate, compared
The Jabal Akhdar region contains excellent exposures of rocks with the better-bedded Fiq. Rabu et al. (1986), however, warned
belonging to the Abu Mahara and Nafun Groups of the Huqf that locally, in particular in Wadi Mu’aydin, the sandstones of
Supergroup (Fig. 20.1; see Allen 2007, fig. 4, p. 145, for a the Fiq interfinger with conglomerate similar to that of the
summary of stratigraphic nomenclature, and fig. 8, p. 149, for Ghubrah, rendering the two virtually indistinguishable. Leather
the composite stratigraphic column). Whereas the base of the (2001) stated that the facies within the Ghubrah and Fiq For-
Huqf Supergroup is not seen, the top is an erosional, angular mations were ‘notably different’, with the Ghubrah Formation
unconformity beneath the carbonate rocks of the Permian Saiq, lacking bedding for much of its thickness.
which cuts down variably. In some locations (e.g. Wadi Mistal), McCarron (2000), Brasier et al. (2000) and Leather (2001) all
the Saiq cuts down deeply into the Abu Mahara Group, whereas followed Rabu et al. (1986) in believing the Ghubrah to be an
in others a full succession of Nafun Group is preserved. In Wadi older stratigraphic unit separated from the Fiq Formation by the
Bani Awf the Nafun Group passes up into the Fara Formation, a volcanogenic Saqlah Member. Le Guerroué et al. (2005) identified
surface equivalent of part of the Ara Group, which is truncated an angular relationship between the two formations, proposing an
by the base– Saiq unconformity. The Abu Mahara Group com- unconformity at the base of the Saqlah. However, the relationship
prises a lower diamictite-rich Ghubrah Formation, separated of the Ghubrah to the Fiq, and the sedimentology and structural
from an upper diamictite-rich Fiq Formation by an intervening geology of the Ghubrah, have never been given the attention
254 P. A. ALLEN ET AL.

they deserve. The U –Pb zircon date reported by Brasier et al. occurs as c. 10-cm-thick boudinaged beds in Wadi Mistal consist-
(2000), and more accurate U –Pb dates on similar material reported ing of quartz grains and lithic clasts in a dolomite matrix of
in Bowring et al. (2007), need to be viewed in this context (see ,50 mm rhombs. The tuffaceous and ashy sediments that have
below). provided the U – Pb date reported by Brasier et al. (2000) and
The facies of the Ghubrah Formation are described briefly by Bowring et al. (2007) indicate nearby volcanism during deposition
Leather (2001). The Formation comprises a distal glaciomarine of the Ghubrah Formation.
facies association made of massive diamictites, massive and
graded siltstones, dropstone laminites, very minor carbonate
rocks and tuffaceous sandstones. Sedimentology of the Fiq Formation
The massive diamictites are up to hundreds of metres thick, with
gradational upper and lower boundaries into massive and graded The Fiq Formation is well exposed in a number of wadis in the
siltstones. The diamictites appear structureless, with some striated Jabal Akhdar of northern Oman, particularly in Wadi Sahtan,
and faceted clasts. These massive diamictites may be thick amal- Wadi Hajir and Wadi Mistal (Figs 20.2 & 20.3). Although the stra-
gamated debris flows that reworked glacially transported clasts, tigraphy, first described by PDO geologists in 1965, was mapped
but the association with dropstone laminites and the large lateral and interpreted as partly glaciomarine (Glennie et al. 1974; Beur-
extent (.40 km) suggest a glaciomarine origin by fallout from rier et al. 1986; Rabu et al. 1986; Hughes Clark 1988), the first
suspension and rainout of ice-rafted debris, possibly in an outer detailed sedimentological analysis is in the doctoral thesis of
shelf/slope setting (Anderson et al. 1984; Deynoux 1985; Eyles Leather (2001).
1988; Eyles & Lagoe 1990; Moncrieff & Hambrey 1990; Brodzi-
kowski & van Loon 1991). The associated siltstones indicate quiet Facies associations. The facies present in the Fiq Formation can be
marine sedimentation when ice rafting was reduced, either because assigned to four different facies associations: (i) distal glaciomar-
of glacial retreat (Eyles & Lagoe 1990) or movement of iceberg ine, (ii) proximal glaciomarine, (iii) non-glacial sediment gravity
transport lanes (Moncrieff & Hambrey 1990; Miller 1996). The flow and (iv) non-glacial shallow marine. The lithofacies associ-
dropstone laminites found in Wadi Mu’aydin indicate a clear ations used have been adapted in part from Eyles et al. (1985),
ice-rafted component (Powell & Domack 1995). Carbonate Moncrieff & Hambrey (1990), Brodzikowski & van Loon (1991)

Fig. 20.3. Summary sedimentological logs through the Fiq Formation from 10 localities (see Fig. 20.2), hung from the level of the Hadash Formation. Diamictites
are highlighted to illustrate their lateral continuity and vertical superimposition. F1 to F7 are stratigraphic units of Leather (2001), and T1 to T6 are transgressive surfaces.
Diamictites in F5 and F7 are basin-wide deposits, whereas older diamictites, such as that in F1, are locally developed. (After Allen et al. 2004.)
THE ABU MAHARA GROUP 255

and Miller (1996). Lithofacies names have been partly adapted large intraformational rafts, representing debris flows or hypercon-
from Eyles et al. (1983), Moncrieff (1989) and Moncrieff & centrated flows (Nardin et al. 1979; Costa 1988); (ii) common
Hambrey (1990). sharp-based, massive and graded, sheet-like, commonly amalga-
mated sandstones, locally with rippled tops and fluted bases,
Distal glaciomarine facies association. The distal glaciomarine which are turbidites; (iii) current-rippled sandstones and siltstones
facies association represents distal glaciomarine environments with ubiquitous water-escape structures representing deposition
where ice rafting and non-glacial processes dominate, and is from more dilute turbidity currents, with local wave-generated
found in Wadi Sahtan, Wadi Bani Awf and Wadi Hajir. It is domi- ripples at the top of shallowing-up cycles; (iv) massive and lami-
nated by massive, sheet-like diamictites (,30 m thick) laterally nated mudstones and convoluted siltstones, which settled slowly
and vertically transitional to deep-water mudstones and siltstones. from turbiditic plumes/tails or river plumes/underflows (Stow
Clasts include striated and faceted examples indicating subglacial et al. 1996); (v) thin, brown carbonate on a transgressive surface
transport (Wentworth 1936; Boulton 1978; Dowdeswell et al. overlying glacial diamictites (Wadi Sahtan, log WS1, unit F6)
1985). An ice-rafted origin is inferred on the basis of the glacially suggesting carbonate precipitation at a time of deglacial sea-
transported clasts, common transitional lower boundaries and level rise.
lack of internal stratification (Anderson et al. 1984; Eyles 1988;
Kellerhals & Matter 2003). The clast-poor nature of many of the Non-glacial shallow marine facies association. The non-glacial
diamictites, their sheet-like geometry and their association with shallow marine facies association forms a minor part of the non-
deep-water deposits suggests that they formed as a result of a com- glacial sediments of the Fiq Formation. It comprises (i) laterally
bination of settling of suspended sediment and ice rafting, possibly persistent conglomerates composed of well-rounded pebbles and
in an outer shelf/slope setting relatively distant from active ice cobbles, representing transgressive lag (ravinement) deposits win-
margins (Deynoux 1985; Eyles & Lagoe 1990; Moncrieff & nowed of their finer material during flooding; (ii) extensively
Hambrey 1990; Brodzikowski & van Loon 1991). wave-rippled sandstones, particularly common in Wadi Sahtan,
Associated dropstone laminites (,50 m thick), found in Wadi probably representing shoreface deposition; (iii) rippled wave-
Hajir, comprise grey shales and siltstones with sharp-based turbi- generated heterolithics, passing up into sharp-based, massive sand-
ditic interbeds and rare outsized clasts, indicating a distal shelf stones probably representing flows generated by storms.
to slope setting affected by ice rafting, suspended sediment
fallout and dilute turbiditic underflows. Underflows may have
been released by subglacial river output (Eyles 1988; Moncrieff Depositional history of the Fiq Formation
& Hambrey 1990) or brine-rich currents, as occurs on the
Barents Shelf today (Elverhoi et al. 1989). The Fiq Formation is spatially and temporally highly variable
(Fig. 20.3). This variability results from the fundamental bathy-
Proximal glaciomarine facies association. The proximal glaciomar- metric effects produced by the rifted crustal structure. However,
ine facies association formed close to the grounded ice margin distinct genetic units reflecting relative sea-level change can be
(Moncrieff & Hambrey 1990; Miller 1996) where rainout of traced across the entire Jabal Akhdar area. Glacially influenced
coarse ice-rafted debris and sediment gravity flows interact. Sedi- sedimentary rocks occur at a number of different levels in the
ment gravity flows are also found in distal glaciomarine and non- Fiq Formation, commonly followed abruptly by a facies change
glacial environments. In summary, the proximal glaciomarine indicating deepening. The facies evolution through time and the
facies association comprises (i) massive diamictites (from a few relative sea level recorded by these facies can therefore be evalu-
metres thick to 100 m thick) containing striated, faceted and ated (Leather et al. 2002; Allen et al. 2004; Allen & Etienne 2008).
‘flat-iron’-shaped clasts, indicating subglacial transport prior to
ice rafting; (ii) subordinate but widespread stratified diamictites, Unit F1. Unit F1 is a transitional unit between the underlying vol-
formed by a combination of winnowing of ice-rafted material canics of the Saqlah Member and the Fiq Formation. Volcaniclas-
(Eyles et al. 1985; Moncrieff & Hambrey 1990), or re-deposition tic detritus was derived from a source to the east, probably in the
of massive diamictites by sediment gravity flows in ice-proximal Saih Hatat area (Le Métour et al. 1986; Villey et al. 1986), and
settings; (iii) clast-supported, laterally discontinuous conglomer- was deposited in an already-deep marine basin. The occurrence
ates, representing local concentrations of clasts derived from of diamictites (F1b in Wadi Mistal) containing volcanic clasts indi-
iceberg overturn (Ovenshine 1970), lag deposits indicating inten- cates glaciation in F1, though its effects appear to be restricted to
sive current, wave or tide reworking of ice-rafted diamictite depos- the east of the outcrop area.
its in relatively shallow marine conditions (Eyles et al. 1985; Eyles
1988; Moncrieff & Hambrey 1990), and debris flows of remobi- Units F2 and F3. In the west of the Jabal Akhdar, quiet, deep-water
lized ice-rafted material where the fine sediment has been expelled conditions became established in F2 times, as shown by increasing
into an overlying turbulent suspension (Wright & Anderson 1982); interbedding of turbidites derived from the west, followed by shal-
(iv) massive or graded, rippled, sharp-based sandstones and silt- lowing into a wave-rippled shoreface. The overlying F3 proximal
stones, with lonestones/dropstones at bed boundaries, are turbidite diamictite indicates a resumption of glaciation. In the east of Jabal
deposits formed in a glaciomarine stetting extending some distance Akhdar, the F1b diamictite was abruptly flooded and succeeded by
from the grounding line; (v) minor dropstone laminites, due to gravity-flow deposits derived from non-volcanic sources in the
minor ice rafting in a quiet distal glaciomarine environment; (vi) east. An overlying diamictite correlates well with F3 in the west
massive and laminated mudstones and siltstones lacking drop- of Jabal Akhdar, indicating that the entire outcrop area was influ-
stones, derived from dilute turbidity currents or overflow plumes enced by glacial processes at this time.
(Mackiewicz et al. 1984; Powell & Molnia 1989). Taken together,
the facies association suggests an ice-proximal environment Unit F4. Evidence for an important flooding event following depo-
subject to outflows from subglacial streams, debris flows, sandy sition of the F3 diamictite is found across the whole Jabal Akhdar
turbidites, fallout from suspension from turbid plumes, winnowing outcrop area, and there is no record of any glacial processes in the
and reworking, and ice rafting. basin. Subsequently, sediment gravity flows derived from the east
filled the eastern sector of the basin, whereas shallow shoreface
Non-glacial sediment gravity flow facies association. The non-glacial conditions existed in the west, indicating asymmetry in subsidence
sediment gravity flow facies association, which may form units and/or sediment supply. In both the east and west, sediment
,250 m thick, comprises (i) erosive-based, commonly lenticular gravity flows from multiple sources then resumed and dominated
conglomerates and pebbly sandstones, with rounded clasts and sedimentation across the entire basin. A shutdown in sediment
256 P. A. ALLEN ET AL.

supply in the west resulted in the deposition of sediments during Chemostratigraphy


transgressive phase T3.
Carbon-isotopic data
Unit F5. Renewed glacial activity across the whole outcrop area
is recorded by proximal diamictites and sediment gravity flows In total, 271 samples were analysed from seven stratigraphic pro-
in the east and distal glaciomarine deposition in the west. files through the Hadash Formation in the Jabal Akhdar (Fig. 20.4).
‘Least altered’ samples were selected for being well away from
areas of intense veining and for having good microscopic fabric
Unit F6. The end of F5 glaciation is indicated by a major flooding
preservation revealed from petrographic and cathodoluminescence
event in the western and central parts of the outcrop area, whereas
data. In total, 81 samples from five stratigraphic sections were ana-
sediment gravity flows initially dominated the eastern sector of the
lysed for major and minor elements (details in Leather 2001).
basin. In the western and central parts of the basin, deposition of
Elemental and isotopic data were used to screen for the effects
unit F6 continued as turbidites derived primarily from western
of diagenesis and metamorphism, following techniques described
sources, and shallow-water wave-rippled sandstones were depos-
by Brand & Veizer (1980, 1981), Marshall (1992) and Kaufman &
ited in the east. A locally developed diamictite (F6a) indicates a
Knoll (1995). Further details can be found in Leather (2001) and
weak glacial advance before significant basin-wide flooding (T5
Allen et al. (2004).
transgression). The overlying lag conglomerate and thin carbonate
The vertical pattern of C-isotope values (Hadash 1 type section)
rocks and mudstones indicate transgression and establishment of
shows an initial shift to – 4‰ to –8‰ in the basal metre, followed
deep-marine conditions before a gradual shallowing took effect
by a gradual drift back to –1‰ at the top of the carbonate, although
towards the top of F6.
with considerable scatter. This trend is in common with other Neo-
proterozoic cap dolostones (Kennedy 1996; Hoffman et al. 1998b;
Unit F7. Coarse-grained gravity-flow deposits are overlain by Frimmel et al. 2002; Fölling & Frimmel 2002). However, corre-
proximal glaciomarine diamictites and debris flows and, locally, lation of chemostratigraphic facies from east to west across the
by subglacial stream outwash, representing the last of the glacial Jabal Akhdar is problematical. The variation in d13C values
phases in the Fiq Formation. The overlying Hadash Formation is (c. 2‰) between closely spaced stations is as great as the amplitude
a classic, transgressive postglacial carbonate. This T6 transgres- of the isotopic excursions used for correlation. Leather (2001)
sion covered the rift margins and flooded other basement highs therefore concluded that high-resolution intrabasinal correlation
in the Oman region. using d13C curves is highly uncertain.
Further C-isotopic data from the Nafun Group are found in
Burns & Matter (1993), Cozzi & Al Siyabi (2004), Fike et al.
The Hadash Carbonate (2006), Le Guerroué et al. (2006a, b, c) and Le Guerroué &
Cozzi (2010).
The ,1.5-km-thick Fiq Formation is overlain by a transgressive,
13
C-depleted dolostone (Hadash Formation) that deepens up into
the marine shales and siltstones of the Masirah Bay Formation Chemical Index of Alteration
(Allen & Leather 2006) (Fig. 20.4). The Hadash Formation is
dominated by dolostones, and despite being ,15 m thick is later- The use of the Chemical Index of Alteration (CIA) is described in a
ally extensive over at least 80 km throughout the Jabal Akhdar. separate paper in this volume (Bahlburg & Dobrzinski 2011,
Subsurface well penetrations (PDO internal reports) suggest that Chapter 6), and details of the technique are not repeated here.
the Hadash Formation is found throughout the Oman area. Throughout the Fiq and Masirah Bay formations there are signifi-
Surface outcrop studies also demonstrate that a transgressive car- cant compositional and mineralogical variations. When plotted
bonate unit overlies basement crystalline or volcanic rocks in the in Al2O3 – CaO þ N2O – K2O and quartz-plagioclase-K-feldspar
Huqf region (Loosveld et al. 1996; McCarron 2000; Leather space, the data define trends roughly parallel to the Al2O3 –
2001; Allen & Leather 2006; Allen 2007). Directly overlying CaO þ N2O and plagioclase-quartz boundaries, suggesting varia-
glacial diamictites of the Fiq Formation and with a negative bility in the extent of chemical weathering of the sediment in the
C-isotopic signature, the 8-m-thick carbonate-dominated part of source area (Fedo et al. 1995; Nesbitt et al. 1996). The possible
the Hadash Formation shares much in common with cap dolo- influence of changes in grain size, hydrodynamic sorting, prove-
stones worldwide (Williams 1979; Fairchild & Hambrey 1984; nance and diagenetic alteration on the composition of the sedi-
Tucker 1986; Narbonne et al. 1994; Kennedy 1996; Hoffman ments is discussed in Rieu et al. (2007b).
et al. 1998a; Myrow & Kaufman 1999; Hoffman & Schrag Both uncorrected CIA values and those corrected for a
2002; Halverson et al. 2004). The facies of the Hadash Formation maximum amount of K-metasomatism are plotted as a function
are described in Allen et al. (2004). Although the broad deposi- of their stratigraphic height (Fig. 20.5). In both cases, a similar
tional environment of the cap carbonate in the Jabal Akhdar is first-order trend is revealed, comprising three intervals during
deep water, the high proportion of siliciclastics and palaeocurrent which chemical weathering was reduced as indicated by rela-
directions in Wadi Bani Jabir, the orientation of roll-up structures tively low CIA and Mineralogical Index of Alteration (MIA)
(Simonson & Carney 1999) at Hadash and thickness trends values. Reduced chemical weathering in these intervals is in
between the logged sections collectively suggest that the deposi- agreement with the presence of distinctive sedimentary facies
tional environment of the carbonate deepened to the centre-west. (diamictites, dropstone-bearing laminites) that suggest a cold
The cap carbonate therefore partially ‘shales out’ into the deepest- climate. These intervals alternate with units characterized by
water lithofacies deposited in the centre of the Fiq basin. The most relatively high CIA and MIA values and that lack evidence for
important transition in the basin is between the carbonate- any glacial influence during sedimentation, which are interpreted
dominated eastern Jabal Akhdar area (Hadash 1, Hadash 2 and to represent interglacial periods. Importantly, the end of the
Wadi Mu’aydin) and the western Jabal Akhdar area, which entire glacial epoch corresponds to a major increase in CIA
exhibits a higher concentration of fine-grained siliciclastics and MIA values in the lowermost Masirah Bay Formation.
(Wadi Hajir 1, Wadi Hajir 2, Wadi Bani Awf, Wadi Sahtan). These values are the highest found in the succession
This was also an important transition during the deposition of (CIA . 80; MIA, 100). A discussion of these trends, and a justi-
the Fiq Member, with thicker, more shale-prone sections evident fication of their interpretation as due to temporal changes in the
in the west. It is therefore likely that the basement structure chemical weathering of contemporary land surfaces, are given by
controlled water depth and accommodation across this line. Rieu et al. (2007b, c).
N 10 km
WBJ1
Wadi Hedak
Nakhl

Wadi Sahtan
S Wadi Bani
Jabir
Wadi Mistal Legend
Fashah Wadi Bani Awf Al Awabi
Thin dolomite beds in fine-grained siliciclastic rocks (Facies W3)
Amq WS1
Sahtan Bowl Wadi Wadi Bani Ghubrah
Hajir Kharus
WS2 WBA2 WM2 WM1 Sand-grade siliciclastics (Facies E5)
WM3
WBA1 Hajir Ghubrah
WBK1
WH2 Bowl Crystalline limestone and dolospar (Facies E4)
WH1

Permian to Cretaceous cover WM4 Wadi Hajir 2 Dolomite-dominated deposits of Facies W2, E1, E2 and E3
Hadash
Huqf Supergroup Thinly-interbedded dolomite and siltstone (Facies W1)
Al Ayn
Quaternary outside Jabal Akhdar
Log locality and section line MU1 Massive glacial diamictites (Fiq unit 7)
Entrance to wadi
Village or Town
Wadi Mu'aydin
Carbon isotope data: Dolomite Calcite

11 Wadi Sahtan
Wadi Bani Awf Hadash 2 Hadash 1

THE ABU MAHARA GROUP


Height in m above base of Hadash Formation

10 (Wadi Mistal) (Wadi Mistal)


W3 Wadi Bani Jabir
9
Wadi Hajir 1
8
W3
7 W3
Wadi Mu'aydin E4
6 W3
E4 E5
5 W2 E4
4 E3
W2 E3 E3
3 E2
E2
W2 W2 E2
2 W1
W1
1 E1 E1 E1 E1
W1 W1 E1
0 -6 -2 0 -2 0
-4 -4 -2 0 -4 0 -6 -4 0 -4 -2 0 -8 -4 0 -3 -2 -1 0
F7 F7 d13C
F7 F7 F7 F7 F7
d13C d13C d13C F7 d13C d13C d13C

Western Jabal Akhdar Eastern Jabal Akhdar

Fig. 20.4. Summary logs of Hadash Formation from eight localities across the Jabal Akhdar region (see Fig. 20.2), with C-isotopic data shown for seven. Facies labels described in Leather (2001) and Allen et al. (2004). Note the
variable horizontal scale for d13C, which demonstrates the significant variation in C-isotopic ratios between sections. Isotopic and geochemical data are in the supplementary material of Allen et al. (2004).

257
258 P. A. ALLEN ET AL.

Fig. 20.5. Variations in chemical and


mineralogical indices of alteration (CIA,
Al2O3/(Al2O3 þ CaO þ Na2O þ K2O);
MIA, Qtz/(Qtz þ Kfs þ Pl)) with
stratigraphic height for section in Wadi
Sahtan (a) and critical section across the
glacial –post-glacial transition at Hadash,
Wadi Mistal (b). Numbering of diamictites
and flooding surfaces as in Figure 20.3.
Errors in CIA due to uncertainties in major
element concentrations are ,1.5% (,1 CIA
unit). 2s error bars are indicated for MIA
values. MB Fm., Masirah Bay Formation.
CIA values are corrected to show the
maximum possible effect of potassium
metasomatism during burial diagenesis, as
explained by Rieu et al. (2007c, p. 301).
(After Rieu et al. 2007c.)

Boundary relations with overlying and underlying in the stratigraphy in Wadi Mistal (inclination of þ498) indicate
non-glacial units a magnetic reversal during Fiq times. For palaeolatitude determi-
nation, a set of samples from the Hadash Formation provide
The non-sedimentary basement beneath the Huqf Supergroup is more data (12 sites), with tilt-corrected inclinations between
poorly exposed in Oman. No basement rocks are found in the –7.68 and –39.38 for reversed polarities, and between þ218 and
Jabal Akhdar. Crystalline basement is exposed in south Oman in þ298 for normal polarities (Kilner et al. 2005, Data Repository
the Mirbat region (Platel et al. 1992a) and on the Al Hallaniyah Table 1). The average palaeopole and palaeolatitude based on 25
Islands (Platel et al. 1992b), and in small outcrops in the north sites throughout Oman quoted by Kilner et al. (2005) is unfortu-
of the Huqf area of east-central Oman (at Al Jobah, Allen & nately a statistical average of samples of ages that are spread
Leather 2006; Allen 2007; Bowring et al. 2007) and near Sur in from c. 800 Ma to c. 640 Ma (Allen 2007; Allen et al. 2011). Elim-
the Jabal Ja’alan (Roger et al. 1992) further to the north. Most inating all but the Fiq and Hadash samples, it is evident that the
knowledge on the pre-Huqf Supergroup geological history of variation in the B component is large, and the average palaeo-
Oman is derived from the Mirbat area, recently summarized by latitude of 138 for the Muscat region is therefore open to doubt.
Mercolli et al. (2006) and Allen (2007) (see Allen et al. 2011).
The rocks overlying the Abu Mahara Group, termed the Nafun
Group, have been correlated confidently between the Jabal Geochronological constraints
Akhdar, the Huqf area of east-central Oman and the subsurface
of the Oman salt basins (Cozzi & Al-Siyabi 2004; Allen & Geochronological data (Brasier et al. 2000; Amthor et al. 2003;
Leather 2006; Allen 2007; Le Guerroué & Cozzi 2010), in terms Allen & Leather 2006; Rieu 2006; Bowring et al. 2007; Rieu
of major lithological trends and carbon isotopic profiles. The et al. 2007a) were integrated to obtain the best available con-
base of the overlying Nafun Group is marked by a major strati- straints on the ages of stratigraphic units within the Huqf Super-
graphic overstep due to transgression of basin margins, but the group by Allen (2007) (Fig. 20.6).
age of the Hadash Formation and the bulk of the Nafun Group is Brasier et al. (2000) provided a U –Pb date on zircons recovered
not well known. There is no evidence of glaciogenic sedimentary from a tuffaceous ash interbedded with thin (centimetre-scale)
rocks in the Nafun Group, although it contains a very large nega- sandstones and thick, cleaved diamictites from the Ghubrah For-
tive C-isotopic excursion known as the Shuram anomaly (Fike mation in Wadi Mistal. The sample yielded an intercept age of
et al. 2006; Le Guerroué et al. 2006b; Bristow & Kennedy 2008; 723(þ16/– 10) Ma. Resampling and analysis (Bowring et al.
Le Guerroué & Cozzi 2010). 2007) revealed a zircon population of uniform colour and size,
and good crystal forms lacking any obvious evidence of reworking.
Furthermore, of the 25 crystals analysed, 22 defined a discordia
Palaeolatitude and palaeogeography with an upper intercept age of 713.7 + 0.5 Ma, whereas the final
three grains are variably discordant, perhaps due to a small xeno-
Kilner et al. (2005) sampled from the Jabal Akhdar, Huqf and crystic component. The lack of an admixture of older zircons, com-
Dhofar regions. Six sites yielding 20 samples were taken from pared to their abundance in nearby detrital beds (sandstone clast
the Fiq Formation in the Jabal Akhdar, and the Hadash Formation and diamictite matrix, see below), suggests that 714 Ma is a
was sampled at 21 sites giving 60 samples. Kilner et al. (2005) good estimate of the depositional age of the tuffaceous ash. This
found a component of magnetization (Component B) that passed is a rare and valuable constraint on the age of a Neoproterozoic
the fold test and was therefore imparted prior to tectonic defor- glaciation because the tuffaceous ash is interbedded with glacial
mation in the Cambro-Ordovician (Blendinger et al. 1990; Gass diamictites.
et al. 1990). After correction for tectonic tilt, the declination and A number of detrital zircons have been analysed from the
inclination for a sample close to the top of the Fiq Formation in same Ghubrah Formation locality in Wadi Mistal (Bowring et al.
Wadi Sahtan (inclination of –218) and for a sample 300 m lower 2007). The sandstone clast yielded five grains in the age range
THE ABU MAHARA GROUP 259

ABU MAHARA GROUP, JABAL AKHDAR


Neoproterozoic Mesoproterozoic Palaeoproterozoic

c.664 (1 grain) 675 810-860 (10 grains


c.920 (1 grain)
645 715-718 840-860 Fiq n=307 (167)
790-810
750
700-720 1050

750-826 (detrital grains in ash) Fig. 20.6. Zircon geochronology, Abu


c.750 (3 grains) 800-830 (4 detrital grains) Saqlah n=95 (28) Mahara Group, Jabal Akhdar (after Allen
860 2007, supplemented by Bowring et al.
810 910 2007). Histograms of zircon ages derived by
735 LA-ICPMS, reported in Rieu (2006) and
Rieu et al. (2007a). Grey shaded area shows
zircons passing first filter, solid line is
755 (matrix) zircons passing stricter second filter (see
713.7±0.5 (n=22) Ghubrah n=79 (18)
812-826 (5 detrital grains) text). Sample number n represents zircons
710
contributing to the histogram; bracketed
term is the number passing the first filter but
810 failing the stricter filter. Stars indicate
detrital zircon data in Bowring et al. (2007).
Solid bars beneath the time scale show the
0.5 1.0 1.5 2.0 2.5 main age ranges of detrital zircons. Note the
progressively older average age of the
Bowring et al. (2007) data (Lahan core) Age (Ga)
detrital zircon distribution in the younger
Detrital grains (diamictites, turbidites) stratigraphy and the appearance of Meso-
Detrital grains (tuffs, ashes) and Palaeoproterozoic sources in the Fiq
Syn-magmatic (Ghubrah date) Formation.

812– 826 Ma, and a sample of sandy diamictite matrix yielded concordant of Rieu et al.’s (2007a) dataset (28 dates) give a
three grains dated as 755 Ma. main peak at 860 Ma, and subpopulation peaks at 730, 815 and
Rieu et al. (2007a) sampled a graded siltstone situated 1 m 910 Ma. The youngest zircons may have been derived from
above the thin tuffaceous ash yielding the 713.7 + 0.5 Ma date. igneous rocks associated with the accretion of island arc and gneis-
Of the 79 detrital zircon ages (derived by LA-ICPMS) compiled sic terranes in the Arabian –Nubian megasuture (c. 760 Ma), late
(selected as ,25% discordant and 2s errors ,10% for post-tectonic plutons such as those of the Mirbat area (c. 700–
235
U – 207Pb and 238U – 206Pb), 18 passed a stricter filter (,10% 750 Ma), or the start of continental extension at c. 720 Ma. The
discordant and 2s errors ,20% for 235U – 207Pb and ,10% for older zircons reveal the same sources of plutons formed during a
238
U – 206Pb). These most concordant data yield a youngest sub- major period of subduction-related crustal growth at .800 Ma.
population with a peak at 710 Ma, which is in excellent agreement Although disappointing in not providing a syndepositional age
with the age of the underlying tuffaceous ash. An older subpopu- for the Saqlah, the results confirm older syn-Ghubrah and juvenile
lation, which makes up 10% of all analysed zircons, ranges from Neoproterozoic basement sources.
750 to 910 Ma, but only two grains pass the stricter filter, with Bowring et al. (2007) analysed a turbiditic sandstone from the
ages at 810 Ma. It is clearly hazardous to infer the detrital geochro- Fiq Formation collected from the northern part of Wadi Mistal.
nology of the entire Ghubrah Formation from samples collected Of 12 single grains analysed, one grain was c. 920 Ma, ten
from several lithologies at just one locality, and further sampling grains were in the range 810– 860 Ma and one grain yielded an
and analysis is required. With the data currently available, we age of c. 664 Ma. A further sample of an ashy sediment inter-
infer an exclusively Neoproterozoic juvenile source for the bedded with dropstone-bearing laminites and diamictites from
Ghubrah zircons, with a syndepositional crystallization age of core recovered from the Lahan-1 well, thought to be equivalent
714 Ma and a reworked contribution from older sources aged to a position very near the top of the Fiq, yielded zircons with
755– 826 Ma. The c. 755 Ma crystals may represent magmatism ages in the range c. 645– 718 Ma, with significant scatter. Only
associated with the beginning of continental extension responsible five young grains grouped around 645 Ma were present. Conse-
for the Ghubrah basins, or derivation from c. 760 Ma intrusives quently, it is safe to merely conclude that the top of the Fiq
associated with the welding together of arc and gneiss terranes Formation must be younger than 645 Ma. Rieu et al.’s (2007a)
in the Arabian – Nubian Shield to the west. The older zircons larger data set of Fiq samples from Wadi Sahtan and Wadi
(c. 810–825 Ma) have identical ages to basement rocks exposed Mistal (307 analyses, 167 of which passed the stricter filter)
at Al Jobah (Allen & Leather 2006; Allen 2007) at the northern reveals important Mesoproterozoic (1%), Palaeoproterozoic
end of the Huqf region and to the ages of peak metamorphism (5%) and even Archaean (,1%) components. Main peaks of the
and magmatism in the gneisses of Mirbat (Mercolli et al. 2006; most concordant data are at c. 790 –810 Ma and c. 840– 860 Ma,
Allen et al. 2011). with secondary peaks or shoulders at c. 700– 720 Ma and
A number of attempts have been made to date the volcanic and c. 750 Ma.
volcaniclastic Saqlah unit at the base of the Fiq Formation. Combining all data, we conclude that the depositional age of the
Bowring et al. (2007) analysed a sample from the Saqlah in glaciogenic Fiq Formation must be younger than 645 Ma, but that
Wadi Mistal, recovering seven zircons, three of which yielded an the formation mostly contains detrital zircons derived from the
age of c. 750 Ma and four of which yielded ages in the range same Neoproterozoic sources as seen in older parts of the Abu
c. 800 –830 Ma. Further samples of ashy material yielded dates Mahara Group, with two additional components: (i) first cycle
between 750 and 826 Ma, indicating reworked zircons from derivation of very old sources, suggesting progressively deep
sources indistinguishable from those of the Ghubrah. The most exhumation/unroofing of accreted terranes in the catchment
260 P. A. ALLEN ET AL.

areas of the Fiq basins, or inheritance of older zircons derived from We are grateful for the financial and logistical support of Petroleum Develop-
Neoproterozoic terranes (Hargrove et al. 2006); (ii) minor mag- ment Oman over many years and specifically the help of H. al Siyabi, M.
matic sources (.645 Ma) that are most likely contemporaneous Newall, J. Amthor and J. Schreurs. We are also grateful for the additional
with Fiq basin development. financial support of Enterprise Ireland and the Research Committee of ETH-
Zürich. A large number of individuals and field parties have provided input
on the Abu Mahara Group of the Jabal Akhdar for which we are grateful:
B. Levell, A. Heward, P. Hoffman, G. Halverson, A. Matter, and the partici-
Discussion and uncertainties pants of the field trip associated with the 24th IAS Meeting of Sedimentology
in Muscat, January 2005. We are grateful for the reviews of N. Eyles and I.
The Fiq Formation has been used as an example of repeated glacial Fairchild, and the editing of E. Arnaud. This represents a contribution of the
advance and retreat before post-glacial transgression, signifying IUGS- and UNESCO-funded IGCP (International Geoscience Programme)
dynamic glaciation and vigorously acting sediment routing Project #512.
systems throughout the glacial epoch (Allen & Etienne 2008).
Glacially influenced strata, principally proximal and distal
rainout diamictites bearing dropstones, lonestones and faceted
and striated clasts, are embedded in non-glacial stratigraphy domi- References
nated by hemipelagics, sandstone turbidites and debrites, but Allen, P. A. 2007. The Huqf Supergroup of Oman: Basin development
including wave-rippled sandstones requiring an ice-free sea and context for Neoproterozoic glaciation. Earth Science Reviews,
surface. Further details of the sedimentology of the Fiq Formation 84, 139– 185, doi: 10.1016/j.earscirev.2007.06.005.
and its use as a test for the Snowball Earth hypothesis can be found Allen, P. A. & Leather, J. 2006. Siliciclastic marine sedimentation in
in Leather (2001), Leather et al. (2002), Allen et al. (2004) and the aftermath of a Marinoan glacial epoch: The Masirah Bay For-
Allen & Etienne (2008). mation, Huqf Supergroup, of Oman. Precambrian Research, 144,
Although the glacial intervals in the Sultanate of Oman have 167– 198, doi: 10.1016/j.precamres.2005.10.006.
received some prominence in the literature, there remain areas of Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to snowball
uncertainty that would benefit from further work. Chief among Earth. Nature Geoscience, 1, doi: 10.1038/ngeo355.
these are the following: Allen, P. A., Leather, J. & Brasier, M. D. 2004. The Neoproterozoic
Fiq glaciation and its aftermath, Huqf Supergroup of Oman. Basin
† The age of the Fiq Formation is only constrained to terminate at Research, 16, 507– 534, doi: 10.1111/j.1365-2117.2004.00249.x.
younger than 645 Ma based on the presence of a small number Allen, P. A., Rieu, R., Etienne, J. L., Matter, A. & Cozzi, A. 2011.
of detrital zircons in core. A more systematic study of detrital The Ayn Formation of the Mirbat Group, Dhofar, Oman. In:
zircon ages in the Fiq Formation is warranted. Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
† The relationship between the Fiq and Ghubrah formations is logical Record of Neoproterozoic Glaciations. Geological Society,
poorly known and requires structural mapping in the key London, Memoirs, 36, 239–249.
wadis, especially Wadi Mistal. The absence of a major strati- Amthor, J. E., Grotzinger, J. P., Schröder, S., Bowring, S. A.,
graphic gap would imply glaciation took place intermittently Ramezani, J., Martin, M. W. & Matter, A. 2003. Extinction of
throughout the period from c. 720 to 640 Ma. The presence of Cloudina and Namacalathus at the Precambrian – Cambrian boundary
a significant stratigraphic gap and/or angular unconformity in Oman. Geology, 31, 431–434.
would support the idea of discrete glacial epochs in the older Anderson, J. B., Brake, C. F. & Myers, N. C. 1984. Sedimentation on
and younger Cryogenian. the Ross Sea continental shelf, Antarctic. Marine Geology, 57,
† The Ghubrah Formation is almost entirely unknown sedimento- 295– 334.
logically. Notwithstanding its commonly highly cleaved nature, Bahlburg, H. & Dobrzinski, N. 2011. A review of the Chemical Index
there is scope for a sedimentological study of the Formation of Alteration (CIA) and its application to the study of Neoproterozoic
across the Jabal Akhdar region, including Wadi Mu’aydin. glacial deposits and climate transitions. In: Arnaud, E., Halverson,
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
terozoic Glaciations. Geological Society, London, Memoirs, 36,
81 – 92.
Conclusions Beurrier, M., Bechennec, F., Rabu, R. & Hutin, G. 1986. Geological
Map of Rustaq, Sheet NF 40-3D, scale 1:100,000. Directorate General
The .400-m-thick Ghubrah Formation is dominated by thick, of Minerals, Oman Ministry of Petroleum and Minerals.
mostly unstratified marine rainout diamictites, with an interbedded Boulton, G. S. 1978. Boulder shapes and grain-size distribution of debris
tuffaceous ash that has yielded a U –Pb zircon age of 713.7 + as indicators of transport paths through a glacier and till genesis. Sedi-
0.5 Ma. The overlying ,1.5-km-thick Fiq Formation commences mentology, 25, 773– 799.
locally with the volcanogenic Saqlah unit of submarine pillow Blendinger, W., van Vliet, A. & Hughes Clarke, M. W. 1990.
basalts, ashes and volcaniclastic rocks, which passes up into the Updoming, rifting and continental margin development during the
well-bedded cycles of the Fiq Formation, comprising a wide Late Palaeozoic in northern Oman. In: Robertson, A. H. F.,
range of glaciogenic and non-glacial sediment gravity flow and Searle, M. P. & Ries, A. C. (eds) The Geology and Tectonics of
shallow marine facies. Detrital zircons from ashy beds in the the Oman Region. Geological Society, London, Special Publications,
upper part of the Fiq Formation are as young as 645 Ma. The 49, 27 – 37.
exact relationship between the Ghubrah Formation and the Fiq Bowring, S. A., Grotzinger, J. P., Condon, D. J., Ramezani, J.,
Newall, M. J. & Allen, P. A. 2007. Geochronological constraints
Formation is not fully resolved and requires further study.
on the chronostratigraphic framework of the Neoproterozoic Huqf
Calculation of CIA from mudstones and the muddy matrix of
Supergroup, Sultanate of Oman. American Journal of Science, 307,
diamictites reveals large excursions attributable to secular vari- 1097– 1145, doi: 10.2475/10.2007.01.
ations in the intensity of chemical weathering on contemporary Brand, U. & Veizer, J. 1980. Chemical diagenesis of a multicomponent
land surfaces. Such variations were most likely driven by pulses carbonate system 1: Trace elements. Journal of Sedimentary Petrol-
of glaciation separated by interglacial periods. The deposition of ogy, 50, 1219– 1236.
the transgressive cap carbonate coincides with a strong increase Brand, U. & Veizer, J. 1981. Chemical diagenesis of a multicomponent
in CIA, indicative of climatic transit following glaciation. carbonate system 2: Stable isotopes. Journal of Sedimentary Petrol-
Carbon-isotopic ratios in the Hadash carbonate range from –1 to ogy, 51, 987– 997.
–8‰, but variability between sample stations is large, making Brasier, M. D., McCarron, G., Tucker, R., Leather, J., Allen, P. A.
intrabasinal correlation problematic and extrabasinal correlation & Shields, G. 2000. New U– Pb zircon dates for the Neoproterozoic
hazardous. Ghubrah glaciation and for the age of the top of the Huqf Supergroup,
THE ABU MAHARA GROUP 261

Oman. Geology, 28, 175–178, doi: 10.1130/0091-7613(2000)028 Gorin, G. E., Raacz, L. G. & Walter, M. R. 1982. Late
,0175:NUPZDF.2.2.CO;2. Precambrian –Cambrian sediments of Huqf Group, Sultanate of
Bristow, T. F. & Kennedy, M. J. 2008. Carbon isotope excursions and Oman. American Association Petroleum Geologists Bulletin, 66,
the oxidant budget of the Ediacaran atmosphere and ocean. 2609–2627.
Geology, 36, 863–866, doi: 10.1130/G24968A.1. Gray, D. R., Miller, J. M. & Gregory, R. T. 2005. Strain state and
Brodzikowski, K. & van Loon, A. J. 1991. Glaciogenic Sediments. kinematic evolution of a fold-nappe beneath the Samail Opiolite,
Developments in Sedimentology, 49, Elsevier, Amsterdam, 674. Oman. Journal Structural Geology, 27, 1986– 2007.
Burns, S. J. & Matter, A. 1993. Carbon isotopic record of the Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
latest Proterozoic from Oman. Eclogae Geologicae Helvetiae, 86, glaciation (Neoproterozoic) in northeast Svalbard. Basin Research,
595– 607. 16, 297– 324.
Costa, J. E. 1988. Rheologic, geomorphic, and sedimentologic differen- Hargrove, U. S., Stern, R. J., Kimura, J.-I., Manton, W. I. & Johnson,
tiation of water floods, hyperconcentrated flows, and debris flows. P. R. 2006. How juvenile is the Arabian– Nubian Shield? Evidence
In: Baker, V. R., Kochel, R. C. & Patton, P. C. (eds) Flood Geo- from Nd isotopes of pre-Neoproterozoic inherited zircon in the Bi’r
morphology. Wiley, New York, 113–122. Umq suture zone, Saudi Arabia. Earth & Planetary Science Letters,
Cozzi, A. & Al Siyabi, H. 2004. Sedimentology and play potential of 252, 308–326.
the late Neoproterozoic Buah carbonates of Oman. GeoArabia, 9, Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
11 – 36. testing the limits of global change. Terra Nova, 14, 129– 155,
Deynoux, M. 1985. Terrestrial or waterlain diamictites? Three case doi: 10.1046/j.1365-3121.2002.00408.x.
studies from the Late Precambrian and Late Ordovician glacial Hoffman, P. F., Kaufmann, A. J., Halverson, G. P. & Schrag, D. P.
drifts in West Africa. Palaeogeography, Palaeoclimatology, 1998a. A Neoproterozoic snowball Earth. Science, 281, 1342– 1346,
Palaeoecology, 51, 97 –142. doi: 10.1126/science.281.5381.1342.
Dowdeswell, J. A., Hambrey, M. J. & Wu, R. 1985. A comparison of Hoffman, P. F., Kaufman, A. J. & Halverson, G. P. 1998b. Comings
clast fabric and shape in Late Precambrian and modern glaciogenic and goings of global glaciation on a Neoproterozoic tropical platform
sediments. Journal of Sedimentary Petrology, 55, 691– 704. in Namibia. GSA Today, 8, 1– 9.
Elverhoi, A., Pfirman, S. L., Solheim, A. & Larssen, B. B. 1989. Hughes Clarke, M. W. 1988. Stratigraphy and rock nomenclature in the
Glaciomarine sedimentation in epicontinental seas exemplified by oil-producing area of interior Oman. Journal of Petroleum Geology,
the northern Barents Sea. Marine Geology, 85, 225– 250. 11, 5– 60.
Eyles, C. H. 1988. Glacially and tidally influenced shallow marine sedi- Kapp, H. E. & Llewellyn, P. G. 1965. The geology of the Central Oman
mentation of the Late Precambrian Port Askaig Formation. Palaeo- Mountains. Report S00005-9, Geological Group, Petroleum Develop-
geography, Palaeoclimatology, Palaeoecology, 68, 1– 25. ment Oman.
Eyles, C. H. & Lagoe, M. B. 1990. Sedimentation patterns and facies geo- Kaufman, A. J. & Knoll, A. H. 1995. Neoproterozoic variations in the
metries on a temperate glacially-influenced continental shelf: the C-isotopic composition of seawater: stratigraphic and biogeochem-
Yakataga Formation, Middleton Island, Alaska. In: Dowdeswell, ical implications. Precambrian Research, 73, 27– 49.
J. A. & Scourse, J. D. (eds) Glaciomarine Environments: Processes Kellerhals, P. & Matter, A. 2003. Facies analysis of a glaciomarine
and Sediments. Geological Society, London, Special Publications, sequence, the Neoproterozoic Mirbat Sandstone Formation, Sultanate
53, 363– 386. of Oman. Eclogae Geologicae Helvetiae, 96, 49– 70.
Eyles, N., Eyles, C. H. & Miall, A. D. 1983. Lithofacies types and ver- Kennedy, M. J. 1996. Stratigraphy, sedimentology and isotope geochem-
tical facies models: an alternative approach to the description and istry of Australian Neoproterozoic postglacial cap dolostones: degla-
environmental interpretation of glacial diamict and diamict ciation, d13C excursions, and carbonate precipitation. Journal of
sequences. Sedimentology, 30, 393– 410. Sedimentary Research, 66, 1050–1064.
Eyles, C. H., Eyles, N. & Miall, A. D. 1985. Models of glaciomarine Kilner, B., MacNiocaill, C. & Brasier, M. D. 2005. Low-latitude gla-
sedimentation and their application to the interpretation of ancient ciation in the Neoproterozoic of Oman. Geology, 33, 413– 416,
glacial sequences. Palaeogeography, Palaeoclimatology, Palaeoe- doi: 10.1130/G21227.1.
cology, 51, 15 –84. Leather, J. 2001. Sedimentology, chemostratigraphy and geochronology
Fairchild, I. J. & Hambrey, M. J. 1984. The Vendian succession of of the lower Huqf Supergroup, Oman. PhD thesis, Trinity College
northeastern Spitzbergen: petrogenesis of a dolomite-tillite associ- Dublin, vols 1 and 2.
ation. Precambrian Research, 26, 111– 167. Leather, J., Allen, P. A., Brasier, M. D. & Cozzi, A. 2002. Neoproter-
Fedo, C. M., Nesbitt, H. W. & Young, G. M. 1995. Unravelling the ozoic snowball Earth under scrutiny: evidence from the Fiq glaciation
effects of potassium metasomatism in sedimentary rocks and of Oman. Geology, 30, 891– 894.
paleosols, with implications for paleoweathering conditions and pro- Le Guerroué, E. 2010. Duration and synchroneity of the largest negative
venance. Geology, 23, 921–924, doi: 10.1130/0091-7613(1995)023 carbon isotope excursion on Earth: The Shuram/Wonoka
,0921:UTEOPM.2.3.CO;2. anomaly. Comptes Rendus Geosciences. doi: 10.1016/j.crte.2009.
Fike, D. A., Grotzinger, J. P., Pratt, L. M. & Summons, R. E. 2006. 12.008.
Oxidation of the Ediacaran Ocean. Nature, 444, 744– 747, doi: Le Guerroué, E. & Cozzi, A. 2010. Veracity of Neoproterozoic
10.1038/nature05345. negative C isotope values: The termination of the Shuram
Fölling, P. G. & Frimmel, H. E. 2002. Chemostratigraphic correlation of negative excursion. Gondwana Research, 17, 653–661, doi:
carbonate successions in the Gariep and Saldania Belts, Namibia and 10.1016/j.gr.2009.11.002.
South Africa. Basin Research, 14, 69– 88. Le Guerroué, E., Allen, P. A. & Cozzi, A. 2005. Two distinct
Frimmel, H. E., Fölling, P. G. & Eriksson, P. G. 2002. Neoproterozoic glacial successions in the Neoproterozoic of Oman. GeoArabia, 10,
tectonic and climatic evolution recorded in the Gariep Belt, Namibia 17– 34.
and South Africa. Basin Research, 14, 55 –68. Le Guerroué, E., Allen, P. A. & Cozzi, A. 2006a. Chemostratigraphic
Gass, I. G., Ries, A. C., Shackleton, R. M. & Smewing, J. D. 1990. and sedimentological framework of the largest negative carbon isoto-
Tectonics, geochronology and geochemistry of the Precambrian pic excursion in Earth history: the Neoproterozoic Shuram Formation
rocks of Oman. In: Robertson, A. H. F., Searle, M. P. & Ries, (Nafun Group, Oman). Precambrian Research, 146, 68 – 92, doi:
A. C. (eds.) The Geology and Tectonics of the Oman Region. Geologi- 10.1016/j.precamres.2006.01.007.
cal Society, London, Special Publications, 49, 585– 599. Le Guerroué, E., Allen, P. A., Cozzi, A., Etienne, J. L. & Fanning,
Glennie, K. W. 1977. Outline of the geology of Oman. Mémoires de la M. 2006b. 50 million year duration negative carbon isotopic
Societé Géologique de France, hors series, 8, 25 – 31. excursion in the Ediacaran ocean. Terra Nova, 18, 147–153, doi:
Glennie, K. W., Boeuf, M. G. A., Hughes Clark, M. W., 10.1111/j.1365-3121.2006.00674.x.
Moody-Stuart, M., Pilaar, W. F. H. & Reinhardt, B. M. 1974. Le Guerroué, E., Allen, P. A. & Cozzi, A. 2006c. Parasequence devel-
Geology of the Oman Mountains. KSEPL, Rijswijk, The Netherlands, opment in the Ediacaran Shuram Formation (Nafun Group, Oman):
423. high resolution stratigraphic test for primary origin of negative carbon
262 P. A. ALLEN ET AL.

isotopic ratios. Basin Research, 18, 205– 220, doi: 10.1111/ Powell, R. D. & Molnia, B. F. 1989. Glaciomarine sedimentary pro-
j.1365-2117.2006.00292.x. cesses, facies and morphology of the south –southeast Alaska Shelf
Le Métour, J., Villey, M. & de Gramont, X. 1986. Geological map of and fjords. Marine Geology, 85, 359– 390.
Quryat, Sheet NF 40-4D, scale 1:100,000. Directorate of Minerals, Powell, R. & Domack, E. 1995. Glaciomarine processes and sediments.
Oman Ministry of Petroleum and Minerals. In: Menzies, J. (ed.) Modern Glacial Environments. Butterworth-
Loosveld, R., Bell, A. & Terken, J. 1996. The tectonic evolution of Heinemann, Oxford.
interior Oman. GeoArabia, 1, 28 –50. Rabu, D. 1988. Géologie de l’authochthon des montagnes d’Oman, la
Mackiewicz, N. E., Powell, R. D., Carlson, P. R. & Molnia, B. F. fenetre du Jabal Akhdar. PhD thesis, Université Pierre et Marie
1984. Interlaminated ice-proximal glaciomarine sediments in Muir Curie, Paris 6, and documents BRGM 130.
Inlet, Alaska. Marine Geology, 57, 113– 147. Rabu, D., Bechennec, F., Beurrier, M. & Hutin, M. 1986. Geological
Mann, A. & Hanna, S. S. 1990. The tectonic evolution of pre-Permian map of Nakhl. Sheet NF-40-3E, scale 1:100,000. Oman Ministry of
rocks, Central and Southeastern Oman Mountains. In: Robertson, Petroleum and Minerals, Directorate General of Minerals.
A. H. F., Searle, M. P. & Ries, A. C. (eds) The Geology and Rabu, D., Nehlig, P. et al. 1993. Stratigraphy and structure of the Oman
Tectonics of the Oman Region. Geological Society, London, Mountains. Bureau de Récherches Géologiques et Minières, 221.
Special Publications, 49, 307– 325. Rieu, R. 2006. Sedimentology, Stratigraphy and Geochemistry of the
Marshall, J. D. 1992. Climatic and oceanographic isotopic signals from Glacially Influenced Neoproterozoic Mirbat Group, Oman. PhD
the carbonate rock record and their preservation. Geological Maga- thesis, ETH-Zürich.
zine, 129, 143– 160. Rieu, R., Allen, P. A., Cozzi, A., Kosler, J. & Bussy, F. 2007a. A com-
McCarron, G. M. E. 2000. The Sedimentology and Chemostratigraphy posite stratigraphy for the Neoproterozoic Huqf Supergroup of Oman:
of the Nafun Group, Huqf Supergroup, Oman. PhD thesis, University integrating new litho-, chemo- and chronostratigraphic data of the
of Oxford. Mirbat area, south Oman. Geological Society London Journal, 164,
Mercolli, I., Briner, A. P., Frei, R., Schönberg, R., Nagler, T. F., 997– 1009.
Kramers, J. & Peters, T. 2006. Lithostratigraphy and Rieu, R., Allen, P. A., Plötze, M. & Pettke, T. 2007b. Compositional
geochronology of the Neoproterozoic crystalline basement of and mineralogical variations in a Neoproterozoic glacially influenced
Salalah, Dhofar, Sultanate of Oman. Precambrian Research, 145, succession, Mirbat area, south Oman: Implications for paleoweather-
182– 206. ing conditions. Precambrian Research, 154, 248– 265, doi: 10.1016/
Miller, J. M. G. 1996. Glacial sediments. In: Reading, H. G. (ed.) Sedi- j.precamres.2007.01.003.
mentary Environments: Processes, Facies and Stratigraphy. Black- Rieu, R., Allen, P. A., Plötze, M. & Pettke, T. 2007c. Climatic cycles
well Science Ltd, Oxford, 454– 484. during a Neoproterozoic ‘snowball’ glacial epoch. Geology, 35,
Moncrieff, A. C. M. 1989. Classification of poorly sorted sedimentary 299– 302, doi: 10.1130/G23400A.1.
rocks. Sedimentary Geology, 65, 191–194. Roger, J., Béchennec, F., Janjou, D., Le Métour, J., Wyns, R. &
Moncrieff, A. C. M. & Hambrey, M. J. 1990. Marginal marine glacial Buerrier, M. 1992. Explanatory Notes to the Geological Map of
sedimentation in the late Precambrian succession of east Greenland. Ja-alan. 1:100,000 Sheet NF 40-8E. Directorate General of Minerals,
In: Dowdeswell, J. A. & Scourse, J. D. (eds) Glaciomarine Ministry of Petroleum and Minerals.
Environments: Processes and Sediments. Geological Society, Simonson, B. M. & Carney, K. E. 1999. Roll-up structures: Evidence of
London, Special Publications, 53, 387– 410. in situ microbial mats in Late Archaean deep shelf environments. In:
Myrow, P. M. & Kaufman, A. J. 1999. A newly discovered cap carbonate Hagadorn, J. W., Pflueger, F. & Bottjer, D. J. (eds) Unexplored
above Varanger-age glacial deposits in Newfoundland. Canadian Microbial Worlds. Palaios. 14, 13 –24.
Journal of Sedimentary Research, 69, 784– 793. Stow, D. A. V., Reading, H. G. & Collinson, J. D. 1996. Deep
Narbonne, G. M., Kaufman, A. J. & Knoll, A. H. 1994. Integrated che- seas. In: Reading, H. G. (ed.) Sedimentary Environments: Pro-
mostratigraphy and biostratigraphy of the Windermere Supergroup, cesses, Facies and Stratigraphy. Blackwell Science Ltd, Oxford,
northwestern Canada: implications for Neoproterozoic correlations 395– 453.
and the early evolution of animals. Geological Society of America Tucker, M. E. 1986. Formerly aragonitic limestones associated with
Bulletin, 106, 1281–1292. tillites in the Late Proterozoic of Death Valley, California. Journal
Nardin, T. R., Hein, F. J., Gorsline, D. S. & Edwards, B. D. 1979. A of Sedimentary Petrology, 56, 818– 830.
review of mass movement processes, sediment and acoustic charac- Tschopp, R. H. 1967. The general geology of Oman. Proceedings of the
teristics, and contrasts in slope and base-of-slope systems versus 7th World Petroleum Congress, Mexico, 2, 231.
canyon-fan-basin systems. In: Doyle, L. J. & Pilkey, O. H. (eds) Villey, M., Le Métour, J. & Gramont, X. 1986. Geological map of
Geology of Continental Slopes. Society of Economic Paleontologists Fanjah, sheet NF 40-3F, scale 1:100,000. Oman Ministry of Pet-
and Mineralogists, Special Publication. 27, 61 –73. roleum and Minerals, Directorate General of Minerals.
Nesbitt, H. W., Young, G. M., McLennan, S. M. & Keays, R. R. 1996. Wentworth, C. K. 1936. An analysis of the shape of glacial cobbles.
Effects of chemical weathering and sorting on the petrogenesis of Journal of Sedimentary Petrology, 6, 85 –96.
siliciclastic sediments, with implications for provenance studies. Williams, G. E. 1979. Sedimentology, stable isotope geochemistry and
Journal of Geology, 104, 525– 542. palaeoenvironment of dolostones capping late Precambrian glacial
Ovenshine, A. T. 1970. Observations of iceberg rafting in Glacier Bay, sequences in Australia. Journal of the Geological Society of Austra-
Alaska, and the identification of ancient ice-rafted deposits. Bulletin lia, 26, 377–386.
Geological Society America, 81, 891– 894. Wright, R. & Anderson, J. B. 1982. The importance of sediment gravity
Platel, J. P., Roger, J., Peters, T., Mercolli, I., Kramers, J. D. & Le flow sediment transport and sorting in a glacial marine environment:
Métour, J. 1992a. Geological Map of Salalah, Sheet NE 40-09, 1: eastern Weddell Sea, Antarctica. Bulletin Geological Society
250000, with explanatory notes. Directorate General of Minerals, America, 93, 951–963.
Oman Ministry of Petroleum and Minerals. Wright, V. P., Ries, A. C. & Munn, S. G. 1990. Intraplatformal
Platel, J. P., Le Métour, J., Berthiaux, A., Buerrier, M. & Roger, J. basin-fill from the Infracambrian Huqf Group, east-central Oman.
1992b. Geological map of Juzor Al Halaaniyat, sheet NE 40-10, scale In: Robertson, A. H. F., Searle, M. P. & Ries, A. C. (eds) The
1:250,000. Directorate General of Minerals, Oman Ministry of Geology and Tectonics of the Oman Region. Geological Society,
Petroleum and Minerals. London, Special Publications, 49, 601–616.
Chapter 21

The Tambien Group, Northern Ethiopia (Tigre)

NATHAN R. MILLER1*, DOV AVIGAD2, ROBERT J. STERN3 & MICHAEL BEYTH4


1
Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, TX 78712-0254, USA
2
Institute of Earth Sciences, Hebrew University of Jerusalem, Jerusalem 91904, Israel
3
Geosciences Department, University of Texas at Dallas, Richardson, TX 75083-0688, USA
4
Geological Survey of Israel, Jerusalem, Jerusalem 95501, Israel
*Corresponding author (e-mail: nrmiller@mail.utexas.edu)

Abstract: The Tambien Group of northern Ethiopia (Tigre), with probable correlatives in Eritrea, is a 2 –3-km-thick siliciclastic–
carbonate succession that was deposited in an intra-oceanic arc platform setting within the southern Arabian– Nubian Shield (ANS)
area (southern extension of the Nakfa Terrane) of the Mozambique Ocean. Its deposition occurred prior to ocean closure between con-
verging fragments of East and West Gondwana and concomitant structural emergence of the East African Orogen (EAO). The Tambien
Group is well exposed and best studied in the Mai Kenetal and Negash synclinoria, where litho- and chemostratigraphy (including
d13Ccarb, 87Sr/86Sr) provide the basis for a composite reference section. Two glaciogenic intervals have been suggested from exposures
within the Didikama and Matheos Formation in the Negash Synclinorium. No reliable palaeomagnetic data exist to constrain the palaeo-
latitude of Tambien Group deposition and the southern ANS, but palaeogeographic reconstructions and evaporite pseudomorphs in lower
carbonate units (Didikama Formation) imply low to intermediate latitudes (,458). Integration of available geochronological information
(regional magmatism and detrital zircon) suggests c. 775– 660 Ma as a plausible window constraining deposition of the prospective
glacial intervals.
The Tambien Group appears to preserve a coherent chemostratigraphic framework that can be effectively subdivided according to
shifts in d13Ccarb polarity [polarity intervals A (þ), B (–), C (þ), D (–)]. Slates underlying and interstratified with polarity interval A
carbonate preserve evidence of extreme chemical weathering that lessened prior to deposition of polarity interval B carbonate.
Tambien Group carbonate units have sedimentological characteristics consistent with both shallow and deeper marine depositional set-
tings. The lower prospective glacial interval lacks diagnostic sedimentological evidence of synglacial deposition, but is overlain by nega-
tive d13C carbonate (polarity interval B) with sedimentological characteristics consistent with well-documented cap-carbonate
successions. The upper prospective glacial interval in the Negash Synclinorium (Matheos Diamictite) best exhibits characteristics con-
sistent with glaciogenic deposition (matrix-supported polymictic clasts, possible dropstones, possible bullet-nosed and striated clasts). In
contrast to pericratonic rift margin settings that are common for Cryogenian glaciogenic deposits, palaeogeographic reconstructions for
the 775 –660 Ma timeframe place northern Ethiopia within an intra-oceanic setting that was likely far removed from cratonic hinterlands.
More work on Tambien Group sedimentology, geochronology and palaeogeography is required to better evaluate the extent and timing of
glacial conditions associated with the prospective glaciogenic intervals.

Supplementary material: Supplementary Table 21.1 of Tambien Group geochronological age constraints is available at http://www.
geolsoc.org.uk/SUP18462.

The Tambien Group is exposed (Fig. 21.1) throughout portions of Matheos Formation in the core of the Negash Synclinorium
northern Ethiopia (Tigre Province) and Eritrea (NNE extensions (Fig. 21.2a), and this interval may be equivalent to arkosic sand-
from Figure 21.1; Bizen domain and Adobha Abi terrane of stone and conglomerate of the Dugub Formation that similarly
Beyth et al. (2003) and De Souza Filho & Drury (1998)), in tops the Tambien Group in the western Shiraro area (Fig. 21.1,
greenschist-grade terranes comprising the southern portion of the Avigad et al. 2007). The upper diamictite unit at Negash was orig-
Arabian –Nubian Shield (ANS). It may be equivalent to similar inally defined as the ‘Pebbly slate’ (Beyth 1972). It was
carbonate-rich units in NE Sudan (Bailateb Group; Stern et al. subsequently assigned within the Matheos Formation (Garland
1994), SW Saudi Arabia (Hali Group; Greenwood et al. 1976), 1980) and informally described as ‘Pebbly Slate (diamictite)’ in
and possibly western Yemen (inferred from its Red Sea conjugate Miller et al. (2003), ‘Slate/Pebbly slate (Diamictite Slate)’ in
position). Its deposition marks mainly marine siliciclastic and car- Alene et al. (2006) and ‘Negash Diamictite’ in Avigad
bonate sedimentation within the Mozambique Ocean, an extinct et al. (2007). Miller et al. (2009) subdivide the Matheos Formation
Neoproterozoic ocean basin destroyed during the late Neoprotero- into three members, the youngest corresponding to the diamictite
zoic consolidation of Greater Gondwana with the emergence of the facies. In consideration of a possible lower glaciogenic interval
East African Orogen (EAO) (Stern 1994). The Tambien Group is in the Negash Synclinorium, the informal term ‘Matheos Diamic-
best studied in northern Ethiopia from the Mai Kenetal (western tite’ is suggested for the upper prospective glacial interval. In
flank near 138550 N, 388500 E) and Negash (southern extent near addition to these northern Ethiopian localities, Eritrea hosts poss-
138500 N, 398370 E) synclinoria, where much of the Tambien ible glaciogenic units that have not been systematically studied.
Group is continuously exposed, and these localities provide the The southern ANS is still very much a frontier region in need of
substantial basis for regional litho- and chemostratigraphy. systematic sedimentological, geochemical, and geochronological
Two possible glaciogenic intervals have been suggested within studies of Neoproterozoic units. Much of what is known about the
the Tambien Group (Beyth et al. 2003; Miller et al. 2003, 2009). Cryogenian Period for the region has been learned in only the past
The lower interval occurs as a greywacke-conglomerate interval decade. The earliest suggestion of a Late Proterozoic glaciation
within slate of the lower Didikama Formation in the Negash was by Bibolini (1920), who described faceted clasts (facce piane)
Synclinorium (Fig. 21.2a, column D), and the top of this interval within an unnamed pebbly mudstone-conglomerate unit (conglom-
may correlate with the base of the Assem Limestone in the Mai erati poligenici) in northern Eritrea. The basic Neoproterozoic
Kenetal Synclinorium (see discussion). The upper prospective stratigraphic framework for northern Ethiopia was established in
glacial interval tops the Tambien Group as diamictite of the regional mapping by Beyth (1972), including description of the

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 263– 276. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.21
264 N. R. MILLER ET AL.

Fig. 21.1. Location of key Tambien Group


exposures (unshaded units) (A, Shiraro area; B,
Mai Kenetal Synclinorium; C, Negash
Synclinorium; D, Samre area) within northern
Ethiopia (Tigre Province) modified from Miller
et al. (2009). Metavolcano–sedimentary block
boundaries (Tadesse et al. 1999) occur only
within the Tsaliet Group, and are inferred to
represent accreted arc terranes or slivers in a
supra-subduction zone setting. Stars show
geochronological localities for syn-tectonic
intrusives (white), post-orogenic intrusives
(black) and detrital zircons (white stars with black
dots) discussed in the text and shown in
Figure 21.5 (also in Supplementary Table 21.1).
Prospectively equivalent upper Tambien Group
strata of the Gulgula Group occur in western
Eritrea (arrow) just west of the Shiraro Area, as
well as in the Adobha Abi terrane and Bizen
Domain (arrows) to the north in Eritrea. Dashed
box east of Mai Kenetal outlines the Werii study
area of Sifeta et al. (2005). M. Alem marks the
western limb locality (Madahne Alem) within the
Negash Synclinorium bearing 774.7 + 4.8 Ma
zircons in slate c. 16 m below lowest Tambien
Group carbonate beds (Avigad et al. 2007). The
Negash Synclinorium is structurally bounded to
the east by the Atsbi Horst (AH ). Inset map
(below) shows the location of an unnamed pebbly
mudstone unit in northeastern Eritrea (Cecioni
1981) that could correlate to the Tambien Group.

units now posited to have glacial associations. Motivated by the ANS consists of a patchwork of Neoproterozoic tectonostrati-
Snowball Earth hypothesis, a number of reconnaissance studies graphic terranes, now bisected by the Oligocene and younger
have since explored the depositional context of the Tambien Red Sea rift. ANS terranes include significant volumes of
Group (Beyth et al. 2003; Miller et al. 2003; Alene et al. 2006; see juvenile Neoproterozoic crust generated within spreading centers,
also Stern et al. 2006). More comprehensive regional investigations, arc and back-arc settings of the Mozambique Ocean (Stern 1994).
involving U–Pb zircon geochronology, chemical weathering Sutures between terranes, many with ophiolites and dated meta-
indices, and higher resolution sampling for integrated C- and morphic assemblages, document collisional deformation and help
Sr-isotope stratigraphy, have further refined the age and range of constrain the timing of terrane amalgamation. Many sutures have
litho- and chemostratigraphic variations in the Tambien Group appreciable strike – slip offsets, and these may broadly relate to
(Sifeta et al. 2005; Avigad et al. 2007; Miller et al. 2009). c. 600 Ma escape tectonics (Burke & Sengör 1986), during which
ANS terranes were progressively sandwiched by, and offset
between, obliquely converging Gondwana cratonic blocks (de
Structural framework Souza Filho & Drury 1998, and references therein). Tambien
Group exposures in Tigre occur within the presumed southern
(See additional region-specific structural information in the extension of the Nakfa terrane of Eritrea. The Nakfa terrane is
‘Glaciogenic deposits and associated strata’ section.) The greater one of several suture-bounded low-grade volcano-sedimentary
THE TAMBIEN GROUP, NORTHERN ETHIOPIA 265

(a)
Relation A. B. C. D.
to syn- Shiraro Plains Mai Kenetal Negash Negash
& post- Region Synclinorium Synclinorium Synclinorium
tectonic
Tadesse Beyth 1972 Miller et al. 2009
intrusives 1999 Beyth 1972 (Tadesse 1999) Beyth 1972
(Garland 1980)
Pz & Mz
Adigrat Ss Adigrat Ss Enticho Ss Enticho Ss Enticho Ss
Seds

Dugub Fm - - - Mentebtab Fm
Algal* Ls Mai Kenetal Ls (200m) Pebbly slate (200m)
Diamictite Mbr
(Tselim Imni Ls) ?

Matheos Fm
---- Tambien Group ----
Shiraro Group ?
Black detrital ls
Enda Zebi Fm Transitional Mbr
Shiraro Cgl Tsedia Slate (500m) (300m)
? (subdivided into
Post-orogenic granitoids

Black Ls Mbr
Arkosic ss upper Bilato and Mica dolomite &
lower Logmiti Slates) slate (700m)
Mentebtab Ss
Assem LS (300m) Slate & dolomite
Dolomite > Slate

Didikama Fm
Bedded ls (Filafil Ls) (200m)
Didikama Fm

Mica dolomite
Slate > Dolomite
Purple slate w/grn
Purple slate w/grn Werii Slate (1000m) reduction spots; GCI ?
reduction spots (Segali Slate) greywacke (300m) Lower Slate
? (Atsbi Horst)

Syn- Undiff. Tsaliet Tsaliet Tsaliet


Tsaliet Group
Metavolc. Metavolcanics Metavolcanics Metavolcanics
(b)
d13C carb Polarity

Mai Kenetal Synform Negash Synform Afro-Arabian Peneplain


~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
Mai Kenetal LS C (++) Matheos Fm D (-) Diamictite Mbr ? ? ?
Transition Mbr 3
2 Tsedia Slate C (+) C (++) Black LS Mbr
~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ unconformity?
Assem LS B (-) Didikama Fm B (-) Upper (dol>slate)
~~~~~~~~~~~~~~~ GCI ? ? ? 1
? ? ? paraconformity?
Werii Slate A (+?) A (+) Lower (slate >dol)
Lower Slate (Atsbi Horst)

Tsaliet Metavolcanics Tsaliet Metavolcanics

Fig. 21.2. (a) Lithostratigraphic subdivision of the Tambien Group in previous work. Prospective glacial intervals in columns A, C and D are superimposed from Beyth
et al. (2003) and Miller et al. (2009). The upper Algal Limestone unit in the Shiraro plains region (column A), originally termed the ‘Algal, stromatoporoidea (?)
limestone’ by Beyth (1972) was not mapped in the Axum Sheet (Tadesse 1999) and may occur in the Badme region (north of Shiraro, Fig. 21.1). (b) Chemo- and
lithostratigraphic correlations between the Mai Kenetal and Negash synclinoria used to create a composite Tambien Group stratigraphic section (see discussion).
Stratigraphic intervals in each section are alphabetized (A –D) according to the succession of associated d13Ccarb polarity (þ or – ) intervals. See Figure 21.3 for associated
chemostratigraphic characteristics (circled data sets: d13CTOC, 87Sr/86Sr, [Sr] and chemical weathering index) that support the correlations. Numbers 1–3 correspond to
the different segments used in constructing the composite stratigraphy of Figure 21.6. Abbreviations (A&B): Fm, formation; GCI, Greywacke –Conglomerate Interval,
the lower prospective glacial interval in the Negash Synclinorium; grn, green; Ls, limestone; Mbr, member; Metavolc, metavolcanics; Mz, Mesozoic; Pz, Palaeozoic;
Ss, sandstone; Undiff, undifferentiated. (Modified from Miller et al. 2009.)

terranes comprising the greater Nubian Tokar superterrane et al. 2005). An unconformable basal contact also was postulated
(Kröner et al. 1991). from considerable lateral variations observed among basal
The Nakfa terrane supracrustal assemblage records a lower Tambien metasediments (Arkin et al. 1971; Beyth 1972; Tadesse
sequence of igneous rocks associated with arc magmatism in and 1999). Tambien Group deposition, including the two prospective
around the southern ANS sector of the Mozambique Ocean, as glacial intervals, is thought to have occurred in an intra-oceanic
exemplified by calc-alkaline plutons and associated metavolcanic platform setting (above a substratum of consolidated arc terranes)
rocks of the Tsaliet Group. In western Tigre, the Tsaliet Group may following the main phase of Tsaliet arc magmatism in the region
be preserved as a series of accreted volcanic arc terranes within a (c. 780 + 30 Ma; Avigad et al. 2007). The extent to which
suprasubduction setting (Fig. 21.1; volcano-sedimentary blocks of Tsaliet arc subterranes were fully accreted prior to Tambien
Tadesse et al. 2000). Cessation of Tsaliet arc volcanism was fol- Group deposition is uncertain, and some syndepositional relief
lowed by deposition of weathered arc detritus (Sifeta et al. 2005) differentiation related to ongoing shortening and/or extension is
and in turn increasing proportions of carbonate sediments, as possible (Miller et al. 2009).
exemplified by the Tambien Group. The contact between the The entire Nakfa basement complex was deformed after depo-
Tsaliet and Tambien Group is poorly understood, ranging from sition of the Tambien Group in conjunction with closure of the
seemingly conformable and gradational to fault-bounded (Sifeta Mozambique Ocean between fragments of east and west Gondwana,
266 N. R. MILLER ET AL.

and the concomitant structural emergence of the EAO, beginning area, splitting Beyth’s informal Tambien Group facies into the
c. 630 Ma. As a consequence, NNE-trending upright/overturned basal Didikama Formation and overlying Shiraro Group
folds and faults dominate the regional structural grain, with Tambien (Fig. 21.2a, column A). Designated within the latter were three
Group exposures best preserved in synclinoria and grabens. Inter- metasiliciclastic formations (Enda Zebi, Dugub and Mentebtab
nal minor or secondary folds and faults complicate measured sec- formations), each with respective lateral facies variations.
tions in the Mai Kenetal and Negash synclinoria. Restoring In the eastern limb of the Negash Synclinorium, Beyth (1972)
shortening (200%) associated with these collisional structures mapped the base of the Tambien Group as an interval of arkosic
greatly extends the minimal aerial extent of the Tambien basin or sandstone and conglomerate, above a possible tectonic contact
basins. The deformed Tsaliet-through-Tambien Group succession with the Astbi Horst (Fig. 21.1, Fig. 21.2a column C). This
was subsequently intruded by c. 610 Ma (‘Mareb’) granitoids coarse metasiliciclastic interval corresponds to the lower prospec-
associated with crustal thickening and differentiation of the matur- tive glacial interval, the Greywacke-Conglomerate interval (GCI)
ing EAO. Whether or not, and for how long, Tambien Group depo- of the Didikama Formation. The basal contact of the GCI has not
sition continued prior to emplacement of the ‘Mareb’ post-orogenic been studied in detail, but the underlying depositional sequence
granitoids is unknown. Despite these post-depositional orogenic involves a minimum of several hundred metres of conformably
processes, temperatures and pressures did not exceed low- bedded fine-grained marine metasedimentary units (mainly slate
greenschist grade metamorphism and well-preserved primary sedi- with episodic thin carbonate interbeds up to a few centimetres
mentary structures are common in many Tambien Group carbonate thick). This lower slate succession contrasts from coarse volcanic
successions (Alene et al. 2006; Miller et al. 2009). agglomerate and conglomerate of the Tsaliet Group exposed at the
Extensive erosion of the EAO resulted in cutting of a widespread top of the Atsbi Horst (Fig. 21.1), but the contact between these
regional unconformity prior to Cambro-Ordovician time (Avigad units has not been systematically mapped or studied. Miller
et al. 2005), herein termed the Afro-Arabian Peneplain (AAP). et al. (2003, 2009) include this lower slate interval (below the
Uppermost Tambien Group exposures in the core of the Negash GCI) within the Tambien Group, as the informal Lower Slate
Synclinorium (Matheos Diamictite) and possibly equivalent silici- member of the Didikama Formation (Fig. 21.2a, column D).
clastic deposits of the Dugub Formation in the western Shiraro Above the GCI, the Didikama Formation transitions into varie-
area (Fig. 21.1) are thought to be northern Ethiopia’s youngest gated slate that includes the first prominent dolomite beds (infor-
preserved Neoproterozoic metasediments below the AAP. mal Slate . Dolomite member) and thicker bedded dolomite at
the top of the formation (informal Dolomite . Slate member).
The GCI and overlying purple slate (with green reduction spots;
Stratigraphy the ‘Purple Egg Slate’ of Beyth 1972), with estimated combined
thickness of 300 m, forms a distinctive dark (grey-purple)
Beyth (1972) documented Tambien Group outcrops in the Shiraro marker unit that can be traced laterally throughout much of the
area, in several en echelon synclinoria to the east (i.e. Mai Kenetal, c. 30-km-long exposed eastern synclinorial limb.
Tsedia, Chemit and Negash), and the Escarpment area east of The Didikama Formation is overlain sharply, and very likely
Adigrat (Fig. 21.1, study areas A– C). Apparent regional facies unconformably (Beyth 1972, p. 74; Garland 1980, p. 14), by dis-
variations within the internal stratigraphy of the Tambien Group tinctive black limestone of the Matheos Formation. Miller et al.
were described, and a preliminary correlation was proposed for (2009) subdivide the Matheos Formation into three members.
three areas: Shiraro (west), Mai Kenetal (centre) and Negash The basal Black Limestone Member is well laminated and partly
(east). Subsequent workers have modified the stratigraphy in detrital (intraclasts), and transitions upward into non-calcareous
various ways as described below and shown in Figure 21.2a. slate (Transitional Member) and eventually pebbly slate (Diamic-
Figure 21.2b shows a recent correlation scheme proposed for the tite Member).
Tambien Group based on integrated litho- and chemostratigraphy
of the Mai Kenetal and Negash synclinoria (Miller et al. 2009),
which is further elaborated in the discussion section. Prospective glaciogenic deposits and associated strata
Above the Tsaliet Metavolcanics, Beyth (1972) assigned four
formational designations within the Tambien Group based on the Lower prospective glaciogenic unit
Mai Kenetal type section (in stratigraphic order): Werii Slate,
Assem Limestone, Tsedia Slate and Mai Kenetal Limestone. In Greywacke –Conglomerate Interval, lower Didikama Formation (Lower
the Negash Synclinorium, Beyth defined informal facies (in strati- Tambien Group). Beyth (1972) described the GCI within the
graphic order: greywacke and purple slate with green reduction eastern flank of the Negash Synclinorium as poorly sorted, light-
spots, slate and dolomite, mica dolomite and slate, black detrital grey to black greywacke, containing mostly medium grained to
limestone, and pebbly slate), which he suggested might correlate pebble-sized, subrounded quartz, in a mud (slate) supported
with the Mai Kenetal section (Fig 21.2a, columns C v. B). matrix (Fig. 21.2a, column C). Thin conglomeratic layers and
Dolomite-slate successions similar to those of the lower Tambien ripple marks (exposed in bedding planes) also occur within the
Group in the Negash Synclinorium were subsequently documented unit. The unit fines upward into purple slate with distinctive
in the Escarpment east of Adigrat, as well as in synclines in the ovate greenish-yellow patches, interpreted as reduction spots.
Shiraro area. Garland (1980) formalized Beyth’s Tsaliet Metavolca- This lower interval (c. 300 m thick, Beyth 1972) is overlain by var-
nics as the Tsaliet Group, and subdivided the Negash Tambien iegated slate (orange and green), which in turn grades upward
Group succession (Fig. 21.2a, column D) into the lower Didikama (with poor outcrop) into the lowest exposed dolomite beds of the
Formation (after similar dolomitic rocks in the Shiraro Area) and Didikama Formation. Similar purple spotted slate is reported
overlying Matheos Formation (equivalent to Beyth’s informal below the Didikama Formation in the Shiraro area (Beyth 1972;
black detrital limestone and pebbly slate facies). In mapping of Fig. 21.2a, column A). The GCI is underlain by well-layered
the Mai Kenetal synclinorium, Tadesse (1999) used local names slate (with occasional sub-decimetre-thick carbonate interbeds
to address the possibility that some of Beyth’s original type areas and minor foliation) that crops out continuously down-section
(i.e. Werii Slate, Tsedia Slate) might reside in different volcano- (eastward) for at least several hundred metres in the hanging
sedimentary arc terranes (Fig. 21.1). The main difference between wall flank of the Atsbi Horst (Fig. 21.1).
the two lithostratigraphic schemes is differentiation of Beyth’s
Tsedia Slate into a lower Logmiti Slate and overlying Bilato Lime- Upper Werri Slate (Lower Tambien Group). In the NW flank of the
stone and Slate (Fig. 21.2a, column B). Tadesse (1999) also formal- Mai Kenetal Synclinorium, the Assem Limestone lies sharply
ized a new Neoproterozoic stratigraphic scheme for the Shiraro and conformably above the Werii Slate and shares
THE TAMBIEN GROUP, NORTHERN ETHIOPIA 267

sedimentological and isotopic characteristics of well-studied cap- rocks overlie carbonate units assigned to the Didikama For-
carbonate sequences (i.e. abrupt basal lithological transition to mation (Tadesse 1999) and crop out within the Shiraro Block
high-energy carbonate facies, basal dolomite bed (c. 1 m thick) (Fig. 21.1), a graben-like depression that is fault-bounded to
transitioning upward into limestone, consistently negative the east against the Adi Hageray Block and extends westward
d13Ccarb, primary and/or early diagenetic (pre-compaction) pris- beyond the studied area into Eritrea. The areally most extensive
matic and radial fibrous cements (Miller et al. 2009, figs S2D – unit is the Dugub Formation, which we summarize following
G; Hoffman et al. 2007)). The observation that many Cryogenian Tadesse (1999) and our own studies 4.5 km ENE of the town
carbonate successions with negative d13C compositions are of Shiraro (Fig. 21.1).
affiliated with glaciogenic intervals (e.g. Yoshioka et al. 2003; The Dugub Formation consists of weakly metamorphosed con-
Halverson et al. 2005; Corsetti et al. 2007 and other arguments, glomerate, arkosic sandstone and siltstone (Tadesse 1999), which
see discussion) raises the possibility that the pre-Assem Limestone are regionally intercalated at different scales. Well-preserved
depositional sequence could have a glacial affiliation. The under- primary sedimentary structures are common as graded bedding,
lying Werri Slate is mainly finely laminated, well-foliated, non- cross-lamination, ripple marks, and flame and slump structures.
calcareous slate that does not exhibit obvious glaciogenic sedimen- Elliptical (rounded to subrounded) and well-sorted metaconglome-
tological characteristics. However, the nature of the upper Werri rate clasts (8 cm) comprise up to 40% of the rock volume. Clast
Slate and its contact with the Assem Limestone has not been compositions include low-grade (Tsaliet-like) volcanic rocks
studied regionally and poorly sorted metasiliciclastic intervals (chlorite schist, tuffaceous metasediments), phyllite, granite,
are documented, at least locally, in the Werri Slate. For example, quartz pegmatite, quartzite and chert. Fine- to medium-grained
Beyth (1972) reported occurrences of greywacke within the chlorite, muscovite, feldspar and quartz are notable components
Werii Slate regionally, and encountered poorly sorted rhyolitic of the metaconglomerate groundmass. Carbonate is notably
agglomerate with well-rounded metre-scale quartzitic fragments absent, but is reported (as scarce marble layers a few to tens of
at the southern end of Mai Kenetal syncline (around 138470 N metres thick) in the underlying Enda Zebi Formation (Tadesse
and 388510 E). Greywacke is also observed near its gradational 1999). Dugub metasandstone contains quartz pebbles (up to a
contact with Tsaliet Group (Alene et al. 2006). few millimetres across) and grades laterally and vertically into
non-pebbly sandstone and siltstone. Cross bedding exhibits set
heights ranging between 0.1 and 0.5 m; coset heights range up to
Upper prospective glaciogenic unit 1.5 m. Orientations of asymmetrical ripple marks and slump
structures suggest west-southwestward transport and palaeoslope
Diamictite Member, Matheos Formation (Upper Tambien Group). The directions (Tadesse 1999; Avigad et al. 2007).
Matheos Formation, comprising the interior of the Negash Syncli-
norium, records a conformable transition from black limestone to
diamictite. Ranging from faintly layered (massive) to well-bedded, Prospective glaciogenic deposits in Eritrea
with near vertical dips, the c. 250-m-thick basal Black Limestone
Member (Fig. 21.2a, column D) forms a distinctive ridge that The high-grade Arag terrane in NE Eritrea (c. 400 km NNW of
rims the interior of the Negash Synclinorium. Individual beds Negash) has a widespread unnamed pebbly mudstone unit, consist-
range in thickness from decimetre- to metre-scale and are typically ing of silicic clasts within an argillaceous-siliceous matrix (Verri
finely laminated in thin section. Intraclastic grainstone intervals are 1909; Bibolini 1920, 1921, 1922). The unit is reported as widely
common. Towards the synclinorial axis, limestone beds thin exposed between 178N and 178450 N along its north-northwesterly
(becoming sub-decimetre thick) and interstratify with increasing trend (c. 90 km) and penetrated by post-orogenic granitoids (acid
proportions of light to dark grey phyllitic slate. The overlying c. aplitic intrusions, Cecioni 1981), which have not been dated.
100 m Transitional Member includes non-calcareous phyllitic Cecioni (1981) described the pebbly mudstone matrix as a hard,
slate that passes upward into pebbly slate, without interbedded violet, clay-quartz cement, and the clasts as pebbles and boulders
carbonate. The primary depositional fabric involves mainly fine of quartzite and silicified schist. Some of the small pebbles
(,1 to 3 cm scale) horizontal beds. Initial disparate sedimentary (,1 cm) are pitted and knotted; others are angular. Bibolini
clasts are sub-centimetre scale (compositions have not been (1920, 1921) noted the occurrence of faceted clasts and postulated
studied). The overlying c. 200-m-thick Diamictite Member a glaciogenic origin. To our knowledge this is the earliest suggestion
occupies the tightly folded core of the Negash Synclinorium and of a possible Late Proterozoic (‘algonkiano’) glaciation within the
is distinguished by an overall upward increase in clast abundance ANS. Cecioni (1981) preferred a gravity flow origin for this unit,
and size (typically ,10 cm, but up to 20 cm in diameter). Clasts but recognized a possible glaciogenic association due to the
are matrix-supported and the finer matrix retains horizontal layer- occurrence of faceted pebbles. This unit trends more or less along
ing, with some lateral pinching and swelling, throughout the strike with Tambien Group exposures in northern Ethiopia and
member. Syn-orogenic compression of the synclinal core has prospective Bizen Domain equivalents in southern Eritrea, but has
deformed the diamictite to varying extents. More pelitic parts of apparently not been further studied in relation to a glacial
the diamictite display foliation and associated clasts are often some- association.
what elongated with pressure shadows. Despite these superimposed The Neoproterozoic supracrustal succession of western Eritrea
features, clasts that appear to preferentially deform underlying (just west of the Shiraro area in northern Ethiopia, Fig. 21.1) con-
matrix laminae are relatively common. Diamictite clasts have sub- sists of volcano-sedimentary assemblages (Augaro Group) uncon-
rounded, elongate and angular shapes (including some bullet-nosed formably overlain by basin-fill metasediments of the Gulgula
clasts), and some planar clast surfaces may be striated (Miller et al. Group (Teklay et al. 2003; Teklay 2006). The Gulgula succession
2003). Clast lithologies are polymictic, including felsic volcanic involves greenschist-grade polymictic conglomerate, shale, phyllite
rocks, fine-grained black limestone and dolomite (including well- with interstratified marble lenses, in addition to carbonaceous sand-
rounded clasts that retain primary sedimentary structures; e.g. stone, quartz arenite and allodapic carbonate, which, like the
oolite), low-grade semipelitic sediments, and rare volcanic con- Tambien Group, are deformed as upright, recumbent, and isoclinal
glomerate consistent with the upper Tsaliet Group. folds. Teklay (2006) describes the polymictic conglomerates as
clasts, both matrix- and clast-supported, ranging from a few centi-
Dugub Formation, Shiraro Group (Upper Tambien Group). Conglo- metres to a half-metre of granite, slate, phyllite and epidotite. The
merate and arkosic metasediments comprise youngest Tambien Gulgula Group merges southwards into the Shiraro block in north-
Group exposures in the lowland Shiraro area of western Tigre ern Ethiopia, suggesting a composite Gulgula-Shiraro depositional
(Shiraro Group of Tadesse 1999; Fig. 21.2a, column A). These basin and a lithostratigraphic affiliation with the Shiraro Group
268 N. R. MILLER ET AL.

(upper Tambien Group). The origin of Gulgula Group polymict data comprising most (c. 80%) analyses (Fig. 21.3). This compi-
conglomerates, in addition to those described above in the lation reveals significant chemostratigraphic trends, from which
Shiraro Group (Dugub Formation) require further study to clarify regional correlations are proposed (Fig. 21.2b, discussion).
whether or not glacial processes were involved. Carbonate d13C compositions for the Tambien Group range
between –8 and þ8‰ and can be characterized as fluctuating
repeatedly upsection between positive and negative polarity
Boundary relations with overlying and underlying (Fig. 21.3). Low-grade Bizen Domain stromatolitic carbonates
non-glacial units in SE Eritrea, considered to be Tambien Group equivalents
(Beyth et al. 2003), are among the most negative d13Ccarb values
Lower prospective glaciogenic unit reported (range, – 7.2 to –4.8‰; 3 cal, 5 dol) but sample preser-
vation was not critically assessed. Highest d13Ccarb compositions,
Greywacke-Conglomerate Interval (GCI) of the Didikama Formation, with maxima near þ7‰, derive from the Mai Kenetal Limestone
Negash Synclinorium. Where examined in the eastern limb of the and Matheos Formation Black Limestone Member. The negative
Negash synclinorium, the GCI occurs within the Didikama For- d13Ccarb compositions of the Assem Limestone (c. –1 to – 4‰)
mation (Fig. 21.2a, columns C and D) above thick slate (Lower are similar to those comprising the lower negative d13Ccarb interval
Slate Member) and below variegated (purple/green/orange) in the Negash Synclinorium (c. – 1 to – 3‰) (despite their contrast-
slate that, in turn, grades upward into lowest dolomite beds of ing mineralogies), as well as negative d13Ccarb intervals mapped as
the Didikama Formation (Slate . Dol Member). The continuity Assem Limestone in the Chemit ( –4.4 to –3.0‰) and Tsedia
of the transition from the Lower Slate member to the GCI is uncer- (– 4.5 to –0.7‰) synclinoria (Alene et al. 1999, 2006; Miller
tain due to minor folding and possible faulting associated with the et al. 2009). Bizen Domain stromatolitic dolomites with negative
eastward structural transition to the Atsbi Horst (Fig. 21.1), d13Ccarb (Beyth et al. 2003) may correlate with the negative
whereas the transition above the GCI appears to be continuous d13Ccarb interval of the Didikama Formation. The Negash
and conformable. carbonate sequence is so far unique in recording two negative
d13Ccarb excursions.
Upper Werri Slate, Mai Kenetal Synclinorium. Miller et al. (2009) Carbonate d18O compositions range between 0 and –16‰, with
make chemostratigraphic arguments that the equivalent of the dolomite typically enriched (by 2– 5‰) relative to stratigraphi-
GCI may occur in the Mai Kenetal Synclinorium below the cally proximal (or prospectively equivalent) limestone. Although
Assem Limestone, either within the upper Werii Slate or substantial scatter is apparent, median compositions are mainly
between the Werii Slate and Assem Limestone as a paraconfor- between –4 and –10‰, similar to other Cryogenian datasets
mity. The abrupt Werii Slate-Assem Limestone contact can be (Fig. 21.3; Jacobsen & Kaufman 1999; Robb et al. 2004; Halver-
followed laterally in the field as well as traced (.13 km) in son et al. 2005, supplementary information). Values more negative
satellite imagery. than –11‰ may be below the modal range of Cryogenian samples
(Kaufman et al. 1993), and could be particularly altered. The Mai
Kenetal Limestone and Matheos Formation Black Limestone
Upper prospective glaciogenic unit (units considered best preserved for d13Ccarb and 87Sr/86Sr, see
discussion) have median compositions between –7.6 and
Matheos Formation Diamictite Member, Negash Synclinorium. The –3.5‰, with Matheos Black Limestones underlying the transition
Matheos Formation Diamictite Member is mapped only within to diamictite deposition most enriched.
the folded core of the Negash Synclinorium. Within the Matheos Carbon-isotopic compositions of organic matter (d13CTOC) in
Formation, it overlies a distinctive but gradational transition the Tambien Group, principally from the Mai Kenetal and
from black limestone (Black Limestone Member) to non- Negash synclinoria, are mainly in the range of –20 to –30‰,
calcareous slate (Transitional Member). Although a direct with extremely light values (e.g. , –40‰) suggesting contri-
contact has not been mapped, the oldest overlying sediments are butions from methanogenic biomass (Fig. 21.3). Both Mai
Ordovician Enticho Sandstone, the base of which marks the AAP. Kenetal and Negash successions show similar stratigraphic enrich-
ment trends in d13CTOC with pre-diamictite values in uppermost
Dugub Formation, Shiraro Area. The Dugub Formation in the Matheos Formation limestones about 3‰ heavier than uppermost
western Shiraro Area (Fig. 21.1) occupies a grossly similar Mai Kenetal Limestone exposures.
stratigraphic position within the Tambien Group that could be The Tambien Group has a large compilation of 87Sr/86Sr com-
equivalent to the Matheos Diamictite, but boundary relations for positions in studies by Miller et al. (2003, n ¼ 9; 2009, n ¼ 71)
this low-lying unit are poorly known. and Alene et al. (2006, n ¼ 5). Results obtained from the same
units are complementary among the studies despite contrasts
in sample processing approaches. Least-altered 87Sr/86Sr compo-
Chemostratigraphy sitions (normalized to SRM 987 ¼ 0.710240) are mainly
between 0.7055 and 0.7068 (Fig. 21.3). The Mai Kenetal succes-
Chemostratigraphic data for Tambien Group exposures in Ethiopia sion shows a stratigraphic enrichment trend that compares with a
(Alene et al. 1999, 2006; Miller et al. 2003; Sifeta et al. 2005) and more extensive enrichment trend in the Negash succession.
likely equivalents in Eritrea (Beyth et al. 2003) have become The lower negative d13Ccarb interval at Negash has 87Sr/86Sr
available in only the last decade (Fig. 21.3). For the most part, che- compositions (avg: 0.706178 + 40, n ¼ 2) within error of those
mostratigraphic surveys have been limited in terms of types of ana- for Assem Limestone (avg: 0.706175 + 14, n ¼ 14)
lyses performed and stratigraphic sampling frequency within a (Miller et al. 2009). Negative d13Ccarb compositions in the
given lithological unit, with fewer than 30 sample intervals Assem Limestone may initiate with 87Sr/86Sr compositions near
considered in any one study (Fig. 21.3b). Miller et al. (2009) sub- 0.70597 (Miller et al. 2009).
stantially expanded the regional chemostratigraphic database, Tambien Group carbonate units have Sr compositions ranging
assessing 100 carbonate and 30 slate intervals from the Shiraro, up to about 4000 ppm, with pronounced stratigraphic enrichment
Samre, Mai Kenetal and Negash regions. The integrated Tambien evident in both the Mai Kenetal and Negash synclinoria. Sr com-
Group record now constitutes a significant chemostratigraphic data positions less than 1000 ppm typify limestone and dolostone in
set for evaluating Cryogenian marine secular variations (including the Assem Limestone and Didikama Formation (also Didikama
71 stratigraphic intervals with 87Sr/86Sr measurements on Formation in Shiraro and Samre areas), whereas much higher con-
least-altered units), with the Mai Kenetal and Negash synclinoria centrations averaging 2000– 3000 ppm occur in the Mai Kenetal
(a)

THE TAMBIEN GROUP, NORTHERN ETHIOPIA


(b)

Fig. 21.3. Summary of chemostratigraphic data sets for the Tambien Group. (a) Statistical summaries of chemostratigraphic (d13Ccarb, d18Ocarb, d13CTOC, 87Sr/86Sr, Sr concentration and chemical weathering index) data
for Tambien Group lithostratigraphic units by locality and in stratigraphic order. Letters A –D mark stratigraphic units of alternating d13Ccarb polarity; corresponding circled data sets indicate possible correlations between the
Mai Kenetal and Negash synclinoria (black vertical bars) and other Tambien Group localities (see discussion); line types of circled data sets clarify vertical correlations. Black triangles indicate prospective glaciogenic intervals.

269
Data compiled from Miller et al. (2009) and Alene et al. (2006, vein calcite samples excluded) are restricted to samples considered least altered for 87Sr/86Sr. Chemical Index of Weathering (CIA or PIA) data exclude
samples containing lithic fragments or carbonate minerals. (b) Distribution of chemostratigraphic analyses for each study used to compile the upper graph statistical summaries. The histogram (top-to-bottom) bar order
corresponds to analysis type, as listed from left to right under the ‘Total Analyses/Least Altered Analyses’ compilation.
270 N. R. MILLER ET AL.

Limestone and Matheos Formation Black Limestone Member. derivations from comparable juvenile crustal sources. Associated
Intermediate Sr compositions (avg: 1198 + 91 ppm, n ¼ 2, chemical weathering indices (Fig. 21.3), which proxy the degree
Miller et al. 2009) occur in the Tsedia Slate in the Mai Kenetal Syn- of weathering from fresh rock (50) to full clay conversion (100),
clinorium. Similar transitional compositions are missing in the are generally between 70 and 90 for the Chemical Index of Altera-
Negash Synclinorium and instead there is a sudden stratigraphic tion (CIA, Nesbitt & Young 1982) and .75 for the Plagioclase
jump in Sr concentrations between the Didikama Formation dolo- Index of Alteration (PIA, Fedo et al. 1995).
mite and directly overlying Matheos Formation Black Limestone.
A range of Sr concentrations have been measured for matrix car-
bonate (980 + 712 ppm, n ¼ 11; Miller et al. 2003) and carbonate Palaeolatitude and palaeogeography
clasts in the Matheos Formation Diamictite Member (512 ppm,
oolite cobble; Miller et al. 2009). There are no reliable palaeomagnetic constraints for Tambien
The petrological and chemical transition from largely metavol- Group deposition within the southern ANS. Strike –slip displace-
canic and metasiliciclastic (slate) units in the Tsaliet Group and ment and structural shortening associated with closure of the
lower Tambien Group (Werri Slate) to predominant carbonate Mozambique Ocean and formation of the EAO constitute
deposition in the higher Tambien Group (Assem Limestone and significant challenges for early Cryogenian palaeogeographic
Didikama Formation) was investigated by Sifeta et al. (2005) in reconstructions within the ANS.
the Werii area east of the Mai Kenetal Synclinorium (Fig. 21.1). Palaeogeographic reconstructions spanning the 775–660 Ma
Metavolcanic rocks are sub-alkaline, with chemical fingerprints interval (Fig. 21.4a–c) generally place the ANS (as inferred
that are compatible with island arc (primitive or evolved) and/ oceanic arcs) within the greater Mozambique Ocean between sub-
or MORB settings. Lower Werri Slate compositions indicate sequently flanking Gondwana fragments (e.g. between India and

(a) (b) (c)

Au
Au -
Continents with In
c. 750 Ma
Pac
Ka Pa

-M
palaeomagnetic ific In Ka cif

Ma
e ic
Au data iqu
mb

a
In SC S Mo
za
M
EA Az

r
e

r
mbiqu

aste

aste
Moza
L Az

m
Ada

Ada
EAO
Mozambique Co SF RP RP

La
Ocean Co
Adola

La
Ka o SF o
n
Co RP ilian Sah ilia
SF
Am Br
az Br
az
Brasiliano
WA
Am

Ocean

Am
Postulated Schematic
WA subduction WA continental
zone collision

800 Ma 750 Ma 630 Ma


After Meert 2003; Meert & Torsvik 2003 Collins & Pisarevsky 2005
(d)
West Gondwana Tambien Group ( )
Mz-Cz Orogen
?
Pz-Mz Orogen c. 550 Ma
+
+
+

Pz Orogen Saharan
+

AAP
+

Neoproterozoic Orogen West Meta-


+
+

ANS
African craton
+
+

Cration
AAP
+

East Gondwana
+

AO
+

+
+
+

Congo
+
+
+

Indian
tic ~ E

SF
+

Craton
+

+
+
+

Mad Shield
+

Amazonian Australia
+
+

+
+

Craton
+

SL
~ Antarc
+

+ +
+
+

W.
+
+

+
+

+
+

Kalahari East
+

RP Craton N
+

Antarctic
+

S
+

Shield
+
+

+
+

Subduction Zone
Southern limit of Early Palaeozoic sandstone
marking African-Arabian peneplain (AAP)
1000 km
Modified: Meert & Lieberman 2008

Fig. 21.4. Global palaeogeographic reconstructions for (a) 800 Ma, (b) 750 Ma, (c) 630 Ma and (d) 550 Ma, showing the inferred palaeogeographic location and
structural-tectonic setting of the ANS/Tambien Group (black star) during the Cryogenian Period. Prospective glaciogenic intervals in the Tambien Group were likely
deposited during the earlier Cryogenian (.630 Ma) prior to the emergence of the East African Orogen (EAO). (a, b) Rifting and break-up of Rodinia (c. 900–750 Ma)
was associated with sea-floor spreading, arc and back-arc basin formation, and terrane accretion in the Mozambique Ocean. (c) Accommodation space in the southern
ANS basin likely inverted by 630 Ma in response to closure of the Mozambique Ocean between converging elements of West and East Gondwana and emergence of the
EAO. Occurrence of undeformed post-orogenic intrusives of this age or older that puncture the deformed Neoproterozoic supracrustal sequence in Ethiopia and Eritrea,
suggests that the Tambien Group at 630 Ma was likely deformed and uplifted within the EAO. (d) Location of the Tambien Group at c. 550 Ma within the context of
Gondwana amalgamation and emergence of the Antarctic-EAO. The modern southern limit of Early Palaeozoic sandstone (superimposed from Avigad et al. 2005)
documents the minimal extent of the Afro-Arabian Peneplain (AAP) and extent of regional uplift and Cambro-Ordovician erosion associated with the emergent EAO
(Avigad et al. 2005). Reconstructions modified from (a) Meert 2003; Meert & Torskvik 2003; (b, c) Collins & Pisarevsky 2005; (d) Meert & Lieberman 2008; Grey et al.
2008. Abbreviations for cratons and continents: Am, Amazonia; AuMa, Australia/Mawson Block; Az, Azania; B, Baltica; Co, Congo/Tanzania/Bangweulu Block; EA,
East Antarctica; Ka, Kalahari Block; La/L, Laurentia; In, India; Mad, Madigascar; RP, Rio de la Plata; Sah, Saharan Metacraton; SC, South China; SF, Sao Francisco;
S, Siberia; WA, West Africa; Adola, Adamastor, Braziliano, Mozambique, Pacific, oceanic basins; Mz, Mesozoic, Cz, Cenozoic; Pz, Palaeozoic.
THE TAMBIEN GROUP, NORTHERN ETHIOPIA 271

500 PHZ Enticho Ss (Ord)


Ediacaran C detrital zircon
550 2 2
Post-orogenic plutons Fig. 21.5. Radiometric age constraints
age distribution 2 Mai
Mareb R. Kenetal bearing on the age of the Tambien Group
Eritrea Sibta Shire 6 7
600 1 1 6 within the magmatic evolution of the
8
9 Hawzien Negash southern ANS and EAO. Tsaliet arc
650 E. Sudan Youngest detrital zircons in upper magmatic products and ‘Mareb’
Age (Ma)

NEOPROTEROZOIC

prospective glacial intervals Matheos Tambien post-orogenic pluton ages arranged


700
Cryogenian

Dugub Fm Diamictite Group according to their relative west-to-east


(Shiraro) Zircons below 1st dolomite beds (Negash) Carbonate positions in Tigre. Magmatic equivalents
750 1 2 2 1 1 Deposition for localities in Eritrea and Sudan (open
1
800 Azeho Deset 2 Madahne BSS squares) are shown to the left. The Enticho
5 4 Eritrea
2
Hawzien Alem Sandstone detrital zircon age spectrum
9 3 Chila (Negash)
850 10 Rama (Avigad et al. 2007) shown at the left of the
Ton.

E. Sudan Tsaliet Group syn-tectonic granites and metavolcanics figure derives from the Enticho area as
900 shown in Figure 21.1. The shaded
Relative W-to-E position (Tigre) horizontal bar shows the age range (827+6
to 770 + 7 Ma) for the Bitter Springs stage
STUDY DATE TYPE/METHOD
(BSS) negative d13Ccarb excursion of
Australia (Halverson et al. 2005, 2007; and
1. Avigad et al. (2007) 6. Miller et al. (2003) Pb-Pb zircon U-Pb zircon
2. Tadesse et al. (2000) 7. Asrat et al. (2004)
references therein). Note that this range
Rb-Sr Zircon evaporation
3. Tadesse et al. (1999) 8. Teklay et al. (2001) overlaps with the main phase of Tsaliet
Sm-Nd Youngest detrital zircon
4. Teklay (1997) 9. Kröner et al. (1991) syn-tectonic arc magmatism; see text for
CHIME zircon (SHRIMP U-Pb)
5. Teklay et al. (2002) 10. Teklay et al. (2003) discussion.

Congo cratons). For example, Collins & Pisarevsky (2005) place Geochronological constraints
ophiolite-bearing strata of the Adola Belt (southern Ethiopia) out-
board of the Congo Craton (Fig. 21.4b). As the Adola Belt is gener- The age of Tambien Group deposition is bracketed by magmatism
ally considered to mark the southern extent of the EAO, it is reasonable associated with formation of the underlying Nakfa basement
to infer a similar outboard location for the Tambien Group. complex (Tsaliet Group) and orogenic thickening and differen-
The earliest reliable regional palaeomagnetic data derive from tiation related to maturation of the EAO (‘Mareb’ granitoids).
the time of Gondwana amalgamation (Fig. 21.4d); late Cryogenian Affiliated magmatism is widely dated throughout the ANS (e.g.
(593 + 15 Ma) Dokhan volcanics of Egypt (Davies et al. 1980; Meert 2003; Johnson & Kattan 2007, and references therein).
Wilde & Youssef 2000), interpreted to have a subtropical palaeo- Our geochronological review is restricted to southern ANS
latitude (P-lat: 20.6 + 5.08, A95: 10.08, Trindade & Macouin studies in eastern Sudan (Kröner et al. 1991), Eritrea (Teklay
2007; but see Nairn et al. 1987). The Dokhan palaeopole has 1997; Teklay et al. 2001, 2003; Andersson et al. 2006), northern
been widely used in subsequent palaeogeographic reconstructions Ethiopia (Tadesse et al. 1997, 2000; Miller et al. 2003; Asrat
(e.g. Meert 2003; Meert & Torsvik 2003; Macouin et al. 2004; et al. 2004; Avigad et al. 2007) and western Ethiopia (Ayalew
Trindade & Macouin 2007). et al. 1990) closest to Tambien Group exposures and its prospec-
Low subtropical palaeolatitudes (9 –138) were also reported for tive equivalents (Figs 21.1, 21.5, Supplementary Table 21.1).
Huqf Supergroup units in Oman thought to represent deposition Detrital zircon age spectra from mature Ordovician sand (Enticho
during and after an upper Cryogenian (Fiq Formation) glaciation Sandstone) capping the AAP in Tigre are interpreted by Avigad et al.
before 544 Ma (Kempf et al. 2000; Kilner et al. 2005), when (2007) to reflect the magmatic history of the regional Neoprotero-
Oman was likely amalgamating with the ANS (detrital zircon zoic–Ordovician basement complex, as it was uplifted and eroded
data in Rieu et al. 2007). Subsequent work (Rieu et al. 2006; within the EAO. The Enticho Sandstone zircon age distribution
Allen 2007, and references therein) demonstrates that Mirbat (Fig. 21.5) reinforces that Neoproterozoic magmatism occurred in
Group localities in both palaeomagnetic studies correspond to an two main episodes: (i) arc-related calc-alkaline magmatism
older Cryogenian (Ghubrah-Ayn Formations) glaciation, con- c. 850–740 Ma (c. 780 Ma peak), largely related to petrogenesis
strained radiometrically to be ,722 Ma and ongoing at of the Tsaliet Group, and (ii) post-orogenic magmatism c. 660–
711.8 + 1.6 Ma. The fact that palaeolatitudes for Oman units 580 Ma (c. 630 Ma peak) after Tambien Group deposition.
associated with both glacial intervals (between c. 722 and Additional geochronological constraints elaborated below and
544 Ma) are similarly low, and also close to published c. 550 Ma shown in Figure 21.5 further suggest that the lower and upper
palaeopoles from Gondwana, has yet to be fully reconciled with prospective glacial intervals within the Tambien Group were depos-
the available geologic data (Allen 2007). However, even if low ited during the lull between these two magmatic episodes, with 775
palaeolatitudes are confirmed, Oman had yet to accrete with the to .660 Ma as a plausible depositional window (see Discussion).
ANS at the time of the older Cryogenian (Ghubrah-Ayn For-
mations) glaciation, and therefore these constraints could not be
precisely extended to the Tambien Group. Pre-Tambien Group magmatism
To the extent that the ANS occupied a gross latitudinal range
similar to the Congo-Sào Francisco and/or East-Sahara cratons Deformed (syn-tectonic) plutons and metavolcanic rocks of the
during the early Cryogenian (c. 750 Ma), as implied by various Tsaliet Group (lower calc-alkaline magmatic components of the
palaeogeographic reconstructions (e.g. Fig. 21.4a,b; Trindade & Nakfa terrane) have been dated by a number of techniques (e.g.
Macouin 2007, fig. 3a), the Tambien Group may have occupied Pb/Pb zircon, Rb –Sr, Sm –Nd, CHIME zircon, U –Pb zircon,
low to intermediate latitudes (,458). Occurrence of evaporite zircon evaporation) with varying associated uncertainties (Sup-
pseudomorphs in lower Didikama Formation dolomite (interpreted plementary Table 21.1 – Geochronological Data). These ages
as ,774.7 + 4.8 Ma, Miller et al. 2009) may support a subtropical provide an indirect lower age limit for Tambien Group carbonate
palaeolatitude, as palaeomagnetically constrained evaporite basins deposition. The oldest rocks in this region include an
as old as 2.25 Ga have statistically significant volume-weighted 862 + 6 Ma granite clast within the Gulgula Group (Fig. 21.1,
concentrations in the palaeo-subtropics (Evans 2006). Teklay et al. 2003), 854 + 3 Ma deformed volcanic rocks and
272 N. R. MILLER ET AL.

811 + 11 Ma granite from Eritrea (Teklay 1997). In the Axum continuity of the slate-to-carbonate transition, leading to negative
area (western Tigre), syn-tectonic pluton ages range between d13Ccarb compositions above the lower prospective glacial interval
806 + 21 Ma and 756 + 33 Ma (Tadesse et al. 2000). A (GCI) in the eastern limb and positive d13Ccarb compositions
784 + 14 Ma syn-tectonic granitoid occurs SSW of Hauzien above variegated slate in the western limb (see Miller et al. 2009
(Fig. 21.1; Avigad et al. 2007). A conformable felsic interval for details). The available regional data suggest that Tambien
shortly (c. 17 m) below initial dolomite beds of the Didikama For- Group carbonate deposition began with positive d13Ccarb compo-
mation, in the western limb of the Negash Synclinorium (Madahne sitions (interval A). For Negash this interval (c. 250 m thick) in
Alem), produced a zircon U – Pb date of 774.7 + 4.8 Ma (Avigad the western limb may crop out within an east-verging thrust
et al. 2007; Figs 21.4 and 21.5). block, whereas the interval in the eastern limb may be faulted
out or yet unrecognized below the GCI. The nature and continuity
of the transition from Tsaliet Group agglomerates to lower Didi-
Post-Tambien Group magmatism kama Formation slate and dolomite remains poorly understood
and a systematic stratigraphic transition from lower positive (inter-
Post-orogenic ‘Mareb’ granitoid plutons (Fig. 21.1) penetrate the val A) to higher negative (interval B) d13Ccarb compositions has yet
older deformed Neoproterozoic complex (including the Tambien to be documented in a conformable sequence. Based on similar
Group) throughout northern Ethiopia and Eritrea. In Tigre, ages lithostratigraphic position and character (slate with negligible car-
for these commonly rounded and undeformed intrusives range bonate content and advanced chemical weathering indices;
from 613.4 + 0.9 Ma (Mai Kenetal Granite, Avigad et al. Fig. 21.3), the informal lower slate member of the Didikama
2007) to 545 + 24 Ma (Mareb Granite, Tadesse 1997). The Formation could be equivalent to the Werii Slate and these units
Negash Pluton, just west of the Negash syncline, is c. 606 Ma are tentatively included in polarity interval A.
(606.0 + 0.9 Ma, Miller et al. 2003; 607 + 7 Ma, Asrat et al. The lower prospective glacial interval in both sequences
2004). A small granitoid body puncturing the western Negash underlies a lower carbonate interval with negative d13Ccarb (interval
limb (lower Didikama Formation) is interpreted to have a compar- B) that is in turn overlain by a distinctive black limestone succession
able age, but did not render suitable zircons for geochronology with the most enriched d13Ccarb compositions of the entire Tambien
(Miller et al. 2009). Ages of post-orogenic intrusives are somewhat Group (interval C; Fig. 21.2b). Polarity interval B and C units in
older in Eritrea (628 + 4 Ma, 622 + 1 Ma; Teklay et al. 2001) and both localities have highly complementary d13CTOC, 87Sr/86Sr and
Sudan (SE of Tokar: 652 + 14 Ma, Kröner et al. 1991). The good Sr compositions that support their regional correlation (Fig. 21.3).
agreement between individually dated syn-tectonic and post- The lithological and chemostratigraphic transition from interval B
orogenic magmatic products with corresponding modal peaks in to C appears to be continuous in Mai Kenetal, whereas this transition
the Enticho Sandstone zircon age distribution (Fig. 21.5), and in Negash is abrupt and likely associated with an unconformity
the lack of detrital zircons younger than c. 739.2 + 6.3 Ma in (Miller et al. 2009). Above the black limestone interval in the
either of the prospective glaciogenic units capping the Tambien Negash Synclinorium is the second ‘upper’ negative d13Ccarb excur-
Group (Dugub Fm, Matheos Diamictite) are strong evidence that sion (interval D) associated with a conformable transition to the
the Tambien Group is older than EAO post-tectonic granitic upper prospective glacial unit (Matheos Diamictite).
rocks (Avigad et al. 2007). In the vicinity of Mai Kenetal and
Negash, this constraint is c. 610 Ma, but its upper age is likely
much older considering that regional deformation and crustal Composite chemostratigraphic reference section
thickening must have preceded the ‘Mareb’ intrusives. The
oldest dated post-tectonic pluton in the region (652 + 14 Ma; Figure 21.6 presents a composite chemostratigraphic reference
Kröner et al. 1991) and Enticho sandstone detrital zircon record section for the Tambien Group based on the correlations between
(initiation of the second magmatic phase at c. 660 Ma; Avigad the Mai Kenetal and Negash synclinoria shown in Figure 21.2b
et al. 2007) suggest that the Tambien Group may have been depos- and the reasoning above. The Mai Kenetal section above the Werii
ited prior to c. 660 Ma (Fig. 21.5). Slate (Assem Limestone, Tsedia Slate and Mai Kenetal Limestone)
essentially replaces the Negash sequence above the basal positive
d13Ccarb interval of the lower Didikama Formation and below
Discussion (or within) the lower portion of the Matheos Formation Black
Limestone Member. This composite effectively replaces dolomite
Regional chemostratigraphic context of the prospective lithologies below a probable unconformity (Negash) with a continu-
glacial intervals ous and conformable succession of limestone lithologies; these seem
to show gradational chemostratigraphic trends into highly enriched
The observation that many Tambien Group exposures in Tigre values that correlate well with the lower Matheos Formation Black
exhibit a similar succession of lithofacies that have comparable che- Limestone Member (Miller et al. 2009). The abrupt transition
mostratigraphic characteristics suggests that the Tambien Group pre- from positive (polarity interval A) to negative (polarity interval
serves a regionally coherent chemostratigraphic framework (Beyth B) d13Ccarb compositions may be a structural contact, but we
1972; Alene et al. 2006; Miller et al. 2009). Tambien Group deposi- suggest that the relative stratigraphic order of d13Ccarb polarity is
tional history, including the depositional context of the prospective maintained in the composite. The 87Sr/86Sr evolution of Cryogen-
glaciogenic intervals, can be effectively considered relative to ian seawater (albeit poorly delineated) involves a general trend of
stratigraphic changes in the polarity of associated d13Ccarb. These increasing 87Sr/86Sr (Halverson et al. 2007). That least altered
relationships are best demonstrated in the Mai Kenetal and Negash samples from polarity interval A dolomite beds have lowest
87
synclinoria (Figs 21.2b, 21.3) but are consistent with the stratigraphic Sr/86Sr compositions is consistent with early Cryogenian depo-
succession in other areas (e.g. Shiraro and Samre). sition within the Tambien Group carbonate platform. The evolution
The carbonate successions in the Mai Kenetal and Negash syn- of Tambien Group depositional environments is now considered
clinoria both initiate above thick marine slate sequences; however, based on the composite reference section (Fig. 21.6).
the nature of this transition to carbonate deposition differs in each
locality. In Mai Kenetal the transition from Werii Slate to the
Assem Limestone is a sharp conformable or paraconformable Palaeoenvironmental changes
contact, the Assem Limestone beginning with and maintaining
negative d13Ccarb compositions. In Negash, faulting in the lower The Tambien Group involves mainly a transition from weathered
portion of each synclinorial limb interrupts the stratigraphic Tsaliet arc detritus (lower slate intervals; Sifeta et al. 2005) to
THE TAMBIEN GROUP, NORTHERN ETHIOPIA 273

d13Ccarb Sr (ppm) TOC


Enticho
Sandstone -8 -4 0 4 8 0 2000 4000 0.0 0.1 1.0
2000

Tr Diam.

matrix
oolite
(d)
Interval devoid of bedded carbonate
Matheos Fm
cobble
1800
Black LS

1600
3
MK
LS

1400
(c)
Stratigraphic Height (m)

Bilato
Tambien Group

1200
Tsedia Slate
Logmiti

1000

800
Assem

(b)
LS

MTS
600
CC ?
2 ?
Slate
>Dol

400
(a)

774.7±4.7 Ma
Lower

200
Slate

Extreme weathering indices (PIA: 92-99)


0
1
Tsaliet Gp 0.705 0.706 0.707 -15 -10 -5 0 -45 -35 -25
(Meta-
volcanics) 87Sr/86Sr d Ocarb
18
d CTOC
13

MTS CC
Sheet crack cements Evaporite pseudomorphs Stromatolites Internal slump Molar tooth structures Probable dropstones Cap carbonate-like textures

Fig. 21.6. Composite chemostratigraphic reference sequence for the Tambien Group based on the Mai Kenetal and Negash synclinoria. Segments 1 –3 denote the
basis for constructing this composite sequence (cf. Fig. 21.2b) and the black triangles denote the prospective glacial intervals. The composite sequence is differentiable by
four changes (a –d, separated by horizontal dashed lines) in the polarity of associated d13Ccarb and consists of limestone units except for segment 1 (dolomite comprising
the lower portion of the Didikama Formation in the western limb of Negash). The lower horizontal shaded interval indicates the stratigraphic range of slates having
extreme chemical weathering indices (PIA, plagioclase index of alteration of Fedo et al. 1995). The upper horizontal shaded interval corresponds to the upper portion of
the Matheos Formation lacking bedded carbonate; the darker grey d13Ccarb field in the Matheos Diamictite shows the range for diamictite matrix carbonate (Miller et al.
2003), whereas the uppermost data point corresponds to an oolite cobble (Miller et al. 2009). Abbreviations: Ls, Limestone; Dol, dolomite; Tr, transitional,
Diam-Matheos diamictite; MK, Mai Kenetal.

predominant carbonate deposition in a marine arc-accretion plat- associated with polarity interval A are interpreted to have been
form setting. The lack of obvious ash beds, volcaniclastic intervals deposited in high alkalinity tidal flat and intertidal settings. Mod-
and syn-tectonic intrusive rocks in the main carbonate succession erate energy shallow intertidal or subtidal settings are interpreted
(polarity interval B and higher) suggests deposition during a mag- for the Assem Limestone and upper Didikama Formation
matically quiet interval (Avigad et al. 2007), and the upward (polarity interval B) on the basis of prominent domal and interdi-
waning of slate deposition in favour of carbonate may represent gitate stromatolites, coarse rip-ups, and cross-bedded grainstones
a phasing out of arc magmatic activity in the region. (Miller et al. 2009, figs S2, S3). The base of the Assem Lime-
Chemical weathering indices of slates in polarity interval A stone occurs abruptly above the Werii Slate as a metre-thick
suggest an overall stratigraphic trend of increasing chemical weath- dolomite bed, with sedimentary features similar to those
ering of source areas to extreme levels, which lessened somewhat described for well-studied transgressive cap-carbonate sequences.
prior to deposition of polarity interval B (Fig. 21.6). The association Lower carbonate units of the Tambien Group (polarity intervals
of extreme weathering indices with increasing 87Sr/86Sr is consist- A –B) have low TOC contents with highly variable d13CTOC,
ent with an interval of rapid chemical weathering of arc terranes in including light compositions, which together with the sedimen-
the southern ANS portion of the Mozambique Ocean. The fine tary characteristics are consistent with nearshore environments
grained nature of slates and absence of shallow water indicators (Miller et al. 2009, figs 7B,C and 9).
suggests a moderately deep-water depositional setting. Deeper subtidal environments with lower (but still variable)
Above the lower slate intervals, the Tambien Group carbonate energy levels and ubiquitous micrite are interpreted for upper
pile is broadly differentiable between lower carbonate units carbonate units of the Tambien Group (polarity interval C).
(polarity intervals A –B) suggestive of shallow marine deposi- The transition appears to begin at the base of the Tsedia Slate in
tional settings and upper carbonate units (polarity interval C) sug- Mai Kenetal (equivalent to the proposed unconformity between
gestive of deeper marine environments. Based on the occurrence the Didikama and Matheos Formations in Negash), and this bound-
of evaporite pseudomorphs, sheetcrack cements, microbialami- ary could mark the base of a transgressive system following
nates and stromatolites (Miller et al. 2009, fig. S5), carbonates regression. Both the Mai Kenetal Limestone and the Matheos
274 N. R. MILLER ET AL.

Formation Black Limestone are distinctive dark grey to black lime- 827 Ma), are substantially older than ages determined by Rb –Sr
stone units characterized by fine horizontal layering and high lateral whole rock methods (c. 770– 670 Ma). The younger ages have
continuity (Miller et al. 2009, figs S3 and S6). The lack of stroma- been interpreted to reflect isotopic disturbance due to regional
tolites and high-energy sedimentary structures (coarsely graded deformation of the arc complex, possibly as it accreted with the
beds, cross-bedding) but occurrence of dark intraclastic intervals African continent (Kröner et al. 1991; Saharan Metacraton?).
and intraformational slumping may indicate slope deposition Structural deformation of the orogen may thus have begun
below storm wave base. These upper carbonate units are relatively before 670 Ma. As this deformation was post-depositional and pre-
enriched in TOC that has less variable d13CTOC, consistent with ceded the first post-orogenic intrusives (652 + 14 Ma) by tens of
more open marine environments (Miller et al. 2009, fig. 7B,C). millions of years, the Tambien Group minimum age, and the age
Upper carbonate units are also most enriched in d13Ccarb and of the upper prospective glacial interval, could be appreciably
d13CTOC, particularly in the upper part of the Matheos Formation older than 652 + 14 Ma.
Black Limestone Member. This mutual stratigraphic enrichment
pattern, together with distinct stratigraphic increases in 87Sr/86Sr
and Sr concentrations, is consistent with high rates of organic Evidence for glacial influence on sedimentation
matter burial and related fractionation of the upper (photic)
marine d13CDIC pool in the lead up to Matheos Diamictite depo- The lower prospective glacial interval of the Tambien Group,
sition. Although d18Ocarb is the marine proxy most likely to as the GCI in Negash and a possibly equivalent interval or paracon-
undergo post-depositional alteration, it is notable that the intervals formity at the top of the Werri Slate is most controversial.
with most enriched compositions (consistent with cryogenic Direct sedimentologic evidence of syn-glacial deposition (i.e.
sequestering of d16O) occur in association with the two prospective definitive dropstones, striated clasts) is so far lacking, with the
glacial intervals (Fig. 21.6). The possibility that d18O records fluc- possible exception of matrix supported horizons in the GCI
tuations in ice volume is more likely for the upper prospective (Beyth 1972). On the other hand, both of these intervals are over-
glacial interval because the lower prospective glacial interval is lain by negative d13Ccarb carbonates with similar chemostrati-
underlain by dolomite, which is typically enriched by 2 –4‰ graphic characteristics. Where the base of this carbonate interval
over coeval limestone (Jaffrés et al. 2007, and references therein). is well exposed (Assem Limestone, Mai Kenetal) it exhibits fea-
tures consistent with well-documented Cryogenian post-glacial
cap carbonates.
Timing of the prospective glacial intervals Alene et al. (2006) discounted a glacial association for the
Assem Limestone, but did not document or sample the base of
Integration of available geochronological information suggest c. this unit above its abrupt contact with Werri Slate. They suggested
775–660 Ma as a plausible window for Tambien Group deposition the Assem Limestone may correspond to a non-glacial negative
bracketing the two prospective glacial intervals. This depositional d13Ccarb interval possibly equivalent to the Bitter Springs stage
window is not definitive, but corresponds to the lull between (BSS) of Australia (as suggested by Halverson et al. (2005,
Tsaliet arc magmatism and later magmatism associated with matu- 2007) for similar negative d13Ccarb intervals within early Cryogen-
ration of the EAO, on the basis of regional stratigraphy and cross- ian sections from NW Canada (Little Dal Group) and Svalbard
cutting relationships. Direct dates on magmatic products from each (Akademikerbreen Group)). The age for the BSS in Australia
magmatic episode vary locally within Tigre but concur with the requires interbasinal correlations that are difficult to confirm, and
detrital zircon age distribution from the immediately overlying the age range has a corresponding large degree of uncertainty
Ordovician Enticho Sandstone (Fig. 21.5). Somewhat older arc between 827 + 6 Ma (Gairdner Dyke Swarm) and 777 + 7 Ma
magmatism is indicated in Eritrea and eastern Sudan. (Boucat Volcanics) among various studies (Halverson et al.
The 774.7 + 4.8 Ma date obtained from zircons recovered from 2005, 2007). Although a correlation with the BSS is possible, the
variegated (tuffaceous?) slate, c. 17 m below the first positive c. 827 –770 Ma age range would correspond to a time of active
d13Ccarb (polarity interval A) dolomite beds of the Didikama For- arc magmatism in the southern ANS (Fig. 21.5). The Assem Lime-
mation, is a possible maximum age constraint for significant stone and higher Tambien Group succession lacks evidence of
(bedded) carbonate deposition in the Tambien Group (Fig. 21.5). active magmatism (ash beds/slate intervals/syn-tectonic intru-
The interval was originally interpreted as either a volcaniclastic sives) suggesting it was deposited after the c. 780 Ma acme of
interval or sill injected into a volcaniclastic interval. Subsequent arc magmatism in the region. The lower Negash negative
petrographic analysis of this interval revealed probable sedimen- d13Ccarb interval occurs several hundred metres above the varie-
tary textures, including rounded quartz grains and a large (3 cm) gated slate interval dated at 774.7 + 4.8 Ma (zircon). This
rounded quartzose clast. If this date holds, the lower prospective suggests that the lower negative d13Ccarb interval could be appreci-
glacial interval occurring at least c. 250 m higher is likely to be ably younger than c. 775 Ma. As there are definitively younger
appreciably younger. The upper prospective glacial interval is post-glacial cap carbonates (e.g. above the c. 755 Ma Kaigas For-
older than deformation and later magmatism associated with the mation, southern Namibia, Gariep Belt, Hoffmann et al. 2006)
EAO. The oldest indication of EAO magmatic activity in the we suggest that a lower prospective glacial association within
southern ANS is from a 652 + 14 Ma post-orogenic pluton in the Tambien Group remains a viable interpretation. It is conceiva-
eastern Sudan (southern Red Sea Hills, c. 500 km NNW of ble that the glaciogenic association could be indirect. For example,
Negash; Kröner et al. 1991), whereas locally dated ‘Mareb’ gran- the cap-carbonate features and negative d13Ccarb excursion could
itoids in Tigre are c. 610 Ma (avg: 609.8 + 3.8 Ma, Mai Kenetal, be products of enhanced upwelling associated with global
Negash, Hauzien plutons). cooling and/or vigorous circulation associated with post-glacial
The dated Red Sea Hills locality in eastern Sudan occurs climate change in a setting that did not earlier accumulate glacially
c. 50 km NW of the area of prospective glacial deposits (pebbly transported clastic sediments. Both of these intervals in the Mai
mudstones with flattened clasts) in northern Eritrea (Fig. 21.1) Kenetal and Negash synclinoria warrant additional study to
(Cecioni 1981). If the pebbly mudstones are glaciogenic and evaluate glacial associations.
correlative with the upper prospective glacial interval of the Of the proposed upper glaciogenic intervals for the Tambien
Tambien Group, the Red Sea Hills age offers an important Group, the Matheos Formation Diamictite Member best exhibits
minimum age constraint for the Tambien Group. The most sedimentological characteristics consistent with glacially influ-
precise age estimates on older (pre-pebbly mudstone) associated enced deposition. In addition to matrix supported polymict
arc metavolcanic and plutonic rocks in the Red Sea Hills area, clasts, clasts that appear to preferentially deform underlying
obtained by single zircon 207Pb/206Pb methods (c. 870– bedding are interpreted as ice-rafted debris. Some bullet-nosed
THE TAMBIEN GROUP, NORTHERN ETHIOPIA 275

and possibly striated clasts are consistent with mechanical abrasion Pan-African orogens: implications for global environment. Earth
during glacial entrainment (Kuhn et al. 1993). This interval and Planetary Science Letters, 240, 818–826.
is further significant because of its gradational contact with the Avigad, D., Stern, R. J., Beyth, M., Miller, N. & McWilliams, M.
underlying Matheos Black Limestone Member, which records a 2007. Detrital zircon U– Pb geochronology of Cryogenian diamictites
decline in d13Ccarb and possibly also 87Sr/86Sr, prior to diamictite and Lower Palaeozoic sandstone in Ethiopia (Tigrai): age constraints
deposition. If the latter is not the product of diagenetic alteration, on Neoproterozoic glaciation and crustal evolution of the southern
the declining Sr-isotope compositions could signal a strong Arabian– Nubian Shield. Precambrian Research, 154, 88– 106.
decrease in continental weathering input relative to oceanic Ayalew, T., Bell, K., Moore, J. M. & Parrish, R. R. 1990. U-Pb and
Rb-Sr geochronology of the Western Ethiopian Shields. Geological
hydrothermal input, as consistent with increasing ice cover.
Society of America Bulletin, 102, 1309– 1316.
The possibly equivalent Dugub Formation in the Shiraro area Beyth, M. 1972. The geology of central western Tigre, Ethiopia. PhD
occupies a gross stratigraphic position similar to the Matheos thesis, Bonn, University of Bonn.
Diamictite as well as similar detrital zircon age distributions Beyth, M., Avigad, D., Wetzel, H. U., Matthews, A. & Berhe, S. M.
(Avigad et al. 2007), but the underlying stratigraphy is poorly 2003. Crustal exhumation and indications for snowball Earth in the
known. Palaeocurrent proxies suggest a north-northeastern East African Orogen: North Ethiopia and East Eritrea. Precambrian
source area for clasts, which considering the prospective glacial Research, 123, 187–201.
intervals in northern and eastern Eritrea could be consistent with Bibolini, A. 1920. Risultali preliminary delle osservazioni faite nel
a regional phase of glaciation. Nord-est della Colona Eritrea. Asmara.
Bibolini, A. 1921. Sui conglomerati di Rore Babla e dei Monti Haggar
in Colonia Eritrea. Bollettino della Società Geologica Italiana, 40,
Tectonic and palaeogeographic setting 169– 176.
Bibolini, A. 1922. Contributions a l’étude de la géologi de l’Afrique
The intra-oceanic setting of the Tambien Group contrasts distinctly orientale Italienne. 13th International Geological Congress (1922,
from pericratonic rift margin settings that are often associated with Brussels, Belgium), Title Comptes rendus de la XIIIe session, en
Cryogenian glaciogenic units. Later Cryogenian intervals are unli- Belgique, parts 1 –3, 797– 814.
kely to be directly recorded throughout much of the ANS because Burke, K. & Sengör, C. 1986. Tectonic escape in the evolution of
the emergence of the EAO (by 660 Ma) would likely have either continental crust. In: Reflection seismology – The continental crust
American Geophysical Union, Geodynamic series, 14, 41 –53.
destroyed accommodation space or uplifted related strata to
Cecioni, G. 1981. Precambrian pebbly mudstones in Eritrea, northeastern
levels planed by subsequent erosion. The EAO could have hosted Ethiopia. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-
alpine/continental glaciation, however, perhaps even instigating Pleistocene Glacial Record. Cambridge University Press, A24, 150.
initial downcutting of the vast AAP (Fig. 21.4d; Stern et al. 2006). Collins, A. S. & Pisarevsky, S. A. 2005. Amalgamating eastern Gond-
wana: the evolution of the Circum-Indian Orogens. Earth-Science
The US– Israel Binational Science Foundation (grant no. 2002337) supported this Reviews, 71, 229–270.
study. We thank K. Mehari, D. Küster and T. Tadesse for field expertise and Corsetti, F. A., Stewart, J. H. & Hagadorn, J. W. 2007. Neoprotero-
assistance, A. Abraham (Chief Geologist, Geological Survey of Ethiopia) and zoic diamictite-cap carbonate succession and d13C chemostratigraphy
Mekele University for logistical support during field seasons, and M. Abdel- from eastern Sonora, Mexico. Chemical Geology, 237, 129–142.
Salem for helpful feedback on geological aspects of the ANS. B. Schilman and Davies, J., Nairn, A. E. M. & Ressetar, R. 1980. The palaeomagnetism
A. Ayalon of the Geological Survey of Israel provided stable isotope expertise. of certain late Precambrian and early Palaeozoic rocks from the Red
The manuscript was improved by thoughtful reviews from P. Johnson, M. Pope Sea Hills, eastern desert, Egypt. Journal of Geophysical Research,
and E. Arnaud. This represents a contribution of the IUGS- and UNESCO- 85, 3699–3710.
funded IGCP (International Geoscience Programme) Project #512. De Souza Filho, C. R. & Drury, S. A. 1998. A Neoproterozoic supra-
subduction terrane in northern Eritrea, NE Africa. Journal of the
Geological Society, London, 155, 551– 566.
References Evans, D. 2006. Proterozoic low orbital obliquity and axial-dipolar geo-
magnetic field from evaporite palaeolatitudes. Nature, 444, 51 – 55.
Alene, M., Conti, A., Sacchi, R. & Zuppi, G. 1999. Stable isotope com- Fedo, C. M., Nesbitt, H. W. & Young, G. M. 1995. Unraveling the
position (13C and 18O) of Neoproterozoic limestones and dolomites effects of potassium metasomatism in sedimentary rocks and palaeo-
from Tigre, North Ethiopia. Bollettino della Società Geologica sols, with implication for palaeoweathering conditions and prove-
Italiana, 118, 611– 615. nance. Geology, 23, 921– 924.
Alene, M., Jenkin, G. R. T., Leng, M. J. & Darbyshire, D. P. 2006. The Garland, C. R. 1980. Geology of the Adigrat area. Ministry of Mines
Tambien Group, Ethiopa: an early Cryogenian (ca. 800– 735 Ma) Memoir No. 1, 51, Addis Ababa, Ethiopia. 1:250,000 map.
Neoproterozoic sequence in the Arabian– Nubian Shield. Precam- Gray, D. R., Foster, D. A., Meert, J. G., Goscombe, B. D., Armstrong,
brian Research, 149, 79 –89. R., Truow, R. A. J. & Passchier, C. W. 2008. A Damaran Perspec-
Allen, P. A. 2007. The Huqf Supergroup of Oman: basin development tive on the Assembly of Southwestern Gondwana. Geological Society,
and context for Neoproterozoic glaciation. Earth Science Reviews, London, Special Publications, 294, 257–278.
84, 139– 185. Greenwood, W. R., Hadley, D. G., Anderson, R. E., Fleck, R. J. &
Andersson, U. B., Ghebreab, W. & Teklay, W. 2006. Crustal evolution Schmidt, D. L. 1976. Late Proterozoic cratonization in southwestern
and metamorphism in east-central Eritrea, south-east Arabian-Nubian Saudi Arabia. Philosophical Transactions of the Royal Society of
Shield. Journal of African Earth Sciences, 44, 45– 65. London, 280, 517– 527.
Arkin, Y., Beyth, M., Dow, D. B., Levitte, D., Haile, T. & Hailu, T. Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
1971. Geological map of Mekele Sheet area ND 37-11 Tigre Pro- Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
vince. Imperial Ethiopian Government, Ministry of Mines, Geologi- isotope record. Geological Society of America Bulletin, 117,
cal Survey of Ethiopia, scale 1:250,000. 1181–1207.
Asrat, A., Barbey, P., Ludden, J. N., Reisberg, L., Gleizes, G. & Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A. 2007.
Ayalew, D. 2004. Petrology and isotope geochemistry of the Evolution of the 87Sr/86Sr composition of Neoproterozoic seawater.
Pan-African Negash Pluton, northern Ethiopia: mafic-felsic magma Palaeogeography, Palaeoclimatology, Palaeoecology, 256,
interactions during the construction of shallow-level calc-alkaline 103– 129.
plutons. Journal of Petrology, 45, 1147–1179 Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
Avigad, D., Sandler, A., Kolodner, K., Stern, R. J., McWilliams, Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran
M. O., Miller, N. & Beyth, M. 2005. Mass-production of Cambrian (635 Ma) post-glacial ‘cap dolostones’ diachronous? Earth and
quartz-rich sandstone as a consequence of chemical weathering of Planetary Science Letters, 258, 114– 131.
276 N. R. MILLER ET AL.

Hoffmann, K.-H., Condon, D. J., Bowring, S. A., Prave, A. R. & of the Mirbat area, southern Oman. Journal of the Geological
Fallick, A. 2006. Lithostratigraphic, carbon (d13C) isotope and Society, London, 164, 997–1009.
U –Pb zircon age constraints on early Neoproterozoic (ca. 755 Ma) Robb, L. J., Knoll, A. H., Plumb, K. A., Shields, G. A., Strauss, H. &
glaciation in the Gariep Belt, southern Namibia, Snowball Earth Veizer, J. 2004. The Precambrian: the Archean and Proterozoic
Conference, 16 – 21 July 2006, Ascona, Switzerland [abstract], 51. Eons. In: Gradstein, F., Ogg, J. & Smith, A. G. (eds) A
Jacobsen, S. B. & Kaufman, A. J. 1999. The Sr, C and O isotopic evol- Geologic Time Scale 2004. Cambridge University Press, Cambridge,
ution of Neoproterozoic seawater. Chemical Geology, 161, 37 –57. 129– 140.
Jaffrés, J. B. D., Shields, G. A. & Wallmann, K. 2007. The oxygen Sifeta, K., Roser, B. P. & Kimura, J. I. 2005. Geochemistry, provenance,
isotope evolution of seawater: a critical review of a long-standing and tectonic setting of Neoproterozoic metavolcanic and metasedi-
controversy and an improved geological water cycle model for the mentary units, Werri area, Northern Ethiopia. Journal of African
past 3.4 billion years. Earth-Science Reviews, 83, 83– 122. Earth Sciences, 41, 212–234.
Johnson, P. R. & Kattan, F. H. 2007. Geochronologic dataset for Stern, R. J. 1994. Arc assembly and continental collision in the Neopro-
Precambrian rocks in the Arabian peninsula. Saudi Geological terozoic East African Orogen: implications for the consolidation of
Survey Open-File Report SGS-OF-2007-3, 21, tables. Gondwanaland. Annual Reviews of Earth and Planetary Sciences,
Kaufman, A. J., Jacobsen, S. B. & Knoll, A. H. 1993. The Vendian 22, 319– 351.
record of C- and Sr-isotopic variations: implications for tectonics Stern, R. J., Kröner, A., Bender, R., Reischmann, T. & Dawoud,
and paleoclimate. Earth and Planetary Science Letters, 120, A. S. 1994. Precambrian basement around Wadi Halfa, Sudan: a
409– 430. new perspective on the evolution of the East Saharan Craton. Geolo-
Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes, gische Rundschau, 83, 564–577.
ice ages, and terminal Proterozoic earth history. Proceedings of the Stern, R. J., Avigad, D., Miller, N. R. & Beyth, M. 2006.
National Academy of Sciences USA, 94, 6600– 6605. Geological Society of Africa Presidential Review: Evidence for
Kempf, O., Kellerhals, P., Lowrie, W. & Matter, A. 2000. Palaeo- the Snowball Earth Hypothesis in the Arabian –Nubian Shield and
magnetic directions in late Precambrian glaciomarine sediments of the East African Orogen. Journal of African Earth Sciences, 44,
the Mirbat Sandstone Formation, Oman. Earth and Planetary 1 –20.
Science Letters, 175, 181– 190. Tadesse, T. 1997. The Geology of Axum Area (ND 37-6). Ethiopian Insti-
Kilner, B., Conall, M. N. & Brasier, M. 2005. Low-latitude glaciation tute of Geological Surveys, Addis Ababa (Memoir No. 9).
in the Neoproterozoic of Oman. Geology, 33, 413–416. Tadesse, T. 1999. Axum sheet geological map. Geological Survey of
Kröner, A., Linnebacher, P., Stern, R. J., Reischmann, T., Manton, Ethiopia, Addis Ababa, Ethiopia. 1:250,000 map.
W. & Hussein, I. M. 1991. Evolution of the Pan-African island arc Tadesse, T., Hoshino, M. & Sawada, Y. 1999. Geochemistry of low-
assemblages in the southern Red Sea Hills, Sudan, and in south- grade metavolcanic rocks from the Pan African of the Axum area,
western Arabia as exemplified by geochemistry and geochronology. northern Ethiopia. Precambrian Research, 99, 101– 124.
Precambrian Research, 53, 99 –118. Tadesse, T., Hoshino, M., Suzuki, K. & Izumi, S. 2000. Sm– Nd, Rb– Sr
Kuhn, G., Melles, M., Ehrmann, W. U., Hambrey, M. J. & Schmiedl, and U– Pb zircon ages of syn and post-tectonic grantoids from the
G. 1993. Character of clasts in glaciomarine sediments as an indicator Axum area of northern Ethiopia. Journal of African Earth Sciences,
of transport and depositional processes, Weddell and Lazarev Seas, 30, 313– 327.
Antarctica. Journal of Sedimentary Petrology, 63, 477– 487. Teklay, M. 1997. Petrology, geochemistry and geochronology of Neopro-
Macouin, M., Besse, J., Ader, M., Gilder, S., Yang, Z., Sun, Z. & terozoic magmatic arc rocks from Eritrea: implications for crustal
Agrinier, P. 2004. Combined palaeomagnetic and isotopic data evolution in the southern Nubian Shield. Department of Mines,
from the Doushantuo carbonates, South China: implications for the Asmara, Eritrea, Memoir 1, 125.
‘snowball Earth’ hypothesis. Earth and Planetary Science Letters, Teklay, M. 2006. Neoproterozoic arc –back-arc system analog to
224, 387– 398 modern arc –back-arc systems: evidence from tholeiite – boninite
Meert, J. G. 2003. A synopsis of events related to the assembly of eastern association, serpentinite mudflows and across-arc geochemical
Gondwana. Tectonophysics, 362, 1 –40. trends in Eritrea, southern Arabian –Nubian shield. Precambrian
Meert, J. G. & Lieberman, B. S. 2008. The Neoproterozoic Assembly Research, 15, 81 – 92.
of Gondwana and its relationship to the Ediacaran – Cambrian Teklay, M., Kröner, A. & Metzger, K. 2001. Geochemistry, geochro-
radiation. Gondwana Research, 14, 5– 21. nology and isotope geology of Nakfa intrusive rocks, northern
Meert, J. G. & Torsvik, T. H. 2003. The making and unmaking of a Eritrea: products of a tectonically thickened Neoproterozoic arc
supercontinent: Rodinia revisited. Tectonophysics, 375, 261–288. crust. Journal of African Earth Sciences, 33, 283–301.
Miller, N. R., Alene, M., Sacchi, R., Stern, R., Conti, A., Kröner, A. Teklay, M., Kroner, A. & Mezger, K. 2002. Enrichment from plume
& Zuppi, G. 2003. Significance of the Tambien Group (Tigre, interaction in the generation of Neoproterozoic arc rocks in northern
N. Ethiopia) for snowball Earth events in the Arabian –Nubian Eritrea: implications for crustal accretion in the southern Arabian-
Shield. Precambrian Research, 121, 263– 283. Nubian Shield. Chemical Geology, 184, 167– 184.
Miller, N. R., Stern, R. J., Avigad, D., Beyth, M. & Schilman, B. Teklay, M., Haile, T., Kröner, A., Asmerom, Y. & Watson, J. 2003. A
2009. Neoproterozoic carbonate-slate sequences of the Tambien back-arc palaeotectonic setting for the Augaro Neoproterozoic mag-
Group, N. Ethiopia (I): pre-‘Sturtian’ chemostratigraphy and regional matic rocks of western Eritrea. Gondwana Research, 6, 629–640.
correlation. Precambrian Research, 170, 129– 156. Trindade, R. I. F. & Macouin, M. 2007. Palaeolatitude of glacial depos-
Nairn, A. E. M., Perry, T. A., Ressetar, R. & Rogers, S. 1987. A its and palaeogeography of Neoproterozoic ice ages. Comptes Rendus
palaeomagnetic study of the Dokhan volcanic formation and younger Geoscience, 339, 200–211.
granites, eastern desert of Egypt. Journal of African Earth Sciences, Verri, P. 1909. Contributo allo studio geografio della Colonia Eritrea. Bol-
6, 353–365. lettino della Società Geografica Italiana, 10, 251–320. Carta Geolo-
Nesbitt, H. W. & Young, G. M. 1982. Early Proterozoic climates and gica 1:1,500,000.
plate motions inferred from major element geochemistry of lutites. Wilde, S. A. & Youssef, K. 2000. Significance of SHRIMP dating of the
Nature, 299, 715–717. Imperial porphyry and associated Dokhan volcanics, Gebel Dokhan,
Rieu, R., Allen, P. A., Etienne, J. L., Cozzi, A. & Wiechert, U. 2006. northeastern desert, Egypt. Journal of African Earth Science, 31,
A Neoproterozoic glacially influenced basin margin succession and 403– 413.
‘atypical’ cap carbonate associated with bedrock palaeovalleys, Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003.
Mirbat area, southern Oman. Basin Research, 18, 471– 496. Systematic variations in C, O, and Sr isotopes and elemental
Rieu, R., Allen, P. A., Cozzi, A., Kosler, J. & Bussy, F. 2007. A concentrations in Neoproterozoic carbonates in Namibia: impli-
composite stratigraphy for the Neoproterozoic Huqf Supergroup of cations for a glacial to interglacial transition. Precambrian Research,
Oman: integrating new litho-, chemo- and chronolstratigraphic data 124, 69– 85.
Chapter 22

Evidence for Early and Mid-Cryogenian glaciation in the Northern Arabian –Nubian Shield
(Egypt, Sudan, and western Arabia)

ROBERT J. STERN1*, PETER R. JOHNSON2, KAMAL A. ALI1,3 & SUMIT K. MUKHERJEE1,4


1
Geosciences Department, U Texas at Dallas, Richardson TX 75080-3021, USA
2
Johnson and Vranas Associates, Ltd., Geological Consulting, 6016 SW Haines Street, Portland, Oregon 97219-7046, USA
3
Present address: Faculty of Earth Sciences, King Abdulaziz University, Jeddah 21589, Saudi Arabia
4
Present address: BP America Exploration & Production Company, Houston, TX, USA
*Corresponding author (e-mail: rjstern@utdallas.edu)

Abstract: Evidence of Early- to Mid-Cryogenian (c. 780 Ma and c. 740 Ma) glacial activity is summarized for the northern
Arabian–Nubian Shield (ANS), including structural framework, stratigraphy, lithological descriptions and relationships with
younger and older units, banded iron formation chemostratigraphy, other characteristics, geochronological constraints, and discus-
sion. The ANS is a broad tract of juvenile continental crust, formed from accreted arc-backarc basin terranes developed around
the margins of the Mozambique Ocean. As a result, these successions formed in marine environments at some distance from con-
tinental margins. Deposits include banded iron formation (BIF) and possibly glacial diamictite scattered over broad regions of the
Central Eastern Desert of Egypt, NW Arabia and possible correlative units in NE Sudan. The older (c. 780 Ma) examples (Meritri
group, NE Sudan; basal Mahd group, Arabia) occur in the central ANS, on the southern flank of an important lithospheric boundary,
an ophiolite-decorated suture zone. Mahd group diamictite is thin (1– 5 m thick) and rests above the earliest (Cryogenian) ANS
unconformity. The Meritri group interval near Port Sudan is much thicker and part of a deformed passive margin. Both Mahd
and Meritri group deposits need further study before they are accepted as glaciogenic; confirmation of this interpretation would indi-
cate that Neoproterozoic glacial activity began at least as early as 780 Ma ago. The younger (c. 740 Ma) glacial deposits include
diamictite and BIF: the Atud diamictite and BIFs of the Central Eastern Desert of Egypt and the correlative Nuwaybah diamictite
and BIF of NW Arabia. Northern ANS-BIF is a well-layered chemical sediment of interlaminated hematite-magnetite and jasper. A
glacial origin for the Atud-Nuwaybah diamictites is inferred because large clasts and matrix zircons have ages (Palaeoproterozoic
and Neoarchean) and compositions (especially quartzite, arkose, and microdiamictite) that require transport from outside the ANS
Cryogenian basin. Northern ANS-BIF may also reveal glacial influence, having been deposited in response to reoxygenation of a
suboxic ocean. The 740 Ma diamictite and/or BIF may correlate with Tambien Group diamictites in Ethiopia (Miller et al.
2011). Northern ANS diamictite and BIF were deposited in an oceanic basin of unknown size, as indicated by association with
abundant ophiolites; they are strongly deformed, obscuring many primary features.
There is no strong evidence for or against Ediacaran glaciation in the ANS, largely because the region was uplifted at this time.
The c. 600 Ma ANS peneplain may have been partly cut by Ediacaran glaciation. Some of the post-accretionary basins of Arabia
could preserve glaciogenic deposits of Ediacaran age, but assessing this possibility requires further investigation.

The Arabian –Nubian Shield (ANS) consists of mostly Neoproter- possible glacial deposits from the southern ANS as well as Allen
ozoic outcrops around the Red Sea in NE Africa and West Arabia, et al. (2011a, b) for chapters covering successions in Oman, to
exposed by Oligocene and younger uplift and erosion. The ANS is the east of the ANS. The units summarized here include (i)
one of the largest tracts of juvenile continental crust of Neoproter- Meritri group metaconglomerate (E. Sudan) and basal Mahd
ozoic age on Earth; its evolution accompanied a supercontinent group diamictite (central Arabian Shield); (ii) Atud diamictite
cycle that defined Neoproterozoic tectonics, beginning with the (E. Egypt) and Za’am group Nuwaybah diamictite (NW Arabian
break-up of the end-Mesoproterozoic supercontinent Rodinia in Shield); and (iii) BIF from E. Egypt, NW Saudi Arabia
early Neoproterozoic time (Stern 2008). ANS juvenile crust (Sawawin deposit) and NE Sudan (Fodikwan). Locations of
(intra-oceanic arcs and oceanic plateaux) was generated around these units are shown in Figure 22.1 and co-ordinates are listed
and within the Mozambique Ocean and coalesced as this ocean in Table 22.1. We know less about the Meritri group and basal
closed (Stern 1994). The tectonic cycle culminated in a protracted Mahd group and are less confident that they are glaciogenic than
collision beginning c. 630 Ma, forming the East African Orogen we would like to be, but if confirmed, they would be the oldest
(EAO), an important weld in the end-Neoproterozoic supercontinent known Neoproterozoic glaciogenic deposits. These are discussed
of ‘Greater Gondwana’ (Stern 1994) or ‘Pannotia’ (Dalziel 1997). In here to encourage further detailed studies.
reconstructed Gondwana, the EAO extends from the Mediterranean The deposits of E. Sudan and the central Arabian Shield have not
(Tethys) southward along the eastern margin of Africa and across been the subject of focused sedimentological study. The Meritri
East Antarctica. Sedimentary evidence of early Cryogenian glacial group conglomerate of Sudan was first identified and briefly
episodes is likely to be preserved in ANS sequences, because described by Abdelsalam & Stern (1993). The basal Mahd group
these episodes occurred when ANS crustal components were diamictite was first mentioned by Johnson et al. (2003) and then
mostly below sea level (Stern et al. 2006). In contrast, evidence of discussed in somewhat greater detail by Stern et al. (2006).
late Cryogenian to Ediacaran glacial episodes may be absent The diamictite and banded iron formations to the north in Egypt
because this was a time of collision and uplift in the ANS. Some evi- and NW Arabia (Fig. 22.1) have been studied for some time. The
dence for Ediacaran climate may be preserved in post-accretionary Atud Formation was first described as conglomerate (El-Essawy
basins of Ediacaran age in Arabia (Johnson 2003). 1964). Since its first recognition, the Atud Formation has only
In this chapter we describe possible glaciogenic units from both been recognized in eastern Egypt between 258N and 268N.
flanks of the Red Sea in the northern ANS; the interested reader At the type locality near Gebel Atud, it includes schist, metamud-
should see the chapter by Miller et al. (2011) for an overview of stone, and metagreywacke in addition to diamictite, all of which

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 277– 284. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.22
278 R. J. STERN ET AL.

Cairo Jordan suffered greenschist –facies metamorphism (El-Essawy 1964).


The term ‘conglomerate’ is not appropriate for the poorly sorted,
29°
matrix-supported deposits of the Atud Formation. These better
fit the description of diamictite by Flint et al. (1960) as poorly
Area of Figure 3 sorted, heterolithic, and very coarse terrigenous sediments.
Saudi Arabia
Accordingly, in this chapter we refer to the Atud diamictite
(Stern et al. 2006).
Area of Figure 3

Al Muwaylih
Hurghada
27°
Safajah
Nil
e

Quseir Al Wajh
Structural framework

The Meritri group conglomerate (Sudan) is preserved within the B’ir


25° Marsa Alam Umq-Nakasib Suture Zone, whereas the Mahd group diamictite
Egypt (Saudi Arabia) is preserved just south of this suture zone. This is
Al Madinah
one of the longest and best-defined ophiolite-decorated suture
zones in the ANS (Johnson et al. 2003) and extends (with the Red
Sea closed) ENE–WSW over 600 km from the Nile in Sudan into
23°
Red Sea the central Arabian Shield. The suture zone itself consists of rocks
that originated in a variety of juvenile oceanic environments and
Biír Umq suture Mahd gp. include strongly deformed ophiolite nappes, and metavolcanic,
Sudan F Jiddah
diamictite metasedimentary, and intrusive rocks. Dating of the ophiolites, vol-
canic rocks, and pre- and syntectonic plutons indicates that oceanic
21°
magmatism in the region was active c. 870–830 Ma, whereas sutur-
ing occurred c. 780–760 Ma (Hargrove III et al. 2006). Structural
utu
re Port Sudan complexities (folding, faulting, shearing, etc.) are especially severe
sibs for the Meritri group, which is part of a southwards-directed nappe
N aka
19° stack (Fig. 22.2a), discussed in the following section (Abdelsalam
200 km Meritri gp. & Stern 1993). The Mahd group diamictite is less folded and
diamictite faulted because it lies south of the suture zone.
31° 33° 35° 37° 39° 41° The Atud and Nuwaybah diamictites and associated BIF in
Suture Egypt and NW Saudi Arabia are also folded and faulted. BIF is
Phanerozoic Shear zone
Exposed Neoproterozoic crust
strongly deformed although much of this deformation may have
Diamictite
Ophiolitic complexes Banded Iron Formation occurred as a result of slumping of dense, weak sediments in a tec-
tonically unstable basin. BIF and surrounding sediments are meta-
Fig. 22.1. Locality map of the northern ANS showing the approximate morphosed to greenschist facies. Tectonic deformation began as a
locations of geological units described in this chapter. The dark, dashed square result of collision between arc terranes prior to c. 680 Ma (Ries
designates the area shown in Figure 22.3. Early Cryogenian (c. 780 Ma) et al. 1983). Other deformation resulted from pervasive left-lateral
diamictite sequences (Meritri and Mahd) are confined to the Nakasib and Bir strike-slip shearing along the Najd fault system, which was active
Umq suture areas in the south, whereas Atud and Nuwaybah diamictite and BIF during Ediacaran time (Sultan et al. 1988). Najd deformation was a
sequences are confined to Egypt and NW Saudi Arabia. far-field manifestation of collision between fragments of east and
west Gondwana (Abdelsalam & Stern 1997). Najd faulting
imparted a penetrative shear fabric to most Cryogenian supracrus-
tal units in Egypt and NW Arabia.
Table 22.1. Locations of Neoproterozoic deposits of possible glacial origin in
the northern ANS
Stratigraphy
Nation Location and lithology Age (Ma) Latitude (N) Longitude (E)
Cryogenian supracrustal sequences in the northern ANS are
Egypt dominated by variably deformed immature clastic metasediments
Wadi Abu Marwat BIF c. 750 268310 338380 (greywackes) and metavolcanic rocks; rare sedimentary carbon-
Wadi Kareim BIF c. 750 258560 500 348020 ates also occur. Cryogenian stratigraphic reconstructions are
Wadi Kareim diamictite c. 750 258560 500 348020 complicated because of strong deformation as well as the pres-
Wadi Muweilha diamictite c. 750 258520 348540 3000 ence of several accreted terranes (Johnson & Woldehaimanot
Wadi El Dabbah BIF c. 750 258490 348090 2003). Cryogenian metasedimentary successions are also litho-
Um Gerifat BIF c. 750 258400 348200 logically monotonous. Distinctive units such as sedimentary car-
Um Ghamis BIF c. 750 258370 348190 bonates are uncommon, although these become increasingly
Wadi Sitra BIF c. 750 258310 348140 important in the ANS farther south in Sudan, Eritrea, Ethiopia,
Wadi Mubarak diamictite c. 750 258260 348340 and SW Arabia. As a result, ANS Cryogenian supracrustal
El Hadid BIF c. 750 258210 348080 units often lack a useful formal stratigraphic framework. This
El Imra BIF c. 750 258170 348270 is especially true for Egypt, where Stern (1981) informally
Um Nar BIF c. 750 258160 348160 divided the Cryogenian supracrustal succession of the Central
Gebel Atud diamictite c. 750 258010 348270 Eastern Desert into a basal ‘older (ophiolitic) metavolcanics’,
Saudi Arabia ‘metasediments’, and ‘younger (arc-like) metavolcanics’. Egyp-
Wadi Sawawin BIF c. 750 278540 358460
tian Cryogenian successions are unconformably overlain by
Nuwaybah diamictite c. 750 268270 368240
Ediacaran clastic sediments of the Hammamat Group and the
Mahd diamictite c. 780 238240 2200 408460 5200
Dokhan Volcanics. A similar situation exists for Sudan. In con-
Sudan
Meritri Gp. conglomerate c. 780 198380 368420
trast, there are a plethora of stratigraphic names for Cryogenian
Fodikwan BIF – 218440 368420
supracrustal successions of Saudi Arabia, largely based on the
results of quadrangle mapping.
EARLY AND MID-CRYOGENIAN GLACIATION IN THE NORTHERN ARABIAN –NUBIAN SHIELD 279

Fig. 22.2. Stratigraphic summaries of possibly glaciogenic Cryogenian units described in this chapter, locations presented in Table 22.1. (a) Stratigraphic chart for
possibly glaciogenic conglomerates of the Meritri group, Nakasib suture, NE Sudan. Column A is a measured section for the lower part of the Arba’at volcanic rocks
along Khor Arba’at. Column B is a representative section for the sedimentary part of the Salatib group along Khor Salatib. Column C is a representative section for
diamictite of the Meritri group along Khor Meritri. Column D summarizes the stacking order observed in the nappes of the Nakasib suture (from Abdelsalam & Stern
1993). (b) Simplified stratigraphic column for Mahd group rocks showing stratigraphic position of basal diamictites (after Lowther 1994). (c) Simplified stratigraphic
column for Atud diamictite and associated BIF in Wadi Kareim, Central Eastern Desert of Egypt, from Ali et al. (2010). (d) Simplified stratigraphic column for
Nuwaybah diamictite in NW Saudi Arabia (after Davies 1985).
280 R. J. STERN ET AL.

0 50 100 km east of Gebal Atud (Table 22.1, Fig. 22.3). The Wadi Kareim
33˚E Banded Iron Formation
occurrence is especially significant (Figs 22.2c & 22.3),
28˚N Diamictite because this is the only locality where the Atud diamictite is
Ophiolitic melange found in clear stratigraphic relationship with BIF. The diamictite
Egypt is part of a supracrustal succession in which metavolcanic rocks
Ophiolites
HU
are conformably overlain by immature metasedimentary rocks
Major city (c. 100 m thick of wackestone-sandstone, siltstone and diamic-
Local Road tite) and BIF. Metavolcanic rocks at the base of the section
27˚N Highway are c. 100 m thick and are truncated below by a thrust fault.
36˚E
Egyptian diamictite and BIF are found in regional association
SA with Cryogenian ophiolites, a tripartite association that indicates
37˚E
Abu Marawat that both sedimentary units were deposited in an oceanic basin.
Sawawin Deformation obscures small-scale lateral variations of the Atud
28˚ Diamictite; the BIF is restricted to the Central Eastern Desert
ALM
QU
26˚N Qift - Quseir Ro
ad
Kareim of Egypt, with nine occurrences between 258150 N and 268350 N
Muweilih Saudi Arabia (Fig. 22.3, Table 22.1).
Dabbah In NW Saudi Arabia, the diamictite and BIF are geographically
Um Gerifat separate parts of the Za’am group, the oldest known unit in the
Um Ghamis Mobarak Midyan terrane. The diamictite (Nuwaybah Formation) is in the
Gebal EL Hadid El Imra 27˚ upper part of the Zaam group, which is an assemblage of basaltic,
MA Nuwaybah
andesitic, and subordinate rhyolitic flows and tuffs, abundant
25˚N Idfu - Marsa Alam Road
Um Nar volcaniclastic sandstone and siltstone, and volcaniclastic conglom-
Atud erate (Fig. 22.2d). The Zaam group was deformed and metamor-
AL
phosed (greenschist facies) prior to c. 660 Ma, and neither
HU: Hurghada MA: Marsa Alam
SA: Safajah AL: Al Wajh
original top nor bottom is exposed. It is overlain unconformably
QU: Quseir ALM: Al Muwaylih
by the younger Thalbah group. The significance of the basal
26˚
Thalbah group unconformity is unknown but could correspond
to Cryogenian glacial erosion. The lateral extent of the Nuwaybah
Fig. 22.3. Location of the Atud and Nuwaybah diamictite and associated BIF, diamictite is not yet known.
as well as ophiolitic rocks in Egypt and Saudi Arabia, northern ANS, with the
Further north, in Wadi Sawawin, the BIF-bearing succession
Red Sea closed. Diamictites in this area are thought to be early Cryogenian in
consists of Ghawjah Formation metavolcanic rocks overlain by
age. See Figure 22.1 for regional location (modified after Ali et al. 2010)
and Table 22.1 for geographical coordinates of sites.
Silasia Formation metasedimentary rocks. Ghawjah metavolcanic
rocks are correlated with the ‘younger metavolcanic’ rocks of
Egypt. The Silasia Formation consists of c. 1-km-thick immature
clastic metasedimentary rocks with BIF-bearing units near the
Meritri group metaconglomerate of possible glacial origin in top; the sequence is intruded by diabase sills. The stratigraphic
E. Sudan is associated with five informal groups, separated by and age relationships shown by the successions at Wadi Kareim
thrusts so that original depositional relationships are unresolved and Wadi Sawawin are remarkably similar except that the diamic-
(Fig. 22.2a). The c. 2-km-thick Arbaat (metavolcanic) group lies tite is absent from the latter succession.
at the base of the succession and appears to be autochthonous.
Subsequent units are thrust southwards on top of the Arbaat
group and one another and include from the base: the Salatib Possible glaciogenic deposits and associated strata
group, a 1.2-km-thick succession of clastic metasediments, car-
bonates, conglomerate and felsic tuffs, the Meritri group composed (A1) Meritri group metaconglomerate, E. Sudan
of coarse conglomerate, lithic greywacke, limestone, red sandstone
and felsic metavolcanics, the Nakasib ophiolite and at the top, the Abdelsalam & Stern (1993) inferred that the original thickness for
arc-like Shalhout group (Fig. 22.2a; Abdelsalam & Stern 1993). the Meritri group in Khor Meritri was c. 2 km. They subdivided the
Lateral variations within the Meritri group show changes from group into four formations comprising, from oldest to youngest:
granitic-clast dominated in the east to volcanic-clast dominated conglomerate (possibly glaciogenic); lithic greywacke; inter-
in the west. calated limestone, red sandstone and felsic tuff; and felsic volcanic
Diamictite at the base of the c. 780 Ma Mahd group (Fig. 22.2b) rocks. The coarsest clastic beds are made up of abundant polymict
unconformably overlies diorite and tonalite of the c. 810 Ma conglomerate intercalated with minor lithic greywacke and lime-
Dhukhr batholith, indicating an episode of possibly glacial stone. The polymict conglomerate is matrix supported, with the
erosion at c. 780 –810 Ma (Johnson et al. 2003). Continuity of matrix made up of lithic greywacke and minor carbonate. Clasts
unit to the north is truncated by terrane boundary; possible conti- include granite, granodiorite, diorite, rhyolite, ignimbrite and car-
nuity to the south and east is unknown. The basal Mahd group bonate as well as subordinate clastic metasediments. Intermediate
unconformity is one of the oldest unconformities known in the to felsic volcanic clasts become more abundant to the west. In the
ANS. Another early Cryogenian unconformity, which might be NE, along Khor Meritri, plutonic clasts are most abundant, com-
related to the basal Mahd group unconformity, is found farther prising c. 50% of total clasts, whereas volcanic clasts are c. 35%
south in the Arabian Shield, between plutonic rocks of the and clasts of metasediments make up c. 15%. Abdelsalam (pers.
c. 800 Ma An Nimas batholith and metamorphosed conglomerate, comm. 2007) measured 76 of these clasts, finding a maximum
limestone, and sandstone of the 780 –795 Ma Hali group (Cooper size of 70  40  35 cm3 (clasts are deformed and stretched);
et al. 1979). This encourages speculation that heretofore unrecog- clasts .40 cm are common. The lithic greywacke unit is made
nized Cryogenian glacial deposits might exist in the southern up of lithic greywacke intercalated with minor felsic volcanic
Arabian Shield. layers and limestones. Locally the lithic greywacke grades into
The Atud diamictite is a distinctive lithology found in the conglomerate with smaller clasts of mostly felsic volcanic rocks.
Egyptian metasedimentary succession. This has been reported Sedimentary structures include graded bedding in the sandstone,
from four areas in the Central Eastern Desert of Egypt: Wadi cross-bedding and channels. Channel structures and cross-bedding
Kareim, Wadi Mobarak, Wadi Muweilih, and the type locality indicate that the palaeocurrent direction was from SE to NW.
EARLY AND MID-CRYOGENIAN GLACIATION IN THE NORTHERN ARABIAN –NUBIAN SHIELD 281

(A2) Basal Mahd group diamictite, central Arabian Shield Boundary relations with overlying and underlying
non-glacial units
Although dominated by volcanic rocks, the Mahd group has a 1–
5-m-thick diamictite resting on the unconformity. The diamictite is (A1) Meritri Group metaconglomerate, E. Sudan
matrix-supported, with a dark-grey, immature, arkosic matrix con-
taining abundant, angular to sub-angular clasts (up to 30 cm Contacts with underlying and overlying units are faulted.
across) of granitic and felsic volcanic rocks.
(A2) Basal Mahd Group diamictite, central Arabian Shield
(B1) Atud diamictite, E. Egypt
Contacts with underlying units are unconformable. Contacts with
The Atud diamictite consists of massive poorly sorted and rounded overlying units are conformable.
clasts, from gravel to boulder, in a sheared grey matrix. Atud clasts
include quartzite, highly altered granitoid, and a distinctive arkosic (B1) Atud diamictite, E. Egypt
breccia (microdiamictite) (Ali et al. 2010). Clasts and matrix
zircons in the Atud diamictite are dominated by c. 750 Ma granitic Contacts with underlying units are only clear at Wadi Kareim,
rocks and microdiamictite but with a significant amount of where they appear conformable.
Palaeoproterozoic and Neoarchean granitic rocks and quartzite
(Ali et al. 2010). (B2) Nuwaybah Formation (Za’am group) diamictite,
We have also examined Atud diamicite from the eastern half of
the Wadi Mobarak belt in Egypt. Basement exposures around NW Saudi Arabia
Wadi Mobarak are dominated by highly deformed ophiolitic frag-
ments of serpentinites, metagabbro, and greenschist-facies mafic The basal contact of the Nuwaybah Formation is thought to be con-
metavolcanic rocks, whereas the sedimentary sequence includes formable on the underlying Umm Ashsh Formation (part of the
tuff, shale, schist and Atud diamictite (Akaad et al. 1995). Diamic- Za’am group). The top of the Formation is obscured by granite
tite clasts are similar to those observed in Wadi Kareim. intrusions and by an angular unconformity with younger sedimen-
tary rocks (Thalbah group).
(B2) Nuwaybah Formation (Za’am group) diamictite, (C1) BIF, E. Egypt
NW Saudi Arabia
This BIF is found in conformable stratigraphic relationship with
The Nuwaybah locality is located within the Al Wajh quadrangle underlying immature, possibly tuffaceous greywackes. It is associ-
in the northwestern part of the Arabian Shield. The diamictite ated with metavolcanic rocks and intrusive diabase.
locality we studied is beautifully exposed in a roadcut along
the Red Sea highway (Table 22.1). Clasts and matrix zircons in (C2) Silasia Formation BIF, NW Saudi Arabia
the Nuwaybah diamictite are dominated by c. 750 Ma granitic
rocks and distinctive arkose/micro-diamictite but with a signifi- The BIF-bearing sediments are conformable above metavolcanic
cant amount of Palaeoproterozoic and Neoarchaean granitic rocks of the Ghawjah Formation. The top of the formation is not
rocks and quartzite (Ali et al. 2010), similar to lithologies and exposed at the Sawawin locality because it is intruded by metadiorite.
ages of the Atud diamictite in Egypt.

(C1) Banded iron formation (BIF), E. Egypt Chemostratigraphy

BIF of early Cryogenian age is found in the Central Eastern Desert This section is applicable only for the BIF. There are no significant
of Egypt, with nine occurrences between latitudes 258150 and carbonate sediments in the region, although these become increas-
268350 N. The Fodikwan BIF in NE Sudan may be related ingly important to the south (in Sudan, S. Saudi Arabia, Eritrea,
(Table 22.1, Fig. 22.1), but more work is needed to test this possi- and Ethiopia, see Allen et al. 2011a, b; Miller et al. 2011). NW
bility. BIF occurs as fairly regular bands interbedded with meta- Saudi Arabia (Sawawin) BIF has been sampled systematically
sediments and metavolcanics in a zone that originally had a and stratigraphically, in order to identify the nature of its sources
stratigraphic thickness of 100– 200 m, within which the aggregate (Mukherjee 2008). Samples of Egyptian BIF were also studied in
BIF thickness is about 10 –20 m (Sims & James 1984). Egyptian order to understand regional variations. Rare earth elements
BIF is mostly an oxide facies, consisting of interlaminated hema- (REE) are particularly useful for studying the BIF because these
tite and jasper, and containing 40– 46% Fe (Sims & James 1984). record the composition of equilibrium seawater that the BIF
In several locales, BIF-bearing metasediments are intruded by precipitated from.
metadiabase sills. REE patterns for the Sawawin BIFs are similar to a mixture of
modern shallow suboxic seawater (German et al. 1991; Webb &
Kamber 2000) with low-T hydrothermal vent fluid solution (Bau
(C2) Silasia Formation BIF, NW Saudi Arabia & Dulski 1999) suggesting a dominant hydrothermal input of
REEs and by analogy Fe into the BIFs (Mukherjee 2008). BIF
BIFs in NW Saudi Arabia occupy a more restricted region than do iron input sources are isotopically dominated by hydrothermal
their Egyptian counterparts. Arabian BIF occurs within the Silasia vent fluids but continental runoff is also significant as revealed
Formation, which, like the Egyptian section, consists of volcano- by Ce/Ce* and Eu/Eu* (Mukherjee 2008).
genic greywackes that appear to rest conformably on metavolcanic
rocks (Ghawjah Formation). The exposed thickness of the Silasia
Formation is estimated to be c. 1160 m in the reference area of Other characteristics
Wadi Sawawin (Goldring 1990). Arabian BIF is mostly present
as an oxide facies, consisting of interbedded hematite and jasper, Gold mineralization may be associated with BIF at Abu
and containing 40–46% Fe (Goldring 1990). Similar to the Egyp- Marawat (Botros 2002). BIF here occurs as sharply defined hor-
tian section, the Silasia Formation is intruded by metadiabase sills izons within a volcanic-sedimentary succession, which is region-
up to 100 m thick. ally metamorphosed into greenschist facies. Gold concentrations
282 R. J. STERN ET AL.

of up to 2.15 ppm occur in the BIF, either enclosed in the flaky SHRIMP U –Pb zircon ion probe age of c. 750 Ma for the meta-
hematite crystals of hematite-rich layers or as fine inclusions in volcanics that lie beneath the Atud diamictite at Wadi Kareim.
magnetite-rich bands. One metavolcanic sample yielded a weighted mean 206Pb– 238U
El-Habaak & Mahmoud (1995) identified spherical bodies as age of 769 + 29 Ma.
Eosphaera tyleri in jasper-rich layers from Wadi Kareim BIF, As noted above, the Atud diamictite at Wadi Kareim is
Egypt. Those occur as more or less clear spherules of quartz sur- dominated by Cryogenian clasts and matrix material but con-
rounded by thin veneers of very fine hematite granules. They tains abundant clasts of Palaeoproterozoic and Neoarchaean
recognized two distinct varieties of E. tyleri, one of which is granitic rocks and pre-Neoproterozoic quartzite. Clasts in the
c. 15 mm in diameter and the other c. 60 mm; El-Habaak & diamictite are as young as 754 + 15 Ma (Ali et al. 2010).
Mahmoud (1995) suggested that biological activity played a role Because of stratigraphic relationships shown on Figure 22.2c,
in spherule formation and BIF deposition. the maximum age of c. 750 Ma for the metavolcanics and
The Mahd group basal diamictite is overlain by hypabyssal intru- 754 + 12 Ma for the Atud diamictite at Wadi Kareim also pro-
sive to volcaniclastic rocks (caldera complex) that hosts a major vides a maximum age for BIF deposition at this locality (Ali
epithermal gold deposit (Mahd adh Dhahab Mine). The age of gold et al. 2009, 2010).
mineralization is not known with certainty, but is likely to have Geochronological data support the inference that clasts in
been syngenetic with Mahd Group igneous activity (c. 760– the Atud diamictite sample much older rocks than are exposed in
c. 780 Ma; Hargrove III et al. 2006). The deposit consists of quartz the Eastern Desert of Egypt and so must have been transported
veins and stockwork that contains copper, zinc, iron, and lead sul- some distance. Two granitic cobbles from NW of Marsa Alum
phides and very fine-grained gold and silver (averaging 5–30 mm), (also referred to as the Wadi Mobarak metasedimentary unit)
mostly as tellurides, associated with the sulphides (Moore 1979). yielded highly discordant conventional multigrain U –Pb zircon
upper intercept ages of 1120 and 2060 Ma (Dixon 1981). Dixon
(1979) obtained a discordant U –Pb zircon upper intercept of
Palaeolatitude and palaeogeography 2.3 Ga for a granitic cobble from Atud conglomerate outcrops
west of Quesir. SHRIMP geochronological studies by Ali et al.
Extensive deformation and metamorphism makes palaeomagnetic (2010) confirm and extend Dixon’s conclusions that many Atud
determinations unreliable, and for this reason few have been clasts are pre-Neoproterozoic.
carried out (Reischmann et al. 1992).

(B2) Nuwaybah Formation (Za’am Group) diamictite,


Geochronological constraints
NW Saudi Arabia
(A1) Meritri Group metaconglomerate, E. Sudan
Like the Atud diamictite, this unit is dominated by Cryogenian
material but contains abundant clasts of Palaeoproterozoic and
The age of the Meritri group is constrained by the age of underlying
Neoarchaean granitoids and quartzite. The youngest clasts in the
metavolcanics and intrusive plutonic rocks, reported by Stern &
Nuwaybah diamictite yield SHRIMP U –Pb zircon ages of
Abdelsalam (1998). Conventional multigrain analyses of two size
765 + 22 Ma (Ali et al. 2010). Another clast is an arkose with a
fractions of zircons separated from a metarhyolite lava of the appar-
SHRIMP U – Pb zircon age of 766 + 5 Ma, which we interpret
ently underlying Arba’at Formation yielded a nearly concordant
as the age of the basement supplying the arkose, which was
age of 790 + 2 Ma. The supracrustal sequence, including the
rapidly eroded and deposited, lithified, and re-eroded, all appar-
Meritri group, was intruded by the Arba’at quartz diorite, which
ently within a very few millions of years.
yielded a single fraction, nearly concordant U–Pb zircon age of
The formation has a minimum age of deposition constrained by
779 + 3 Ma (Stern & Abdelsalam 1998). Faulting leads to uncer-
the age of intrusions in the Zaam Group near the diamictite. These
tainties about the exact relationship between the intrusion and
include the Buwaydah complex, dated by the conventional U –Pb
Meritri group, but these ages suggest that deposition of the Meritri
zircon method at 725 + 4 Ma (Hedge 1984) and the Imdan
diamictite may be constrained between 779 + 3 and 790 + 2 Ma.
complex, dated by the SHRIMP U –Pb zircon method at
676 + 6 Ma (15 data points (Kennedy et al. 2011) and convention-
(A2) Basal Mahd Group diamictite, central Arabian Shield al U – Pb zircon method at 660 + 4 Ma (Hedge 1984). Zaam Group
felsic tuff 175 km SE of the diamictite locality yields U– Pb zircon
The age of this diamictite is constrained by the age of subjacent and SHRIMP ages of 711 + 10 Ma and 708 + 4 Ma (Kennedy et al.
superjacent igneous rocks. The Dhukhr complex, which uncon- 2004, 2005). However, the stratigraphic relationship between the
formably underlies the diamictite, has robust conventional multi- diamictite and these tuffs is poorly understood, thus the signifi-
grain U –Pb zircon crystallization ages of 811 + 4 Ma (Stoeser cance of these data for constraining the depositional age of the dia-
& Stacey 1988) and 816 + 4 Ma (Calvez & Kemp 1982). The mictite is not clear.
Hufayriyah batholith, which also unconformably underlies the
Mahd group c. 60 km north –NE of the diamictite locality, has a
U –Pb zircon SHRIMP age of 785 + 6 Ma (Hargrove III et al. (C1) BIF, E. Egypt
2006). Rhyolite in the Mahd group upsection from the basal dia-
mictite has been dated by U –Pb zircon SHRIMP techniques at BIF age is best known from Wadi Kareim, where ages of conform-
777 + 5 Ma (Hargrove III et al. 2006). In addition to the U –Pb ably underlying metavolcanics and diamictite indicate that it is
zircon SHRIMP age, the Hufayriyah batholith also yields a con- younger than c. 750 Ma (Ali et al. 2009, 2010), but perhaps not
ventional multigrain U –Pb zircon age of 760 + 10 Ma (Calvez much younger.
& Kemp 1982); this result may represent a younger pulse of
Hufayriyah intrusion and is not necessarily a robust constraint on (C2) Silasia Formation BIF, NW Saudi Arabia
the age of rocks at the base of the Mahd group.
The Silasia Formation, which hosts this BIF, is intruded by pluto-
(B1) Atud diamictite, E. Egypt nic rocks of the Muwalylih suite, dated by U –Pb zircon techniques
at 710–725 Ma (Hedge 1984). Unfortunately, Hedge (1984) did
The age of this unit is constrained to be younger than both the age not provide analytical details, the number of zircon populations
of the underlying metavolcanics (Fig. 22.2c) and the age of the or statistics about the results (uncertainty and mean square of the
youngest clast within the diamictite. Ali et al. (2009) report a weighted deviates). Ali et al. (2011) report a U –Pb zircon ion
EARLY AND MID-CRYOGENIAN GLACIATION IN THE NORTHERN ARABIAN –NUBIAN SHIELD 283

probe weighted mean 206Pb/238U age of 763 + 25 Ma for of the post-accretionary basins of Arabia may preserve glaciogenic
Ghawjah metavolcanics beneath the BIF. Silasia Formation was deposits of Ediacaran age, but focused investigations are needed to
deformed and then intruded by Sawawin complex diorite, which establish or refute this possibility (Johnson 2003).
yields a U – Pb zircon concordia age of 661.5 + 2.3 Ma (Ali
et al. 2011). Silasia Formation BIF is thus , 763 + 25 Ma, prob- We are grateful for support by many individuals and agencies over the years
ably c. 750 Ma, similar in age to Egypt BIF. including the Sudan Geologic Survey, Saudi Geological Survey, Nuclear Materials
Agency of Egypt, NASA, NSF and the US–Egypt Joint Technical Program. We
also thank M. Abdelsalam, D. Avigad, M. Beyth, N. Miller, F. Kattan, V. Pease,
M. Whitehouse and A. Kröner for discussions on this topic, A. Collins, V. Pease
Discussion
and D. Avigad for thoughtful reviews of the manuscript, and E. Arnaud for
careful final editing. This represents a contribution of the IUGS- and UNESCO-
The foregoing summary indicates that there is compelling evi-
funded IGCP (International Geoscience Programme) Project #512.
dence of a c. 740 Ma age glaciation and plausible evidence of a
c. 780 Ma glaciation in the northern ANS. Indistinguishable brack-
eting ages of the Meritri group (790 + 2 Ma to 779 + 3 Ma) and
Mahd group (785 + 6 Ma to 777 + 5 Ma) indicate that deposition References
of these units may have occurred about the same time. A glacio-
genic interpretation for c. 780 Ma Mahd-Meritri groups might Abdelsalam, M. G. & Stern, R. J. 1993. Tectonic evolution of the
help explain the enigmatic Bitter Springs d13C excursion of Nakasib suture, Red Sea Hills, Sudan: evidence for a late Precambrian
Wilson cycle. Journal of the Geological Society, London, 150,
approximately the same age, thought to be associated with glacia-
393– 404.
tion but for which no definitive evidence of glaciation has been
Abdelsalam, M. G. & Stern, R. J. 1997. Sutures and shear zones in the
identified to date (Halverson et al. 2005, 2007). The association Arabian-Nubian Shield. Journal of African Earth Sciences, 23,
of the Mahd diamicite with the oldest unconformity known from 289– 310.
the Arabian Shield is also consistent with a glacial interpretation Abdelsalam, M. G., Liégeois, J. P. & Stern, R. J. 2002. The Saharan
and is tentative evidence of a c. 780 Ma local continental Metacraton. Journal of African Earth Sciences, 34, 119–136.
glaciation, probably sourced to the south (present coordinates). Akaad, M. K., Noweir, A. M. & Abu-El-Ela, A. M. 1995. The volcano-
Certainly, more work is needed to map, study sedimentology and sedimentary association and ophiolites of Wadi Mubarak, eastern
stratigraphy, date clasts and matrix, and examine intervening Desert, Egypt. Proc. Inter. Conf. 30 Years Cooper., 69. Geological
sedimentary and possible igneous environments of Mahd and Survey of Egypt Special Publication, 231–248.
Meritri groups deposits. Ali, K. A., Stern, R. J., Manton, W. I., Kimura, J.-I. & Khamis, H. A.
Evidence for a c. 740 Ma glaciation is found in the ophiolite- 2009. Geochemistry, Nd isotopes, and U –Pb SHRIMP zircon dating
Atud diamictite-BIF basin and includes far travelled exotic clasts of Neoproterozoic volcanic rocks from the Central Eastern Desert of
(based on age and composition) and association of diamictite Egypt: new insights into the c. 750 Ma crust-forming event. Precam-
with BIF. By virtue of their age, these deposits may correlate brian Research, 171, 1 –22.
with early Cryogenian glaciogenic successions elsewhere (Fair- Ali, K. A., Stern, R. J., Manton, W. I., Johnson, P. R. & Mukherjee,
child & Kennedy 2007). The Atud-Nuwaybah diamictite may be S. K. 2010. Neoproterozoic diamictite in the Eastern Desert of Egypt
strictly correlative, with similar ages of deposition, clast lithology and Northern Saudi Arabia: evidence of c. 750 Ma glaciation in the
and distribution of Cryogenian, Palaeoproterozoic and Neoarch- Arabian– Nubian Shield. International Journal of Earth Sciences,
aean clast ages (Ali et al. 2010). The association of diamictite 99, 705– 726.
Ali, K. A., Stern, R. J. et al. 2011. Geochemical, U– Pb zircon and Nd
with BIF is consistent with the genetic coupling suggested by
isotopic investigations of the Ghawjah metavolcanics of northwestern
Snowball Earth concepts (Hoffman 2005), although alternative
Saudi Arabia: the 750 Ma crust-forming event correlated across the
interpretations exist. Based on similar age constraints (the young- Red Sea. Lithos, 120, 379– 392.
est detrital zircons in the Ethiopian diamictites are also c. 0.75 Ga), Allen, P. A., Rieu, R., Etienne, J. L., Matter, A. & Cozzi, A.
Atud diamictite may also be broadly synchronous with Negash and 2011a. The Ayn Formation of the Mirbat Group, Dhofar, Oman.
Shiraro diamictites in northern Ethiopia (Avigad et al. 2007; Miller In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds)
et al. 2003, 2011). Neoproterozoic diamictites from the ANS (this The Geological Record of Neoproterozoic Glaciations. Geological
chapter and Miller et al. 2011) do not seem to correlate directly Society, London, Memoirs, 36, 239–249.
with diamicitites of the Abu Mahara Group of the Huqf Super- Allen, P. A., Leather, J. et al. 2011b. The Abu Mahara Group
group, Oman, which are younger in age (700– 735 Ma) and are (Ghubrah, and Fiq Formations), Jabal Akhdar, Oman. In: Arnaud,
dominated by c. 860 Ma zircons that are not common in the E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
Atud/Nuwaybah diamictites (Rieu et al. 2007). Record of Neoproterozoic Glaciations. Geological Society, London,
Because pre-Neoproterozoic basement is unknown in Egypt east Memoirs, 36, 251–262.
of the Nile, Dixon (1981) concluded that these clasts were derived Avigad, D., Stern, R. J., Beyth, M., Miller, N. & McWilliams, M. O.
from older crust to the west or south, perhaps from the Saharan 2007. Detrital zircon U– Pb geochronology of Cryogenian
Metacraton (Abdelsalam et al. 2002). Dixon (1979) suggested diamictites and Lower Paleozoic sandstone in Ethiopia (Tigrai):
that Atud clasts were transported such great distances by ice age constraints on Neoproterozoic glaciation and crustal evolution
rafting, but other possible explanations (e.g. meteorite impact of the southern Arabian– Nubian Shield. Precambrian Research,
ejecta blanket, far-travelled submarine debris flow) are possible. 154, 88– 106.
There is controversy regarding how the Egyptian BIFs formed, Bau, M. & Dulski, P. 1996. Distribution of yttrium and rare-earth
elements in the Penge and Kuruman iron-formations, Transvaal
although ideas published in the geological literature were mostly
Supergroup, South Africa. Precambrian Research, 79, 37 –55.
developed prior to the Snowball Earth hypothesis (Fairchild &
Bau, M. & Dulski, P. 1999. Comparing yttrium and rare earths in hydro-
Kennedy 2007). Sims & James (1984) suggested that BIF thermal fluids from the Mid-Alantic Ridge: Implications for Y and
formed as chemical precipitates during lulls in dominantly subaqu- REE behaviour during near-vent mixing and for the Y/Ho ratio of
eous, calc-alkaline volcanism, apparently within an intraoceanic Proterozoic seawater. Chemical Geology, 155, 77 – 90.
island-arc environment. The close association in time and space Botros, N. S. 2002. Metallogeny of gold in relation to the evolution of the
between volcanic activity and deposition of the BIF suggests a Nubian Shield in Egypt. Ore Geology Reviews, 19, 137– 162.
genetic relation (Sims & James 1984). Calvez, J.-Y. & Kemp, J. 1982. Geochronological investigations in the
Finally, it should be noted that there is no strong evidence for or Madh adh Dhahab quadrangle, Central Arabian Shield. BRGM-TR-
against an Ediacaran glaciation in the ANS. The ANS peneplain 02-5, Saudi Arabian Deputy Ministry for Mineral Resources.
may have been partly cut during this time, but regional uplift Cooper, J. A., Stacey, J. S., Stoeser, D. G. & Fleck, R. J. 1979. An
resulted in generally poor preservation (Stern et al. 2006). Some evaluation of the zircon method of isotopic dating in the Southern
284 R. J. STERN ET AL.

Arabian Craton. Contributions to Mineralogy and Petrology, 68, Kennedy, A., Kozdroj, W., Johnson, P. R. & Kattan, F. H. 2011.
429– 439. SHRIMP geochronology in the northern Arabian Shield. Part III.
Dalziel, I. W. D. 1997. Neoproterozoic –Paleozoic geography and tec- Data Acquisition, 2006. Saudi Geological Survey, Open-File
tonics: review, hypothesis, environmental speculation. Geological Report SGS-OF-2007-9.
Society of America Bulletin, 109, 16 –42. Lowther, J. M. 1994. Mahd adh Dhahab gold deposit. In: Collenette,
Davies, F. B. 1985. Geologic map of the Al Wajh quandrangle, sheet 26B, P. & Grainger, D. J. (eds) Mineral Resources of Saudi Arabia. Saudi
Kingdom of Saudi Arabia. Saudi Arabian Deputy Ministry for Arabian Directorate General of Mineral Resources, Special Publi-
Mineral Resouces, Geoscience Map 83A, Jeddah. cation SP-2, Jeddah, 105– 111.
Dixon, T. H. 1979. The evolution of continental crust in the Late Precam- Ludwig, K. R. 2000. SQUID 1.00. A User’s Manual, Berkeley Geochro-
brian Egyptian Shield. Ph.D. Thesis, UC San Diego. nology Center, Berkeley.
Dixon, T. H. 1981. Age and chemical characteristics of some pre-Pan- Miller, N. R., Alene, M., Sacchi, R., Stern, R., Conti, A., Kröner, A.
African rocks in the Egyptian Shield. Precambrian Research, 14, & Zuppi, G. 2003. Significance of the Tambien group (Tigrai,
119– 133. N. Ethiopia) for Snowball Earth events in the Arabian– Nubian
El-Essawy, M. A. 1964. Geology of the area east of Gabal Atud, Shield. Precambrian Research, 121, 263– 276.
Eastern Desert, Egypt. MSc thesis, Assiut University, Assiut, Miller, N. R., Avigad, D., Stern, R. J. & Beyth, M. 2011. The Tambien
Egypt, 235. Group, Northern Ethiopia (Tigre). In: Arnaud, E., Halverson, G. P.
El-Habaak, G. H. & Mahmoud, M. S. 1995. Carbonaceous bodies of & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
debatable organic provenance in the banded iron formation of the Glaciations. Geological Society, London, Memoirs, 36, 263–276.
Wadi Kareim area, Eastern Desert, Egypt. Journal of African Earth Mukherjee, S. K. 2008. Petrography, age (uranium-lead zircon), geo-
Sciences, 19, 125– 133. chemical and isotopic studies of the Sawawin Banded Iron Formation
Fairchild, I. J. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the (BIF), Northwestern Saudi Arabia: Implications for understanding
Earth System. Journal of the Geological Society, London, 164, Neoproterozoic climate change. PhD thesis, The University of
895– 921. Texas at Dallas, 147.
Flint, R. F., Sanders, J. E. & Rodgers, J. 1960. Diamictite, a substitute Reischmann, T., Bachtadse, V., Kröner, A. & Layer, P. 1992. Geo-
term for symmictite. Geological Society of America Bulletin, 71, chronology and palaeomagnetism of a Late Proterozoic island arc
1809– 1810. terrane from the Red Sea Hills, northeast Sudan. Earth and Planetary
German, C. R., Holiday, B. P. & Elderfield, H. 1991. Redox cycling of Science Letters, 114, 1– 15.
rare earth elements in the suboxic zone of the Black Sea. Geochimica Ries, A. C., Shackleton, R. M., Graham, R. H. & Fitches, W. R. 1983.
et Cosmochimica Acta, 55, 3553–3558. Pan-African structures, ophiolites and mélange in the Eastern Desert
Goldring, D. C. 1990. Banded iron formation of Wadi Sawawin district, of Egypt: a traverse at 268N. Journal of the Geological Society of
Kingdom of Saudi Arabia. Transactions of the Institution of Mining London, 140, 75 –95.
and Metallurgy (Sect. B: Applications for Earth Science), 99, Rieu, R., Allen, P. A., Cozzi, A., Kosler, J. & Bussy, F. 2007. A com-
B1 –B14. posite stratigraphy for the Neoproterozoic Huqf Supergroup of
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Oman: integrating new litho-, chemo-, and chronostratigraphic data
Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon- of the Mirbat area, southern Oman. Journal of the Geological
isotope record. Geological Society of America Bulletin, 117, Society, London, 164, 997–1009.
1181– 1207. Sims, P. K. & James, H. L. 1984. Banded iron-formations of late Protero-
Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A. 2007. zoic age in the central Eastern Desert, Egypt; geology and tectonic
Evolution of the 87Sr/86Sr composition of Neoproterozoic seawater. setting. Economic Geology, 79, 1777– 1784.
Palaeogeography, Palaeoclimatology, Palaeoecology, 256, 103–129. Stern, R. J. 1981. Petrogenesis and tectonic setting of late Precambrian
Hargrove III, U. S., Stern, R. J., Griffin, W. R., Johnson, P. R. & ensimatic volcanic rocks, Central Eastern Desert of Egypt. Precam-
Abdelsalam, M. G. 2006. From Island Arc to Craton: Timescales brian Research, 16, 195–230.
of Crustal Formation along the Neoproterozoic Bi’r Umq Suture Stern, R. J. 1994. Arc assembly and continental collision in the Neopro-
Zone, Kingdom of Saudi Arabia, Saudi Geological Survey. terozoic East African orogen: implications for the consolidation of
Hedge, C. E. 1984. Precambrian Geochronology of part of northwestern Gondwanaland. Annual Reviews of Earth and Planetary Sciences,
Saudi Arabia, Kingdom of Saudi Arabia, U.S. Geological Survey. 22, 319– 351.
Hoffman, P. F. 2005. On Cryogenian (Neoproterozoic) ice-sheet Stern, R. J. 2008. Neoproterozoic crustal growth: the solid Earth system
dynamics and the limitations of the glacial sedimentary record. during a critical time of Earth history. Gondwana Research, 14,
28th DeBeers Alex. Du Toit Memorial Lecture. South African 33 – 50.
Journal of Geology, 108, 557– 576. Stern, R. J. & Abdelsalam, M. G. 1998. Formation of continental
Johnson, P. R. 2003. Post-amalgamation basins of the NE Arabian shield crust in the Arabian –Nubian shield: evidence from granitic rocks
and implications for Neoproterozoic III tectonism in the northern of the Nakasib suture, NE Sudan. Geologische Rundschau, 87,
East African orogen. Precambrian Research, 123, 321– 338. 150– 160.
Johnson, P. R. & Woldehaimanot, B. 2003. Development of the Stern, R. J., Avigad, D., Miller, N. R. & Beyth, M. 2006. Evidence
Arabian –Nubian Shield: perspectives on accretion and deformation for the snowball Earth hypothesis in the Arabian –Nubian Shield
in the East African Orogen and the assembly of Gondwana. In: and the East African Orogen. Journal of African Earth Sciences,
Yoshida, M., Windley, B. F. & Dasgupta, S. (eds) Proterozoic 44, 1 – 20.
East Gondwana: Supercontinent Assembly and Breakup. Geological Stoeser, D. B. & Stacey, J. S. 1988. Evolution, U– Pb geochronology,
Society, London, Special Publications, 206, 289–325. and isotope geology of the Pan-African Nabitah orogenic belt of
Johnson, P. R., Abdelsalam, M. G. & Stern, R. J. 2003. The Bi’r Umq- the Saudi Arabian Shield. In: El-Gaby, S. & Greiling, R. O. (eds)
Nakasib suture zone in the Arabian-Nubian shield: a key to under- The Pan-African Belt of NE Africa and Adjacent Areas. Friedr.
standing crustal growth in the East African Orogen. Gondwana Vieweg & Sohn, Braunschweig, 227– 289.
Research, 6, 523–530. Sultan, M., Arvidson, R. E., Duncan, I. J., Stern, R. J. & Kaliouby,
Kennedy, A., Johnson, P. R. & Kattan, F. H. 2004. SHRIMP geochro- B. E. 1988. Extension of the Najd shear system from Saudi Arabia to
nology in the northern Arabian Shield. Part I: Data acquisition. Saudi the central Eastern Desert of Egypt based on integrated field and
Geological Survey Open File Report, SGS-OF-2004-11. Landsat observations. Tectonics, 7, 1291–1306.
Kennedy, A., Johnson, P. R. & Kattan, F. H. 2005. SHRIMP geochro- Webb, G. E. & Kamber, B. S. 2000. Rare earth elements in Holocene
nology in the northern Arabian Shield. Part II: Data acquisition 2004. reefal microbialites; a new shallow seawater proxy. Geochimica et
Saudi Geological Survey Open File Report, SGS-OF-2005-10. Cosmochimica Acta, 64, 1557– 1565.
Chapter 23

Glacial deposits of the Bokson Group, East Sayan Mountains, Buryatian Republic,
Russian Federation

NICKOLAY M. CHUMAKOV
Geological Institute of the Russian Academy of Sciences, Pyzhevsky per. 7, Moscow 119017, Russia (e-mail: chumakov@ginras.ru)

Abstract: The Bokson Group (Gr.) forms the platform cover in northern part of the Tuva-Mongolian Massif, south of the East Sayan
Mountains. Widespread diamictites occur in the lower part of the Bokson Gr., in the Zabit Formation (Fm.). These contain erratic, faceted
and striated stones along with dropstones. They have disconformable or gradual lower boundaries and conformable or transitional upper
boundaries with overlying deposits. The isotopic d13Ccrb curve records some maximum and minimum values as a background of insig-
nificantly varying values (from – 3 to 2‰). Maximum values (up to 5.5‰), are recorded in deposits underlying diamictites, whereas the
overlying beds of the Zabit Formation (Fm.) yielded three peaks or minima down to 24‰. A fourth negative anomaly derived from the
middle part of the Bokson Gr. in the lower part (close to base) of the Tabinzurt Fm. is as low as – 5.7‰. The deposits underlying the
diamictites contain Ediacaran (Vendian) microfossils. Cloudina sp. was discovered in the granule of diamictites, whereas beds just
above include small shelly fossils (SSF) of the Nemakit-Daldynian type (Cambrotubulus decurvatus Miss., Anabarites trisulcatus
Miss.), which is overlain by a member with abundant fossils of the Tommotian assemblage. The diamictites represent glaciomarine
deposits formed in the uppermost Ediacaran or lowermost Cambrian stage (Upper Vendian).

The Bokson Gr. occurs in the Oka River basin, in the southern part volcanic rocks have yielded Rb – Sr ages of 718 + 30 Ma (ero-
of the East Sayan Mountains, Buryatian Republic of the Russian chrona, Buyakayte et al. 1989). Both thickness and composition
Federation. Intensive investigation of the region began in the mid- of the lower part of Bokson Gr. vary noticeably laterally. In the
twentieth century in search of bauxite (Il’ina 1958) and phosphor- type section the lowermost Bokson Gr. is the Zabit Fm. A compo-
ite (Volkov et al. 1972), with detailed stratigraphic work published site section of the lower part of the Bokson Gr. at the type section
later (e.g. Khomentovskiy et al. 1985; Terleev & Zadorozhnyi consists of the following from the base upwards (Kheraskova &
1996; and others). Reconstruction of sedimentation history Samygin 1992; Osokin & Tyzhinov 1998; Kuzmichev 2004):
(Kheraskova & Samygin 1992) with specific emphasis on the
Sarkhoy Gr.
nature of the diamictites (Osokin & Tyzhinov 1998) was made
Dzhundzhugur Fm., Sarkhoy Gr.
later. Initially the Bokson Gr. was defined as a formation but
further studies resulted in dividing it into five, sometimes six, for- (1) Deformed basic volcanogenic rocks, dark-coloured clayey and
mations. Diamictites are usually described from the lower part of siliceous shales, diabases.
the Zabit Fm. (Kheraskova & Samygin 1992; Pokrovskiy et al. (2) Weathering crust (a few metres).
1999; and others) or from the basal, mainly terrigenous, Khushatay
Fm. (Osokin & Tyzhinov 1998; Kuzmichev 2004). The type Angular unconformity
section of the diamictic member of the Zabit Fm. occurs in the The Boxon.
Bokson River basin (Fig. 23.1, I) and that of the Khushatay Fm. The Zabit Fm.
in the Sarkhoy River basin (Fig. 23.1, II). (3) Boulder and large pebbled conglomerates composed mainly
of fragments of underlying deposits, less frequently granites
(3 –40 m).
Structural framework (4) Diamictites with a mudstone matrix with scattered stones of
different size, composition and roundness (0 –100 m).
The Bokson Gr. forms a platform cover of the Tuva-Mongolian (5) Grey and dark grey dolomites, often clastic and oncolitic with
Massif, unconformably lying above the Tonian-Cryogenian? abundant sedimentary structures indicative of subaqueous
(Upper Riphean) folded basement composed of deposits shed in slumping (200 –225 m).
an island arc and active continental margin settings. During the (6) Breccia consisting of both angular and rounded fragments of
early Palaeozoic Caledonian folding, the Bokson Gr. experienced dolomite and carbonate matrix; frequent structures indicative
significant deformation, which makes correlation of units between of underwater slumping (5– 15 m).
different sites difficult: deformation includes folding, disjunctive (7) Massive, locally sandy, grey and dark grey dolomites (15 m).
faulting and overthrusting (Terleev & Zadorozhnyi 1996; Fedo- (8) Yellow-grey dolomitic phosphorites and phosphatic dolomites
tova & Khain 2002). In the north, the deposits were metamor- (30 –35 m).
phosed to chlorite-sericite phase, less frequently than quartz- (9) Dark grey dolomites, partly brecciated and stromatolitic
biotite shales (Kheraskova & Samygin 1992; Kuzmichev 2004). (140 –200 m).
The Bokson Gr. fills the Bokson-Sarkhoy Synclinorium in the
northern part of the Tuva-Mongolian Massif (Fig. 23.1). The Hiatus and palaeokarst
lower part of the group is of flyschoid type and the upper part of The Tubinzurta Fm.
the group consists of typical carbonate platform sediments. (10) Siltstone member containing a bauxite bed.
(11) Variegated and black siltstones, breccia-like dolomites.
(12) Grey and black silicified dolomites.
Stratigraphy
Hiatus and palaeokarst
Diamictites form the lower part of the Bokson Gr. and uncon- The Khuzhirtai Fm.
formably overlie Tonian-Cryogenian? (Upper Riphean) ophiolites, (13) Dolomites passing into dark grey limestones with archae-
diabases, turbidites and conglomeratic breccias of the Dzhund- otheans, trilobites and brachiopods characteristic of the Atda-
zhugur Fm. or the sedimentary-volcanogenic Sarkhoy Gr. whose banian Stage of the Lower Cambrian (.500 m).

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 285– 288. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.23
286 N. M. CHUMAKOV

o
100

Irkutsk

52o Fig. 23.1. Geographic position of the


Bokson Gr. and its main sections. (a) Map
Baikal L. showing general location of study area
0 100 km
(box). (b) Geological map of described area.
1, granitoides; 2, Middle Cambrian–
Khubsugul L. Ordovician; 3, the Bokson Group, latest
(a)
100 o o Ediacaran (Upper Vendian) to Middle
104
52 o Cambrian; 4, Tonian-Cryogenian? (Upper
1 5
Riphean); 5, pre-Mesoproterozoic (pre-
Riphean) crystalline basement; 6, main
2 6
thrusts; 7, main faults; 8, boundary of the
3 7
Russian Federation and Mongolia; I,
4 8 0 10 20 km Bokson River Basin; II, Sarkhoy River
(b) Basin.

In the western part of the Bokson-Sarkhoy Synclinorium, along the horizon (Kuzmichev 2004). In the Sorok River basin and else-
Sarkhoy River (Fig. 23.1, II), the Zabit Fm. is separated from the where, the boundary between the underlying deposits and the dia-
Cryogenian basement by a thick (up to 500 m) red clastic mictites is gradational (Osokin & Tyzhinov 1998). The upper
sequence, the Khushatay Fm. of the Bokson Gr. Some researchers boundary of the diamictite member with overlying dolomites is
(Osokin & Tyzhinov 1998; Kuzmichev 2004) reported the pres- sharp but conformable and locally this boundary is gradational
ence of diamictites at the base and top of Khushatay Fm. In the (Osokin & Tyzhinov 1998; Kuzmichev 2004).
middle part of the Khushatay Fm., black clayey limestones and
shales up to 150 m thick occur (Kheraskova & Samygin 1992).
Sections of the Khushatay Fm change significantly over short dis-
Chemostratigraphy
tances. It is likely, that the basal diamictite member of the Zabit
Isotopic analyses (oxygen and carbon) of the Bokson Gr. carbonate
Fm. is coeval to the upper part of the Khushatay Fm.
rocks were carried out by B.G. Pokrovskiy and coauthors
(Pokrovskiy et al. 1999). Results obtained are illustrated in
Glaciogenic deposits and associated strata Figure 23.2. Carbonate samples were dissolved in H2PO4 at
25 8C for 1.5 h in the case of the calcitic fraction; dissolution
The diamictites of the Zabit and the Khushatay Fm. have a wide was made at 100 8C for 1 h for the dolomitic fraction. A correction
lateral distribution (more than 130  360 km2, taking into to isotopic oxygen values was introduced on different fractionation
account the presence of Bokson Gr. equivalents in the northern of calcite and dolomite (þ1.1‰ for the latter).
Mongolian Khubsugul area). Lithology of the diamictite The d13C range determined for calcite varied from 24.4 to
members of the Bokson Gr. was briefly described by Osokin & 24.9‰ in beds that underlie the diamictites; from 22.9 to 1.8‰
Tyzhinov (1998) and Kheraskov & Samygin (1992). These for the Zabit Fm. units that overlie the diamictites; from – 4 to
diamictites are massive, but sometimes crudely bedded. Thickness 1.5‰ for the Tabinzurt Fm.; from –2.1 to 0.3‰ for the Khuzhirtay
of the massive beds varies from 20 to 55 m. These are composed of Fm.; and from –2.1 to 0.7‰ for the lower part of the Nyurga Fm.
greenish and cherry-coloured shale or carbonate shale matrix with The isotopic d13C curve demonstrates distinct positive values cor-
scattered pebbles, and small and large boulders derived from rocks responding to the beds under the diamictites: three negative peaks
underlying the Bokson Gr.: sandstones, quartzites, jaspers, diabases for the lower carbonate part of the Zabit Fm.; an interval of mainly
and gabbro, as well as pebbles and boulders of sandstones and dolo- positive low values for the upper Zabit – lower Tabinzurt Fm. beds,
mites containing stromatolites or fossils of the algae Renalcis. a sharp drop in value for the Tabinzurt beds that overlie the bauxite
Lenses of dolomitic breccias occur locally. Dolomitic boulders member; change with some oscillations to almost zero for the
and breccias are confined to the sections where the Zabit Fm. is uppermost Tabinzurt Fm.; and low negative values for the
separated from the Cryogenian basement by the Khushatay Fm., Khuzhirtay Fm.
which includes dolomites. The diamictites contain erratic stones
similar to the granitoids of pre-Mesoproterozoic crystalline base- Other characteristics
ment that crop out on the Gargan Massif to the east. Semi-angular
and poorly rounded stones dominate over rounded ones. Some clast The member of dolomites with interbeds of microcrystalline dolo-
are of flat-iron shape, with flattened and striated faces. Stones are mitic phosphorite occurs in the upper part of the Zabit Fm. in the
dispersed irregularly and sometimes form ‘stone nests’. The dia- Bokson area. The thickness of this member varies from 35 to
mictites include rare conglomerate interbeds and sandstone lenses. 50 m. Total thickness of phosphorite beds is on average 5.5 m.
The massive diamictites are associated with lenses and sub- Average contents of P2O5 in this phosphorite is 8.3% (Yanshin
ordinate beds of thin and rhythmically bedded shales and mud- & Zharkov 1986).
stones containing scattered dropstones, which press through or The basal bauxite member of the Tabinzurt Fm. overlies a karst
break underlying beds and are themselves in turn draped by over- surface of the Zabit Fm. in the Bokson area (Kuzmichev 2004).
lying beds. Dolomites, which overlie the diamictites, are grey or This member is 4– 23 m thick and consists of massive, brecciated
black, massive or brecciated, and sometimes stromatolitic. and laminated reddish bauxite consisting of diaspore, boehmite
and a higher proportion of Fe2O3 and SiO2 (Il’ina 1958).
Boundary relations with overlying and underlying units
Palaeolatitude and palaeogeography
In most sections, the diamictites and basal conglomerates uncon-
formably overlie the Sarkhoy Gr. (Osokin & Tyzhinov 1998). In No valid results of palaeomagnetic investigations have been
the Bokson River basin their contact is marked by a weathering obtained and published for the Bokson Gr. and adjacent
BOKSON GROUP, EAST SAYAN MOUNTAINS 287

Fig. 23.2. Lithological and


chemostratigraphic succession of the lower
part of Bokson Gr. (from Pokrovskiy et al.
1999, with additions). 1, limestones; 2,
massive dolostones; 3, stromatolitic
dolostones; 4, dark dolostones, calcarenites
and calcirudites; 5, phosphatic calcarenites;
6, limey shales; 7, bauxites; 8, argillites and
siltstones; 9, diamictites; 10, eluvial debris;
11, silicification; 12a, calcite; 12b,
dolomite; ]1, Lower Cambrian; V2,
Ediacaran-Cambrian (Upper Vendian); R3,
Cryogenian (Upper Riphean) (Sorkhoy
Gr.); SS, Nemakit-Daldyn SSF (small
shelly fossils); Cl, Cloudina;
Mf, microfossils.

deposits. Following regional palaeotectonic reconstructions for Discussion


the late Cryogenian and Ediacaran (Early and Late Vendian),
the Tuva-Mongolian Massif and its constituents seems to be Sedimentary environments
in low latitudes of around 08 (Fedotova & Khain 2002) and
2308 (Smith 2001) respectively. A glacial origin of the Zabit diamictites is suggested by the wide
distribution of massive and bedded diamictites at the same strati-
graphic interval, the presence of erratic and ‘flat iron’ shaped
Geochronological constraints stones with striated faces and dropstones in thin-bedded rocks
(Osokin & Tyzhinov 1998). Osokin & Tyzhinov (1998) proposed
The Rb –Sr age of volcanic rocks in the Sorkhoy Gr. unconform- that the diamictites originated in glacial terrestrial, lacustrine and
ably underlying the Bokson Gr. has been estimated to be shelf environments. The opinion of these and other authors
718 + 30 Ma (erochrona, Buyakayte et al. 1989). The basal (Kuzmichev 2004 and others) about the glacial origin for the dia-
Khushatay Fm. of the Bokson Gr. has yielded the acritarch Gran- mictites seems correct. However, the presence of continental
omarginatasphaeria judomica Pjat. and multiserial sheaths of glacial deposits in the Bokson Gr. is doubtful. There are no signs
organic-walled microfossils Polytrichoides lineatus Hermann, of glaciodislocations, zones of assimilation or glacial pavements
suggestive of an Ediacaran (Vendian) age (Veis & Vorob’eva at the base of massive diamictites. The preferred orientation of
1993). Cloudina sp. (attributed by A. Y. Zhuravlev) was recovered elongated stones typically found in basal tillites does not occur
from granule of the diamictites (Kheraskova & Samygin 1992), in the Bokson diamictites. On the contrary, the stones are oriented
and the SSFs Cambrotubulus decurvatus Miss. and Anabarites tri- in a random fashion and often form ‘stone nests’. Gradual
sulcatus Miss., characteristic of the Nemakit-Daldynian (Many- transitions between alternating diamictites and thin-bedded sedi-
kaian) Stage of Siberia, were recovered 30 m above the ments and numerous indicators of underwater slumping suggest
diamictitic member (Terleev & Zadorozhnyi 1996; Postnikov & that the Bokson diamictites are glaciomarine deposits, partly
Terleev 2004). Higher up the section, the phosphoritic member reworked by slumps and debris flows, as noted by Kheraskova &
of the Zabit Fm. contains abundant assemblages of SSF (Cambro- Samygin (1992).
tubulus decurvatus Miss., Anabarites trisulcatus Miss., Tisitheca There are other interpretations of the Bokson Gr. diamictites.
sp., Igorella sp., Fomitchella infundibuliformis Miss.), silicified Some researchers suggest that these diamictites represent sedimen-
algae (Obruchevella magna Golov., O. parva Reitl., Heliconema tary olistostromes formed by reworking of Sarkhoy Gr. deposits
sp., Eosynechococcus meorei Hoffm. and other forms), radio- (Khomentovskiy et al. 1985). Other geologists consider the dia-
larians (Palaeocenosphaera parva Nazarov, Entasitinia sp.) and mictites to be tectonic olistostromes formed in front of the nappe
other fossils characteristic of the Lower Cambrian Tommotian where the Bokson Gr. overthrusts Palaeozoic deposits (Terleev
stage (Postnikov & Terleev 2004). & Zadorozhnyi 1996; Fedotova & Khain 2002; and others). This
288 N. M. CHUMAKOV

tectonic interpretation contradicts the fact that the Bokson Group References
has a gradational contact with the underlying beds in some areas,
convincingly demonstrated by regional mapping (Khomentovskiy Buyakayte, M. I., Kuzmichev, A. B. & Sokolov, D. D. 1989. 718
et al. 1985; Kheraskova & Samygin 1992; Kuzmichev 2004). million year – Rb –Sr erochrona Sarkhoy Group of East Sayana.
Another argument against tectonic olistostromes is the fact that Doklady Academy of Sciences of USSR, 309, 150–154.
the carbonate stones in the diamictites have high positive d13C, Chumakov, N. M. 1981. Upper Proterozoic glaciogenic rocks and their
similar only to carbonate rocks below the diamictites. Carbonate stratigraphic significance. Precambrian Research, 15, 373–396.
rocks of the Bokson Gr. that overlie the diamictites have only Chumakov, N. M. 2009a. The Baykonurian Glaciohorizon of the Late
negative d13C values. Vendian. Stratigraphy and Geological Correlation, 17, 373–381.
Osokin & Tyzhinov (1998) suggested a glacial origin for the Chumakov, N. M. 2009b. Neoproterozoic Glacial Events in Eurasia. In:
Gaucher, C., Sial, A. N., Halverson, G. P. & Frimmel, H. E.
carbonate breccia of the Zabit Fm. (see bed 6 in stratigraphy
(eds) Neoproterozoic-Cambrian Tectonics, Global Change and
section described above). They base this on scattered carbonate Evolution: a focus on southwestern Gondwana. Developments in
gravel and small pebbles that occur in some lenses of rhythmi- Precambrian Geology, 16, Elsevier, The Netherlands, 389– 403.
cally thin-bedded limestones and dolomites related with the brec- Chumakov, N. M. 2011. Glacial deposits of the Baykonur Formation,
cias. The description of the composite type section presented Kazakhstan and Kyrgyzstan. In: Arnaud, E., Halverson, G. P. &
above, notes that the breccias and enclosing deposits bear abun- Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
dant traces of underwater slumping, suggestive that breccias were Glaciations. Geological Society, London, Memoirs, 36, 303–307.
most likely products of slides and slumps on slopes of the car- Fedotova, A. A. & Khain, E. V. 2002. Tectonics of the south East Sayan
bonate platform. The rhythmically and thin bedded limestones and its position in the Ural– Mongolian Belt (in Russian). Nauchnyi
and dolomites appear to represent sediments deposited from Mir, Moscow.
related turbidity currents. A non-glacial origin is supported by Guan Baode, Wu Ruitang, Hambrey, M. J. & Geng Wu. 1986. Glacial
the absence of any foreign clasts in the breccias and sediments and erosional pavement near the Cambrian – Precambrian
thin-bedded rocks. boundary in western Henan Province, China. Journal of Geological
Society, London, 143, 311–323.
Il’ina, A. V. 1958. Geology and origin of Boxon bauxites in East Sayan.
Age In: Dolgopolov, N. N. (ed.) Bauxites their Mineralogenesis Origin
(in Russian). Academy of Sciences of USSR Press, Moscow, 267–281.
In terms of the framework proposed by the International Strati- Kheraskova, T. N. & Samygin, S. G. 1992. Tectonic environments of
graphic Committee, the diamictites of Bokson Fm. can be formation of Vendian-Middle Cambrian clastic-carbonates sequence
referred to the uppermost part of the Ediacaran or the lowermost of the East Sayan. Geotectonics, 6, 18 –36.
part of the Lower Cambrian (Nemakit-Daldyn Stage). In terms of Khomentovskiy, V. V., Pak, K. L., Postnikov, A. A. & Skopintsev, V. G.
the Russian stratigraphic scale, diamictites of the Bokson Fm. 1985. Geology structure of the basin of Ukha-Gol River (in Russian). In:
can be referred to the uppermost part of the Kotlin or the lower- Khomentovskiy, V. V. (ed.) Stratigraphy of the Late Precambrian and
most part of the Nemakit-Daldyn Stages of the Vendian. The dia- Lower Palaeozoic of Siberia. Riphean and Vendian. Institute of the
mictites of the Zabit Fm. occur above the Khushatai Fm., which Geology and Geophysics, Novosibirsk, 76–106.
contains microfossils, similar to those of the latest Ediacaran Kuzmichev, A. B. 2004. Tectonic History of the Tuvino-Mongolian
(Late Vendian) (Veis & Vorob’eva 1993). The diamictites contain Massif (in Russian). Probel-2000, Moscow.
Osokin, P. V. & Tyzhinov, A. V. 1998. Precambrian tilloids Oka-
Cloudina sp. (Kheraskova & Samygin 1992) and lie below beds
Khubsugul phosphorite basin (East Sayan, North-West Mongolia).
containing Cambrotubulus decurvatus Miss. and Anabarites tri-
Lithology and Mineral Resources, 2, 162–176.
sulcatus Miss. (Terleev & Zadorozhnyi 1996; Postnikov & Pokrovskiy, B. G. & Missarzhevskiy, V. V. 1993. Isotopic correlations
Terleev 2004), indicative of a Nemakit-Daldynian age. These boundary sections of Precambrian and Cambrian of Siberian
data together with the Tommotian fossils present higher in the Platform. Doklady Academii Nayk SSSR, 329, 768– 771.
Zabit Fm. allow the diamictites to be referred to the latest Edia- Pokrovskiy, B. G., Letnikova, E. F. & Samygin, S. G. 1999. Isotopic
caran (Late Vendian) (Kuzmichev 2004), and more specifically, stratigraphy of Boxon Group, Vendian-Cambrian of East Sayan.
either a slightly antedated Nemakit-Daldyn age or concurrent Stratigraphy and Geological Correlation, 7, 23 –41.
to its commencement (Chumakov 2009a, b). The Cloudina in Postnikov, A. A. & Terleev, A. A. 2004. Stratigraphy of Neoproterozoic
the diamictites was likely redeposited from underlying succes- of the Altay-Sayan folded region. Geology and Geophysics, 45,
sion. The lowermost negative d13C anomaly in the Zabit Fm. 295–309.
can be correlated to the negative anomaly of d13C at the base Smith, A. G. 2001. Palaeomagnetically and tectonically based global
of the Nemakit-Daldyn stage (Pokrovskiy & Missarzhevskiy maps for Vendian to Mid-Ordovician time. In: Zuravlev, A. Yu &
1993). On an interregional scale, it is possible to correlate the Riding, R. (eds) The Ecology of the Cambrian Radiation. Columbia
Bokson diamictites to the Baykonur Glacial Horizon in Kazakh- University Press, New York, 11– 46.
stan and Kyrgyzstan (Chumakov 1981, 2011) and to the Hankal- Terleev, A. A. & Zadorozhnyi, V. M. 1996. Discovery of Palaeozoic
chough and Luoguan formations of northern China (Baode et al. Foraminifera in ‘Precambrian’ of East Sayan (Sarkhoy River).
1986), based on their similar stratigraphic position. Doklady Academy of Sciences of Russia, 351, 373–374.
Veis, A. F. & Vorob’eva, N. G. 1993. First finds wall-organic microfossils
Research projects were supported by grants of the Russian Fund of Basic Inves- in Upper Precambrian of Boxon-Sarkhoy basin (East Sayan). Strati-
tigations No. 11-05-00232 and Program No. 25 of the Presidium of the Russian graphy and Geological Correlation, 1, 27 –32.
Academy of the Sciences. The author would like to acknowledge stimulating Volkov, R. I., Zaytsev, N. S., Il’in, A. V. & Osokin, P. V. 1972.
discussions arising from participation in IGCP project 512 ‘Neoproterozoic Ice Ukhangol phosphorite field of East Sayan. Soviet Geology, 2,
Ages’ (http://groups.google.com/IGCP-512?nl¼enIGCP-512@?hl¼en). This 94 – 107.
represents a contribution of the IUGS- and UNESCO-funded IGCP (International Yanshin, A. L. & Zharkov, M. A. 1986. Phosphorus and Potassium in
Geoscience Programme) Project #512. Nature (in Russian). Nauka, Novosibirsk.
Chapter 24

The Neoproterozoic glacial formations of the North and Middle Urals

NICKOLAY M. CHUMAKOV
Geological Institute Russian Academy of Sciences, Pyzhevskiy per. 7, Moscow 109017, Russia (e-mail: chumakov@ginras.ru)

Abstract: In the North and Middle Urals, four Neoproterozoic diamictite-dominating units are known: the Churochnaya, Tany and
Koyva formations and Lower Starye Pechi Subformation. A glacial origin for the diamictites is indicated by a number of characteristic
features such as erratic stones, striated and faceted clasts, shales with dropstones, preferential orientation of elongated stones, the wide
distribution and confinement to certain stratigraphic levels of the deposits, and the association with distinct post-glacial dolostones. Most
diamictites were deposited in a glaciomarine environment and sourced from an ice sheet on the East European craton. The Churochnaya
Formation also contains subordinate terrestrial and probably seasonal sea ice deposits. This formation accumulated on a glaciated con-
tinental shelf, while the Tany, Koyva and Starye Pechi formations were deposited on the outer shelf and continental slope of the eastern
margin of the East European Craton. The Middle Ural sections contain the White Sea (Ediacaran) Metazoa assemblage, typical late
Neoproterozoic microfossils, stromatolite associations, and radiometric dates obtained from volcanic tuffs, granosyenites, trachytes
and trachyandesites. The combined data suggest that the Lower Starye Pechi Subformation was deposited in the middle of the Ediacaran
Period (Early Vendian). The Churochnaya Formation of the North Urals and the Tany and Koyva formations of the Middle Urals were
likely deposited in the late Cryogenian period (Early Vendian).

Neoproterozoic glacial deposits are known in the western parts of three times thicker than those of the Polyudov Ridge and
the North, Middle and South Urals. They are best exposed and contain a higher percentage of siliciclastics in the form of turbi-
studied along the Pulyudov Ridge and Churochnaya River in the dites and shales (i.e. flysch). The Neoproterozoic deposits of the
northern Urals and along tributaries of the Chusovaya River in middle Urals, along with younger deposits, were locally heavily
the Middle Urals (Fig. 24.1). This chapter focuses on glacial deformed during the Hercynian, with isoclinal folds and numer-
deposits of the North and Middle Urals; glacial deposits of the ous faults and overthrusts. The Tany, Koyva and Starye Pechi
South Urals were recently described and are reviewed elsewhere formations were metamorphosed to lower greenschists facies
(Chumakov 1981a, b, c, 1992, 1998; Maslov 2000). at this time. The Wil’va Formation was metamorphosed to
Until the mid-1960s, the late Precambrian strata of the North and greenschist facies.
Middle Urals were regarded as Palaeozoic in age and the occur-
rence of associated ‘tillite-like conglomerates’ (diamictites) was
explained by subaqueous slumping. This view was reflected in Stratigraphy
many publications and geological maps of the Urals and USSR
territory. Long regional investigations resulted in the elaboration Neoproterozoic stratigraphy, lithologies and pertinent radiometric
of the late Precambrian timescale as reviewed by Ablizin et al. ages of the northern and middle Urals under consideration are
(1982). Modern stratigraphic investigations of Neoproterozoic schematically shown in Figure 24.2. The Polyudov Ridge section
sedimentary sequences in the Urals, including glacial deposits, begins with the thick (800 –900 m) siliciclastic Rassolny For-
began in the second half of the twentieth century during lithologi- mation (not shown in Fig. 24.2), which is overlain by pink lime-
cal investigations (the Polyudov Ridge; Borovko 1967) and stones and marls of the Deminskaya Formation. Above is the
geological mapping (the Middle Urals, Ablizin et al. 1982) that Niz’va Formation, comprising mainly calcareous dolomites with
was verified by Ehlakov & Morozov (2006). stromatolites similar to those of the late Mesoproterozoic –early
Glacial deposits of the Polyudov Ridge in the North Urals are Neoproterozoic (Upper Riphean) Min’yar and Uk formations of
confined to the Churochnaya Formation. Three glacial levels the South Urals (Raaben 2007). The Niz’va Formation is uncon-
were established in the Middle Urals: the Tany and Koyva formably overlain by the clastic Ust-Churochnaya Formation,
formations of the Serebryanka Group and in the Lower Starye which, in turn, is unconformably overlain by the Churochnaya For-
Pechi Subformation of the overlying Sylvitsa Group (Fig. 24.2). mation. The latter consists of red sandstones in the lower half and
The Tany and Koyva formations are replaced eastward by the diamictites with an interval of dolomite and black shales in the
Wil’va Formation, which contains some diamictite beds. The upper part. The black shales are overlain by a thick sandy-shaly
possible occurrence of diamictites in the Us’va, Fedotovka and sequence, which is divided by an unconformity into the Il’ya-Vosh
Kernos formations has not been confirmed by recent investi- and Kocheshor formations.
gations. The Us’va and Fedotovka shales locally contain single The late Precambrian succession of the Middle Urals is more
lonestones. complicated. Sequences that are approximately coeval with those
of the Polyudov Ridge are considerably thicker, more lithologi-
cally diverse, and divisible into four groups, each separated by a
Structural framework depositional hiatus. Each group is further subdivided into for-
mations as shown in Figure 24.2.
The Neoproterozoic deposits of the Polyudov Ridge accumu-
lated in the marginal shelf zone of the Russian platform and
deformed during the Timanian and, more significantly, during Glaciogenic deposits and associated strata
the Hercynian orogeny. Now they are exposed in the cores of
marginal brachyanticlines of the Uralian folded zone, disrupted The Churochnaya Formation
by many faults, and variably foliated. In the Middle Urals,
Neoproterozoic sedimentation occurred in the outer shelf and This formation is 350–500 m thick. Its lower part (c. 150 m)
continental slope environments. Sections of this region are consists of red sandstones. The upper part of the formation

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 289– 296. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.24
290 N. M. CHUMAKOV

o o
56 58

R.
ra
Po

she
ly
d

Vi
ov
Moscow Perm

Ri
Krasnovishersk

dg
e
Ka
m
aR
.

o
60

Phanerozoic
Solikamck Sylvitsa Gr., Kocheshor,
Il’ay-Vozh Fms
Serebryanka Gr., Churochnzya,
Ust-Churochnaya Fms
Riphean

main disjunctive
dislocations
Main Neoproterozoic sections
.
aR Mezhevaya Utka R.
’v
Kos
. Serebryanka R.
va R
Us’ Sylvitsa R.
Koyva R.
Vil’va R.
Us’va R.
R. Churochnaya R.
ovaya
C hus
58 o
Perm

0 25 50 km Fig. 24.1. Late Mesoproterozoic to


Neoproterozoic (Vendian and Upper
Riphean deposits of the North and Middle
Urals).

(c. 200–350 m) is composed of massive and bedded green-grey and encrust fragments of breccia. Thin, intercalated black chert
or reddish diamictites containing some beds of sandstones, (2– 3 cm) overlies the dolostone. The chert is followed by
shales and conglomerates (Fig. 24.3). The diamictites are over- .20 m of black shale with a relatively high phosphorous content
lain by the dolomite member (10 m) and above it by black (0.1 –1%). The black shales are overlain by multicolour shales
shales (.20 m). The diamictites have a sand-silty matrix with with sandstone intercalations.
dispersed stones (2 –10% of the diamictite volume) of variable
size, roundness and composition. Some diamictites show indis-
Boundary relations with overlying and underlying non-glacial
tinct bedding, thin lenses and interbeds of sandstones and
units. Conglomerates at the base of the Churochnaya Formation
shales, and nested clasts. In addition to the predominant sand-
rest on a probable unconformity. Breccias in the upper part of
stone and quartzite, there are many erratic stones, including
the cap dolomite are apparently related to weathering and karstifi-
granites, gneisses and schists derived from the crystalline base-
cation. Nevertheless, the hiatus between the dolomite and over-
ment of the Russian plate. The proportion of erratic stones
lying chert was probably not significant, as a 1.5-m-thick
increases upsection. Between 3 and 7% of stones are striated,
diamictite bed occurs 12 m above the base of the black
grooved and faceted. Rare diamictites beds contain a large pro-
shale member.
portion of well-rounded stones of medium size. Some diamic-
tites preserved a preferred orientation among elongated stones.
Rhythmically banded shales with dropstones occur at the base Chemostratigraphy. Dolomites and limestones of the Niz’va
of the diamictites. Formation, which unconformably underlies the Churochnaya
The diamictites are overlain by a dolomite member (up to 10 m). Formation, are characterized by slightly variable but on
This dolostone is pale, slightly calcareous (1 –7% CaCO3), distinct average low, positive d13C values ( –0.8 to þ2.5‰) and stable
in structure and texture, and commonly massive or brecciated. In d18O values ( –10 to – 6‰). Dolomites overlying the diamictites
places it is thinly bedded and distorted by subaqueous slumps. of the Churochnaya Formation have stable negative d13C values
Abundant carbonate, chert and quartz veins cut the dolomite (from –3 to – 5‰).
NEOPROTEROZOIC OF THE NORTH AND MIDDLE URALS 291

Polyudov Ridge Central Ural


O3 O3

. .. .. .. .. .
. .. .. .. .. . Ust’-Sylvitsa Fm. . .. .. .. .. . diamictite
. .. .. .. .. .
350 m .....
..... . .. .. .. .. . . .. .. .. .. .

Sylvitsa Gr.
Kocheshor Fm. Chernyi Kamen’ Fm. ..... ..... sandstone
500 m .. .. .. .. .. 557 ± 13 U-Pb, z 1700 m .. .. .. .. ..
..... shales
545 K-Ar, g Perevolok Fm.
570 K-Ar, g ..... 300 m dropstone
..... .....
..... Starye Pechi Fm. .....
.....
limestone
300 m
.....
Kernos Fm. ..... dolstone
Il’ay-Vozh Fm. 300 m .....
.....
480 m .. .. .. .. .. Buton Fm. volcanic rock
610; 620 K-Ar, g
Serebrayanka Gr.

400 m
cap dolomite
Koyva Fm.
Churochnaya Fm. 700 m
stromatolite
350 m 569 ± 42 Sm; 559 ± 16 Rb
. .. .. .. .. .
..... .....
Garevka Fm. .....
200-700 m .....

Tany Fm. .....


.....

800 m
.....
..... Us’va Fm. .....
Ust’- . .. .. .. .. .
1200 m . . . . .
Churochnaya Fm. .....
.....
..... Fedotovka Fm.
Basegi Gr.

720 m . . . . .
1200 m
635 K-Ar, g .....
. .. .. .. .. .
Shcherogrovitsy Fm.
0-900 m
671+24 Rb; 671+24 U-Pb, z
.....
Oslyanka Fm. .....
300 m .....
Kedrovka Gr.

Niz’va Fm. Klyktan Fm.


950 m Fig. 24.2. Stratigraphic position of Vendian
750 K-Ar, g 1300 m
..... glacials at the North and Middle Urals
Deminskaya Fm. .....
Sinii Gory Fm. ..... (Ablizin et al. 1982; Chumakov 1998;
300 m 1000 m ..... Maslov et al. 2007). Radiometric dating
techniques: g, glauconite; z, zircon.

Tany Formation Boundary relations with overlying and underlying non-glacial


units. The lower boundary of the Tany Formation is probably
The Tany Formation (up to 800 m thick) consists of two members erosional judging from the abundance of clasts from the underlying
of massive dark grey diamictites divided by a sandstone member. Kedrovka and Basegi groups within the basal diamictites.
The base of the thick Lower Member (c. 350–450 m) consists of The upper boundary of the Tany Formation is sharp and
alternating mafic volcanics, schists, limestones and diamictites conformable.
(Ablizin et al. 1982). The lower member of diamictites in some
section is capped by a dolomite bed 8 m thick (Fig. 24.4). The
lithology of diamictites clasts is diverse and includes quartz are- Koyva Formation
ntites, granites, gneisses and carbonates. A considerable portion
(up to 45%) of the diamictite clasts are felsic igneous rocks. The The Koyva Formation contains diamictites only to the north of
abundance and size of granites and gneiss clasts increase to the Sylvitsa River. In general, diamictites are similar to those of
the SW (towards the Russian platform) and in the upper part of the Tany Formation but differ in their reddish colour and con-
the lower member (Ablizin et al. 1982). Some blocks of granite finement of bedded diamictites and laminated shales with drop-
and gneiss are up to 3.5 m in diameter. Nests of medium-sized stones to the upper part of the formation. Father to the north, at
and large stones are common. Some rare shale interbeds contain the Vil’va, Us’va and Kos’va rivers, the Koyva Formation con-
dropstones. The Middle Member consists of bedded quartz- tains alkali basalts flows. In the stratotype section at the Koyva
feldspathic sandstones (80– 150 m). The Upper Member includes River, diamictites are overlain by a transitional member of thin-
massive and bedded diamictites and thin intercalations of lami- bedded variegated shales with small dropstones and then by the
nated shales with dropstones (Chumakov 1992). dolomite member (up to 6 m thick; Fig. 24.5). Thin (1– 2 cm)
292 N. M. CHUMAKOV

Lower Starye Pechi Subformation


massive diamictites
In the Lower Starye Pechi Subformation, diamictites alternate with
sedimentary and quartzites stones
sandstones and are overlain by thin-bedded shales containing scat-
granitoid and gneisses stones tered sand grains and pebble dropstones (Fig. 24.6). The matrix of
the diamictites is dark grey sandy-silt, and the diamictites are
diamictites with well rounded stones bedded in the upper part of the formaton. Clasts in the diamictite
sandstones with pebbles vary in size and shape; they are dominated by fragments of the
underlying rocks, but some erratic pebbles of quartz, quartzites
sandstones and plagio-granites occur as well. Striated and grooved clasts
have been found. Diamictite wedges are found at the base of
shales
some of the diamictite beds.
dolomites
Boundary relations with overlying and underlying non-glacial
chert units. The contact between the Lower Starye Pechi Subformation
black rocks and the underlying Kernos Formation is generally assumed to be an
unconformity (Ablizin et al. 1982). Nevertheless, locally the lower
brown-red rocks Starye Pechi Formation contains interbedded sandstones, which
striated stones
are typical of the Kernos Formation, and thus imply a transitional
contact (Chumakov 1998). The upper boundary of the Starye Pechi
dropstones Subformation is transitional.
stone fabric
ch .. stone nests
.. Geochronological constraints
== vague bedding
rhythmic lamination Polyudov Ridge
==
subaqueous slump structure No precise radiometric data are available for the Neoproterozoic
formations of the Polyudov Ridge. Numerous radiometric ages
higher content of phosphorus
were obtained in the middle of last centenary by K –Ar dating
of glauconites (Garris et al. 1964; Borovko 1967; Keller &
Chumakov 1983). They show a rather wide range of ages. The
average K –Ar ages recalculated with recent 40K decay constants
== (l40Kb- ¼ 4.962  10210 a21; l40Ke- ¼ 0.581  10210 a21) are
30 750 + 25 Ma for the Niz’va Fm., 660 + 30 Ma for the Ust’-
Churochnaya Fm., 610 + 20 Ma for the Il’ay-Vozh Fm., and
== 570 + 17 Ma for the lower and 545 + 15 Ma for the upper part
.. of the Kocheshor Formation, respectively.
The Niz’va Fm. contains abundant stromatolites. Their assem-
blage consists of Poludia mutabilis Raab., Gymnosolen ramsayi
Steinm., Conophyton miloradovichi Raab., Poludia polimorpha
Raab., Minjaria uralica Kryl., Linella ukka Kryl., and other
0 m forms typical of the Cryogenian (uppermost Upper Riphean) of
the Middle and South Urals, the Timans and Svalbard (Raaben
2007). In general, K –Ar ages from the Niv’zva Formation are
Fig. 24.3. Representative stratigraphic section of the upper Ghuroghnaya consistent with the chronology implied by stromatolite
Formation, Ghurochnaya River. biostratigraphy.

interbeds of dolomites occur in rhythmically laminated shales Middle Urals


overlying the dolomite.
Abundant non-skeletal Metazoa (Cyclomedusa davdi Sprigg,
Dickinsonia tenuis Glaessner and Wade, Irridinitus multiradiatus
Boundary relations with overlying and underlying non-glacial uni-
ts. The lower and upper boundaries of the Koyva Formation are
Fedonkin, Vaizitsinia sophia Sokolov and Fedonkin and other
transitional (Ablizin et al. 1982). forms) characteristic of the White Sea and Eriacaran biota occur
in the Chernyi Kamen’ Formation in the upper Sylvitsa Group
(Grazhdankin et al. 2005, 2007; Maslov et al. 2007). A U –Pb
Chemostratigraphy. In the stratotype section at the Koyva River the (SHRIMP-II) zircon age of 557 + 13 Ma was obtained on a tuff
upper dolomite member yielded d13C values from –3 to –5‰. within the Chernyi Kamen’ Formation (Maslov et al. 2007). This
South of the Koyva River, where there are no diamictites or age provides a maximum constraint for all of the Neoproterozoic
shales with dropstones within the Kyva Formation, the basal bed diamictites of the Middle Urals.
of the upper dolomite member is 2 m thick and has d13C values More controversial dates were obtained for the individual
from – 4 to –7‰ (A.V. Maslov, pers. comm.). glacial formations of the succession using less reliable radiometric
techniques. The diamictites of the Starey Pechi Subformation seem
to be younger than 569 + 42 or 559 + 16 Ma based on Sm– Nd
Other characteristics. At the Us’va River, the lower part of the (clinopyroxene) and Rb – Sr (whole rock) ages from the Koyva tra-
Koyva Formation contains alkali basalts, tuffs and banded hemati- chyandesites (Petrov et al. 2005). The Sm –Nd age is imprecise,
tic rocks and ores, the latter of which vary from 10 to 40 m thick. whereas the Rb – Sr age is similar to an age of 557 + 13 Ma
NEOPROTEROZOIC OF THE NORTH AND MIDDLE URALS 293

Mezhevaya Utka R. Serebryanka R. Us’va R.


Garevka Fm.

3
dimictites,
crystalline clasts
dimictites,
sedimentary clasts
2
bedded diamictites
laminated shales
with dropstones
Tany Fm.

sandstones

shales
1
limestones

dolomites

striated stones
dropstones

rhythmic lamination
0m bedding
50

100

150

Us’va Fm. Fig. 24.4. Stratigraphic sections of the Tany


Formation.

obtained by Maslov et al. from 1.5 km upsection (Sylvitsa at the latitude of around 608S between 630 and 600 Ma, then
Group). Macrofossil associations lower in the Sylvitsa Group are displaced to 208S between 600 and 550 Ma.
similar to the Miaohe biota of the uppermost Doushantou For-
mation of South China (Grazhdankin et al. 2007), implying a
late Ediacaran age. Discussion
It was long accepted that grano-syenites of the Troitsk massif
(the Kos’va River) cut the Tany and Garevka formations of the Sedimentary environments
Serebryanka Group and that they were emplaced between depo-
sition of the Serebryanka and Sylvitsa groups. This hypothesis Massive and bedded diamictites contain shale interbeds with drop-
implies that the diamictites of the Tany Formation would be stones, clasts with typical glacier striae, grooves and facets, erratic
older than 630 + 20 (Pb/Pb zircon) or 621 + 12 Ma (Rb –Sr stones (including very large ones), and nests of stones. They are
whole rock; Petrov et al. 2005). More recent mapping and overlain by carbonates resembling typical Neoproterozoic ‘cap
dating, however, suggest that the Troitsk grano-syenites cut only dolomites’. In combination, these features suggest that most dia-
the much older Shechegrovittsy Formation (Basegi Group) and mictites of the Tany and Koyva formations and the Lower Starye
have an age of 671 + 24 Ma (SHRIMP U –Pb zircon)(Ronkin Pechi Subformation are glaciomarine deposits. The same con-
et al. 2007). The differences between the ages may be related to clusion can be drawn for most diamictites of the Churochnaya
two phases of magmatic activity in the Troitsk massif. In any Formation, except for members characterized by till-like fabric
case, the diamictites of the Koyva and Tany formations are indicated by preferred orientations in elongated stones and the
constrained to between 557 + 13 and 671 + 24 Ma, which is con- occurrence of varved shales, both of which suggest a continental
sistent with biostratigraphic data. origin. Some other members of the Churochnaya Formation
contain mainly well-rounded clasts that may have been deposited
by seasonal ice.
Palaeolatitudes and palaeogeography The Niz’va, Ust-churochnaya and Churochnaya formations
were deposited in inner shelf environments on the margin of the
Palaeomagnetic investigations of the late Proterozoic rocks of the Russian Plate. The flysch-like nature of sediments lying under,
South Urals have failed to reveal primary palaeomagnetic direc- between and over the diamictites of the Serebryanka Group and
tions. According to palaeogeographic reconstructions by Li et al. the Starye Pechi Formation show that they were deposited on the
(2008), the eastern margin of the Russian Platform was located outer shelf to continental slope.
294 N. M. CHUMAKOV

Possible correlations
m
The glaciogenic deposits of Starye Pechi Formation are older
diamictites than those of the Chernyi Kamen’ Formation (which contain
Edacarian Metazoa and a tuff dated as 557 + 13 Ma; Maslov
bedded diamictites et al. 2007). The Lower Starye Pechi Subformation may
therefore be correlated tentatively to middle Ediacaran glacial
sandstones horizons of the East European craton (the Glussk and Mor-
tensnes formations) and the glacial Gaskiers Formation in
70 Newfoundland.
multicolour shales The Koyva Formation shares similar stratigraphy and facies
with the Tany Formation, from which it is separated by a turbidite
black shales sequence (the Garevka Formation) that is noticeably reduced in
thickness in some places and pinches out near the Koyva River.
dolomites Therefore, the Koyva, Garevka and Tany formations were likely
deposited during a single glacial epoch. This composite unit is
very large boulders slightly older than Buton Formation containing Ediacaran Obru-
chevella (Golovenok et al. 1989) and much younger than the
Shchegrovitsy Formation and Troitsk grano-syenites massif
raft of diamictites (671 + 24 Ma). These are tentatively correlated to the lower
glacial horizon of the Lower Vendian (upper Cryogenian) on the
. .. stone nests East European craton (the Blon’ and Smalfjord formations), the
glacial Nantuo Formation in South China, and the Yerelina Sub-
26 dropstones group of South Australia.
The age and correlation of the Churochnaya Formation is
lamination subject to debate. The overlying Il’ya-Vozh and Kocheshor for-
mations are separated by an erosional unconformity (Fig. 24.2),
and it is possible that the Kocheshor Formation is equivalent to
rhythmic lamination the lower Sylvitsa Goup. In this case, the Churochnaya and Il’ya-
Vozh formations together may correlate with the upper Sereb-
6 ryanka Group (Maslov 2004). It follows from this hypothesis
that the Churochnaya Formation may be coeval with the glacial
Koyva Formation (Fig. 24.2). The traditional correlation of the
Niz’va Formation to the middle Neoproterozoic of the Karatavian
(Upper Riphean) stratotype of the South Urals and the Klytkan
20 Formation of the Middle Urals is supported by similar stromato-
lites assemblages (Raaben 2007) and isotopic data (d13C values
from –0.5 to 2.0‰ for the Niz’va Formation).

6 Regional palaeogeography

The lithology of diamictite clasts suggests that crystalline base-


ment and sedimentary cover of the Russian plate were the
main sources for the Churochnaya, Tany and Koyva formations.
Local sources for the diamictite clasts have also been estab-
50 lished, particularly for the Lower Starye Pechi Subformation
(Ablizin et al. 1982) and Vil’va Formation in the eastern
Urals (Suslov & Teterin 1997). The regional palaeogeographical
reconstructions (Chumakov & Sergeev 2004) suggest that an
inland ice sheet existed during Tany time on the eastern part of
the Russian Plate, whereas a marginal belt of shelf glaciers lay
to the east of it (Churochnaya, Tany and Koyva formations).
. .. Further east, sedimentation was dominated by turbidity currents
and ice rafting in an outer shelf –slope environment (Tany,
Koyva and Wil’va formation). The existence of extensive inland
glacier sheets in the northeastern part of East European Craton
300 is also evident in a number of other Early Vendian (late Cryogen-
ian) glacial units on the craton. The most complete late Cryogenian
glaciogenic sequences are located in the marginal parts of the
... East European craton and are further described elsewhere in this
volume. Within the cratonic interior, glacial sedimentation
occurred in aulocogens that developed in the middle-Late Riphean
(late Mesoproterozoic– Neoproterozoic) (Bessonova & Chumakov
1969; Chumakov 1971, 1981b, c, 1992).
Fig. 24.5. Stratigraphic section of the upper Koyva Formation, Koyva River.
Our research was supported by grant 11-05-00232 from the Russian Fund of Basic
Investigations and Program No 25 of the Presidium of the Russian Academy of
NEOPROTEROZOIC OF THE NORTH AND MIDDLE URALS 295

m diamictites
bedded diamictites

sedimentary and quartzite stones

Sp2 30 granitoid stones

sandstones with small pebbles

calcareous sandstones
massive quartz-feldspathic sandstones
Sp 4 shales
1,5
shales with dropstones
5,5
dropstones
. ..
rhythmic lamination
Sp1 11
vague bedding
. .. till wedges
. .. stone nests
6

10
20 m

6
10
5

Kr
0
15
Fig. 24.6. Stratigraphic section of lower
Starye Pechi Formation, Sylvitsa River. Sp,
Starye Pechi Formation; Kr, Kernos
Formation.

Sciences. The authors would like to acknowledge stimulating discussions arising Chumakov, N. M. 1981b. Late Precambrian glacial deposits of
from participation in IGCP project 512 ‘Neoproterozoic Ice Ages.’ the Vilchitsy Formation of western regions of the USSR.
In: Hambrey, M. & Harland, W. B. (eds) Earth’s Pre-
Pleistocene Glacial Record. Cambridge University Press, Cam-
References bridge, 655– 659.
Chumakov, N. M. 1981c. Late Precambrian glacial deposits of the Blon
Ablizin, B. D., Klyuzhina, M. L., Kurbatskaya, F. A. & Kurbatskiy, Formation, Belorussia, USSR. In: Hambrey, M. & Harland,
A. M. 1982. Upper Riphean and Vendian of the West Slope of the W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge
Middle Ural (in Russian). Nayka, Moscow. University Press, Cambridge, 660–662.
Bessonova, V. Ya. & Chumakov, N. M. 1969. Upper Precambrian Chumakov, N. M. 1992. The Problems of Old Glaciations (Pre-
glacial deposits of western regions of USSR. Lithology and Mineral Pleistocene Glaciogeology in the USSR). Harwood Academic
Resources, 2, 73– 89. Publishers, Pennsylvania.
Borovko, N. G. 1967. Vendian and Lower Paleozoic of Polyudov Ridge Chumakov, N. M. 1998. The key section of Vendian glacial deposits in
of the North Urals (in Russian). VSEGEI, Leningrad. the South Urals (Kurgashly Formation, Krivoluksky graben) in
Chumakov, N. M. 1971. Vendian glaciation of the Europe and North Russian. In: Knipper, A. L., Kurenkov, C. A. & Semikhatov,
Atlantic. Doklady Academii Nauk SSSR, 198, 419– 422. M. A. (eds) The Urals: Fundamental Problems of Geodynamics
Chumakov, N. M. 1981a. Late Precambrian Kurgashglya tilloids, and Stratigraphy. Nauka, Moscow, 138 –153 (in Russian).
soutern Urals. In: Hambrey, M. & Harland, W. B. (eds) Earth’s Chumakov, N. M. & Sergeev, V. N. 2004. Problems of climatic zonality
Pre-Pleistocene Glacial Record. Cambridge University Press, of the Late Precambrian. Climate and biotic events (in Russian)
Cambridge, 674–677. In: Semikhatov, M. A. & Chumakov, N. M. (eds) Climate During
296 N. M. CHUMAKOV

the Epochs of Principal Biosphere Rearrangements. Transactions of Maslov, A. V. 2004. Riphean and Vendian sedimentary sequences of the
Geological Institute of RAS, 550, 271. Timanides and Uralides, the eastern periphery of the East European
Ehlakov, Yu. A. & Morozov, G. G. 2006. Stratigraphy (in Russian). Craton. In: Gee, D. G. & Pease, V. L. (eds) The Neoproterozoic
In: Kudryashov, A. I. (ed.) Mineral-Stuff Resources of Permian Timanide Orogen of Eastern Baltica. Geological Society, London,
territory. Knizhnaya Ploshchad’, Perm, 49– 63. Memoirs, 30, 19– 35.
Garris, M. A., Kasakov, G. A. & Keller, B. M. 1964. Geochronological Maslov, A. V., Grazhdankin, D. V. et al. 2007. U –Pb (SRIMP II) age
scale of Upper Proterozoic (Riphean and Vendian). In: Absolute Age of the zircons from ash tuffs of the Chernyi Kamen’ Formation,
of Geological Formations. Nauka, Moscow, 431– 455 (in Russian). Sylvitsa Group of Vendian (Middle Ural). Doklady Russian
Golovenok, V. K., Belova, M. Iu. & Kurbatskaia, F. A. 1989. First Academy of Sciences, Geosciences, 411, 354– 359.
find of Obruchevella Reitlinger in Vendian sediments of Middle Petrov, G. A., Maslov, A. B. & Ronkin, Yu. L. 2005. Pre-Paleozoic
Urals. Doklady Academii Nauk SSSR, 309, 701–705. magmatic complexes of the Kvarkush-Kamennogorsk anticlinorium
Grazhdankin, D. V., Maslov, A. V., Mustill, T. M. R. & Krupenin, (the Middle Ural): new data on the geochemistry and the geody-
M. T. 2005. The Ediacaran White Sea biota in the Central Urals. namics (in Russian). Lithosphere, 4, 42 –69.
Doklady Earth Sciences, 401, 382– 385. Raaben, M. E. 2007. Stromatolitic formations of the Riphean of East-
Grazhdankin, D. V., Nagovitsin, K. E. & Maslov, A. V. 2007. Late European plateform. Stratigraphy and Geological Correlation, 1,
Vendian Miaohe-type ecological assemblage of the East European 35 – 46.
platform. Doklady Earth Sciences, 417, 1183–1187. Ronkin, Yu. L., Maslov, A. V., Petrov, G. A., Matukov, D. I. &
Keller, B. M. & Chumakov, N. M. (eds) 1983. Stratotype of Riphean. Suslov, S. B. 2007. In situ U– Pb (SRIMP II) date of zircons from
Stratigraphy. Geochronology. Nauka, Moscow (in Russian). granosyenites of Tritsk Massiv (Kvarkush-Kamennogorsk Meganti-
Li, Z. X., Bogdanova, S. V. et al. 2008 Assembly, configuration, and clinorium, Middle Ural). Doklady Russian Academy of Sciences,
break-up history of Rodinia: a synthesis. Precambrian Research, Geosciences, 412, 87 –92.
160, 179– 210. Suslov, S. B. & Teterin, I. P. 1997. On the origin ‘exotic’ stones of
Maslov, A. V. 2000. Some features of sedimentation in Early Vendian ‘tillite like’ conglomerates of Upper Wil’va Subformation. In: Ibla-
at the South and Middle Ural. Lithology and Mineral Resources, minov, R. G. (ed.) Geology and Mineral Resources of the Western
6, 624–639. Ural. Permian State University, Perm, 30 – 31 (in Russian).
Chapter 25

Glacial deposits of the Nichatka Formation, Chara River basin and review
of Upper Precambrian diamictites of Central Siberia

NICKOLAY M. CHUMAKOV
Geological Institute of the Russian Academy of Sciences, Pyzhevsky per. 7, Moscow 119017, Russia (e-mail: chumakov@ginras.ru)

Abstract: Outcrops of the Nichatka Formation are located in the southern centroclinal part of the marginal Berezovskaya basin of the
Siberian craton. Deposits of the gently sloping southeastern limb of the asymmetrical basin are unaltered, whereas those of the steep,
partly overturned western limb exhibit cleavage and are locally slightly metamorphosed. In the Late Precambrian the Berezovskaya
basin was a shallow marine area of the passive margin of the Siberian craton. The Nichatka Formation is the basal unit of the Dal’nyaya
Tayga Group of the Patom Supergroup. On the southeastern limb of the Berezovskaya Basin, the formation is composed of massive and
bedded diamictites, conglomerates, sandstones and thin-bedded rocks with dropstones. The diamictites contain erratic blocks of different
size and roundness. The formation thins out eastward. On the western limb, the formation is represented predominantly by graded-bedded
and less frequently massive diamictites, sandstones and rare thin bedded members with dropstones. The diamictites contain clasts with
glacial facets and striae. The formation unconformably lies on the Ballaganakh Group of the Patom Supergroup or older rocks and grades
into a marker member of thin-bedded red dolomitic marls and dolomites. Overlying carbonate-terrigenous formations of the Patom
Supergroup are characterized by two negative (down to –6 and – 10.5‰) d13C anomalies and one positive (up to 3.9‰) anomaly.
87
Sr/86Sr ratios in the carbonates above the Nichatka Formation increase upsection from 0.70725 in the overlying deposits to
0.70837 in the Nemakit-Daldyn (Fortunian) horizon of the early Cambrian. The Nichatka Formation as well as the host deposits are
assigned to the Lower Vendian (Late Cryogenian) by reliable geological and isotopic correlations to biostratigraphically better
studied Upper Precambrian sequences of the Ura Uplift in the Lena River basin. Stratigraphically similar glacial units were traced
along the margins of the Baikal-Patom Highland up to the southern end of Lake Baikal and further to the NW towards East Sayan.
The Siberian Craton was the main source area for these glacial units, which points to the existence of the great inland glaciers at the
southern part of this craton. Some older Upper Precambrian diamictites of unclear age and genesis are also recorded in Central Siberia.

Diamictites, sandstones and conglomerates, which lie unconform- carbonate platforms. Sedimentation was cyclic. The Nichatka For-
ably on Lower Proterozoic–Archaean granites and quartzites in mation is a basal part of the second Dal’naya Taiga cycle. During
the Lake Nichatka area, are named the Nichatka Formation sedimentation of the Patom Supergroup, the basin transgressed
(Zhuravleva et al. 1959). From this area the formation extends to over the craton. This resulted in overlapping of older deposits by
the NE up to the eastern bank of the Chara River and to the north younger ones with unconformities at the margins. Because of thin-
up to the Dzhelinda River and its inflows (Fig. 25.1b). Later, ning and probably erosional processes, the Nichatka formation is
some researchers (Ivanov et al. 1995 and others) incorrectly referred missing in sections along the left bank of the Chara River.
to the Nichatka Formation along the western limb of the Berezovs-
kaya Basin as the Dzhemkukan Formation, based on long-distance
Stratigraphy
correlations with deposits of different structural and facies zones
and composition. The stratotype section of the formation is
A sequence of the Patom Supergroup in the southern Berezovskaya
located on the northwestern bank of Lake Nichatka, 3.5 km to the
Basin is shown in Figure 25.2. Analogues of the groups and
NW of the mouth of the Shirik River (578470 3600 N, 1178370 0900 E–
formations of the Patom Supergroups were established there by
578480 0600 N, 1178380 3500 E). Although a glacial origin for the for-
means of successive correlation of sections (Zhuravleva et al.
mation was assumed in the mid-twentieth century, geological
1959), which was supported by geological mapping (Ivanov
facts confirming this idea have been discovered and published
et al. 1995), similarity of stromatolitic and microphytolitic associ-
only recently (Chumakov 1993).
ations (Dol’nik 2005) and recent data on C and Sr isotopes
(Pokrovsky et al. 2006). The basal Ballaganakh Gr. is exposed in
Structural framework the western limb of the basin. The group is mainly composed of
sandstones and small-pebbled conglomerates. Its upper part is rep-
In the eastern area of distribution of the Nichatka Fm. gently resented by massive and stromatolitic dolomites and limestones of
dipping (5 –158) strata compose the southeastern limb of the asym- the Mariinsky Fm. An erosional surface of this formation is over-
metrical marginal Berezovskaya Basin of the Siberian Craton lapped by the Nichatka Formation, which is a basal part of the
(Fig. 25.1b). The deposits have not been altered by secondary pro- Dal’nyaya Tayga Gr. The Nichatka Fm. is composed of dia-
cesses. In the steep and locally overturned western limb, the mictites, thin-bedded clay-rich siltstones with lonestones (drop-
Nichatka Fm. is exposed in the cores of faulted anticlines. There stones), conglomerates and sandstones. The formation decreases
the deposits are intensely cleaved and experienced lower green- in thickness eastward and thins out or was eroded on the right
schist metamorphism. Deformation of the western limb may bank of the Chara River. The Nichatka Fm. is well correlated
have been caused by the clockwise rotation of the Aldan Shield with the Bol’shoy Patom Fm. of the transverse Ura Uplift, which
during the Middle Palaeozoic (Shatsillo 2006). complicates the northeastern sector of the Patom folded arc.
The Nichatka Fm. accumulated in the shallow eastern margins Both formations are similar in stratigraphic position, lithological
of the Late Precambrian Patom basin, which was a passive composition and genesis. In other sectors of the arc, both for-
margin of the Siberian Craton (Chumakov et al. 2007). The mations are replaced by sands and shales of the Dzhemkukan Fm.
terrigenous-carbonate Patom Supergroup was formed in the The Nichatka Fm. is conformably overlain by the Kumakh-
basin during the Late Precambrian. Towards the east, all units of Ulakh Fm., which begins with a member of red-brown dolomites
the supergroup, including the Nichatka Formation, are reduced marls and dolomites. The rest of the Kumakh-Ulakh Fm. is com-
in thickness, and terrigenous formations are partly substituted by posed of black and variegated shales with limestone interbeds.

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 297– 302. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.25
298 N. M. CHUMAKOV

Fig. 25.1. (a) Geographical position of


Nichatka Fm. and coeval glacial units at
Baikal-Patom Highland: Nch, Nichatka
Fm.; Dz, Dzhemkukan Fm.; Bp, Bol’shoy
Patom Fm.; B.P., Bol’shoy Patom R.
(b) Geological scheme for the south
centroclinal part of Berezovskaya Basin: 1,
Mezsozoic syenite-porphyry; 2, Tinnaya,
Zherba and Torgo formations; 3, Sen’ and
Kumakh-Ulakh formations; 4, Nichatskaya
Fm.; 5, more old Proterozoic deposits; 6,
Palaeoproterozoic and Archaean granites;
7, Archaean gnesses and granites; 8, faults
and thrusts. Numbers in circles indicate type
sections of Nichatka Fm.: 1, upper course of
Dzhelinda R.; 2, west side of Nichatka L.; 3,
Mokryi Kumakh-Ulakh R.

The formation is correlative to the Barakun Fm. of the Patom diamictites, as well as foreign gneisses and pink granites. Some
Supergroup. The Dal’nyaya Tayga Gr. is crowned by the Sen’ stones (a few percent) have striated and grooved surfaces. Longi-
Fm. represented by alternating sandstones and dolomites in the tudinal, variably sized, subparallel striae and grooves are often
lower part and by limestones, massive stromatolitic and oncolitic confined to newly made abrasion facets. Grooves on granite
dolomites in the upper one. Therefore, some geologists divide pebbles sometimes resemble a series of small spear-shaped chat-
the Sen’ Fm. into two formations: Imalyk and Tokko (Petrov termarks (Chumakov 1993, fig. 3). Stone nests (clast clusters)
1976). The Sen’ Fm. is correlated to the shaly Valyukhta Fm. of and variously rounded and then split stones are characteristic.
the Patom folded arc and to the terrigenous Ura and carbonate Elongated stones are predominantly oriented to the north and
Kalancha formations of the Ura Uplift. The Dal’nyaya Tayga Gr. NE. Lower parts of the diamictite are usually bedded, suggesting
is overlain by the Zhuya Gr. represented by a single Torgo Fm., a gradual transition from the underlying thinly bedded rocks
which comprises a variegated marl member below and aphanitic with dropstones to massive diamictites. However, at the base of
algal and oolitic limestones and dolomites in the larger upper some massive diamictites there are signs of erosion or disharmonic
part. Within the Patom folded arc, the Torgp Fm. corresponds to folding and brecciation of underlying thinly bedded rock. In this
the Nikol’skoe Fm. of variegated marls and the Chencha Fm. case, overlying diamictites contain undeformed rafts of underlying
of limestones. rocks and small lenses and boudins of assimilated sandstones.
The Precambrian deposits of the southern Berezovskaya Units between diamictites are composed of alternations of lami-
depression is terminated by the Zherba and Tinnnaya formations, nated red siltstones and mudstones with graded fine-grained
which are identical in composition to synonymous formations of arkosic sandstones. Upper parts of these units are less distinctly
the Patom arc and Ura Uplift. The Zherba Fm. is mostly composed bedded and contain scattered pebbles and boulders (0.5 –150 cm
of quartzite-like, glauconite-bearing sandstones, and the Tinnaya across). These tear underlying beds and are enveloped by over-
Fm. by bituminous limestones, dolomites and dolomitic breccias. lying ones; such dropstones bear abraded faces with striae and
In the Ura Uplift, the uppermost beds of the latter contain Nemakit- grooves. At lateral contacts of pebbles and boulders, synsedimen-
Daldyn fauna (Khomentovskiy et al. 2004). The overlying Yuedey tary deformation structures were observed. In general, lithological
Fm. of variegated marls, dolomites and limestones is correlative changes throughout the section reflect a gradual transition from
to the Lower Cambrian Nokhtuysk Fm. of the Ura Uplift contain- thinly bedded members to bedded diamictites.
ing the Tommotian fauna at the base and to the Pestrotsvetnaya A few kilometres to the NE, diamictites and thinly bedded rocks
Fm. of the Aldan Shield. are replaced by conglomerates with subordinate diamictite inter-
beds. The conglomerates include large boulders consisting of
granites, gneisses, quartzites and boulders and plates of finely
Glaciogenic deposits and associated strata crystalline and oolitic dolomites up to 2.3  2  1.8 m in size.
Further to the north, the conglomerates are rapidly substituted by
Sections of the Nichatka Fm. are essentially variable in lithological a sequence of cross-bedded arkosic sandstones and poorly sorted
composition. The stratotype section on the northwestern bank of small-pebbled conglomerates (Chumakov 1993). The Nichatka
Lake Nichatka (Fig. 25.3, 2) exhibits repeatedly alternating units Fm. demonstrates similar composition and structure throughout
of massive, less frequently bedded diamictites (5 –10 m thick) the southeastern limb of the Berezovskaya Basin.
and thin-bedded clay-rich mudstones and sandstones (1.5 –20 m The character of the formation is different in the western limb
thick). The diamictites consist of lilac-grey clayey-silty-sandy (Figs. 25.3, 1). The most complete and representative section of
matrix containing 10– 15% of scattered stones of different form the Nichatka Fm. is located at the Creek Opornyi, a left inflow
and size (2– 75 cm, singular up to 170  100 cm). Roundness of the Dzhelinda River (588030 2500 N, 1178350 2600 E – 588030 3200 N,
varies from class 0 to 5 on a five-class scale, but falls into 178360 0300 E). There the formation is mainly represented by
classes 1 or 2 on average because of the dominance of only slightly yellowish-grey and grey diamictites with graded bedding, which
rounded or unrounded stones. Most stones are fragments of grey frequently constitute single gradational rhythms with sandstones.
granites and reddish quartzites similar to those underlying the Massive and bedded diamictites are subordinate. There are also
NICHATKA FORMATION, CHARA RIVER BASIN 299

Fig. 25.2. Stratigraphic position of the


Nichatka Fm.: 1, diamictites; 2,
conglomerates; 3, sandstones; 4, shales; 5,
limestones; 6, dolostones; 7, marls; 8,
dolostone breccia; 9, oolitic carbonates; 10,
stromatolitic carbonates; 11, lowermost
value of 87Sr/86Sr of formations
(chemostratigraphic data by Pokrovsky
et al. 2006).

interbeds of cross-bedded sandstones and members of thinly pyrophyllitic shales of the Purpol Fm. On the western limb of
bedded shales with dropstones. Beds of conglomerates and the basin the Nichatka Fm. lies on an erosional surface above the
breccia are rare. Massive diamictites contain predominantly Upper Proterozoic Mariinsky or Bugarikhta formations. The
pebbles, and small and medium-sized boulders, but large boulders base of the Nichatka Fm. is composed in this region of conglomer-
and blocks of granites 0.5  1.2 m in size also occur in some ates and breccia of massive stromatolitic and oncolitic dolomites
members. The stones have a roundness of 1– 2.5 (according to and limestones of the Mariinsky Fm. Locally, dolomite surfaces
the five class scale). Clasts in the diamictites are mostly rep- show buried relief (the Bogayukhta River, 578500 0200 N,
resented by local and erratic granites, quartzites and gneisses. 1178270 5500 E).
Pebbles and poorly rounded boulders of dolomites similar to The upper boundary of the Nichatka Fm. is more uniform. The
those of the Mariinsky Fm. are frequent. Some stones bear sub- formation is everywhere gradually overlain by a basal member of
parallel striae. Massive diamictites enclose small sand lenses the Kumakh-Ulakh Fm., which consists of alternating laminated
with erosional or gradational lower boundaries as well as signs red dolomitic marls and pink dolomites. The thickness of this
of underwater slides in the form of plastic folds and slide rolls. member is more than 3 m in the southeastern limb and up to
10 m in the western limb of the Berezovskaya Basin. Contacts of
the Nichatka Fm. and this carbonate member are observed in the
Boundary relations with overlying and underlying middle course of the Sen’ River, 2 km below the Uraga mouth
non-glacial units (578500 0200 N, 1178270 5500 E) and at the Creek Opornyi. At the
first locality the Nichatka diamictites grade into coarse and then
On the southeastern limb of the Berezovka Basin, the Nichatka Fm. fine-grained clayey sandstones (2 m), which are overlain by
overlies with angular disconformity Archaean (Fig. 25.1, 3) and platy, sandy and then by calcareous mudstones (8 m). The latter
Early Proterozoic gneisses and granites, dark red quartzites and are conformably overlain by red-brown thin-bedded dolomitic
300 N. M. CHUMAKOV

Fig. 25.3. Type sections of Nichatka Fm.:


1, conglomerates; 2, massive diamictites; 3,
bedded diamictites; 4, shales with
dropstones; 5, sandstones; 6, shales; 7,
limestones; 8, dolstones; 9, marls; 10,
granites; 11, striated and faceted stones;
12, dropstones; 13, glaciotectonic
deformations; 14, till-like stone fabric; 15,
stone nests; 16, split, semi-rounded stones;
17, rhythmic bedding and lamination; 18,
erosional channels; 19, small sandstone
lenses; 20, slide and slump structures;
21, slide rolls; 22, lamination; 23,
cross-bedding; 24, conglobreccies. PR1,
Palaeoproterozoic; km, Kumakh-Ulakh
Fm.; mr, Mariinsky Fm.; nč, Nichatka Fm.

marls of the Kumakh-Ulakh Fm. (0.7 m) passing upward into lami- for the dolomite component are slightly higher but exhibit parallel
nated red clayey dolomites (1.5 m apparent thickness). Similar trends. The d13C curve shows two negative anomalies. The first
conformable and gradual transition between the Nichatka and negative anomaly in the Kumakh-Ulakh Fm. ranges from –2.9
Kumakh-Ulakh formations is recorded in the Creek Opornyi to – 6.3‰ Vienna Pee Dee Belemnite (VPDB) standard, reaching
section. There the uppermost Nichatka diamictites grade into red- a minimum value in the lower part of the formation and increasing
brown sandstones (.5 m) with scattered pebbles and boulders. upsection. d18P values vary considerably from 14.9 to 22‰
Alternating with shales, the sandstones are replaced by finely Vienna Standard Mean Ocean Water (VSMOW). The second
bedded variegated shales (15 m) including thin sandstone inter- negative anomaly, showing a decrease to –8.0 to –10.5‰, is
beds and higher up by a basal member of the Kumakh-Ulakh found in the Torgo Fm. d18P varies insignificantly from 21 to
Fm. (10 m). This member begins with marls and, higher, exhibits 23‰ with some values as low as 19‰. Between these two negative
alternating beds (2 –10 cm) of red micritic dolomites and dolomite anomalies there is a moderately positive one (from – 2.7 to
marls. þ3.4‰) with a maximal d13C value of þ3.4 to þ3.9‰ in the
upper part of the Sen’ Fm. d18P values for this formation vary
between 22 and 25‰, reaching 27‰ in its upper part. The
Chemostratigraphy overlying carbonate Tinnaya Fm. is characterized by gradually
increasing d13C values from –0.8 to þ2.1‰ and d18P values
C, O and Sr isotopic compositions of carbonates were studied by from 20 to 23‰.
Pokrovsky et al. (2006) in craton sections of the Patom Supergroup The lowermost 87Sr/86Sr ratio for samples from the Kumakh-
along the Chara River, which runs across the southeastern limb of Ulakh Fm. with low values of 87Rb – 86Sr (0.001 – 0.005), Mn –Sr
the Berezovskaya Basin. Results for the calcite component are (0.07 –0.12) and Fe/Sr (2.05 –3.30) is 0.70725 (three determi-
described below and represented in part in Figure 25.2. Values nations: Sr 881–2013 ppm; range 0.70725 –0.70734). For the
NICHATKA FORMATION, CHARA RIVER BASIN 301

Torgo Fm. the lowermost 87Sr/86Sr ratio is 0.70837 (one deter- glaciolacustrine. Along strike, the basal tillites are replaced by
mination: Sr 158 ppm). For the Tinnaya Fm. the lowermost ratio boulder-sized conglomerates, and further along strike by cross-
is 0.70799 (four determinations: Sr 299–812 ppm; range of bedded arkosic sandstones. These deposits, which can be inter-
0.7079 –0.70832). preted as glaciofluvial, form extensive fans penetrating into the
adjacent basin. Distal sediments of the fans occur on the western
limb of the Berezovskaya Basin. They are mostly represented
Palaeolatitude and palaeogeography there by graded bedded and massive diamictites. Massive diamic-
tites contain rare striated stones and frequent small sand lenses, the
No palaeomagnetic investigations of Upper Precambrian deposits tops of which are marked by thin silt interlayers and the bases of
have been carried out in the Berezovskaya Basin. According to the which exhibit a gradual transition into the sandy-clayey-silty
majority of recent global reconstructions (Powell et al. 2001; matrix of diamictite.
Smith 2001; Meert & Torsvik 2003; Shatsillo 2006; and others), There are also small sand lenses filling gentle incisions, showing
the Siberian craton, including this basin, was located at low lati- that local low-energy currents washed out clay and silt from the
tudes at that time. diamictons. The supposition is that massive and bedded diamic-
tites were deposited by floating shelf glaciers, whereas graded
bedded diamictites represent glaciogenic deposits reworked by
Geochronological constraints mud and debris flows. The direction of cross-bedding in the
sandy interlayers and the lithology of the clasts suggest that
There are no available radiometric and reliable biostratigraphic material was transported both from the east and SE and partly
data on the Upper Precambrian deposits of the Berezovskaya from the west to the western end of the basin. Diamictites
Basin. They can be dated by means of correlation with biostratigra- include rare units of thin and rhythmically bedded siltstones and
phically better studied sections of the Patom Supergroup of the Ura mudstones with dropstones, which may represent distal turbidite
Uplift (the Lena River). The corrrelations are based on successive deposits where stones dropped from melting icebergs.
section-by-section tracing of formations, stromatolites and
isotopic data (Chumakov et al. 2011). d13C anomalies can also
Age
provide confident correlations. The lower Kumakh-Ulakh negative
d13C anomaly corresponds to the negative anomaly for the upper
Reliable stratigraphic correlation between the sections of the
Bol’shoy Patom –lower Barakun interval of the Ura Uplift. The
Berezovskaya Basin and the Ura Uplift along with biostratigraphic
positive Sen’ anomaly can be correlated with the analogous
data existing for Ura Uplift succession (Chumakov et al. 2011)
Barakun –Valyukht anomaly of that uplift, and the significant
allow us to suppose that the Tinnaya and probably Zherba
negative Torgo anomaly can be easily correlated with the analo-
formations can be tentatively assigned to the Nemakit-Daldyn
gous Zhuya anomaly (Pokrovsky et al. 2006). These correlations
Stage of the Upper Vendian, the Zhuya Group to the Kotlin and
are supported by similar 87Sr/86Sr ratios for both regions. These
Redkino horizons of the Middle Vendian, and the Dal’nyaya
data suggest that the Nichatka Fm. is correlative to the glacial
Tayga Group (and the Nichatka Fm. correspondingly) to the
Bol’shoy Patom Fm. and that the lower Sen’ Fm. corresponds to
lower part of Laplandian Horizon of the Lower Vendian. Accord-
the Ura Fm. The last formation contains a rich assemblage of
ing to the timescale of the International Commission on Stratigra-
acanthomorphic palynomorphs (Chumakov et al. 2007; Vorob’eva
phy, the Tinnaya Fm. must be referred to the pre-Tommotian
et al. 2008) very similar to the Ediacaran (lower Vendian) Kel’tma
(Fortunian) interval of the early Cambrian, the Zhuya Group and
microbiota of the Russian plate (Veis et al. 2006) and to ‘late Per-
apparently the upper part of the Laplandian Horizon to the Edia-
tatataka’ [¼ECAP (Ediacaran Complex Acanthomorph Palyno-
caran System, and the lower part of the Laplandian Horizon to
flora), Grey 2005] microbiota of Australia. The upper part of the
the Upper Cryogenian. In the context of distant correlations, the
Tinnaya Formation of the Ura Uplift contains small-shelly fauna
data presented above suggest that the Nichatka Fm., like the
of Nemakit-Daldyn type (Khomentovskiy et al. 2004).
Bol’shoy Patom Fm., is correlative to the Nantuo Fm. of South
China and probably the Yerelina Subgroup of Australia.
Discussion
Regional palaeogeography
Sedimentary environments
As described earlier (Chumakov 1993, 2009), glacial deposits
The Nichatka Fm. has a great number of features that are typical of coeval with the Nichatka Fm. are traceable (with some interrup-
glacial deposits. It contains many massive and bedded diamictites tions) along almost the entire margin of the Baikal-Patom folded
with clasts of variable size, roundness and composition. These zone (Fig. 25.1a). They are known as the Dzemkukan Fm. to the
stones have, on average, roundness falling into classes 1 or 2 north of the Lake Nichatka between the Dzhelinda and Chencha
(using the five classes scale). Erratic and abraded stones (with rivers. Farther on, they are exposed in the Ura Uplift along the
clearly striated and grooved faces and spear-like chattermarks) Lena River, where they form the Bol’shoy Patom (or Dzemkukan
are common. Rounded and subsequently split stones are also fre- Fm.). To the SW, glacial deposits are exposed along the western
quent. Dropstones and stone nests (clast clusters) occur commonly margin of the Baikal-Patom folded zone, where they are also
in laminated rocks, which alternate with bedded diamictites. called the Dzemkukan Fm. The Dzemkukan Fm. can then be
The disharmonic deformation at the base of some massive dia- traced from the Lena –Bol’shoy Patom interfluve up to the
mictites resembles glaciotectonic disruption, and the orientation Vitim –Chaya interfluve. Although slightly thinner, the Dzemku-
of elongated stones is similar to that of tills, suggesting that the kan Fm. of this region is similar in composition and structure to
Nichatka Fm. on the southeastern limb of the Berezovskaya the Bol’shoy Patom Fm. (Chumakov 1993). To the SW of the
Basin contains basal tillites. The prevailing northern and northeast- Malaya Chuya River head, the Ballaganakh and lower part of the
ern orientation of elongated stones and the occurrence of erratic Dal’nyaya Tayga groups with the glacial deposits thin out. At
stones of plagiogneisses and micaceous greisses along with the Chaya River, analogues of the upper part of the Patom Super-
local stones indicate the Aldan Shield as a source for Nichatka group form the Baikal Supergroup extending up to Lake Baikal. At
deposition in the SE of the basin. The basal tillites alternate the southern end of the lake at the base of the supergroup there
with thin and rhythmically bedded siltstones and mudstones appears a sequence of diamictites, conglomerates and sandstones,
with dropstones. Such thin-bedded deposits are likely to be which are called the Bugul’deka Member and are correlated to the
302 N. M. CHUMAKOV

Dzemkukan, Bol’shoy Patom and Nichatka formations (Chuma- (eds) The Geological Record of Neoproterozoic Glaciations. Geo-
kov 1993; Sovetov & Komlev 2005). logical Society, London, Memoirs, 36, 309–316.
These geological correlations have recently been supported by Dol’nik, T. A. 2005. Stromatolites and Mikrophytolites in Stratigraphy
C- and Sr-isotopic studies, which revealed considerable similarity of Riphean and Vendian of Folded Framework of the South Part of
between the Baikal Supergroup and the upper part of the Patom Siberian Craton (in Russian). Geo, Novosibirsk.
Supergroup (Kuznetsov & Letnikova 2005), and indirectly by Grey, K. 2005. Ediacaran Palinology of Australia. Memoir 31. Associ-
Pb/Pb dating of the Uluntuu limestones (550 + 40 Ma, Kuznetsov ation of Australian Palaeontologists, Canberra.
et al. 2006). All of the above-mentioned glacial units were united Ivanov, A. I., Lifshits, V. I., Perevalov, O. V., Strakhova, T. M.,
Yablonovskiy, B. V, Gryzer, M. I., Il’inskaya, Kh. G. & Golove-
into a single Middle Siberian Glacial Horizon (Chumakov 1993).
nok, V. K. 1995. Precambrian of Patom Highland (in Russian).
By means of lithological and sequence stratigraphic data, the Nedra, Moscow.
Bugul’deyka Member can now be correlated to glacial deposits Khomentovskiy, V. V., Postnikov, A. A., Karlova, G. A., Kochnev,
of the Marninsk Fm. of the Sayany region (Sovetov & Komlev B. B., Yakshin, M. S. & Ponomarchuk, V. A. 2004. Vendian of
2005). This makes the area of distribution of the Middle Siberian Baikal-Patom Upland (Siberia). Geology and Geophysics, 45,
Glacial Horizon much wider. 465– 484.
As stated above, the main source for the Nichatka deposits Kuznetsov, A. B. & Letnikova, E. F. 2005. Opening Baikal Branch of
was the Aldan Shield. Adjacent parts of the Siberian Craton also Palaeoasian ocean: Sr and C data (in Russian). In: Koryakin, Y. V.
supplied debris for the Bol’shoy Patom Fm. of the Ura Uplift (ed.) Tectonics of the Earth Crust and Mantle. The Tectonic Regu-
(Chumakov & Krasil’nikov 1991) and the Dzemkukan Fm. at larities of the Occurrences Mineral Recourses. GEOS, Moscow, I,
the western margin of the Baikal-Patom folded zone. Glaciers 352– 355.
moved from the craton to the western Baikal area (Sovetov & Kuznetsov, A. B., Ovchinnikova, G. B., Kaurova, O. K. & Letni-
Komlev 2005). Sometimes, debris was transported to the Sayany kova, E. F. 2006. Pb/Pb age and Sr chemostratigraphy of carbonate
region from the folded framework of the Siberian craton rocks of the Baikal Group, the south-west Pribaykal’e (in Russian).
(Sovetov 2002; Sovetov & Komlev 2005), and also to the Isotopic dating of the ore formation, magmatism, sedimentation and
western part of the Berezovskaya Basin. In general, The Siberian metamorphism. GEOS, Moscow, 1, 362– 365.
Craton was the main sourceland for the Middle Siberian Glacial Meert, J. G. & Torsvik, T. H. 2003. The making and unmaking of a
Horizon. This suggests that the main centres of Middle Siberian supercontinent: Rodinia revisited. Tectonophysics. 375, 261– 288.
glaciation were located on the Siberian Craton, and at least the Petrov, A. F. 1976. Precambrian orogenic complexes of west part of
entire southern part of the craton was covered by glaciers, Aldan shield (in Russian). Nayka, Novosibirsk, 120.
Pokrovsky, B. G., Melezhik, B. A. & Buyakayte, M. I. 2006. Geo-
which, as evinced in the wide distribution of glaciomarine facies,
chemistry of isotopes C, O, Sr and S, chemostratigraphy and environ-
flowed into the Baikal Ocean and surrounding seas. The very ments of sedimentation of Late Precambrian deposits of Patom
extensive area of distribution of the glacial deposits (almost trough. Lithology and Mineral Resourses, 5, 505– 530.
2000 km long) unambiguously indicates the existence of inland Powell, C. McA., Pisarevsky, S. A. & Winwate, M. T. D. 2001. An ani-
glacial sheets at all southern parts of the Siberian Craton during mated history of Rodinia. Geological Society Australia Abstracts,
the Middle Siberian glaciation. 65, 85 – 87.
Shatsillo, A. V. 2006. Palaeomagnetism of the Vendian of the South Part
Research was supported by grants of the Russian Fund of Basic Investigations of the Siberian Craton and Some Aspects of Late Precambrian
(No. 11-05-00232), Program No. 25 of Presidium of Russian Academy of the Palaeogeodynamics (in Russia). Institutes Physics of the Earth
Sciences. The author would like to acknowledge stimulating discussions arising Russian Academy of Sciences, Moscow.
from participation in IGCP project 512 ‘Neoproterozoic Ice Ages’. This rep- Smith, A. G. 2001. Palaeomagnetically and tectonically based global
resents a contribution of the IUGS- and UNESCO-funded IGCP (International maps for Vendian to Mid-Ordovician time. In: Zuravlev, A. Yu
Geoscience Programme) project #512. & Riding, R. (eds) The Ecology of the Cambrian Radiation. Colum-
bia University Press, New York, 11 – 46.
Sovetov, J. K. 2002. Vendian foreland basin of the Siberian cratonic
References margin: Palaeopangean accretionary phases. Russian Journal of
Earth Sciences, 4, 363– 387.
Chumakov, N. M. 1993. Riphean Middle Siberian Glaciohorizon. Strati- Sovetov, Yu. K. & Komlev, D. A. 2005. Tillites at the base of Oselok
graphy and Geological Correlation, 1, 17 –28. Group Prisayan’ya and lower boundary of Vendian in south-west
Chumakov, N. M. 2009. Neoproterozoic glacial events in Eurasia. part of Siberian Craton. Stratigraphy and Geological Correlation.
In: Gaucher, C., Sial, A. N., Halverso, G. P. & Frimmel, H. R. 13, 3 – 34.
(eds) Neoproterozoic– Cambrian Tectonics, Global Change and Veis, A. F., Vorob’eva, N. G. & Golubkova, E. Y. 2006. First find of
Evolution: a Focus on Southwestern Gondwana. Developments in Lower Vendian microfossils at Russian plate: taxonomy composition
Precambrian Geology. Elsevier, The Netherlands, 16, 389– 403. and biostratigraphic significance. Stratigraphy and Geological
Chumakov, N. M. & Krasil’nikov, S. S. 1991. Lithology of Correlation, 14, 28– 46.
Riphean tilloids; Ura Uplift. Lithology and Mineral Resourses, 3, Vorob’eva, N. G., Sergeev, V. N. & Chumakov, N. M. 2008. New
58 –78. occurrences of Lower Vendian microfossils in Ura Formation:
Chumakov, N. M., Pokrovsky, B. G. & Melezhik, V. A. 2007. Geologic problem Patom Supergroup of Middle Siberia. Doclady Academii
history of Patom Supergroup, Late Precambrian, Middle Siberia. Nayk. Geologiya, 419, 782– 787.
Doclady Academii Nayk. Geologiya, 413, 379– 383. Zhuravleva, Z. A., Komar, V. A. & Chumakov, N. M. 1959. Strati-
Chumakov, N. M., Pokrovsky, V. G. & Melezhik, V. A. 2011. The graphic correlations Patom Complex with sedimentary deposits
glaciogenic Bol’shoy Patom Formation, Lena River, central west and north slopes of Aldan Shield. Doclady Academii Nayk
Siberia. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. USSR, 128, 1026– 1029.
Chapter 26

Glacial deposits of the Baykonur Formation, Kazakhstan and Kyrgyzstan

NICKOLAY M. CHUMAKOV
Geological Institute of the Russian Academy of Sciences, Pyzhevsky per. 7, Moscow 119017, Russia (e-mail: chumakov@ginras.ru)

Abstract: The Baykonur Formation extends with some interruptions for over 1700 km along the eastern and northern margins of the
Upper Precambrian Syrdarya and Tarim microcontinents. The Baykonur Formation constitutes the upper part of the Ulutau Group,
which lies on an erosional surface of granosyenite with a U–Pb age of 720 + 20 Ma and is overlain by Lower Cambrian vanadium-
bearing carbonaceous-siliceous shales. The Baykonur Formation is mostly composed of diamictite with erratic and striated stones,
and includes thin-bedded shale beds with lonestones (dropstones) and a ‘cap dolomite’ unit at its top. The lithological composition of
this formation implies a glaciomarine origin, while its stratigraphic position suggests a Late Vendian or an Early Cambrian age (the
latest Ediacaran or Early Cambrian according to scale of the International Commission on Stratigraphy).

The Baykonur Formation represents a horizon of Upper Precam- Tarim microcontinents (Kiselev 2001). The Baykonur Formation
brian diamictites, which extend with some interruptions as an experienced folding, thrusting and other dislocations during
extended arc for over 1700 km from the Ulutau Mountains of Palaeozoic ‘Caledonian’ orogeny. The deposits are usually
Central Kazakhstan (Fig. 26.1, 1) through the Bol’shoy Karatau cleaved (locally significantly) and weakly metamorphosed up to
Ridge and Dzhebagly Mountains of Southern Kazakhstan lower greenschist facies.
(Fig. 26.1, 3, 4, 5), the Chatkal, Sandalash (Fig. 26.1, 6) and
Kokiirim-Too (Fig. 26.1, 7) ridges of Western Kyrgyzstan up to
the Naryn-Too and Dzhetym-Too ridges of Central Kyrgyzstan Stratigraphy
(Fig. 26.1, 8) and the Sarydzhaz Ridge of Eastern Kyrgyzstan
(Fig. 26.1, 9) and further up to the China boundary. The Baykonur One of the type sections of the upper part of the Ulutau Group is a
Formation is the upper formation of the group called Ulutau in section at the Rang River (the Bol’shoi Karatau Ridge), where the
Kazakhstan and Dzhetym in Kyrgyzstan. Its name is derived following members are exposed (from the base upward, Fig. 26.2).
from the Baykonur River flowing from the southwestern slope of
the Ulutau Mountains where the stratotype section of the formation The Aksumbe Formation
is located. The second type section, deposits of which are less (1) Pale limestones alternating with green and black clay-rich
metamorphosed and better studied, is an outcrop along the Rang shales (200 m).
River, which flows from the northeastern slope of the Bol’shoy
Karatau Ridge. Probable stratigraphic analogues of the Baykonur The Baykonur Formation
Formation are the diamictite of the lower part of the Kyrshibakty (2) Dark thin-bedded clay-rich shales with scattered stones of
Formation in the Malyi Karatau Ridge (Korolev & Maksumova quartz, dolomite, quartz porphyres and shales. The deposits
1984) and the upper diamictitic part of the Kopal Formation were disturbed by subaqueous slumps (10 m).
exposed in the Mointy River basin (Fig. 26.1, 2). (3) Diamictite with a muddy matrix and scattered stones of differ-
The Baykonur diamictites were first recovered in 1924 by D. V. ent size, roundness and composition; bedding is disturbed by
Nalivkin, who referred them as Upper Palaeozoic glacial deposits. sliding, ‘quivering’ undulatory lamination and rare dolomite
Some time later, the Baykonur Formation was considered to be olistoliths (200 m).
Cambrian ‘facies of the foot of rocky shores’ (Borovikov 1955), (4) Dark grey, brown on surface, bedded dolomite with intrafor-
proluvium (Makarychev 1967), subsurface sliding deposits mational breccias and signs of sliding (7.5 m).
(Volin 1966; Kholodov 1973; and others) or subaqueous sliding
deposits with an admixture of mountain glacial debris (Knipper The Kurumsak Formation
1963). In the 1960s it became evident that the Baykonur diamic- (5) Black cherts and vanadium-bearing carbonaceous-siliceous
tites belong to the Upper Precambrian (Korolev 1963). A majority shales (250 m).
of geologists agreed with their Precambrian age but continued,
after Knipper, to interpret them as polygenic deposits of subaqu- Sections of the Baykonur Formation in the Ulutau Mountains
eous slides, clastic and grain flows or, partly, as proluvium and (Fig. 26.2) and other areas of distribution are similar to the strato-
mountain glacial deposits (Zaytsev & Kheraskova 1979; Kheras- type section. Thicknesses (from 600 to 100 m or less) and names of
kova 1981a, 1986; Korolev & Maksumova 1984; and others). At units above and below the Baykonur Formation are variable.
the same time, evidence was gradually accumulating to indicate
that the Baykonur diamictites were to a considerable extent
formed by glacial activity (Ankinovich 1961; Zubtsov 1972; Glaciogenic deposits and associated strata
Chumakov 1978; Azerbaev 1988).
The Baykonur Formation is mostly composed of diamictite with
subordinate interbeds and members of shale, sandstone, conglom-
Structural framework erate and breccia. The diamictite retains its main characteristic
features throughout the region, but may vary in appearance and
The Baykonur Formation is confined to the synclinorium of degree of metamorphism. The diamictite matrix usually comprises
the Caledonian basement stretching along the ancient eastern dark grey mudstone with variable portions of sand and dolomite
active margins of the Syrdarya and northern margin of the grains. Increased content of the latter imparts a brownish-grey

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 303– 307. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.26
304 N. M. CHUMAKOV

Fig. 26.1. Outcrops of Baykonur


Formation and coeval glacial units in
Middle Asia. a, Outcrops of Baykonur and
Kopal Formations; b, boundary of states.
Kazakh Republic: 1, Ulutau Mountains; 2,
Mointy River; 3, North Bol’shoy Karatau
Ridge; 4, Central Bol’shoy Karatau Ridge;
5, South Bol’shoy Karatau Ridge and
Dzhebagly Mountains. Kirgiz Republic: 6,
Chatkal Mountain Range; 7, Kokirim
Mountain Range; 8, Naryntau and
Dzhetymtau Mountain Ranges; 9,
Sarydzhaz Mountain Range. I, Uzbek
Republic; II, People’s Republic of China.

colour to the matrix. There may also be admixture of chlorite, In the Rang section, the diamictite contains a great number of
depending on the metamorphic grade. Chlorite imparts a small and large (up to a few metres long) dolomite olistoliths,
greenish-grey colour to the matrix. Pyrite inclusions are also fre- brecciated dolomite and fragments of clay-rich sandstone beds.
quent at times. The matrix contains chaotically scattered clasts In some places olistoliths are deformed, disrupted and associated
of different sizes, roundness and composition that can amount to with smaller fragments of the same rocks (Korolev & Maksumova
between 1 and 20% of the total rock. Their predominant size 1984). In addition to the dolomites, there are fragments and blocks
ranges from 1 to 10–15 cm, although more rare larger fragments of arkosic sandstone. Clast clusters (‘stone nests’) are common.
up to few metres across are also found. Semi-rounded and angular Some parts of the diamictite exhibit a massive texture. Varying
stones dominate over rounded and well-rounded stones. Striated concentrations of clasts in the Baykonur diamictite may make
and faceted clasts occur rarely. The clasts are mainly composed of impressions of coarse and lenticular bedding. In the Rang section,
rocks underlying the diamictite, but there are also erratic rocks that varying dolomite content defines vague bedding exhibiting
are foreign to the underlying successions (Azerbaev 1988). gradual boundaries every 0.2– 5 m within the diamictite, which

Fig. 26.2. Type sections of Baykonur


Formation (Ulutau Mauntins: Knipper
1963; Kheraskova 1986; Dzhetymtau
Range: Korolev & Maksumova 1984).
BAYKONUR FORMATION, KAZAKHSTAN AND KYRGYZSTAN 305

is often characterized by fine plastic crimping. Large-scale soft- Formation, which demonstrates a Middle Cambrian fauna
sedimentary deformation provides evidence of subsurface sliding. (Zaytsev & Kheraskova 1979; Kheraskova 1981b).

Shales
Chemostratigraphy
Greenish-grey thin-bedded shales constitute definable units
(members) 5–40 m thick in the lower part and less commonly in No chemostratigraphic investigations of Neoproterozoic depos-
the middle part of the Baykonur Formation. The shales usually its have been carried out in regions of the distribution of
include scattered small and medium-sized, and more rarely large Baykonur Fm.
pebbles and boulders. The lonestones cut through into the under-
lying beds, while the overlying deposits drape the stones
(Knipper 1963; Azerbaev 1988). The shales bear frequent signs Other characteristics
of subaqueous sliding.
A unit of black, thin and rhythmically bedded vanadium-bearing
carbonaceous chert occurs 5–30 m above the base of the Kurum-
Sandstones sak Formation of the Bol’shoy Karatau. The shales are 10 –12 m
thick and include phosphorite concretions. The chert contains
The sandstones are composed of feldspar and quartz and form 75– 90% SiO2, 1.5 –15% Corg, 0.05–1.5% V2O5 and up to 2%
thin interbeds, lenses and small members within the diamictite BaO (Kholodov 1973). Under the microscope, fossil radiolarians
beds. Locally the sandstones have fine (up to 2 cm) rhythmical and sponge spicules can be observed in the chert (Kholodov 1973).
graded bedding. The rhythms begin with coarse-grained sand-
stones or small-pebbled conglomerates or grits, which gradually
pass into fine-grained sandstones. Many rhythms have an
uppermost thin interbed of siltstones with carbonate cement
Palaeolatitude and palaeogeography
(Knipper 1963).
No valid results for palaeomagnetic investigations have been pub-
lished for the Baykonur Formation and adjacent deposits. Accord-
Dolomite ing to general palaeotectonic reconstructions, the Kazakhstan and
Kyrgyzstan microcontinents were located at low latitudes during
In many sections, the Baykonur Formation is terminated by a dolo- the Cryogenian-Ediacaran (Vendian) (Kheraskova et al. 2003).
mite bed. Its thickness reaches 7.5 –10 m in the Rang section. The
dolomite is calcareous (up to 10% of CaCO3), slightly sandy (up to
11%), and has a dark grey colour, which is pale yellow or light Geochronological constraints
brown on weathered surfaces (Chumakov 1992). Its structure is
massive or bedded, frequently breccia-like, with signs of sub- The Baykonur Formation is the uppermost formation of the Ulutau
aqueous slides. Some dolomite interlayers represent typical intra- (Dzhetym) Group. In the Bol’shoy Karatau Ridge, the group lies
clast breccias consisting of numerous flat dolomite clasts. unconformably on deeply eroded granosyenite with a U – Pb
zircon age of 720 + 20 Ma. (Kiselev 2001). The formation has a
conformable contact with overlying carbonaceous-siliceous
Conglomerates shales of the lower Koktal Formation (the Ulutau Mountains) or
its stratigraphic and lithological equivalent, the Kurumsak For-
Conglomerates occur as separate beds and lenses. They greatly mation (the Bol’shoy Karatau Ridge). The Koktal carbonaceous
increase in amount on the western limb of the Baykonur synclinor- chert (c. 30–50 m higher than the base of the formation) yielded
ium (Kheraskova 1986). The clasts in these conglomerates are the acritarchs Micrhystridium aff. dissimilare Volk., M. tornatum
same as those in the diamictites. Volk., M. lubomlense Kirjan., Cymatiosphaera? membranacea
Kirjan., Leosohaeridia sp. ‘This acritarch assemblage is character-
istic of the upper Lower Cambrian and very similar to the associ-
Boundary relations with overlying and underlying ation from the Vergol Horizon of the East European platform’
non-glacial units (Krylov et al. 1986). The Koktal Formation lies beneath the Kok-
bulak Formation containing Middle and Late Cambrian trilobites
In the Ulutau Mountains, the Baykonur Formation lies conform- (Ergaliev 1965).
ably on underlying deposits in the axial part of the Baykonur syn-
clinorium and on an erosional surface on the synclinorium limbs
(Kheraskova 1981a, 1986). An unconformity at the base of the Discussion
Baykonur Formation can also be observed at the Rang section of
the Bol’shoy Karatau Ridge and at the Dzetymtau Ridge near Sedimentary environments
the Kalmakashu Pass (Korolev & Maksumova 1984).
The upper boundary of the Baykonur Formation is distinct but The wide lateral distribution of Baykonur diamictites in the same
conformable in the Baykonur synclinorium (Kheraskova 1981a) stratigraphic position, as well as the occurrence of erratic and
and at Rang River. Further to the south, in the Tien Shan Moun- striated stones, suggests a significant role for glaciers in their
tains, there is a hiatus at the top of the formation locally associated genesis. Interbeds and whole units of thin-bedded shales with drop-
with a weathering crust (Korolev & Maksumova 1984). stones and frequent subsurface sliding indicate that the diamictites
In the Mointy region, the Kopal Formation contains a diamictite were formed in glaciomarine environments and subsequently
member very similar to the Baykonur Formation (Kheraskova affected by flows of variable density. A characteristic succession
1981b). It lies conformably on top of dolomites of the lower part of post-glacial deposits is worthy of attention: diamictites are
of the Kopal Formation. In some sections, the Basagin Formation directly overlain by a marker bed of dolomites resembling
gradually replaces the diamictites. It contains stromatolites of cap dolomite and this latter unit is overlapped by carbonaceous
Cambrian type and is covered without hiatus by the Kyzylzhar chert containing phosphorite concretions and considerable
306 N. M. CHUMAKOV

concentrations of organic carbon (1.5 –15%). This facies succes- Azerbaev, N. A. 1988. Lithologic features and origin of Vendian con-
sion is similar to many Upper Precambrian post-glacial sequences glomerates of Bol’shoy Karatau (in Russian). Izvestia of Academy
(Chumakov 1992; Zhu et al. 2007). of Sciences of Kazakh SSR. Geological Series, 2, 53 – 63.
Borovikov, L. I. 1955. Lower Paleozoic of Dzhezkazgan-Ulutau Region
of West Part of Kazakhstan (in Russian). Nedra, Moscow.
Age Chen, J., Zhang, H., Xing, Y. & Ma, G. 1981. On the Upper Precam-
brian (Sinian Suberathem) in China. Precambrian Research, 15,
The sharp unconformity between the Ulutau Group and the under- 207– 228.
lying granosyenites, which have a U – Pb zircon age of Chumakov, N. M. 1978. Precambrian Tillites and Tilloids (in Russian).
720 + 20 Ma, provides evidence for a considerably younger age Nauka, Moscow.
for the group as a whole. The Baykonur Formation has a conform- Chumakov, N. M. 1981. Upper Proterozoic glaciogenic rocks and their
able contact with the overlying Kurumsak Formation or its strati- stratigraphic significance. Precambrian Research, 15, 373–396.
graphic analogues, for example the Koktal Formation. The Chumakov, N. M. 1992. The Problems of Old Glaciations (Pre-
erosional disconformity at the top of the Baykonur Formation in Pleistocene Glaciogeology in the USSR). Harwood Academic Pub-
the Tien Shan sections is therefore unlikely to reflect a great lishers, Pennsylvania.
hiatus. This suggestion is supported by persistent, but thin (few Chumakov, N. M. 2009. The Baykonurian Glaciohorizon of the Late
metres) cap dolomite of the Baykonur Formation over a long dis- Vendian. Stratigraphy and Geological Correlation, 17, 373–381.
tance along the Bol’shoy Karatau Ridge and in some other sec- Chumakov, N. M. 2011. Glacial deposits of the Bokson Group, East
tions. The conformable overlapping of the Baykonur Formation Sayan Mountains, Buryatian Republic, Russian Federation. In:
by the Lower Cambrian Koktal Formation indicates a relatively Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
young age for the Baykonur Formation. As mentioned above, the logical Record of Neoproterozoic Glaciations. Geological Society,
microfossils characteristic of the upper Lower Cambrian (the London, Memoirs, 36, 285–288.
Ergaliev, G. Kh. 1965. On stratigraphy of Vendian and Cambrian Bay-
Vergol Horizon correlatable with the Holmia Zone) were found
konur Karatau-Dzhebagly zone (in Russian). Izvestia of Academy of
some tens of metres above the Koktal Formation base. Strata Sciences of Kazakh SSR. Geological series, 6, 31 –43.
between those with Early Cambrian microfossils and the Baykonur Guan, B., Wu, R., Hambrey, M. J. & Geng, W. 1986. Glacial sediments
cap dolomite were formed in a basin experiencing condensed sedi- and erosional pavement near the Cambrian – Precambrian boundary
mentation. This suggests that their accumulation, although occur- in western Henan Province, China. Journal of Geological Society,
ring over a long time, would be unlikely to have taken more than London, 143, 311–323.
several million years. From these considerations an inference Kheraskova, T. H. 1981a. Late Precambrian tilloid of Baykonur
can be made that the Baykonur Formation was formed during Formation in Ulutau Mountains, Central Kazakhstan, USSR. In:
the Late Vendian or initial Early Cambrian, that is, during the Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene
latest the Ediacaran or Early Cambrian according to the scale of Glacial Record. Cambridge University Press, Cambridge,
the International Commission on Stratigraphy. 348– 352.
Glaciogenic deposits similar to the Baykonur Formation depos- Kheraskova, T. H. 1981b. Late Precambrian tilloid of Kopal Formation
its are also known in regions adjacent to Kyrgyzstan in China, in the Atasu-Mointy interfluve, Central Kazakhstan, USSR. In:
namely in the Kuruktag Range (Wang et al. 1981; Chen et al. Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene
1981). The glacial Hankalchough Formation has a very similar Glacial Record. Cambridge University Press, Cambridge,
stratigraphic position to the Baykonur Formation, just below 358– 360.
deposits with Lower Cambrian fossils (Zhu et al. 2011). The Kheraskova, T. N. 1986. Vendian-Cambrian Rock Associations of
basal member of the Lower Cambrian in the Kuruktag Range con- Caledonides of Asia (in Russian). Nayka, Moscow.
sists of black carbonaceous chert containing phosphorite, as in Kheraskova, T. N., Didenko, A. N., Bush, V. A. & Volozh, Yu. A.
Kyrgyzstan and Kazakhstan. A similar stratigraphic position has 2003. The Vendian– Early Paleozoic history of continental margin
of Eastern Paleogondwana, Paleoasian Ocean, and Central Asian
also been reported for the glacial Hongtiegou Formation of
Foldbelt. Russian Journal of Earth Sciences, 5, 165–184.
Chaidam Basin (Shen et al. 2010) and Luoquan Formation Kholodov, B. N. 1973. Sedimentary Ore Genesis and Metallogeny of
of North China (Mu 1981; Guan et al. 1986) and glacial deposits Vanadium (in Russian). Nauka, Moscow.
of the Bokson Formation (Chumakov 2009, 2011). I have com- Kiselev, V. V. 2001. Analogues of the Sinian complex in the central and
bined all enumerated formations as the Baykonurian Glacial northern Tien Shan. Geology and Geophysics, 42, 1453– 1463.
Horizon (Chumakov 1978, 1981). The widespread extent of the Knipper, A. L. 1963. Tectonics of Baikonur Synclinorium (Central
Baykonurian Glacial Horizon points to a significant glacial event Kazakhstan) (in Russian). Publishing Office Academy of Sciences
occurring close to the Precambrian –Cambrian boundary that of the USSR, Moscow.
affected the Kazakhstan, Kyrgyzstan, Tuva-Mongolian and Knoll, A. H. 2000. Learning to tell Neoproterozoic time. Precambrian
Tarim microcontinents as well as the North China Craton. It is Research, 100, 3– 20.
possible that the large negative d13C anomaly that occurred close Korolev, V. G. 1963. About Cambrian boundaries in the Middle Asia
to the lower boundary of the Nemakit-Daldyn Horizon (Knoll (in Russian). Trudy Frunzinskogo politekhnicheskogo institute, Geo-
2000; Zhu et al. 2007; and others) is related to this Baykonurian logiya, Gornoe delo, 10, 16 –21.
Glaciation. Korolev, V. G. & Maksumova, R. A. 1984. Precambrian Tillites and
Tilloids of Tien Shan (in Russian). Ilim, Frunze.
Research was supported by grants from the Russian Fund of Basic Investigations Krylov, N. N., Sergeev, V. N. & Kheraskova, T. N. 1986. Discovery
(Nos 11-05-00232 and 10-05-00294), and Program No. 25 of the Presidium of the of Cambrian microfossils in deposits of Baikonur Synclinorium
Russian Academy of the Sciences. The author would like to acknowledge stimu- (in Russian). Izvestiya of Academy of Sciences of the USSR, Seriya
lating discussions arising from participation in IGCP project 512 ‘Neoproterozoic geologicheskaya, 1, 51– 56.
Ice Ages’. This represents a contribution of the IUGS- and UNESCO-funded Makarychev, G. I. 1967. Stratigraphy of Protorozoic and Lower Paleo-
IGCP (International Geoscience Programme) project #512. zoic deposits of Bol’shoy Karatau (in Russian). Bulleten’ Moskovs-
kogo obshchestva ispytateley prirody, XXXII, 31 –46.
Mu, Y. 1981. Luoquan tillite of the Sinian System in China. In: Hambrey,
References M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial
Record. Cambridge University Press, Cambridge, 402–413.
Ankinovich, S. G. 1961. Lower Paleozoic vanadium-bearing basin of the Shen, B., Xiao, S., Zhou, G., Kaufman, A. J. & Yuan, X. 2010. Carbon
North Tien Shan and West Margin of Central Kazakhstan (in and sulphur isotope chemostratigraphy of the Neoproterozoic Quanji
Russian). Part I, Alma-Ata, 272. Group of the Chaidam Basin, NW China: Basin stratification in the
BAYKONUR FORMATION, KAZAKHSTAN AND KYRGYZSTAN 307

aftermath of an Ediacaran glaciation postdating the Shuram event? Zhu, M. & Wang, H. 2011. Neoproterozoic glacigenic diamictites in
Precambrian Research, 177, 241– 252. the Tarim Block, NW China. In: Arnaud, E., Halverson, G. P. &
Volin, A. V. 1966. Breccias of sliding and tillites in relation with problem Shields-Zhou, G. (eds) The Geological Record of Neoprotero-
glaciations and moving of poles (in Russian). In: Materials k sovesh- zoic Glaciations. Geological Society, London, Memoirs, 36,
chaniyu ‘Obshchie sakonomernosti geologicheskikh yavleniy’. 367– 378.
Leningrad, 1, 31 – 46. Zhu, M., Strauss, H. & Shields, G. A. 2007. From snowball earth to the
Wang, Y., Lu, S., Gao, Z., Lin, W. & Ma, G. 1981. Sinian tillites of Cambrian bioradiation: calibration of Ediacaran –Cambrian earth
China. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre- history in South China. Palaeogeography, Palaeoclimatology,
Pleistocene Glacial Record. Cambridge University Press, Cam- Palaeoecology, 254, 1 –6.
bridge, 386– 401. Zubtsov, E. I. 1972. Precambrian tillites of Tien Shan and their strati-
Zaytsev, Y. A. & Kheraskova, T. N. 1979. Vendian of Central Kazakh- graphic significance. Bulleten’ Moskovskogo obshchestva ispytateley
stan (in Russian). Moscow University Publisher, Moscow. prirody, XLVII, 42 – 56.
Chapter 27

The glaciogenic Bol’shoy Patom Formation, Lena River, central Siberia

NICKOLAY M. CHUMAKOV1*, BORIS G. POKROVSKY1 & VICTOR A. MELEZHIK2


1
Geological Institute Russian Academy of Sciences, Pyzhevskiy 7, Moscow 119017, Russia
2
Geological Survey of Norway, Leiv Eirkssons Vei 39, N-7491 Trondheim, Norway
*Corresponding author (e-mail: chumakov@ginras.ru)

Abstract: The Bol’shoy Patom Formation (Fm.) is part of the Upper Precambrian Patom Supergroup, which comprises a siliciclastic and
carbonate succession divided into (from base to top) the Ballaganakh, Dal’nyaya Tayga and Zhuya Groups (Gr.). The supergroup was
deposited within the bay-like passive margins of the Siberian craton. The Bol’shoy Patom Fm. is the lower unit of the Dal’nyaya Tayga
Gr. Massive and stratified diamictites with subordinate sandstones, mudstones, siltstones, conglomerates and conglo-breccias form the
Bol’shoy Patom Fm. New biostratigraphic (microfossils of Pertatataka type, Ediacaran fossils) and chemostratigraphic data (87Sr/86Sr,
d13C) point to a Vendian (Late Cryogenian and Ediacaran) age for the Dal’nyaya Tayga and Zhuya Groups. Unsorted diamictite matrix, a
very wide range of size and roundness of dispersed erratic clasts, the presence of glacial grooves on boulders, dropstones and till pellets in
laminated mudstones and siltstones are all evidence for intense ice rafting and ice-shelf sedimentation in the northeastern part of the
Patom basin during deposition of the Bol’shoy Patom Fm.

A ‘conglomerate’ (diamictite) unit was first identified in the lower and volcanic rocks (metamorphic andesites and andesito-basalts).
reaches of the Bol’shoy Patom River and on the southern banks of The supergroup succession points to the cyclic progradation of a
the Lena River (10 km upstream of the mouth of the Bol’shoy sedimentary wedge from craton margins into the basin (Chumakov
Patom River) in 1935 by Z. M. Starostina and formally named in et al. 2007).
1941 as the Bol’shoy Patom Fm. by A. A. Predtechensky (Chuma-
kov 1959). This name has been used in many publications up to
today (Keller 1963; Bobrov 1979; Chumakov 1993; Vorob’eva Stratigraphy
et al. 2007), although some authors incorrectly refer to it as the
Dzhemkukan Fm., based on correlation with a unit of very differ- The Patom SGr. comprises a siliciclastic and carbonate succession
ent lithology (shales and sandstones) in the Zhuya River area divided into three parts (from base to top): the Ballaganakh, Dal’-
(Ivanov et al. 1995 and others). The best outcrops of the nyaya Tayga and the Zhuya Groups (Gr.). The general succession
Bol’shoy Patom Fm. are in canyons of the Bol’shoy Patom and and lithostratigraphy of the two upper groups are shown in
Ura River. The upper part of the formation is seen along the Figure 27.2 and Table 27.1. The Bol’shoy Patom Fm. is the
Lena River. The type section of the formation is situated on the lower unit of the Dal’nyaya Tayga Gr. Lateral step-by-step litho-
eastern limb of the Zheday anticline in a canyon of the Bol’shoy stratigraphic correlations of the groups and their formations,
Patom River. using some robust lithological and isotopic markers, allow most
Lungersgauzen (1963) first suggested a glacial origin for the dia- formations to be traced more than 1000 km along the marginal
mictites of the Bol’shoy Patom Fm. (he called it the Ust-Patom folded arch of the Patom fold system. Beside the Ura Uplift, dia-
Fm.). Lungersgauzen described them as lithified basal moraines mictite formations also occur in the lower part of the Dal’nyaya
and fluvioglacial deposits. Subsequently, evidence for bedding, Tayga Gr. in two other uplifted areas of the folded arch: at its
graded bedding and subaqueous slumps in the diamictites led northwestern flank (Bol’shaya Chuya R. – ‘Dzhemkukan’ Fm.)
Lungersgauzen’s interpretation to be modified (Chumakov 1965; and its eastern flank (Dzhelinda and Chara Rs. – Nichatka Fm.).
Chumakov & Krasil’nikov 1991). There were three main cycles of progradation of Patom SGr.
sedimentation represented by the Ballaganakh, Dal’nyaya Tayga
and Zhuya Groups, respectively (Chumakov et al. 2007). The
Structural framework lower Ballaganakh Gr. is up to 6.5 km thick in the inner part of
the Patom arch. It thins to 0.5 km at the craton margin and thins
The Bol’shoy Patom Fm. forms the cores of two anticlines that out completely on the craton. The group consists of conglomerates
exhibit a northeastern trend and an en-echelon-like arrangement. at the base, sandstones in the more substantial middle part and car-
The anticlines form the apex of the transverse Ura Uplift, which is bonates in the upper part (Mariinskiy Fm.). The Mariinskiy Fm.
oriented perpendicular to a huge marginal arch of folds bordering contains limestones and shales in the inner part of the Patom
the Patom fold system. The uplift consists of the simply folded, arch and stromatolitic and oolitic dolomites of carbonate platforms
thick (.6 km) Neoproterozoic Patom Supergroup (SGr.). The at the margins of the craton. The Dal’nyaya Tayga Gr. was depos-
more southerly Zheday Anticline (Fig. 27.1b) crosses the lowermost ited conformably on the Mariinskiy Fm. in the deep anoxic basin of
course of the Bol’shoy Patom River and the middle course of the the Patom zone and bordering arch. Margin disconformities occur
Lena River south of the settlement Chapaevo. The Ura anticline at the base of the Dal’nyaya Tayga Gr. Near the edge of the craton
(Fig. 27.1a) is located to the east of the Lena River along the Ura the Mariinskiy Fm. was preserved as thin erosional relicts only.
River. Both anticlines have flat crests and moderate limb dips The Dal’nyaya Tayga Gr. consists of black slates and limestones
(30–608). Rocks of the Bol’shoy Patom Fm have been slightly and some sandstones in the Patom basin and in its margin fold
altered, mostly by early epigenetic processes. Faint cleavage of the arch. The Dal’nyaya Tayga Gr. is divided in this region into
shale is developed only at the southwestern end of the Zheday fold. three formations (from the bottom): Dzhemkukan, Barakun and
The Patom SGr. was deposited within the bay-like passive Valyukhta. To the NE, in the Ura uplift, black shales and sand-
margins of the Siberian craton. The supergroup disconformably stones of the Dzhemkukan Fm. give way to diamictites and associ-
overlies older rift structures filled by conglomerates, sandstones ated rocks of the Bol’shoy Patom Fm. The lower part of the

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 309– 316. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.27
310 N. M. CHUMAKOV ET AL.

Fig. 27.1. Geographic distribution and structural framework of the Bol’shoy


Patom Fm. Key: 1, Jurassic System and post Jurassic deposits; 2, Ordovician
System; 3, Cambrian System; 4 –8 Neoproterozoic: 4, Tinnaya and Zherba
Fms.; 5, Zhuya Gr.; 6, Upper and middle parts of Dal’naya Tayga Gr.; 7,
Bol’shoy Patom Fm. (lower formation of Dal’naya Tayga Gr.); 8, Ballaganakh
Gr.; 9, main faults; a, Ura Anticline; b, Zheday Anticline.

Valyukhta Fm. is replaced here by siltstones and limestones (Ura


Fm.) and its upper part is replaced by stromatolitic and oncolitic
dolomites and limestones of the carbonate platform (Kalancha
Fm.). Similar lateral facies changes in the lower and upper parts
of the Dal’nyaya Tayga Gr. can also be observed on uplifted
parts of the eastern and western flanks of the Patom arch (Dzhe-
linda and Bol’shaya Chuya Rivers, accordingly) and at the
western edge of the Aldan Shield beside the Patom arch (Chara
River). Carbonate platforms of the upper part of the Zhuya Gr. Fig. 27.2. Lithostratigraphic section, palaeontological data and C- and
drape all older deposits from the Patom arch up to the edge of Sr-isotope variations in carbonate rocks of the upper part of the Patom SGr. and
the Aldan shield, thinning in this direction from 600 m to 200 m. lowermost Cambrian system. Key: 1, diamictites (glacial and reworked glacial
deposits); 2, carbonate conglo-breccias; 3, conglomerates; 4, sandstones; 5,
siltstones; 6, limestones; 7, sandy limestones; 8, oolitic limestones; 9,
stromatolitic limestones; 10, dolomites; 11, marls; 12, Pertatatataka microfossil
Glaciogenic deposits and associated strata
assemblage; 13, Nemakit-Daldyn small shelly fossils; 14, Early Cambrian
fossils (sunnaginicus biozone); 15, erosional boundaries; 16–18, d13C
The thickness of the Bol’shoy Patom Fm. is up to 1100 m. Differ- variations in sections of the Patom SGr. (Pokrovsky et al. 2006): 16, Ura Uplift;
ent kinds of diamictites make up c. 60– 80% of the sections as well 17, Zhuya R.; 18, Boreholes at the left bank of Lena R.; d13C variations in late
as subordinate sandstones, mudstones, siltstones, conglomerates Cryogenian –Cambrian (Vendian-Cambrian) sections: 19, Ura Uplift (Sochava
and conglo-breccias. It is impossible to describe here the detailed et al. 1996; Pelechaty 1998); 87Sr/86Sr: – 0.70725 – 0.70855 (Gorokhov et al.
succession of the formation owing to its great thickness and the 1995; Vinogradov et al. 1996; Pokrovsky et al. 2006).
frequent intercalation of rocks of different types (Fig. 27.3). In
general, the formation can be divided into three informal members.
Following geological sections of the Bol’shoy Patom Fm and
enclosing deposits in the Ura Uplift (from top to base): Member II: frequent intercalation of graded bedded sandstones,
siltstones, carbonate conglo-breccias and dark grey faintly
Lower part of Barakun Fm. bedded diamictites with slump structures, rare massive diamic-
Member II: black shales with intercalations of black limestones tites, graded diamictites; there are intercalations of cross-bedded
(.200 m) fine sandstones; lenses and scours filled by sandstones and con-
Member I: dark grey, thinly bedded pelitomorphic dolomite that glomerates; rare dropstones, slump structures (150 –320 m)
conformably overlies unit 4 (c. 6 m) Member I: intercalation of dark grey massive and faintly bedded
diamictites with graded bedded sandstone and mudstones beds
Bol’shoy Patom Fm. and small sandstone lenses. Rare slump and small cross
Member III: at the upper boundary of the member, thin polymictic bedding structures (450 –700 m)
silty, fine sandstone occurs with rare coarse grains (1 mm);
below are dark-grey massive and sometimes faintly bedded dia- Upper part of Mariinskiy Fm.
mictites; rare slump structures, small sandstone and conglomer-
ate lenses and interbeds (120 –180 m) Grey sandy limestone (.200 m)
BOL’SHOY PATOM FORMATION 311

Table 27.1. Succession of the Patom SGr. in the Ura Uplift

Zhuya Gr.
Chencha Fm. White, pale and pink micritic, stromatolitic and oolitic Limestones, silty and sandy limestone in upper part 450– 600 m
Nikol’skoe Fm. Red and green thin-bedded marls with intercalations of limestone; sandstone member at base 500– 600 m
Dal’nyaya Tayga Gr.
Kalancha Fm. Dark grey limestone and pale stromatolitic, oncolitic and oolitic dolomites; siltstone intercalations 550– 60 m
Ura Fm. Grey bedded mudstones, siltstones with intercalations of dark limestone in upper part and fine sandstones in lower part 350 m
Barakun Fm. Black limestone, partly oolitic and brecciated in upper part; black shales with intercalation of black limestone in lower part; thin-bedded dark
dolomite at the base 600 m
Bol’shoy Patom Fm. Diamictites, shales, sandstones, minor conglomerates 900–1100 m
Ballaganakh Gr.
Mariinsky Fm. Grey limestones, sandy limestones and black shales 500 m
Bugarikhta Fm. Sandstones, grey shales, minor gravelly sandstones .1000 m

Diamictites the northwestern Aldan Shield, which at that time would have
been situated no less than 100 km to the east of the Ura uplift
Diamictites have a grey or dark grey (almost black) fine matrix and crystalline basement of the southern part of Siberian craton.
with randomly dispersed, usually paler clasts. The matrix ranges Clast roundness varied between 0 and 4, but mean roundness is
between silty sandstone to sandy mudstone. Diamictites with a generally between 0.7 to 1.2 (using the Khabokov 1933 scale).
mudstone matrix are darker. Matrix is arkosic (quartz grains, In general, crystalline stones are more rounded (1 –3), although
30– 70%; plagioclase, 20–58%; microcline, up to 18%). Minor some angular and subangular slabs of granites and gneisses are
components include quartzites (1 –12%), gneisses and granitoids found too. Two main types of diamictites can be identified:
(2 –10%), carbonates (1 –6%) and mica (1– 5%). Roundness of massive and stratified (Chumakov & Krasil’nikov 1991).
sand grains varies, but in general is low (less than 1.5 on the
five-number scale of Khabakov 1933, and 0.45 on the scale of
Krumbein & Sloss 1951). Grains are cemented by sericite, rarely Massive diamictites. Massive diamictites are the most typical rocks
carbonate, chlorite and detrital biotite. of the Bol’shoy Patom Fm., particularly in the lower and upper
Clast content varies from 1 to 15%, commonly 3 –10%. Small members. They form 50 –70% of formation sections. The thick-
and medium-sized pebbles are most typical, cobbles and boulders nesses of individual units of these diamictites vary from 3 to
are also common with a few blocks reaching 1.1–1.7 m across. 70 m. Only indistinct clast concentrations can be recognized some-
Clasts are randomly scattered in the matrix, often forming nests times as diffuse horizons or lenticular bodies. Large boulders occur
of 3–4 stones and occasionally thin stone lags. Clast composition individually or as nests at certain stratigraphic levels. Small lenses
is quite uniform: granitoids (45 –65%), granite-gneisses (20– 35%) (0.1 –0.3  1– 2 m) of sandstones or sandy siltstones of slightly
and gneisses (5 –30%). Other clast types (5 –15%) include lime- concavo-convex shape occur rarely in some massive diamictite
stones, dolostones, sandstones, amphibolites, quartzites, quartz units. Commonly, the lenses have a gradational lower boundary
and schists. The source area for the granitoids and gneisses was and a sharp upper boundary sometimes with very thin silt

Fig. 27.3. Composite sections of the


Bol’shoy Patom Formation (parts of
sections not limited by line from right side
are based on eluvial debris and small
dispersed outcrops, white indicates absence
of outcrops). Key: 1, massive diamictites; 2,
faint bedded diamictites and diamictites
with relicts of bedding; 3, carbonate
conglo-breccies; 4, conglomerates; 5,
gritstones; 6, gravelly sandstones; 7,
sandstones; 8, sandstones with traces of
thermal metamorphism; 9, siltstones; 10,
sandy limestones, sandstones and siltstones
containing many carbonate grains; 11,
limestones; 12, dolomites; 13, crystalline
and sedimentary clasts; 14, carbonate clasts;
15, cross-bedding; 16, wavy bedding; 17,
normal graded bedding; 18, reverse graded
bedding; 19, sandstone lenses; 20, erosional
channels and scours; 21, ripple marks; 22,
striated boulders; 23, slump rolls; 24, slump
boudins; 25, slumpfolds; 26, dropstones; 27,
lonestons; 28, stone nests; 29, broken
stones; 30, sole marks; 31, stone fabrics
(hatch lines from Lungersgauzen 1963); 32,
dominant direction of cross-bedding (hatch
lines from Lungersgauzen 1963).
312 N. M. CHUMAKOV ET AL.

laminae. Some lenses have an erosional lower boundary with basal Sandstones and grits. These make up between 10 and 25% of sec-
grits, pebble conglomerates or single boulders (Fig. 27.4a). tions. The mineral composition of sandstones is similar to that of
Some of the massive diamictites contain rare small fragments of the diamictite matrix. Limestone grains occur only in the middle
sandstone or siltstone beds partly reworked by slumps as folds, member of the formation on the western limb of the Zheday
rolls, boudins, ragged fragments and other relicts of sedimentary fold, where beds of fine sandstones occur with between 10 and
beds. Some slump fragments have faint and gradational boundaries 30% limestone clasts. The thicknesses of sandstone beds generally
with the matrix, forming clods, clots and ‘noses’, implying slumps range from 2 to 10 m. In the Lena River section, fine- and medium-
of semi- and unlithified sediments. Massive diamictites containing grained sandstone forms units up to 40 m thick. There are three
slump structures constitute 4– 30% of sections. types of sandstones. The first main type of sandstones exhibits pro-
Stone nests and broken stones are common. Flat surfaces of nounced normally graded bedding: from grits and coarse sand-
single quartzite boulders show one or two systems of longitudinal stones at the base to medium and fine-grained sandstones at the
sub-parallel grooves and striae of different sizes. Elongated stones top. Thin siltstone laminae occasionally overlie these sandstones.
in some massive diamictites had primary orientations in one predo- Graded-bedded sandstones are often associated with graded-
minant direction and one or more crossing it in subdominant direc- bedded diamictites. Thin beds of reverse graded-bedded sand-
tion(s) (Fig. 27.3). At Lena River, the predominant direction has a stones occur rarely.
NW –SE or north –south, or more rarely a NE –SW or WNW – ESE The second type of sandstones has restricted development in
orientation. At Ura River, the predominant direction has a north – the middle member of the Bol’shoy Patom Fm. They are
south orientation. thinly bedded silty, fine-grained sandstones, often with small
climbing ripple cross-bedding. These sandstones form individual
Stratified diamictites. Stratified diamictites are of two types: bedded units up to 40 m thick. The cross-beds dip generally to the south
and graded-bedded. Bedded diamictites have matrices that are at Ura River and to the NW at Lena River. The third type of
muddy, silty or sandy to varying degrees, and exhibit a range of sandstones includes fine, thinly bedded sandstones with abun-
clast sizes. Some bedded diamictites have a laminated muddy dant carbonate grains.
matrix. The thicknesses of individual laminae vary from 0.01 to
1 cm. Rare pebbles in these rocks are angular or subangular.
Another form of bedded diamictites has a silty to sandy matrix Conglomerates. These occur rarely as thin lenses. It seems that they
with bed thicknesses from 0.1 to 0.4 m or more. Most such beds are residual, a result of the winnowing of diamictites by bottom
have a lenticular or irregular form (Fig. 27.4b). currents. Other rare types of coarse sandstones, grits and conglom-
Graded-bedded diamictites are characterized by clear vertical erates fill erosion channels 0.5–10 m deep and several tens of
changes in the granulometry of matrix and clasts. The thicknesses metres wide. In places, the channels have rather steep slopes.
of graded-bedded diamictites vary from 1 to 60 cm. These diamictites
are often related to the lower or middle parts of cyclic alternations of Mudstones and siltstones. These form beds, and more rarely units,
diamictites, sandstones and siltstones (Fig. 27.4c, d). The thickness of between a few metres and a few tens of metres thick. The
of individual cycles varies from 0.5 to 3 m, while the total thickness rocks exhibit rhythmic lamination and contain dispersed lone-
of the cyclic units can reach tens of metres. Some cycles have normal stones and small till pellets. Some lonestones can be identified as
grading (Fig. 27.4c), whereas others have reverse grading in the lower dropstones, displaying splash features, with rupture of laminae
parts and normal grading in the upper parts of the cycles (Fig. 27.4d). below and draping of laminae above the clasts (Fig. 27.4e).
In this latter case there are often rip-up structures with abundant
mudstone chips at the base of cycles. Large stones or fragments of
sedimentary rocks, similar to rafts of debris flow, occur sometimes Carbonate conglo-breccies. These form a number of beds from 0.1
in the upper or middle parts of the cycles. to 0.5 m thick in the middle member of the formation. They have a
siliciclastic, carbonate-bearing sandstone matrix and contain abun-
dant (from 30 to 70%) angular and subangular fragments (from
Associated sedimentary rocks small pebbles to blocks 1.2 m across) of oolitic and stromatolitic
dolomites and limestones. The matrix of the conglo-breccies has
Beside diamictites, the Bol’shoy Patom Fm. contains subordinate normal and reverse graded bedding. Some large stones occur in
rocks: sandstones and grits, conglomerates, siltstones and the middle or upper parts of the beds and are similar to rafts of
conglo-breccias. debris flows.

Fig. 27.4. Some typical structures in


diamictites of the Bol’shoy Patom
Formation. (a) Sandstone lens with boulders
in silty diamictites. (b) Two thin irregular
beds of silty-sandy diamictites occur in silty
diamictite. (c) Normal grading of cycle. (d)
Reverse grading in lower and normal
grading in upper part of cycle. (e) Dropstone
and till pellets in laminated mudstone.1,
stones of granite and sandstone; 2, massive
dimictites; 3, bedded diamictites; 4, shale
chips; 5, sandstone; 6, shale.
BOL’SHOY PATOM FORMATION 313

Boundary relations with overlying and underlying relatively small (0.7079 – 0.7087), but is greater lower down in
non-glacial units the succession: Nikol’skaya Fm. (0.7079 –0.7094); Valukhta Fm.
(0.70765 –0.7092); Barakun Fm. (0.7025 –0.7102); Mariinskiy
The upper boundary of the Bol’shoy Patom Fm. is conformable. It Fm. (0.7075 –0.7092). Taking into account the low measured
was observed in two places. The best locality is situated 7.5 km Rb –Sr (,0.01), Mn –Sr (,0.1) and Fe –Sr (,2) ratios, we con-
from the mouth of the Ura River on its northern bank, and the sider that 87Sr/86Sr (min.) ¼ 87Sr/86Sr (primary) for all formations
second one is located at the beach on the southern side of the excluding, probably, the Mariinskiy Formation.
Lena River 6 km upstream of the Bol’shoy Patom River mouth. In general, isotope stratigraphic data from the upper part of
At Ura River, diamictite with a sandy mudstone matrix and rare the Patom SGr. show an increase in lowermost 87Sr/86Sr ratios
clasts gradually change upsection, at first to silty coarse sandstone from high-Sr limestones of from 0.70725 in the Barakun Fm. to
with dispersed pebbles and then to silty fine polymictic sandstone 0.7079 –0.7080 in the Nikol’sk and Chencha Fms.
with rare coarse grains. Transitional beds are c. 1 m thick. Fine
sandstone is conformably overlain by the first dolomite member
of the Barakun Fm. This dark grey or black laminated dolomite Palaeolatitudes and palaeogeography
is peloidal, with small asymmetric anticline-like structures
(0.5  0.5 m) with sharp angular aggrading crests resembling Systematic palaeomagnetic investigations of the Patom SGr. were
‘tepees’. made twice at the Paleomagnetic Laboratory of the Geological
The lower contact of the Bol’shoy Patom Fm. has not been Institute of the RAS, Moscow (by S. V. Shipunov) and then at
observed in outcrop, but the lower boundary of the Dal’nyaya the Paleomagnetic Laboratory of the US Geological Survey in
Tayga Gr. is believed to be erosional, at least along the basin Flagstaff (by D. Elston). All rocks of the supergroup appear to
margins. This opinion is based on the existence in the Bol’shoy have been remagnetized during Palaeozoic times.
Patom Fm. and in the lower part of the coeval ‘Dzhemkukan’
and Nichatka Fms. of big carbonate clasts of rocks of the Mariins-
kiy Fm. (Chumakov 1993; Ivanov et al. 1995 and others). Geochronological constraints

From the beginning of the 1960s up to 2005, the Patom SGr. was
Chemostratigraphy dated on the basis of stromatolites and microphytolites to be
Middle and Late Riphean (Mesoproterozoic-Cryogenian) (Zhurav-
The C-isotope profile of carbonates of the Patom SGr. exhibits an leva et al. 1961; Keller 1963; Chumakov 1993; and others) or Late
alternation of high and low d13C intervals (Fig. 27.2) with two Riphean (Tonian-Cryogenian) (Khomentovskiy et al. 1998). Pre-
extraordinary features: large-amplitude d13C variation, exceeding sumed correlation between the Russian stratigraphic chart on
21‰ (– 13.5 to 8.3), and an enormous thickness (up to 1000 m) of Late Precambrian and the stratigraphic chart of the International
13
C-depleted strata (d13C, , –7.5) Zhuya Gr. (Pokrovsky et al. Commission on Stratigaphy is shown in Table 27.2.
2006). First suggestions of a Vendian (late Cryogenian-Ediacaran) age
The Mariinskiy Fm., underlying the Bol’shoy Patom (Dzhemku- for the upper part of the Patom SGr. were made by J.K. Sovetov,
kan) Fm., exhibits a 600 m positive ‘plateau’ around d13C  7– who correlated tillites of the Prisayan’ya (South Siberia) with the
8‰. Negative d13C values have been found in the carbonate- Bol’shoy Patom Fm. (Sovetov 2002). The Zherba and Tinnaya for-
bearing sandstone of the upper part of the Bol’shoy Patom Fm. mations overlying the Patom SGr. were suggested to be Vendian
at the Ura River (28.8‰) and in the lower part of Dzhemkukan (Ediacaran –lowermost Cambrian) by all investigators of the
Fm. at the Zhuya River (29.1‰) where diamictites are absent. region. This opinion was recently confirmed by the discovery of
Cap dolostones at the Ura River have moderately low d13C small shelly fossils of the Nemakit-Daldyn stage in the uppermost
values from –3.3 to –4.2‰. Upsection there is a second positive part of the Tinnaya Fm. and fossils of the Cambrian Sunnaginicus
‘plateau’ within the Barakun (d13C  6–8‰) and Valyukhta Biozone of the Tommotian Stage in the lowermost part of the
(d13C  4–6‰) Fms. with an overall thickness near 1.5 km, Nokhtuysk Fm. (Khomentovsky et al. 2004).
which is overlain by the 300-m-thick Nikol’skoe Fm. consisting There are no precise radiometric age data pertaining to the
of variegated micritic and peloidal limestones, marls and oncolitic Patom SGr. Numerous data were obtained using the K –Ar and
limestones, which are all depleted in 13C, with d13C values as low Rb –Sr methods on whole sedimentary rocks of the upper part of
as –13.5‰. The strongly 13C-depleted (in the range d13C  –7.5 Dal’nyaya Tayga Gr., revealing strong epigenetic rejuvenation
to 10‰) carbonates continue upwards, where they also character- (452 –531 Ma; Vinogradov et al. 1996). A wide range of ages
ize the Chencha Fm., a c. 600–700-m-thick unit of micritic lime- was obtained on gabbro-dolerite sheets intruded between the
stones, stromatolitic limestones and pink oncolitic limestones. Bol’shoy Patom and Barakun Fms. The hypabyssal nature of the
Ultrahigh-d13C carbonates of the Dal’niyaya Tayga Gr., and sheets implies that they were formed at depths of 2 –4 km.
ultralow-d13C of the Zhuya Gr. have been traced around the Dating by the K –Ar whole-rock method shows ages between
Patom Basin to the east of this basin (at Chara River and the Chara- 520 and 633 Ma (Oleynikov et al. 1983). These data can represent
Tokko watershed) without significant variation in C-isotope com- only very approximate age constraints. Taking into consideration
position. The thickness of high-d13C dolomites of the Sen’ Fm. the thickness of overlying deposits, it is possible to suggest that
(correlative of Ura and Kalancha Fms.) at the Chara River does
not exceed 200–250 m, and low-d13C limestones of the Torgo
Fm. (correlative of Chencha Fm.) extend over 300– 350 m. Table 27.2. Correlation between Vendian and Ediacaran Systems
Further to the east, the Torgo Fm. (150 m) partly consists of dolo-
mites with the same d13C ( –7.5 to –10), but higher d18O (up to Cambrian Lower Tommotian Stage
Cambrian Lower
Nemakit-Daldynian Horizon
30‰; Pokrovskii et al. 2006). Upper Kotlin Horizon
High Sr contents are typical for limestones of the Patom SGr. Vendian Redkino Horizon
Ediacaran
They reach 11 700 ppm in the Chencha Fm., 2500 ppm in the Laplandian
Glusk Gaskiers

Barakun and Valukhta Fms., 2200 ppm in the Mariinskiy Fm.; Lower Horizon
Upper Blon'
Lower Blon'
only in the Nikol’skaya Fm. do they not exceed 400 ppm. The
range of measured 87Sr/86Sr ratios (data from Gorokhov et al. Cryogenian
Upper Riphean
1995; Vinogradov et al. 1996; Pokrovskii et al. 2006 and their glacial horizons
unpublished data) in the limestones of the Chencha Fm. is
314 N. M. CHUMAKOV ET AL.

the age of the sheets is close to the age of the upper part of the Ultra-low-d13C carbonates of the Zhuya Gr. could be correlated,
Patom SGr. or deposits immediately overlying. A wide range of besides anomaly ‘Dounce’, with the negative d13C anomalies of
ages was also obtained using the K –Ar glauconite method for the Shuram and Buah formations (Oman) and probably the
the Zherba Fm. overlying the Patom SGr., giving 570–604 Ma similar anomaly in the Wonoka Fm. (S. Australia). Sedimentologi-
(Ivanov et al. 1995, corrected with constants from Steiger & cal and geochemical parameters suggest that the Zhuya Gr. had
Jäger 1977). preserved its primary C-isotope composition and was formed
New biostratigraphic and chemostratigraphic data allow us to over at least 5 million years (a similar duration is suggested for
reinterpret radically the ages of the Dal’nayay Tayga and Zhuya the Dounce anomaly). The negative C-isotope anomaly of the
Groups. Ediacaran fossils Beltanelloides sorichevae were found Zhuya Gr. exceeds all known C-isotope anomalies both in ampli-
in the lower part of the Barakun Fm. at Ura River 8.5 km from tude and thickness of strata, providing evidence of unprecedented
its mouth (Leonov & Rud’ko in press). New maceration and rein- ‘contamination’ of the water column by low-13C organic carbon.
vestigation of very rich microfossil assemblages from shales of the If the Nantuo Formation is really coeval with the Yerelina Sub-
Ura Fm. collected by us on the northern bank of the Ura River, group, which remains to be proven, then the upper part of the
3.5 km from its mouth, point to similar age (Vorob’eva et al. Dal’yaya Tayga Gr. and the Zhuya Gr are Ediacaran in age.
2007). This assemblage contains abundant large acanthomorphic According to the Russian stratigraphic chart, both the Dal’yaya
acritarchs of genera Ericiasphaera, Tanarium, Appendisphaera, Tayga and Zhuya groups belong to the Lower and Middle Vendian.
Meghystrichosphaeridium, Sinosphaera and taxa very similar to
the genera Alicesphaeridium, Variomargosphaeridium, Dicrospi-
nasphaera and others. Ericiaspaera adspersa dominates the Discussion
biota, while Tanarium conoideum, Appendisphaera tenuis, ?Sino-
sphaera rupina and Meghistrichosphaeridium hadianensis also Three rock associations can be recognized in the Lower Vendian
occur. There are also numerous morphologically complicated, (Late Criogenian) Bol’shoy Patom Fm. The first and main rock
new acanthomorphic forms. Most of the listed taxa are known associations comprise massive diamictites with subordinate
from a number of post-glacial formations: the post-Olympic bedded diamictites and laminated mudstones and siltstones. This
(?post-Marinoan) Pertatataka Fm. of Central Australia (Gray association is typical of the lower and upper members of the
2005), the post-Nantuo Doushantuo Fm of South China (Zhou & Bol’shoy Patom Fm. Intercalation with laminated mudstones and
Xiao 2007), the post-Blaini Infrakrol Fm. of the Lesser Himalayas siltstones, bedded diamictites and the existence of sandstone
(Tiwary & Knoll 1994), as well as the pre-Redkino Vychegda Fm. lenses related to winnowing of diamictites by local bottom currents
of the Russian Plate (Veis et al. 2006; Vorob’eva et al. 2006) and implies that massive diamictites were deposited in a rather quiet
the upper part of the Nepa Horizon of the Siberian Plate. Associ- basin. Laminated mudstones and siltstones undisturbed by waves
ations of such microfossils are usually referred to as ‘Pertatataka- and traces of slumping show that the basin floor was rather deep
type biota’. If confirmed, the Ura biota would appear to be similar and sloped. It was a large basin, probably open to the sea, as the
to the second microfossil zone of ECAP of K. Grey of Central Aus- Bol’shoy Patom Fm. is gradually replaced towards the south and
tralia (Vorob’eva et al. 2007). The presence of acanthomorphic SW by the thick Dzhemkukan Fm. comprising distal turbidite
acritarchs in the Ura microbiota has been reported previously by slates with subordinate sandstones (c. 1000 m).
some other researchers. They considered them to be Palaeozoic Unsorted matrix, a very wide range of size and roundness of dis-
forms penetrating into the Ura Formation, or interpreted the micro- persed erratic clasts, the presence of glacial grooves on boulders,
biota as Late Riphean (Cryogenian) but with many Vendian dropstones and till pellets in laminated mudstones and siltstones
elements (for a review see Nagovitsin et al. 2004; Vorob’eva are all evidence for intense ice-rafting in the basin. A wide size
et al. 2007). The conclusion on the Late Riphean (Cryogenian) range and low mean roundness of stones are not typical for
age of the Ura microbiota was made by Nagovitsin et al. (2004), debris rafted by seasonal or perennial sea ice, which commonly
particularly from the finding of the form described incorrectly as transports well-rounded and well-sorted stones picked up from
Trachyhystrichosphaera aff. aimika. In the opinion of Vorob’eva coastal beaches. Therefore, it is possible to conclude that the
et al. (2007), the forms described in this paper (Nagovitsin et al. first rock association was deposited mainly by glacier and
2004) are really fragments of Early Vendian taxa, such as Tanar- iceberg rafting of material derived from shelf glaciers and outlet
ium, Alicesphaeridium and others. glaciers. Glacier-like clast fabrics in some massive diamictites
The assignment of the Dal’naya Tayga Group to the Early point to the presence of grounded shelf glaciers moving in
Vendian or Late Cryogenian –Early Ediacaran has been confirmed north –south and NW – SE directions. The basement of the Siberian
by a minimal age of 600 + 10 Ma from detrital zircons from the Craton was the main source of clastic deposits in the Bol’shoy
upper part of the Khomolkho Fm. (U –Th – Pb, LA-ICPMS, Patom Fm. This is clear from the predominance of granitoids,
Meffre et al. 2008). The Khomolkho Fm. is a stratigraphic ana- granite-gneisses and gneisses, which make up almost all of the
logue of the Valyukhta Fm. (Zhadnova 1961; Keller et al. 1967; cobbles and boulders in the massive diamictites. The source of car-
Ivanov et al. 1995, and many others). Khomentovsky & Postnikov bonate conglo-breccias was from carbonate platforms bordering
(2001) even suggest that the Khomolkho Formation is a strati- the Patom basin in Mariinsky time. Their erosion was related to
graphic analogue of the Ura Fm. This date therefore indicates a a fall in sea level during the Bol’shoy Patom glaciation. Cross-
maximal age of Valyukhta and Ura Fm. bedding in sandstones allows us to suggest that in the Lena –
These data allow us to suggest correlation of the Bol’shoy Patom Bol’shoy rivers area, glacial meltwaters flowed towards the
Fm. with the Nantuo, Blaini and Olympic Fms. In this case, the posi- north and NW and in the Ura River area to the SW and south.
tive d13C anomaly in the upper part of the Dal’nyaya Tayga Gr. may The presence of wet-based, tidal and outlet glaciers was probably
be correlated with a similar anomaly in the lower part of Doushantuo the main cause of the rather restricted areal distribution of iceberg
Fm., while the negative anomaly in the Zhuya Gr. may be correla- facies in the Patom Basin. In such environments, the main dis-
tive with a similar anomaly ‘Dounce’ in the uppermost member of charge of glacial debris from shelf glaciers and icebergs occurs
the Doushantuo Fm. (Zhou & Xiao 2007; Zhu et al. 2007). in a proximal position. In general, the first rock association
87
Sr/86Sr ratios of the uppermost part of the Dal’nyaya Tayga Gr. implies glacial and proximal glacial marine palaeoenvironments.
and of the Doushantuo Fm. (Yang et al. 1999) are consistent with The second and third rock associations are found predominantly
this interpretation (minimum values of 0.70765 and 0.7077, respect- in the middle member of the Bol’shoy Patom Fm. The second rock
ively). Correlations with the Nantuo and the Doushantuo formations association comprises intercalations of turbidite, debris and grain
allows us to constrain the age of the Dal’nyaya Tayga and Zhuya flow sandstones and massive or bedded diamictites. Rare cross-
Groups to between 636/660 and 550 Ma. bedded coarse sandstones and conglomerates filled channels cut
BOL’SHOY PATOM FORMATION 315

into graded-bedded strata. Laminated mudstones with dropstones Khabakov, A. V. 1933. A Brief Guide to Field Studies of Conglomerates.
also occur. The second rock association was probably deposited Nauchno-Tekhnicheskoe Izdatel’stvo, Moscow, 11 (in Russian).
as submarine fans. Their formation may be related to high-activity Khomentovskiy, V. V., Postnikov, A. A. & Fayzulin, M. Sh. 1998. Bai-
debris, grain and turbid flows related to sub-glacial or glacial rivers kalian of stratotype locality. Geology and Geophysics, 39, 1505–1517.
during glacial retreat. Khomentovskiy, V. V. & Postnikov, A. A. 2001. Neoproterozoic evol-
The third rock association consists of thinly bedded, fine- ution Baikal-Vilyu branch of Palaeoasian Ocean. Geotectonics, 3,
grained sandstones with small-scale cross-bedding of climbing 3– 21.
ripple-type and finely laminated sandstones with abundant carbon- Khomentovskiy, V. V., Postnikov, A. A., Karlova, G. A., Kochnev,
B. B., Yakshin, M. S. & Ponomarchuk, V. A. 2004. Vendian of
ate grains. These deposits are probably related to interchannel
Baikal-Patom Upland (Siberia). Geology and Geophysics, 45,
interfans and distal parts of submarine fans with bottom currents. 465– 484.
The sandstones with abundant carbonate grains could be distal Krumbein, W. C. & Sloss, L. L. 1951. Stratigraphy and Sedimentation.
turbidites transported by carbonate conglo-breccia debris flows. Freeman, San Francisco.
The third rock association may be related to a rise in sea level. Leonov, M. V. & Rud’ko, S. V. 2010. The find of Ediacarian fossils in
The above-mentioned data allow us to conclude that the Dal’nyaya Tayga Group (Ediacaran-Vendian of Patom Highland),
Bol’shoy Patom Fm. was deposited in the marine Patom basin in press.
during glaciation by and with the strong influence of ice sheets, Lungersgauzen, G. F. 1963. Tillites and tillit-like deposits (in Russian).
iceberg rafting and glacial melt waters. The glaciogenic origin of In: Keller, B. M. (ed.) Stratigraphy of SSSR. Upper Precambrian.
the Bol’shoy Patom Fm. is supported by the presence of a Nedra, Moscow, 566–577 (in Russian).
typical cap dolomite conformably overlying it (first member of Meffre, S., Large, R. R. et al. 2008. Age and pyrite Pb-isotopic
Barakun Fm., see section ‘Boundary relations with overlying and composition of the giant Sukhoi Log sediment-hosted gold
underlying non-glacial units’) that is very similar to such dolomites deposit, Russia. Geochimica et Cosmochimica Acta, doi: 10.1016/
elsewhere in Asia, Svalbard, West Africa, Australia and other j.sca.2008.03.005.
regions of the world and by well-founded correlation of this for- Nagovitsin, K. E., Faizullin, M. Sh. & Yakshin, M. S. 2004. New
mation with the Nichatka Fm., which crops out on the southeastern forms of Baikalian acanthomorhytes from the Ura Formation of
flank of the folded arch and western margin of the Aldan shield. In Patom Uplift, East Siberia. Novosti paleontologii i stratigrafii, 6– 7,
Nichatka Fm., continental glacial facies have been discovered 7– 19.
(Chumakov 1993). During deposition of the Bol’shoy Patom Oleynikov, B. V., Tomshin, M. D. & Kopylova, A. G. 1983. Basic rocks
Fm., the southern part of the Siberian Craton was covered by ice of Ura anticlinorium (in Russian). In: Oleynikov, B. V. (ed.) Petrol-
ogy and Geochemistry Late Precambrian Intrusive Basic Rocks of
sheets flowing into the Patom Sea.
Siberian Platform. Nauka, Novosibirsk, 146– 167 (in Russian).
Pelechaty, S. M. 1998. Integrated chronostratigraphy of the Vendian
We thank M. I. Buyakayte for help in field and isotopic researches, and N. G. Vor-
System of Siberia: implications for a global stratigraphy. Journal of
ob’eva and V.N. Sergeev for the investigation of microfossils from our collection.
Geological Society of London, 155, 957– 973.
This research was supported by grants of the Russian Fund of Basic Investigations
Pokrovsky, B. G., Melezhik, B. A. & Buyakayte, M. I. 2006.
(nos 08-05-00433 and 10-05-00294) and Program No 24 of the Presidium of
Geochemistry of isotopes C, O, Sr and S, chemostratigraphy and
Russian Academy of Sciences. This represents a contribution of the IUGS- and
environments of sedimentation of Late Precambrian deposits of
UNESCO-funded IGCP (International Geoscience Programme) project #512.
Patom trough. Lithology and Mineral Resourses, 5, 505–530.
Sochava, A. V., Podkovyrov, V. N. & Vinogradov, D. P. 1996. Vari-
ation of carbon and oxygen isotopes in Vendian– Lower Cambrian
References carbonate rocks of the Urinsky anticlinorium (South Siberian Plat-
form). Lithology and Mineral Resources, 31, 248–257.
Bobrov, A. K. 1979. Stratigraphy and paleogeography of Upper Pre- Sovetov, J. K. 2002. Vendian foreland basin of the Siberian cratonic
cambrian deposits of South Yakutiya. Yakutian Publishers, Yakutsk margin: Paleopangean accretionary phases. Russian Journal of
(in Russian). Earth Sciences, 4, 363–387.
Chumakov, N. M. 1959. Stratigraphy and tectonics of the south-western Steiger, R. H. & Jäger, E. 1977. Subcommission on geochronology:
part of the Viluy Depression. In: Shatsky, N. S. (ed.) Tektonika convention on the use ofdecay constants in geo- and cosmochronol-
SSSR. IV. AN SSSR, Moscow, 345– 460 (in Russian). ogy. Earth and Planetary Sciences Letters, 36, 359– 362.
Chumakov, N. M. 1965. Certain Precambrian tillite-like rocks of the Tiwary, M. & Knoll, A. H. 1994. Large acanthomorphic acritarchs
USSR. Izvestiya Academii Nauk SSSR, Seriya geologigeskaya, 2, from the Infrakrol Formation of the Lesser Himalaya and their
83 – 101 (in Russian). stratigraphic significance. Journal of Himalayan Geology, 5,
Chumakov, N. M. 1993. Riphean Middle Siberian Glaciohorizon. Strati- 193– 201.
graphy and Geological Correlation, 1, 17 – 28. Veis, A. F., Vorob’eva, N. G. & Golubkova, E. Y. 2006. First find of
Chumakov, N. M. & Krasil’nikov, S. S. 1991. Lithology of Riphean Lower Vendian microfossils at Russian plate: taxonomy composition
tilloids; Ura Uplift. Lithology and Mineral Resourses, 3, 58 –78. and biostratigraphic significance. Stratigraphy and Geological Cor-
Chumakov, N. M., Pokrovsky, B. G. & Melezhik, V. A. 2007. Geologic relation, 14, C28 – 46.
history of Patom Supergroup, Late Precambrian, Middle Siberia. Vinogradov, V. I., Pichugin, L. P., Bikhover, V. N., Golovin, D. I.,
Doclady Earth Sciences, 413, 379–383. Muraviev, V. I. & Bujakayte, M. I. 1996. Isotope features and
Gorokhov, I. M., Semikhatov, M. A. et al. 1995. Strontium isotopic dating of epigenetic alterations of Upper-Precambrian deposits of
composition of Riphean, Vendian and Lower Cambrian carbonate the Urinsky Uplift. Lithology and Mineral Resources, 31, 60 –69.
rocks of Siberia. Stratigraphy and Geological Correlation, 3, 3– 33. Vorob’eva, N. G., Sergeev, V. N. & Semikhatov, M. A. 2006. Unicum
Gray, K. 2005. Ediacaran palynology of Australia. Association of Austra- Lower Vendian Kel’tma microbiota of Timan: new data on palaeon-
lian Palaeontologists. Memoir 31. Association of Australian Palaeon- tology of Vendian and its global features Doclady Academii nayk,
tologists, Canberra, 439. 410, 366–371.
Ivanov, A. I., Lifshits, V. I. et al. 1995. Precambrian of the Patom Vorob’eva, N. G., Sergeev, V. N. & Chumakov, N. M. 2007. New
Upland. Nedra, Moscow (in Russian). occurrences of Lower Vendian microfossils in Ura Formation:
Keller, B. M. (ed.) 1963. Stratigraphy of SSSR. Upper Precambrian. problem Patom Supergroup of Middle Siberia. Doclady Earth
Nedra, Moscow, 716 (in Russian). Science, 419, 782– 787.
Keller, B. M., Semikhatov, M. A. & Chumakov, N. M. 1967. Upper Yang, J., Sun, W., Wang, Z., Xue, Y. & Tao, X. 1999. Variatios in Sr
Proterozoic of Siberian Craton and its frame. In: Stratigraphy of and C isotopes and Ce anomalies in successions from China: evidence
Precambrian and Cambrian of Middle Siberia. Krasnoyarsk for the oxygenation of Neoproterozoic seawater? Precambrian
knizhnoe izdatel’stvo, Krasnoyarsk, 71– 83 (in Russian). Research, 93, 215– 233.
316 N. M. CHUMAKOV ET AL.

Zhadnova, T. P. 1961. Stratigraphy of north-east part of Patom Highland. history in South China. Palaeogeography, Palaeoclimatology,
Trudy Tsentral’nogo nauchno-issledovatel’skogo geologorasvedoch- Palaeoecology, 254, 1 –6.
nogo institute, 38, 103– 123 (in Russian). Zhuravleva, Z. A., Komar, V. A. & Chumakov, N. M. 1961. Succession
Zhou, G. & Xiao, S. 2007. Ediacaran d13C chemostratigraphy of South and correlation of Upper Precambrian deposits of West Yakutiya. In:
China. Chemical Geology, 237, 89 – 108. Bobrov, A. K. (ed.) Data on Geology and Mineral Resources of
Zhu, M., Strauss, H. & Shields, G. A. 2007. From snowball earth to the Yakutiya ASSR. Issue 13. Yakutiya Geological Survey, Yakutsk,
Cambrian bioradiation: calibration of Ediacaran-Cambrian earth 12 – 28 (in Russian).
Chapter 28

Late Cryogenian (Vendian) glaciogenic deposits in the Marnya Formation, Oselok Group,
in the foothills of the East Sayan Range, southwestern Siberian Craton

J. K. SOVETOV
Trofimuk Institute of Petroleum Geology and Geophysics, Russian Academy of Sciences, Siberian Branch,
Koptyug Av. 3, Novosibirsk, 630090, Russian Federation (e-mail: SovetovYK@ipgg.nsc.ru)

Abstract: The glaciogenic deposits of the Ulyakha Member (Mb.) referred to as Ulyakha diamictite (tillite), and the associated Late
Cryogenian (Early Vendian) glaciofluvial deposits, are the lower strata of the Marnya Formation (Fm.) of the Oselok Group (Gr.) in
the southwestern Siberian craton, which rest erosionally over the Cryogenian (Late Riphean) Karagassy Group. The Oselok Group depos-
its fill the Peri-Sayan foredeep and their stratigraphic equivalents immediately overlie the basement of the Siberian Platform. The gla-
ciogenic deposits, including diamictite, breccia, boulder conglomerate and sandstone, had multistage sedimentation, were widespread,
and underlie a cap dolomite of the lower Ozerki Mb. The Late Cryogenian (Early Vendian) age of the Ulyakha Mb. is defined by traces of
shallow-marine soft-bodied animals in the middle part of the Ozerki Mb., and by findings of Ediacaran-like Metazoa in the overlying
Bolshaya Aisa Mb. of the Marnya Fm. Deposition of the Ulyakha tillites and associated rocks has been correlated with the Late Cryo-
genian (Marinoan) glaciation based on the d13C patterns in carbonate deposits in the Marnya and Uda formations and their stratigraphic
position much below the Pre-Manykai and Zhuya d13C negative signatures, which have been equated to the Wonoka anomaly.

Glaciogenic and associated deposits in the Sayan region occur at distinguishing local lithological members (Sovetov & Komlev
the base of the Oselok Group in the Biryusa province (Fig. 28.1). 2005). Besides sedimentology, the Marnya Fm. (and the Oselok
Their stratigraphic equivalents are the basal strata of the Olkha Group as a whole) was sampled for palaeontology in search
Fm. in the Irkutsk province (Baikal Group of the Baikal region), of Ediacaran biota and traces of soft-bodied organisms, and for
and of the Taseeva and Chapa Groups in the Yenisei Range C- and O-isotope compositions in cap carbonates. On this basis,
(Sovetov et al. 2007a). All these sediments fill Late Cryogenian- along with global Late Neoproterozoic correlations, the glacio-
Ediacaran (Vendian) foredeeps echeloned along the southwestern genic deposits of the area have been identified as Late Cryogenian
margin of the Siberian craton. The glacial facies correlate with tillite (Sovetov & Komlev 2005; Sovetov 2007).
poorly sorted sediments in the lowermost sections of the Nepa The discussion below concerns the sedimentology of the
and Vanavara Groups in the craton interior and of the Dalnyaya Marnya Fm. (Oselok Group) in its representative sections, compo-
Taiga Group in the Patom passive margin (Sovetov 2002a, b; sition of glaciogenic deposits, erosional features, pathways
Sovetov & Komlev 2005; Sovetov at al. 2007a). Thus, Late of clastics transported by ice and meltwater, biostratigraphic
Cryogenian (Early Vendian) glacial events left their signature and chemostratigraphic constraints, and palaeomagnetic data
over about 1.8 million square kilometres. with implications for Late Cryogenian –Early Ediacaran (Early
The Oselok Group was originally defined as a formation in the Vendian) palaeogeography of the Siberian craton.
Sayan region (Khomentovsky 1950) and was recognized as a
group after detailed stratigraphic studies in the 1960s when it
was divided into the Marnya, Uda and Aisa formations (Dubin
et al. 1969; Khomentovsky et al. 1972). These three units are Structural framework and Neoproterozoic sedimentary
distinguished from one another and from the underlying rocks by basins of the Sayan region
diamictite and breccia at the base of the Marnya Fm., arkose and
orthoquartzite conglomerate and sandstone at the base of the The study area of the Sayan region occupies the Biryusa catchment
Uda Fm. and lithoclastic massive sandstone near the bottom of along the southwestern margin of the Siberian craton where the
the Aisa Fm. The succession of the three formations appears to basement is exposed or locally buried under Neoproterozoic and
persist continuously from the Tagul River in the NW to the Iya Cambrian sediments (Fig. 28.1). The craton in this region
River in the SE of the Biryusa area (Fig. 28.1). The best outcrops borders a metamorphic belt of Palaeozoic rocks lying at 15 –208
of the Marnya Fm. (400 – 660 m) are found along the sides of the to the basement along the Main Sayan Fault. The province of
Uda valley between the Gladkiy Mys site and the Marnya inlet Palaeozoic rocks corresponds to the East Sayan uplift and includes
(Fig. 28.1). Berzin (1967) discovered unsorted mixitite at the blocks with Precambrian continental crust. The Main Sayan Fault
base of the Marnya Fm. along the Uda River but found no traces is a zone of left-lateral strike– slip faults and mylonites, which
of glacial origin in the diamictite. Dubin et al. (1969) and Khomen- apparently initiated in the Neoproterozoic (Cryogenian) when
tovsky et al. (1972) first distinguished a member of the Oselok several terranes rifted off the Siberian craton (Berzin 1967). The
basal breccia and conglomerate lying over deeply eroded Late zone experienced at least two episodes of compression during
Precambrian (Cryogenian or Upper Riphean) strata of the the Late Precambrian Baikalian (Cadomian) and Early Palaeozoic
Karagassy Group. orogenies that produced a Late Cryogenian (Early Vendian;
The original subdivision of the Marnya Fm. into lithological Sovetov 1977, 2002b; Sovetov & Blagovidov 2004) and Middle
members 1 –4 (Dubin et al. 1969) and I –V (Bragin 1985) was Cambrian –Early Ordovician (Anatol’eva 1972; Sovetov unpub-
later updated (Sovetov 2002a), and the members received their lished) molasse foreland basins.
proper names and genetic identification. Diamictite was jointly The sediments –basement contact is marked by a structural and
assigned to the Ulyakha Mb. and faceted-boulder breccia to the metamorphic unconformity. The stratigraphic gap between the
lower Plity Mb. Diamictite and faceted-boulder breccia in the basement and the Cryogenian Karagassy Group reaches 1–
two members bear numerous traces of glacial origin, including 0.8 Ga in the Biryusa and Urik-Iya fault blocks and 2.9– 2.0 Ga
striated boulders and abrupt facies changes, which has prompted within the Sharyzhalgai terrane where Archaean rocks underlie

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 317– 329. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.28
318 J. K. SOVETOV

Fig. 28.1. Precambrian geology of the Sayan region on the south-western Siberian Craton. (a) Geological map of the region showing structural elements and
lithostratigraphy. (b) Stratigraphic scheme of the Neoproterozoic deposits in the Foothills of the East Sayan Range and their correlation with deposits in the adjacent
regions. Supersequences: Sq 1, Marnya; Sq 2, Greben; Sq 3, Ust-Tagul.
MARNYA FORMATION, OSELOK GROUP 319

the basal diamictite of the Late Cryogenian –Early Ediacaran Stratigraphy


Olkha Fm.
The oldest basement rocks (granulite, gneiss and schist) belong The Karagassy and Oselok Groups (Fig. 28.1) in the Biryusa
to the Archaean Sharyzhalgai Group in the Sharyzhalgai terrane watershed were interpreted jointly as a single complex (Khomen-
(Sez’ko 1988). The Early Proterozoic Biryusa, Kan and Dzhu- tovsky 1950) before a stratigraphic gap between them and a
glyma terranes are composed of migmatized biotite, amphibole, complex stacking pattern was discovered, providing grounds for
garnet granulite and gneiss, granite gneiss and amphibolite. their subdivision (Dubin et al. 1969).
The Palaeo-Mesoproterozoic Urik-Iya trough basin is filled
with 10 000 m of biotite-quartz phyllite, shale, metasandstone,
conglomerate, tuff, and volcanic rocks (Berzin 1967; Sez’ko Karagassy Group
1988). The metamorphic complexes of the Biryusa terrane are
intruded by Palaeoproterozoic granites of the two-stage Sayan The Karagassy Group, varying in thickness from c. 800 to
complex dated between 1870 –1734 Ma (Donskaya et al. 2002; c. 2000 m, consists of the Shangulezh, Izan (currently referred to
Turkina et al. 2003). Unmetamorphosed Neoproterozoic sedi- as Tagul; Egorova et al. 1971; Bragin 1986; Shenfil 1991;
ments fall into two geodynamically different sequences: the Sovetov & Blagovidov 2006) and Ipsit formations (Khomentovsky
rift-related deposits of the Cryogenian Karagassy Group and et al. 1972). The Shangulezh Fm. (250 –600 m) lies with an
the Late Cryogenian –Ediacaran Oselok Group, which filled a angular unconformity over the Urik-Iya folded metamorphosed
foreland basin. sediments and volcanic rocks and over high-grade metamorphic
The Karagassy Group lies, with a large structural unconfor- rocks and Palaeoproterozoic granitoids of the Sayan complex.
mity, over metamorphic and igneous complexes of the Biryusa The three formations are second-order sequences and are separated
fault block and over low-grade metasediments and volcanic com- from one another by gaps and low-stand erosional surfaces. The
plexes of the Urik-Iya basin. The Oselok Group oversteps deeply sequences record facies changes from alluvial fans to a terrigenous
eroded surfaces of the Karagassy formations in the Biryusa catch- carbonate tidal plain and a carbonate tidal shelf with abundant
ment; its stratigraphic equivalents (Olkha and Moty Groups) rest stromatolite buildups, microphytolite dolarenite and dolomicrite.
upon the Archaean Sharyzhalgai metamorphics only outside the Of broad occurrence are the structures of supratidal, subtidal and
rift basin in the Irkut catchment. According to recent data peritidal deposits with bimodal flow directions of terrigenous-
(Sovetov & Blagovidov 2004, 2006; Sovetov et al. 2007a), the carbonate and silt material, minor scouring forms, flat stromatolite
Karagassy sediments fill the Cryogenian Iya-Tumanshet failed mats, mud cracks and halite glyptomorphs.
rift (aulacogen), whereas the Oselok Group is the fill of the
Late Cryogenian –Ediacaran Peri-Sayan foredeep. The fault-
bounded rifts formed at the same time and in the same geody- Oselok Group
namic environment as the Teya-Chapa and Vorogovka failed
rifts 500 km to the NW in the Yenisei Ridge (Sovetov 1997; The Oselok Group overlies the Karagassy Group without structural
Sovetov et al. 2007a). The Karagassy rocks are intruded through- unconformity but with an erosional contact (locally about 1000 m
out the area by small bodies of gabbro dolerite sills, dykes and of the Ipsit and partly Tagul deposits have been removed). The
stocks of the Nersa complex with an Ar/Ar plagioclase age of group includes the Marnya, Uda and Aisa formations (Dubin
741 + 2 Ma (Gladkochub et al. 2006). et al. 1969; Khomentovsky et al. 1972; Sovetov & Komlev
The Late Cryogenian –Ediacaran Peri-Sayan foredeep initiated 2005). The eroded top of the Aisa Fm. underlies, without structural
upon the Iya-Tumanshet rift and the adjacent craton edge after unconformity, the basal conglomerates of the Ust’-Tagul Fm.
regression and deep glacial erosion during the Late Cryogenian
glaciation (Sovetov 2002b; Sovetov & Blagovidov 2004; Marnya Formation. The general succession of members, their
Sovetov & Komlev 2005). The Late Cryogenian –Ediacaran sedi- replacement and pinching out, have been inferred from correlation
mentary fill of the Peri-Sayan, Peri-Baikal and Yenisei foredeeps, and are reported below following Sovetov & Komlev (2005), with
together with the coeval terrigenous and terrigenous-carbonate later additions (Fig. 28.2). The new subdivision updates the earlier
platform cover, belong to an enormous foreland basin on the stratigraphic interpretations (members I to V) by Dubin et al.
western and southwestern margins of the Siberian craton. The (1969) and Bragin (1984, 1985).
Peri-Sayan foredeep basin evolved in two stages recorded in The Karapchatui Member (I), the oldest in the Marnya Fm., was
the Late Cryogenian– Early Ediacaran Marnya (1) and Late Edia- discovered by Bragin (1984), and its stratigraphy was updated
caran Greben (2) regional supersequences (Sovetov et al. 2007a). through studies by Sovetov in 2006 along the right side of the
The Marnya supersequence consists of glaciogenic deposits, Uda River (Karapchatui Brook mouth). The section comprises
shallow-marine terrigenous and carbonate rocks, and less abun- 0.1– 0.2 m of ferruginous silicified cataclastic rocks at the base
dant fluvial facies. The Early Ediacaran was associated with overlain by (1) 3.2 m of massive quartzite faceted-boulder
transgressions from the SW onto the southern and southwestern breccia and white cross-bedded orthoquartzite; (2) 22 m of white
parts of the Siberian craton as a result of the closing Palaeoasian and yellow cross-bedded orthoquartzite with debris in small
ocean and clastic transport from the craton interior. During the fluvial – lacustrine cycles with thin-bedded orthoquartzite and stro-
Late Ediacaran (Greben) syncollisional stage, the basin experi- matolite dolomite; (3) 30–35 m of variegated thinly laminated
enced transgressions from the NE, and the clastic transport was quartz sandstone and stromatolitic dolomite; (4) 20– 30 m of
from flanking orogens in the western, southwestern and southern quartz and massive cross-bedded feldspar-quartzose sandstone.
craton margins. The Cambrian history of the Siberian craton Orthoquartzites give way to breccia in the SW, and also in the
began with the break-up of the Ediacaran (Late Vendian) NE, towards the sides of the underlying channelized unconformity.
orogens, formation of a terrigenous hanging-wall basin, followed The Nersa Mb. (18.7 m thick) (Mb I, Bragin 1984), found on the
by the development of evaporite and carbonate basins (Sovetov west side of the Uda River in the Nersa and Kremenshet interfluve,
et al. 2007a). consists of massive fluvial boulder conglomerate and coarse
The Karagassy and Oselok sediments were deformed by post- pebbly sandstone in the upper section, with clasts of the Karagassy
Cambrian to pre-Devonian west –east compression (Berzin arkose quartzitic sandstone, stromatolite dolomite, chert and doler-
1967) that split the sedimentary basins into blocks displaced ite. A c. 22-m-thick sill of the Nersa complex was found 2.1 m
along strike –slip, thrust and reverse faults. This deformation below the Nersa-Ipsit contact.
event produced shear zones and related folds, but was not The Ulyakha Mb. stratotype occurs along the right side of the
accompanied by metamorphism. Uda River 0.3– 1 km upstream of the Ulyakha mouth, where the
(after Bragin 1985)*

320
Gladky Mys Ulyakha Kirei Ozerki Karapchatui Plity Bogatyr Kedrovyi

Ikei Fm.
Glacio-fluvial quarzite sandstone

Muksut Mb.
with detritus

500
Dolostone of ephemeral
postglacial lake

Mb. Kagat Mb.


Glacio-fluvial litharenite with
single pebbles
Pescherny

Laminated black and grey-green


mudstone with sandstone
lenses
(after Bragin 1985)

400 Eolian (?) quartzite sanstone dunes


Unyl Mb.
Uda Fm.

Laminated cap-dolomite

Current-rippled dolostone
Nizhnyaya Uda Mb.

Plity
Mb. Grainstone sand waves
300 N=31
Quarzite sand waves and bars

Kedrovy
Marnya Fm. (after Dubin et al. 1969)

Quarzite foreshore bars


Bol’shaya Aisa Mb.

Mb.

J. K. SOVETOV
Plity Ozerki Mb. Sandstone-siltstone tempestite
N=53
N=19 Plity Ipsit Ulyakha Mb.
200 Mb. Fm.
Tide-related channelized sandstone
Mb.

170 m -siltstone and micrite limestone


30 m Nersa Mb. tempestite
20 m
N=52 Gravelly sandstone tempestite
Ipsit
Tygnei Mb.

20 m Fm.
Karapchatui Mb.
Ognit Mb.

Hurricane cross-bedded
Ulyakha Mb.

N=27
N=10 sandstone
20 m
Fluvial cross-bedded gravel
100 Ipsit Palaeocurrent directions and sandstone quartzite and
Tagul
Fm.
Tagul
arkose
Fm. Fm.
Delta plain thin bedded siltstone
Red stromatolitic dolostone with gravel and sandstone
Ozerki Mb.

lenses
Kedrovyi 99 00 Sandy tempestite and Ooids and stromatolitic
70
Nizhneudinsk sand-waves with slumps limestone
Sandstone coarse cross-bedded
Dolerite sills bars and channels of deep-water
m0 Clay-rich diamictite with rivers
Uda River

Siberian
r
ive

predominant quarzitic cobbles Trace-fossils


Craton
aR

60 and boulders
Bogatyr
Dolomitic diamictite with Moulds and imprints of soft-body
us

a b c
54 30 predominant dolomite cobbles animals: a.Cyclozoa, b. Bilaterian,
ry

Plity
Bi

and boulders c. Problematics


Karapchatui
Faceted-boulder breccia
al

Ozerki Glacial erosion surface


ik
Ba

Kirei
25 km
ke

Gladky Mys Ulyakha


La

100 110
Boulder conglomerate 20 m Non-cropped parts of sections

Fig. 28.2. Stratigraphy and depositional systems of the Marnya and Uda formations at different locations along the Uda and Biryusa rivers, showing the lateral variability and stratigraphic relationships of the members from
south to north. Note that the Ikei Formation includes the Aisa Fm. and Kagat and Muksut members, which are differentiated on the basis of colour. The Ikei Fm. is generally no longer referred to in the literature. Inset maps
show localities of stratigraphic logs as well as average palaeo-ice flow direction based on sedimentary structures in glacial and glaciofluvial sediments.
MARNYA FORMATION, OSELOK GROUP 321

Member erosively overlies the Tagul Fm. The contact is perfectly dark-grey siltstone with lenses of fine to coarse cross-bedded sand-
exposed in the Uda side 5 km upstream of the Marnya mouth. The stone deposited in subsidiary channels and river mouth bars.
Ulyakha Mb. (up to 53 m thick) (Mb I, Bragin 1985), comprises The Peschernyi Mb. (VIII, Bragin 1985), 7 –30 m, is composed
lenses and upright veins of greenish-grey, dark grey (to black) or of dark-grey and black stromatolitic and microphytolitic limestone
red coarse diamictite and dolomite breccia, and less abundant and limestone tempestite, is locally replaced by epigenetic dolo-
channelized coarse sandstone (Figs 28.2 & 28.3). The member is mite, and varies strongly in thickness, pinching out cratonward.
recognizable by its boulders and basement erratic boulders, abun-
dant glacial striation on clasts, grooves and other traces of erosion
on the surface where diamictite overlies the Tagul Fm. Aisa Fm.
The Tygnei Mb. (80 –90 m thick) (Mb IIa, Bragin 1985), com-
posed of black bituminous pyrite siltstone with thin stromatolite The Aisa (Ikei, Bragin 1985) Fm. is composed of continental
dolomite in the upper part and sandstone interbeds and greenish- molasse of two lithofacies groups: (i) fluvial channel litharenite
grey laminated siltstone in the lower part, crops out along the sandstone and (ii) bar and overbank sandstone, siltstone and,
sides of the Uda River. rarely, mudstone. Estimates of its total thickness vary from
The Plity Mb. (20 –35 m thick) (Mb IIb, Bragin 1985), consists 1800 –2000 m (Egorova et al. 1971) to 1300 –1600 m (Khomen-
of faceted-boulder quartzite breccia and yellowish-white feldspar- tovsky et al. 1972), or 1300 –1400 m according to measurements
quartzose sandstone under massive fluvial and eolian (?) by Sovetov in the most complete section on the sides of the
cross-bedded sandstone. Biryusa and Tagul rivers. The formation has a prominent base
The Kedrovy Mb. (40 m thick), oversteps the Ulyakha diamictite along the roof of the Peschernyi limestone and a very sharp top at
on the west side of the Uda River (near Kedrovy Village) and con- the base of the overlying Ust’-Tagul Fm. conglomerate attributed
sists of grey coarse extralitharenite sandstone with cobbles and to the Late Vendian Nemakit-Daldyn regional stage (Ediacaran-
pebbles at the base of fluvial cycles. It underlies white coarse Early Cambrian) (Sovetov et al. 2007a). The Aisa Fm. includes
arkose sandstone of the Ognit Mb. and is a stratigraphic equivalent four members of different thicknesses: Kagat, Muksut, Serebrovo
of the Plity Mb. and Katalchikov. Overbank deposits contain abundant Arumberia
The Ozerki Mb. (50– 60 m thick) (Mb II, Bragin 1985), lies plant remnants (Sovetov 2006).
conformably over various members of the Marnya Fm. at differ-
ent localities (Fig. 28.2). It pinches out in the Sayan region where
the pre-Marnya erosion was moderate. The sections of the Glaciogenic deposits of the Marnya Fm. and associated strata
member are either (i) yellow laminated, wave rippled and tidal-
current cap dolomite with numerous cavities after pyrite Diamictite of the Ulyakha Mb. was studied in outcrops along the
nodules or (ii) laminated stromatolite and cross-bedded sandy Uda River between the Marnya and Ulyakha mouth areas and
dolomite, and feldspar-quartzose sandstone in the middle part near Kedrovyi Village on the Biryusa River (Nersa section). The
with abundant traces of soft-bodied animals (Sovetov 2007; maximum thickness of diamictite measured from its lower
Sovetov & Solovetskaya 2008). At one locality (Ulyakha, erosion surface to the base of the Ozerki Mb. cap dolomite is
Fig. 28.2), the base of the carbonate Ozerki Mb. overlies the 53 m. The matrix of boulder diamictite (mixtites) of the Ulyakha
diamictite-bearing Ulyakha Mb. and is thought to correlate Mb. (c. 85% of the rock volume) has a sand-silt-clay composition.
with the base of the Ediacaran. The rocks bear evident lithological signatures of tillite (Sovetov
The Ozerki Mb. is overlain by the Ognit and Bolshaya Aisa 2002a; Sovetov & Komlev 2005; Sovetov & Donskaya 2006),
members. The Ognit Mb. (up to 84 m thick) (Mb IV, Bragin namely: (1) iron- and bullet-shaped stones, which are faceted,
1985), consists of dark-grey and black coarse orthoquartzite and striated, polished and range in size from gravel to clasts up to
arkoses, whereas the Bolshaya Aisa Member (105 –269 m thick) 70 cm (occasionally 1.5 m) across; (2) orientation of elongate
(Mb V, Bragin 1984) consists of shoreface greenish-grey cross- stones; (3) lodgment stones pressed into dolomite of the Tagul
bedded sandstone and siltstone and shelf HCS tempestites and tida- Fm.; (4) dropstones in sandy dolomite of outwash lakes; (5) steep-
lites, with less abundant black microphytolite and silty limestone. walled subglacial channels, boulder beds and clustered stones; (6)
The Bolshaya Aisa Mb. contains Ediacaran-like biota (Sovetov & up to 1.5-m-deep nearly vertical crevasses with breccia; (7) inner
Komlev 2005; Sovetov 2007; Sovetov & Solovetskaya 2008). stratification of diamictite in the form of four nested lenses, differ-
ing in coloration, composition of stones, and containing meltout
intercalations; (8) associated fluvial sandstone with debris laterally
replaced by breccia; (9) stones of rocks typical of the underlying
Uda Fm. Tagul and the Ipsit formations (Karagassy Group) and less fre-
quent erratics of gneiss, granite, biotite schist and metarhyolite
The stratigraphic extent of the Uda Fm. long remained unclear derived from the Early Proterozoic basement; (10) erosional
after the lowermost grey rocks of the Oselok Group, originally forms, such as narrow (a few decimetres wide) and shallow
attributed to it, had been included into the Marnya Fm. According furrows, scours tens of metres wide and 3– 4 m deep, local (tens
to Bragin (1985), its lower section comprised the Ognit and to hundreds of metres across) and regional (tens of kilometres
Bolshaya Aisa members but Khomentovsky et al. (1972) and wide) erosional valleys; (11) blunt-end contacts of diamictite
Sovetov & Komlev (2005) identified its lower boundary by with sidewalls of channelized troughs; (12) folds and rupture
arkose and quartz gravelstone of the Nizhnyaya Uda Mb. Together features at the top of the underlying rocks of the Ipsit and Tagul
with the latter, the Uda Fm. is currently divided into three members formations; (13) association with laminated cap dolomite with
(Fig. 28.2). sharp lower contact; (14) association with laminated siltstone
The Nizhnyaya Uda Mb. (VI, Bragin 1985), most often and shale alternated with gravelly-sandy deposits of gravity
17– 30 m, is composed of white and yellow fine-pebble quartz flow; (15) association with boulder conglomerate containing
conglomerate, gravelstone and coarse gravelly cross-bedded striated pebbles; (16) absence of ash fall, impact events, lahars,
fluvial sandstone. The member rests upon the deeply (to 4 m) volcaniclastic debris flows or turbidity flows on the continental
eroded surface of the Bolshaya Aisa Mb. and to the SW its lower slope and rise; (17) absence of indications typical of tectonic
40 m (out of a 70 m) section comprises coastal bar and thrusting or melange; (18) occurrence of diamictite above a
sandwave deposits. regional unconformity.
The Unyl Mb. (VI, Bragin 1985), 25 –100 m, is genetically The architectural elements in Marnya Fm. diamictite were dis-
linked to the Nizhnyaya Uda Mb. and consists of grey and tinguished by mapping outcrops. Figure 28.3 shows cross-sections
322
I II III IV V VI VII VIII IX X XI XII XIII XIV XV

Ozerki Mb.
(a)

IV
I
2m II
III
III
10 m
in Ulakha river location
Glaciogenic deposits

S dolomite

in diamictite
Ulyakha Member
n=111
Marnya Formation

n=141 n=226 n=112


chert

stones
n=649 N arkose and orthoquartzite
20 m basement rocks
(b)

2m
Tagul

10 m
Fm. Ozerki Mb.

N Profile direction 325


n=209
I n=575 II III IV V VI VII
S

J. K. SOVETOV
Plity Member

(c)
Glaciogenic deposits in Ozerki location
Marnya Formation

III

IX
2,5 m VIII
Tygnei Member

5m VIII IX X XII
XI
(d)

20 m

VIII 200

Profile direction 50 Vergence of glacial dislocations


Ulyakha Mb.

Black cross-bedded sandstone in Deglacial fine detrital Outwash-plain cross-bedded glaciofluvial


Key to symbols:

Black argillaceous diamictite and sandstone with cobbles (zandur) III Diamictite lenses separated
stones 20 cm intraglacial and supraglacial channels diamictite by a shear zone
Green-grey argillaceous diamictite Elongation direction of cobbles and Pebbly sandstone from currents Reference sections of outcrop logs
stones
Laminated cap dolomite
(underwater fans and fan-deltas) XI
Tagul
Fm.

Red-brown sand-clay mixitite, Elongation direction of lodgment Massive and layered stromatolite Glacial shears
no stones cobbles and boulders and grainstone dolomite
Shear zones and boundaries Dollutite Percentages of different lithologies of stones
Dolomite diamictite and breccia
of diamictite lenses 4-40 mm
Intraglacial and supraglacial Laminated siltstone and
breccias and gravelstone in Basal shear zone Closed parts of section
mudstone in ephemeral lake
channels and crevasses

Fig. 28.3. Outcrop logs and field sketch showing different facies and lithology of diamictite clasts in the upper and lower Ulyakha Member along the Uda River. (a) Vertical section of an outcrop on the left side of the Uda River
1 km upstream of Ulyakha Brook. (b) Enlarged left side of section (a) showing the geometry of sandstone and conglomerate lenses within black argillaceous diamictite. (c, d) Single vertical section of the lower Ulyakha Mb.
at the Ozerki locality on the left side of the Uda River, 5 km upstream of the Marnya River.
MARNYA FORMATION, OSELOK GROUP 323

and logs of the upper and lower Ulyakha Mb. along the Uda River. diamictites are thought to be buried under black bituminous silt-
For complete stratigraphy of the member, see Figure 28.2. The stone deposited in ephemeral lakes (Tygnei Mb.), which in turn
upper Ulyakha Mb. (Fig. 28.3a) consists of four nested lenses of are overlain by laminated, wave-rippled and stromatolite dolomite
diamictite with shear boundaries (I to IV). They are named black of the Ozerki Mb.
and green argillaceous diamictite after their matrix colour and The Karapchatui Mb. fills a 60–70-m-deep valley, incised into
siliciclastic composition. Diamictite consist of outsized clasts the Ipsit and Tagul Formations (Figs 28.2 & 28.4). The valley is
from sand- and fine-gravel-sized rocks to elongate iron-shaped filled with cross-bedded sandstone, sandy stromatolitic dolomite
boulders with striation oriented at 100 –2158. Lenses III and IV and faceted-boulder breccia. The breccia occur in the valley
are cut with 3–17-m-wide channels filled with either dolomite dia- floor and its east side, and the centre is occupied by an outwash-
mictite, conglomerate or breccia. The lenses are massive or have plain fan oriented at 340– 08 in the NW – north direction, according
inclined bedding with inclusions of black coarse cross-bedded to numerous measurements of oblique foresets in streams, large
sandstone; the channels have steep or overhanging walls, some ripples and dunes.
resemble crevasses (Fig. 28.3b). Palaeocurrent direction measured
on cross-bedding and channel-bed slopes suggests currents moved
towards 230 –3308.
The Ulyakha Mb. at the Ozerki outcrop consists of three lenses Boundary relations with overlying and underlying
separated by shear zones or shear planes (Fig. 28.3c, d). These non-glacial units
deposits host abundant boulders of the Ipsit Fm. quartzite with
striation, less abundant boulders of the Tagul Fm., silicified stro- All members of glaciogenic deposits at the base of the Marnya Fm.
matolitic dolomite, and a few erratic boulders of basement (diamictite, faceted-boulder breccia, conglomerate and sandstone)
granite and gneiss. The boulders and their striation are oriented have sharp contacts with the underlying deposits of the Ipsit and
mainly at 150– 2408. The direction of glacier advance has been Tagul formations in the Sayan region, with local scouring features
inferred from the 150– 2708 orientation of lodgment boulders (Figs 28.2–28.4). The deepest erosion exceeds 500 m in the area
pressed into the Tagul dolomite. A similar interpretative orien- between the Biryusa and Uda rivers (Fig. 28.4) and is c. 1000 m
tation (2008) has been found in folds of the Tagul dolomite and throughout the Sayan region (Dubin et al. 1969; Bragin 1984;
siltstone beds beneath the diamictite. Small folds only occur at Sovetov & Komlev 2005).
the Ozerky location above a horizontal shear zone 8 –10 m The erosive contact of the Ulyakha Mb. black argillaceous dia-
below the Ulyakha Mb., and the underlying part of the Tagul mictite with the underlying dolomite of the Tagul Fm. (Karagassy
Fm. is rather undeformed. The upper Ulyakha Mb. includes Group) was mapped on the left side of the Uda River at the Ozerki
black siltstone and lenses of dolomite diamictite and breccia in locality (Fig. 28.3c). The profile runs across a 90–110-m-long and
U-shaped channels. According to correlation (no visible contact), 4.5-m-deep flat-bottomed channel, which is incised into the Tagul

Fig. 28.4. Cryogenian (Early Vendian) glacial erosional forms in the southwestern Siberian Craton. See Figure 28.2, inset, for map of localities.
324 J. K. SOVETOV

Fm. and is filled with dolomite diamictite from the lower Ulyakha 2007b), similar to a low of –10‰ in the Oskoba Fm., Baikit
Mb. The total erosion depth at the locality approaches 9 m. region (Vinogradov et al. 1994; Khomentovsky et al. 1998;
Grooves and scours with sharp contacts on the surface of the Sovetov et al. 2007b), which both appear to correlate with the
Tagul dolomite, as well as terraces, are oriented at 190– 1958 Zhuya and partly the N2 negative anomalies (Pelechaty 1998;
(Fig. 28.4). Pokrovsky et al. 2006) in the Patom region and with the
A large erosional valley is exposed on the right side of the Uda Wonoka anomaly identified in various parts of the world
River at the mouth of the Karapchatui Brook. The Karapchatui Mb. (Halverson et al. 2005). The grounds for this correlation are
overlies the Tagul Fm., with 10–158 unconformity, in the centre of discussed below.
an erosional valley (Figs 28.2 & 28.4). The eastern valley side dips
at 30 –508 and strikes at 3008, and its upper part is composed of the
Ipsit Fm. rocks (Fig. 28.4). Another exposed erosional valley Palaeolatitude and palaeogeography
occurs along the Uda River at the Plity locality (Fig. 28.2). The
15-m-deep north – south valley is incised into the Ipsit Fm. and Palaeomagnetic directions were measured in diamictite samples
filled with faceted-boulder breccia and minor cross-bedded from the Ulyakha Mb. (Marnya Fm.) (54.338N, 98.878E) and the
sandstone of the lower Plity Mb. Thick-bedded cross-bedded sand- overlying Tygnei black siltstone (54.338N, 98.898E) collected by
stone of the upper Plity Mb. covers the eastern side of the valley Sovetov at the Ozerki locality (Uda River). Diamictite and dolo-
and lies over the Ipsit quartzose arkose siltstone with a sharp mite exhibit nearly horizontal bedding planes at the locality.
contact. The dip of oblique cross-bedding foresets indicates Measurements were performed using the standard procedure
north –south transport of material along the valley. The same (Butler 1992) of stepwise thermal demagnetization on a magnet-
relationship between the Plity Mb. and the Ipsit Fm. is found in ometer at the Paleomagnetic Centre of International Petroleum
outcrops along the right side of the Tagul River 6 km upstream Geology and Geophysics (Novosibirsk, Russia).
of Georgievka Village. The distinct base of the glaciofluvial According to published and preliminary data, the primary
boulder conglomerate of the Nersa Mb. (Marnya Fm.), with silt- palaeomagnetic directions in the rocks of the Marnya Fm. have
stone of the Ipsit Fm., crops out along the Biryusa River (near been overprinted by Early Cambrian geodynamic events (Metelkin
Kedrovyi Village) (Fig. 28.2). et al. 2005; D. V. Metelkin, pers. comm.). The available palaeo-
The upper boundary of the Marnya glaciogenic deposits is well magnetic evidence for Neoproterozoic and Early Palaeozoic
exposed along the sides of the Uda River at the Ulyakha and Ozerki rocks from the Siberian craton (Smethurst et al. 1998; Pisarevsky
localities. The Ulyakha Mb. rocks underlie laminited cap dolomite & Natapov 2003; Shatsillo et al. 2005; Metelkin et al. 2007)
along a sharp contact (Fig. 28.3). The same prominent boundary suggest a sub-equatorial location.
between the Plity sandstone and cap dolomite is exposed at the
Ozerki locality (Fig. 28.2). Dolomite of the Ozerki Mb. pinches Geochronological constraints
out along the sides of the Uda River (Plity and Gladkiy Mys
sites and at Bogatyr’ Hill) and leaves sandstone tempestite of the The minimum age of glaciation comes from Ediacaran (Late
overlying Bolshaya Aisa Mb. Overlying the Plity Mb. (Fig. 28.2). Vendian) –Early Cambrian small-shell fauna found within the
Ostrovnoi Fm., (Khomentovsky et al. 1998), which is correlated
with the upper part of the Ust’Tagul Fm. (Fig. 28.6). The latter ero-
Chemostratigraphy sively overlies the Oselok Group, while its equivalents overlie the
Taseeva and Chapa groups in the Yenisei Ridge (Sovetov 2002b,
Tentative estimates of 87Sr/86Sr ratios were obtained in the fig. 6; Sovetov & Blagovidov 2004). Furthermore, Metazoan
Peschernyi Mb. limestone. These Sr ratios are lower than those trace fossils were found in quartzite in the middle of the Ozerki
reported from the Wonoka Fm. (Calver 2000), and the correlated Mb. and imprints and moulds of Ediacaran-like organisms have
Shuram Fm. of Oman (Burns et al. 1994) (0.70796 – 0.70848 been reported in the Bol’shaya Aisa Mb. (Sovetov & Komlev
v. 0.7087 –0.7088 and 0.7085 –0.7088, respectively; Fig. 28.5). 2005; Sovetov 2007; Sovetov & Solovetskaya 2008).
C- and O-isotope compositions were studied in carbonate rocks The maximum age for the Marnya Fm. glaciogenic deposits
of the Marnya and Uda formations (Fig. 28.5; Sovetov & Komlev in the Sayan region has been constrained by two dates: (i) the
2005; Sovetov et al. 2007b). The resulting C-isotope curve 741 + 2 Ma 39Ar/40Ar age of plagioclase from the Nersa gabbro
includes five characteristic segments A –E (Fig. 28.5; Sovetov & dolerite sills (Gladkochub et al. 2005) that intrude the Shangulezh
Komlev 2005). The d13C values in segment A, which correspond Fm. of the underlying Karagassy Group (gabbro dolerite clasts are
to samples from the middle and upper parts of the Tagul Fm. also found in the Ulyakha Member diamictite), and (ii) the
(Karagassy Group), vary from – 1.5 to þ3‰. All d13C values in 39
Ar/40Ar 696 + 8.5 Ma age of biotite from dolerite and tuff
dolomitic diamictite of the Marnya Fm. (segment B) are positive (Postnikov et al. 2005) among diamictite of the Chivida Fm.
and within þ1.6‰. A negative anomaly occurs in segment C (middle section of Chingasan Group; Yenisei Ridge) that lies
corresponding to the overlying Ozerki dolomite, where d13C stratigraphically lower than the Oselok Gr.
gradually decreases to –5.7‰ and then increases smoothly to
zero towards the top of the Ozerki Mb. A positive d13C excursion
is found in the upper third of the Bolshaya Aisa Mb. (Marnya Fm.; Discussion
from þ6.3‰ decreasing upward to þ4‰) as well as in the
Peschernyi Mb. (Uda Fm.; þ3.1 to þ7.5‰). Based on these Main structural and morphological features of the Marnya basal
data, the Ulyakha diamictite was correlated with the Late strata were discussed in Sovetov & Komlev (2005) and Sovetov
Cryogenian Varanger glacial deposits in Norway (Sovetov & (2008). The formation includes several major glaciogenic sedi-
Komlev 2005). mentary structures and architectural elements (Miall 1985),
The chemostratigraphic chart as a whole has the following fea- namely (i) diamictite with striated boulders and poorly rounded
tures (Fig. 28.5): (i) a d13C low in cap carbonates over the Ulyakha blocks, (ii) faceted-boulder breccia and (iii) intradiamictite- and
Mb., (ii) transition to positive d13C signatures spanning the supra-diamictite glacial channels and crevasses. The subsidiary
Bolshaya Aisa and Peschernyi members in the Sayan area elements are (iv) outwash –plain fluvial debris sandstone with
(Sovetov & Komlev 2005, figs 6–8), comparable to the anomalies thin layers and lenses of stromatolite dolomite, (v) sandstone and
found in the Uluntui Fm. of the Baikal Group, southwestern Baikal gravelstone deposits of alluvial plains, fans and fan deltas, (vi)
area (Khabarov & Ponomarchuk 2005), (iii) a large d13C low in the laminated fine silt and mudstone deposited in ephemeral stagnant
Kachergat Fm. of the Baikal Group, to – 6‰ (Sovetov et al. lakes. Glaciogenic deposition was accompanied by deep erosion of
MARNYA FORMATION, OSELOK GROUP 325

δ13C δ13C Depositional systems


10 8 6 4 2 0 2 4 6 8 10 10 8 6 4 2 0 2 4 6 8 10
Braided alluvial and delta plain
Baikal region, Kurtun river
(isotope data by Sovetov, et al.,

Phytolithic bioherms and

Continental slope and shelf


Сhencha Fm.
Kachergat Fm.=Aisa Fm.

storm-dominated deposits
2007 b)

Nikolsky Fm.
Contour currents deposits

Passive continental margin


(isotope data by Pokrovsky et al.,2006)
Patom Upland Ura river
Deep-water delta
Valukhta Fm.
87 86
Sr/ Sr Siliciclastic
0.70796-0.70848 deep-water fan
Uda Fm.

Pescherny Mb. E
Epicontinental sea, strand plain

Borakun Fm. Siliciclastic and carbonate


(isotope data by Sovetov, Komlev,

Continental rise
Unyl Mb.
Sayan region Uda river

Nizhnyaya Uda Mb. ? deep-water fans


Bol. Aisa Mb. D
Ognit Mb.

nian glaciation
2005)

Late Cryoge-
Marnya Fm.

C Siliciclastic
Ozerki Mb.
Dzhemkukan Fm. deep-water fan
and diamictite
Onset glacigenic deposits

Ulyakha Mb. B Mariinka Fm.


Onset glacigenic deposits
Tagul Fm.
A
Fig. 28.5. Chemostratography of the Oselok Gp. and correlation with deep-water deposits of the Patom Upland. Carbon-isotope curve for the Marnya and Uda formations
is complemented with some data from the Kachergat Fm., which is a stratigraphic equivalent of the Aisa Fm. The uppermost negative anomaly in the Nikolsky and Chencha
formations is thought to correlate with the Wonoka anomaly (Halverson et al. 2005) and the Pre-Manykai anomaly identified by Pokrovsky & Missarzhevsky (1993).

the substrate and multiform scouring, and dislocation of rocks in the Baikal Group of the Baikal region, and the Nepa and Tira for-
the underlying bed (Fig. 28.4). Judging by their morphology, the mations of the Baikit and Nepa-Botuobiya forebulges (Sovetov
channel-like stone clusters formed in ice during deglaciation and et al. 2007a). The regional stratigraphy and correlations are con-
can be identified as kames. Diamictite have all the features of trolled by the following marker units and surfaces (Fig. 28.6):
tillite, and are thought to have been produced by rock crushing (1) transgression and flooding surface at the base of the Katanga
and transport of clasts by ice and through sub-ice channels by melt- and Ostrovnoi formations and its equivalents at the base of the
water. Deglaciation is recorded by deposition of black siltstone (to Kurtun Fm. and in the middle of the Shamanka, Ust’-Tagul, and
0.5 m) in ephemeral lakes and, later, by deposition of dark-grey Uglovoi formations, (2) base of the Ust’-Tagul supersequence
and black laminated cap dolomite of the Ozerki Mb., which with river incision, (3) base of the Greben supersequence with
grade up the section into stromatolitic dolomite, dolomite tidalites river incision, (4) base of the Marnya supersequence defined by
and sand –grainstone sandwaves. a deeply eroded surface under glacial and glaciofluvial deposits.
The Ulyakha tillite, with the associated faceted boulder breccia, Subsidiary correlation levels are associated with third-order
boulder conglomerate and outwash sand, make up a complex suc- sequences (Sovetov 2002b; Sovetov & Komlev 2005). The
cession (Sovetov & Komlev 2005). The palaeocurrent directions Marnya glaciogenic deposits have been correlated with breccia,
suggest the existence of two major glaciers in the Siberian diamictite at the base of the Olkha Fm. (Mordvin 1972; Chumakov
craton: a smaller one marked by the Karapchatui outwash fan in 1993), Baikal Gr. (Sovetov & Komlev 2005) and the base of the
the south and a larger one evident from the Ulyakha Mb. channels Chapa Gr. (Sovetov et al. 2007a), conglomerate and breccia at
and striated lodgment boulders in the NE of the area in the central the base of the Taseeva Gr. (Sovetov et al. 2007a), as well as the
Siberian Craton. The extent of the latter glacier is recorded in Vanavara, and Nepa Formations (Tyschenko 1980) (Fig. 28.6).
clastic transport, erratic stones from the basement of the Siberian Carbonate members in the Marnya, Uda, Goloustnaya and
Platform and in sedimentary structures observed in the Sayan Uluntui formations pinch out cratonward from foredeeps, while
and Baikal regions (Sovetov & Donskaya 2006). carbonate members of the Oskoba and Tira formations are replaced
Lithostratigraphic evidence available for deposits of the in the foredeeps by terrigenous molasse (Fig. 28.6). Sedi-
Precambrian –Cambrian transition from different areas of the mentological studies of the upper Oselok Group showed that depo-
Siberian craton bears some signatures that may serve as ties in sition of the Peschernyi marine limestones was followed by a new
regional correlations. Figure 28.6 shows lithostratigraphic corre- evolutionary stage of the foreland basin, with orogens and the
lation of the Oselok Group with the Taseeva Group of the bordering alluvial plains originating around the southwestern
Yenisei Ridge, the Olkha and Moty Groups of the Irkutsk province, edge of the Siberian craton (Sovetov & Blagovidov 2004;
326
J. K. SOVETOV
Fig. 28.6. (a) Correlation of Late Neoproterozoic deposits in the southwestern Siberian Craton and stratigraphic position of the tillite units, complemented and modified after Sovetov et al. (2007a). Formation name
abbreviations: Ayan, Ayankan Fm.; Bol, Bol’shoi Lug Fm.; Bor, Borakun Fm.; Ch, Chencha Fm.; Kal, Kalancha Fm.; Khuzh, Khuzhir Fm.; Kul, Kulekin Fm.; M, Mariinka Fm.; Nik, Nikol’sky Fm.; Shan, Shanhar Fm.; Sham,
Shamanka Fm.; Tb, Tirbes Mb.; Ul, Uluntui Fm.; Ud, Uda Fm.; Vor, Vorogovka Grp.; Mr, Marnya Fm.; Gol, Goloustnaya Fm.; Tb Mbr, Tirbess Mb. Logs: 1, by (Sovetov et al. 2007a), with additions from Karpinskaya,
Karpinsky & Ustalov (pers. comm.); 2, modified after Vinogradov et al. (1994); 3, after Tyschenko (1980); 4, by Sovetov et al. (2007a) with addition from Shenfil (1991). (b) Map showing location of stratigraphic logs shown in
(a). (c) Map showing distribution of late Cryogenian diamictite and inferred glacial extent for the Siberian Platform. Localities as follows: a, Kutukas river, Enisei Ridge, Chapa Group, Stolbovaya Formation (this paper); b,
Chapa and Teya rivers, Enisei Ridge, Chapa Group, Pod’em Formation (this paper); c, Tagul River, Cis’Sayan region, Oselok Group, Marnya Fm. (Sovetov unpublished data); d, Birusa and Uda rivers, Cis’Sayan region, Oselok
Group, Marnya Fm. (this paper); e, Irkut River, Irkut Cis’Sayan region, Olkha Fm. (Mordvin 1972); f, Buguldeika River, southwestern Cis’Baikal region, Baikal Group, Goloustnaya Fm. (this paper); g, Right Ulkan River,
northwestern Cis’Baikal region, Baikal Group Goloustnaya Fm. (Korobeinikov & Semeikina 1978); h, Patom River, Patom Upland, Dal’nyaya Taiga Group, Dzhemkukan Fm. (this paper); i, Patom River, Patom Upland,
Dal’nyaya Taiga Group, Dzhemkukan Fm.; j, Nichatka Lake, Aldan Shield, Nichatka Fm.; k, Siberian Platform, Markha River, Borehole 225–0, Conglomerate suite (Kraevsky & Shishkin 2000).
MARNYA FORMATION, OSELOK GROUP 327

Sovetov et al. 2007a). Continental molasse deposited during a contribution of the IUGS- and UNESCO-funded IGCP (International
the Ediacaran (Late Vendian) collision belongs to the Greben Geoscience Programme) project #512.
regional stratigraphic stage.
Glaciogenic deposits of the Marnya Fm., interpreted as tillites,
and stratigraphic equivalents of the formation belong to the upper References
diamictite unit in the succession of two regional stratigraphic
stages (Sovetov et al. 2007a). The lower unit was deposited Anatol’eva, A. I. 1972. Pre-Mesozoic Redbed Deposits, Trudy Inst.
during a period that can be correlated with the Early Cryogenian Geol. Geophiz. Sibir. Divis. Akad. Nauk SSSR, Nauka, Novosibirsk,
glacial in a setting of rifting, volcanism and subsidence of deep 190, 323 (in Russian).
basins (Sovetov 1997). Deposition of the Marnya glacial deposits Berzin, N. A. 1967. Main Fault Zone in the East Sayan. Nauka, Moscow
is thought to be coeval with the early stage of the Late Cryogenian (in Russian).
(Early Vendian) foreland basin (Sovetov 2002b; Sovetov & Komlev Bragin, S. S. 1984. On relationship of the Precambrian Karagassy and
2005) and with formation of post-glacial shallow fringing seas in the Oselok Groups in the Sayan region. In: Khomentovsky, V. V.
southwestern part of the craton. The Ulyakha glaciogenic deposits (ed.) The Late Precambrian and the Early Palaeozoic Stratigraphy.
can be equated to a Late Cryogenian glaciation (Halverson et al. Central Siberia. IGG, Novosibirsk, 133–147 (in Russian).
2005), for the following reasons: (i) higher stratigraphic position Bragin, S. S. 1985. Late Precambrian Oselok Group in the Sayan
of the Ulyakha tillite relative to the Chivida diamictite (tillite) of region (stratigraphic division and correlation). In: Khomentovsky,
V. V. (ed.) The Late Precambrian and the Early Palaeozoic of
the Yenisei Range (Sovetov 1997; Sovetov et al. 2007a) constrained
Siberia. The Vendian and the Riphean. IGG, Novosibirsk, 44 –57
by the coeval c. 700 Ma volcanism (Postnikov et al. 2005); (ii)
(in Russian).
occurrences of Vendian-Ediacaran Metazoan in the Ozerki Mb. Bragin, S. S. 1986. Stratigraphy of the Late Riphean Karagassy Group in
(Sovetov & Komlev 2005; Sovetov 2007; Sovetov & Solovetskaya the Sayan region: Some problems. In: Khomentovsky, V. V. &
2008); (iii) post-glacial regional transgression (Sovetov 2007) and Shenfil, W. Yu. (eds) The Late Precambrian and the Early Palaeo-
correlation of related terrigenous deposits (Bolshaya Aisa Mb.) zoic of Siberia. Stratigraphy and Palaeontology. IGG, Novosibirsk,
with the Nepa Fm. containing microfossil assemblages as in the 32– 39 (in Russian).
Ediacaran Pertatatak Fm. in central Australia (Moczydłowska Burns, S. J., Haudenschild, U. & Matter, A. 1994. The strontium
et al. 1993; Golubkova & Raevskaya 2007). isotopic composition of carbonates from the late Precambrian
Chemostratigraphy suggests that the glaciation on the Siberian (c. 560– 540 Ma) Hukf Group of Oman. Chemical Geology, 111,
craton recorded in the Marnya Fm. can be equated to the Marinoan 269– 282.
epoch, which, according to the modern d13C time scale, is slightly Butler, R. F. 1992. Palaeomagnetism: Magnetic Domains to Geologic
older than the base of Doushantuo cap dolomite 635.2 + 0.6 Ma Terrains. Blackwell Science, Oxford.
(Condon et al. 2005). The evidence that the Ulyakha glacial Calver, C. R. 2000. Isotope stratigraphy of the Ediacaran (Neoproterozoic
deposits are stratigraphically lower that the Gaskiers tillite is pro- III) of the Adelaida Rift Complex, Australia, and the overprint of water
vided by: (i) correlation of the Aisa syncollisional molasse in the column stratification. Precambrian Research, 100, 121–150.
Sayan region with the Oskoba Fm. in the Baikit region Chumakov, N. M. 1993. Middle Siberian Riphean glaciohorizon. Strati-
(Fig. 28.6), with the latter marked by an intense d13C low (referred graphy. Geological Correlation. Vol. 1. No 1, 21– 34 (in Russian).
to as the Pre-Manykai negative anomaly by Pokrovsky 1996 Chumakov, N. M., Pokrovsky, B. G. & Melezhik, V. A. 2006. Patom
and Sovetov et al. 2007b) that correlates with the Zhuya d13C nega- Complex geological history, Late Precambrian, Middle Siberia.
Doklady Earth Sciences, 413, 379–383 (in Russian).
tive anomaly (Pokrovsky et al. 2006) in the Patom region and with
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005.
the Wonoka anomaly (Halverson et al. 2005) in South Australia; (ii)
U–Pb ages from the Neoproterozoic Doushantuo Fm., China.
position of a d13C positive anomaly in the Peschernyi Member rocks Science, 308, 95– 98.
(Uda Fm.), which is thought to be correlated with a similar anomaly Donskaya, T. V., Salnikova, E. B. et al. 2002. Early Proterozoic post-
below the Gaskiers glaciation; (iii) tectonothermal events in terranes collisional magmatism at the southern flank of the Siberian craton:
on the craton SW related to their post-glacial accretion at 600– new geochronological data and geodynamic implications. Doklady
550 Ma (Nozhkin et al. 2007), which is confirmed by the available Earth Sciences, 382, 125–129 (in Russian).
87
Sr/86Sr ratios in the Peschernyi Mb. limestone. Dubin, P. V., Khomentovsky, V. V. & Yakshin, M. S. 1969. New
Based on the widespread nature of diamictite and breccia at the data on Late Precambrian geology of the Sayan region. In:
base of the Ediacaran cover of the Siberian Platform, a thick ice Sokolov, B. S. (ed.) Early Cambrian and Late Precambrian Strati-
sheet occupied most of the southwestern craton area (Fig. 28.6). graphy of the Southern Siberian Craton. Nauka, Moscow, 86 –101 (in
Erosion of Neoproterozoic sediments was locally deeper than Russian).
1000 m; erosional valleys reached widths up to 50 km and radiated Egorova, O. P., Urumov, Yu. D., Volynets, Yu. N. & Vernoslova,
centrifugally to the west, SW and south, following the advance of Z. S. 1971. New data on geology and phosphate potential of Upper
the glaciers. At the same time, another glaciation centre, inferred Riphean strata in the Iya-Tumanshet province of the Sayan region.
from the palaeocurrent directions in an outwash fan of the Karap- In: Fainshtain, G. H. (ed.) Geology and Mineral Deposits of the
chatui Mb., was located in the extreme southwestern basement Siberian Craton. Nedra, Moscow, 69– 84 (in Russian).
inliers in the Sayan region, and may record the complex structure Gladkochub, D. P., Donskaya, T. V., Mazukabzov, A. M., Pono-
of a single ice sheet. The continental ice sheet extended as far as marchuk, V. A. & Stanevich, A. M. 2005. Neoproterozoic
remnant deep seas on the Patom passive margin, discharging gabbro dolerites of the Biryusa terrane (southern Siberian craton) as
possible indicators of the Rodinia dispersal, In: Sklyarov, E. V.
gravity flows to produce the deep-water fan deposits of the
(ed.) Geodynamic Evolution of the Lithosphere in the Central Asian
Dzhemkukan Fm. (Chumakov et al. 2006).
Orogen (From Ocean to Continent). Workshop Proceedings. Book
1. IZK, Irkutsk, 59 –62 (in Russian).
Funding for the research was provided by the Russian Foundation for Basic Gladkochub, D. P., Wingate, M. T. D., Pisarevsky, S. A., Donskaya,
Research (grants nos 04-05-65299 and 08-05-00959), which was indispensable T. V., Mazukabzov, A. M., Ponomarchuk, V. A. & Stanevich, A.
for regional sedimentological and stratigraphical studies, and by the Earth M. 2006. Mafic intrusions in southwestern Siberia and implications
Science Department (ESD) of the Russian Academy of Sciences (Project 18 of for a Neoproterozoic connection with Laurentia. Precambrian
the ESD Program ‘Biosphere origin and evolution’ and Project 10.1 of the Research, 147, 260–278.
RAS-ESD Program ‘Central Asian orogen: geodynamics and stages of continen- Golubkova, E. Yu. & Raevskaya, E. G. 2007. Lower Vendian complex
tal crust formation’). I wish to thank D. Metelkin who performed the palaeomag- of microfossils from the interior part of the Siberian platform. In:
netic analysis of samples from the Marnya Fm. and my field companions, Semikhatov, M. A. (ed.) The Rise and Fall of the Vendian (Edia-
postgraduate students D. Komlev, M. Medvedev and A. Kulikova. This represents caran) Biota. Origin of the Modern Biosphere. Transactions of the
328 J. K. SOVETOV

International Conference of the IGSP 493. GEOS, Moscow, 39– 42 Pokrovsky, B. G., Melezhik, V. A. & Buyakaite, M. I. 2006. Carbon,
(in Russian). oxygen, and strontium isotope compositions of Late Precambrian
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & rocks in the Patom complex, Central Siberia. Paper 1. Results, isoto-
Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon- pic stratigraphy, and dating problems. Lithologia i Poleznye Iskopae-
isotope record. GSA Bulletin, 117, 1181–1207. mye, 5, 1 –26 (in Russian).
Khabarov, E. M. & Ponomarchuk, V. A. 2005. Carbon isotopes in the Postnikov, A. A., Nozhkin, A. D. et al. 2005. New data on the age of
Upper Riphean deposits of the Baikal Group in western Cisbaikalia: Neoproterozoic deposits of the Chingasan and Vorogovka Groups in
stratigraphic implications. Russian Geology and Geophysics, 46, the Yenisei Ridge. In: Sklyarov, E. V. (ed.) Geodynamic Evolution
1019– 1027 (in Russian). of the Lithosphere in the Central Asian Orogen (From Ocean to Con-
Khomentovsky, A. S. 1950. Some data on Geology of Tumanshet salt- tinent). Workshop Proceedings, Book 2. IZK, Irkutsk, 71– 74 (in
bearing region (Eastern Siberia). Bulletin of Moscow Society of Russian).
Nature Researchers, 15, 65 – 79 (in Russian). Sez’ko, A. I. 1988. Principal stages of continental crust forming in the
Khomentovsky, V. V., Shenfil’, V. Yu., Yakshin, M. S. & Butakov, Cis-Saya. In: Letnikov, F. A. (ed.) Evolution of Crust in Precam-
E. P. 1972. Late Precambrian and Early Cambrian Reference brian and Palaeozoic. Sayan-Baikal Mountain Region. Nauka Sib.
Sections in the Siberian Platform. Nauka, Moscow (in Russian). Division, Novosibirsk, 7 –41 (in Russian).
Khomentovsky, V. V., Faizulin, M. Sh. & Karlova, G. A. 1998. The Shenfil, V. Yu. 1991. Late Precambrian Stratigraphy of the Siberian
Vendian Nemakit-Daldyn stage in the southwestern Siberian craton. Platform. Nauka, Novosibirsk (in Russian).
Doklady Earth Sciences, 362, 813– 815 (in Russian). Shatsillo, A. V., Didenko, A. N. & Pavlov, V. E. 2005. Two competing
Konstantinova, L. N. 2005. Katanga area. Section II. Vendian. In: palaeomagnetic directions in the Late Vendian: new data for the SW
Melnikov, N. V. (ed.) Stratigraphy of Oil and Gas Basins of region of the Siberian Platform. Russian Journal of Earth Sciences, 7,
Sibera. Riphean and Vendian of Siberian Platform and its Plaited 3 –24.
Border. Academic Publishing House ‘Geo’, Novosibirsk, 187– 193 Smethurst, M. A., Khramov, A. N. & Torsvik, T. H. 1998. The
(in Russian). Neoproterozoic and Palaeozoic palaeomagnetic data for the Siberian
Korobeinikov, N. K. & Semeikina, L. K. 1978. The nature of inter- Platform: From Rodinia to Pangea. Earth Science Research, 43,
relation of Baikal Group and Middle Proterozoic volcanic – plutonic 1 –24.
complex in the north-western Cis’Baikal region. In: Khomentovsky, Sovetov, J. K. 1977. Upper Precambrian Sandstones in the Southwestern
V. V. (ed.) News in Stratigraphy and Palaeontology of Late Precam- Siberian Platform. Nauka, Novosibirsk (in Russian).
brian in the Eastern and Northern Regions of Siberia. Institute Sovetov, J. K. 1997. Late Riphean rifting and the Baikal geodynamic
Geology and Geophysics of SB Academy of Sciences of USSR, cycle of the Siberian Platform. In: Koroteev, V. A. (ed.) The
Novosibirsk, 134– 146 (in Russian). Riphean of Northern Eurasia: Geology and General Problems of
Kraevsky, B. G. & Shishkin, B. B. 2000. Study state and direction of Stratigraphy. Institute of Geology and Geochemistry, Ekaterinburg,
stratigraphy investigation of Riphean deposits in the inner regions 223– 230 (in Russian).
of Siberian Platform. In: Krasnov, V. I. (ed.) Stratigraphy and Sovetov, J. K. 2002a. Vendian glaciation on the Siberian craton. In: Let-
Palaeontology of Siberia. Siberian Scientific Research Institute of nikov, F. A. (ed.) Geology, Geochemistry, and Geophysics at the
Geology, Geophysics and Mineral Deposits, Novosibirsk, 23 –31 Turn of the 20th-to-21st Century. Proceedings, Russian Conference.
(in Russian). IZK, Irkutsk, 122–124 (in Russian).
Metelkin, D. V., Belonosov, I. V., Gladkochub, D. P., Donskaya, T. Sovetov, J. K. 2002b. Vendian foreland basin of the Siberian cratonic
V., Mazukabzov, A. M. & Stanevich, A. M. 2005. Palaeomagnetic margin: Palaeopangean accretionary phases. Russian Journal of
directions from Nersa intrusions of the Biryusa terrane, Siberian Earth Sciences, 4, 363– 387.
craton, as a reflection of tectonic events in the Neoproterozoic. Sovetov, J. K. 2006. Vegetation along Vendian rivers, climate zonation,
Russian Geology and Geophysics, 46, 395–411 (in Russian). and palaeogeography of the Siberian craton in the Late Vendian. In:
Metelkin, D. V., Vernikovsky, V. A. & Kazansky, A. Yu. 2007. The Sklyarov, E. V. (ed.) Geodynamic Evolution of the Lithosphere in
Neoproterozoic stage of the Rodinia evolution in the light of new the Central Asian Orogen (From Ocean to Continent). Proceedings
palaeomagnetic data from the western margin of the Siberian of Workshop on ESD RAS Integration Research Programs, 2. IZK,
craton. Russian Geology and Geophysics 48, 42– 59 (in Russian). Irkutsk, 143– 146 (in Russian).
Miall, A. D. 1985. Architectural-element analysis: A new method of Sovetov, J. K. 2007. New occurrence of soft body Vendian–Ediacaran
facies analysis applied to fluvial deposits. Earth-Science Review, like Metazoa in south-west of the Siberian Platform, Oselok group,
22, 261– 308. Biruysa Cis Sayan. In: Semikhatov, M. A. (ed.) The Rise and Fall
Moczydłowska, M., Vidal, G. & Rudavskaya, V. A. 1993. Neoproter- of the Vendian (Ediacaran) Biota. Origin of the Modern Biosphere.
ozoic (Vendian) phytoplankton from the Siberian platform, Yakutia. Transactions of the International Conference of the IGSP 493.
Palaeontology, 36, 495–521. GEOS, Moscow, 33 –37 (in Russian).
Mordvin, A. P. 1972. The Khuzhir ore occurrence of gold-bearing con- Sovetov, J. 2008. Marinoan glaciation in the Siberian craton: locality,
glomerates, In: Fainshtain, G. H. (ed.) Geology and Gold Potential erosional forms, deposits and constraints to age. CGC-04 Neoproter-
of the Riphean and Vendian Strata in the Southern Periphery of the ozoic ice ages: Quo vadis? – Part 2. International Geological
Irkutsk Amphitheatre. Vostochno-Sibirskoe Izdatelstvo, Irkutsk, Congress. Oslo. 6– 14 August. Abstracts. File://E:/33IGC/
154– 189 (in Russian). 1343311.html.
Nozhkin, A. D., Turkina, O. M., Sovetov, Yu. K. & Travin, A. V. Sovetov, J. K. & Blagovidov, V. V. 2004. Reconstruction of sedimen-
2007. The Vendian accretionary event in the southwestern margin tary basin: an example of Vendian foredeep – foreland basin in south-
of the Siberian craton. Doklady Earth Sciences, 415A, 866–871 (in west of Siberian Platform. In: Leonov, Yu. G. & Volozh, Yu. A.
Russian). (eds) Sedimentary Basins: Methods of Research, Structure and Evol-
Pelechaty, S. M. 1998. Integrated chronostratigraphy of the Vendian ution. Transactions, 543. Scientific World, Moscow, 159– 210 (in
System of Siberia: implications for a global stratigraphy. Journal of Russian)
the Geological Society, London, 155, 957– 973. Sovetov, J. K. & Komlev, D. A. 2005. Tillite at base of the Oselok Group
Pisarevsky, S. A. & Natapov, L. M. 2003. Siberia in Rodinia. Tectono- in the Sayan region and the position of the lower boundary of the
physics, 375, 221–245. Vendian in the southwestern Siberian Platform. Stratigraphy and
Pokrovsky, B. G. 1996. Boundary between Proterozoic and Palaeozoic: Geological Correlation, 13, 3– 34.
isotope anomalies in the sedimentary sections of Siberian Platform Sovetov, J. K. & Blagovidov, V. V. 2006. Late Riphean sedimentation
and global environments change. Lithologia i Poleznye Iskopaemye, in Iya-Tumanshet aulacogene (south-west of the Siberian Platform):
4, 376–392 (in Russian). supersequences and correlation with Riphean stratotype in the
Pokrovsky, B. G. & Missarzhevsky, V. V. 1993. Isotope correlation of Bashkiriyan anticlinorium. In: Maslov, A. V. (ed.) Lithological
boundary Precambrian/Cambrian beds of the Siberian Platform. Aspects of Geology Stratified Environments. Materials of 7th Ural
Doklady Earth Sciences, 329, 768– 771 (in Russian). Lithological Meeting. Institute Geology and Geochemistry of Ural
MARNYA FORMATION, OSELOK GROUP 329

Division of Russian Academy of Sciences, Ekaterinburg, 248–250 Margin to Alleghenian– Variscan Collision. Geological Society of
(in Russian). America, Special Paper 423, 549– 578.
Sovetov, J. K. & Donskaya, T. V. 2006. Basement in south-west of Sovetov, J. K., Blagovidov, V. V. & Talibova, A. G. 2007b. Carbon
Siberian Platform by analysis of erratic stones in the Early Vendian isotopes in Vendian carbonates in the southwestern Siberian craton:
tillite. In: Sklyarov, E. V. (ed.) The Geodynamic Evolution of the correlation with geodynamic and palaeoclimate events and palaeo-
Lithosphere in Central Asian Orogen (From Ocean to Continent). geography. In: Halimov, E. M. (ed.) Proceedings of the Vinogradov
Proceedings of Workshop on ESD RAS Integration Research Pro- XVIII Symposium on Isotope Geochemistry. Moscow, 245– 246 (in
grams, 2. IZK, Irkutsk, 147– 150 (in Russian). Russian).
Sovetov, J. & Solovetskay, L. 2008. Oldest Vendian-Ediacaran Turkina, O. M., Bibikova, E. V. & Nozhkin, A. D. 2003. Stages and
fossils in the Oselok Group: Contribution to Late Neoproterozoic geodynamic settings of Early Proterozoic granite formation on the
(Ediacaran) age of sea transgression and origin of the Siberian plat- southwestern margin of the Siberian craton. Doklady Earth Sciences,
form cover. HPF-07 Rise and Fall of the Ediacaran (Vendian) 389, 159–163 (in Russian).
Biota. International Geological Congress, Oslo, 6– 14 August. Tyschenko, L. F. 1980. Regional correlation of Moty Formation deposits
Abstracts. File://E:/33IGC/1343889.html. in the Irkutsk amphitheatre. In: Karagodin, Yu. N. (ed.) The Pro-
Sovetov, J. K., Kulikova, A. E. & Medvedev, M. N. 2007a. Sedi- blems of Lithostratigraphy. Nauka, Novosibirsk, 149–158 (in
mentary basins in the southwestern Siberian craton: Late Neoproter- Russian).
ozoic –Early Cambrian rifting and collisional events. In: Vinogradov, V. I., Pokrovsky, B. G. et al. 1994. Isotopic geochemistry
Linnemann, U., Nance, R. D., Kraft, P. & Zulauf, G. (eds) The and age of Upper Precambrian deposits in the West Siberian Platform.
Evolution of the Rheic Ocean; From Avalonian-Cadomian Active Lithologia i Poleznye Iskopaemye, 4, 49 –76 (in Russian).
Chapter 29

The Tsagaan Oloom Formation, southwestern Mongolia

FRANCIS A. MACDONALD
Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138, USA (e-mail: fmacdon@fas.harvard.edu)

Abstract: The Tsagaan Oloom Formation (Fm.) in southwestern Mongolia contains two Neoproterozoic glacial deposits, with diamic-
tite in the Maikhan Ul Member (Mb.) and in the Khongoryn Mb., which are separated by over 500 m of limestone. The Maikhan Ul Mb.
ranges in thickness between 5 m and greater than 300 m, expanding in deeper-water sections towards the SW, where it is composed of
two massive diamictites separated by over 100 m of sandstone, siltstone and shale. The basal 10 m of the overlying Tayshir Mb. of the
Tsagaan Oloom Fm. consists of a fine-laminated, dark grey limestone. The Khongoryn Mb. is composed primarily of limestone clasts in a
shale matrix, and is between 0 and 23 m thick. The overlying Ol Mb. contains sedimentary structures characteristic of basal Ediacaran cap
carbonates including micropeloids, tubestone stromatolites, giant wave ripples and former aragonite crystal fans. U– Pb evaporation ages
from zircons in the underlying Dzabkhan Volcanics constrain the Tsagaan Oloom Fm. to ,773 Ma, and tuffs within the Maikhan Ul and
Tayshir members testify to the potential for additional geochronology. The Cryogenian organic-rich limestone of the Tayshir Mb., which
lies between the two glacial deposits, is ideally suited for geochemical studies and has been the subject of several carbon, strontium and
rare earth element investigations. Limited palaeomagnetic studies suggest a mid- to low-latitude position of the Dzabkhan platform
during deposition of the glaciogenic strata, and additional studies are in progress.

Neoproterozoic diamictites are present on the Dzabkhan platform (Khomentovsky & Gibsher 1996; Lindsay et al. 1996). Macdonald
(also referred to as the Zavkhan basin) of southwestern Mongolia, et al. (2009) grouped the lower two diamictites and the intervening
in a .100 km NW – SE trending belt, with additional discontinu- clastic units together in the Maikhan Ul Mb., while referring to the
ous exposures further north. The most complete exposures of the diamictite c. 500 m higher in the sequence as the Khongoryn
Tsagaan Oloom Fm. are between the Khasagty-Nuru ridge and member. Levashova et al. (2010) informally referred to the
the Dzabkhan River (Fig. 29.1). The geology of the Dzabkhan plat- Maikhan Ul diamictite as the Tayshir Fm. Here, we do not
form was first described by Bezzibetsev (1986), who divided the follow this nomenclature because it creates unnecessary confusion,
stratigraphy into three formations (the Dzabkhan, Tsagaan Oloom particularly as the overlying carbonates have been previously
and Bayan Gol). Subsequent work focused on the early Cambrian called the Tayshir Mb. (Macdonald et al. 2009).
palaeontology of the Bayan Gol Fm., with an eye for correlation
with Siberia; the results of these studies were published entirely
in Russian (for a list of these references see Brasier et al. 1996b). Structural framework
The first descriptions in English came in 1996 with the publication
of a Geological Magazine issue dedicated to the Neoproterozoic- The Dzabkhan terrane (also referred to as the Baydaric microcon-
Cambrian stratigraphy of southwestern Mongolia (Brasier et al. tinent when grouped with the Baidrag terrane, Fig. 29.1a) is a com-
1996a). The studies reported therein were the product of two inter- posite Precambrian terrane, hosting a heterogeneous Archaean and
national field excursions, one in 1991 as part of the 21st Joint Proterozoic crystalline basement intruded by c. 805 –770 Ma con-
Soviet – Mongolian Palaeontological Expedition (Zhegallo & tinental arc volcanism (Badarch et al. 2002; Zhao et al. 2006).
Zhuravelev 1991), and a second in 1993 sponsored by IGCP Based on similarities in the Neoproterozoic stratigraphy, radio-
Project 303 and the Mongolian Academy of Sciences (Dorjnamjaa metric ages in the underlying basement (Badarch et al. 1998),
et al. 1993). The results of these excursions included the translation and the continuity of aeromagnetic anomalies associated with
of geological maps and measured sections into English (Khomen- the fringing Neoproterozoic ophiolites (Buchan et al. 2002), the
tovsky & Gibsher 1996), a reconnaissance chemostratigraphic southwestern margin of the Dzabkhan basin can be traced to the
characterization of the Tsagaan Oloom and Bayan Gol formations western margin of the Khubsugul basin along the Tuva-Mongolia
(Brasier et al. 1996b), and a detailed stratigraphic study of the border (Fig. 29.1a). The tectonic events that transformed the south-
Maikhan Ul diamictite at Tsagaan Gol (Lindsay et al. 1996). western and western margins of the Dzabkhan and Khubsugul ter-
Recently, Macdonald et al. (2009) conducted detailed chemo- ranes from continental arcs to thermally subsiding passive margins
and litho-stratigraphic studies on previously unstudied sections remain unclear. On the southern margin of the Dzabkhan terrane,
and discovered an additional diamictite higher in the succession. on the south side of the Khasagty-Nuru ridge, the Dzabkhan and
This work supported the earlier conclusion of Brasier et al. the Tsagaan Oloom Formations are separated by as much as
(1996b) that the Maikhan Ul member is early Cryogenian in age 2 km of canabalizing, rift-related sediments with inter-fingering
and established that the Khongoryn diamictite is an end Cryogen- basalt that are referred to as the Shargyngol complex (Ruzhentsev
ian glacial deposit. With the new C-isotope chemostratigraphic & Burashnikov 1996). Facies patterns and the orientation of cross-
correlations and the documentation of a low-angle unconformity, beds in the Tsagaan Oloom Fm. indicate deepening to the SW
a .40 million year hiatus was identified within the Tsagaan (Macdonald et al. 2009). In the latest Ediacaran to early Cambrian,
Oloom Formation, above the Khongoryn diamictite but below the rifted passive margin began to subside again after a deposi-
the phosphorite horizon (Macdonald et al. 2009). tional hiatus of .40 Ma. It has been proposed that this accommo-
At Tsagaan Gol, the Maikhan Ul Mb. contains two diamictites dation space was created by flexure with the arrival of the
separated by over 100 m of sandstone, siltstone and shale Khantayshir-Dariv arc (Macdonald et al. 2009). With the early
(Lindsay et al. 1996). This creates a bit of confusion in the litera- Cambrian arc – continent collision, the Neoproterozoic stratigra-
ture, particularly with the discovery of an additional diamictite phy was shortened and repeated in thrust blocks with a basal
higher in the Tsagaan Oloom Fm., because the two diamictites detachment beneath the Dzabkhan Fm. The deformation is
within the Maikhan Ul Mb. have also been referred to as the largely brittle and thin-skinned, without involvement of the base-
upper and lower diamictites of the Tsagaan Oloom Fm. ment. Early Palaeozoic granites and narrow NW-trending grabens

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 331– 337. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.29
332 F. A. MACDONALD

Fig. 29.1. (a) Tectonic map of western Mongolia modified from Badarch et al. (2002) and Windley et al. (2007). Teeth on faults indicate the inferred dip of subduction
zones. Key: K-D Arc, Khantayshir-Dariv Arc; AZ, accretionary zone including the arc, metamorphic rocks and ophiolitic assemblages. (b) Geological map of the Tayshir
region, highlighting the members of the Tsagaan Oloom Fm. and the position of measured sections.

cut the Cambrian NNE-vergent structures. On the outcrop scale, and ribbonite that include nodular black chert and bed parallel,
the diamictite and carbonate rocks of the Tsagaan Oloom Fm. meandering ichnogenera (Goldring & Jensen 1996). Above the
are little deformed with no apparent strain. Sedimentary structures Zunne Arts Mb. of the Tsagaan Oloom Fm., the early Cambrian
are typically preserved in limestone, but are often obfuscated by Bayan Gol Fm. is composed of c. 1000 m of mixed carbonate
recrystallization in dolomite. and siltstone with a rich diversity of ichnogenera, small shelly
fossils and calcimicrobial patch reefs (Kruse et al. 1996).

Stratigraphy
Glaciogenic deposits and associated strata
The stratigraphy of the Dzabkhan basin (Fig. 29.2) begins with
.2 km of silicic to intermediate volcanic rocks of the Dzabkhan The Maikhan Ul Member
Fm. On the south side of the Khasagty-Nuru ridge, the Dzabkhan
Fm. is succeeded by as much as 2 km of sandstone turbidites and The Maikhan Ul Mb. progressively thickens to the SW (Fig. 29.3),
conglomerate that are referred to as the Shargyngol complex but also displays considerable variability on individual thrust
(Ruzhentsev & Burashnikov 1996). North of the Khasagty-Nuru blocks. For example, at the easternmost exposures on the
ridge, the Shargyngol complex is ,100 m thick and often absent, Tayshir Block (F718), the Maikhan Ul Mb. is only 6.7 m thick
with the Maikhan Ul Mb. of the Tsagaan Oloom Fm. non- and is composed predominantly of a massive cobble –boulder
conformably overlying the Dzabkhan Fm. The Maikhan Ul Mb. is clast diamictite, whereas just 1 km to the west (F713) it thickens
composed of diamictite, sandstone and shale, and varies in thickness to 81.6 m with multiple diamictite units separated by 57 m of
from 5 m to .275 m. The Maikhan Ul Mb. is overlain with a knife- massive, fine to coarse-grained sandstone. These sandstone
sharp contact by the Tayshir Mb., which consists of c. 570 m of lime- bodies are composed of graded centimetre- to metre-thick beds
stone that included three super-sequences (Macdonald et al. 2009). and contain no evidence of tidal influence (Fig. 29.3).
The Tayshir Mb. is succeeded by the Khongoryn diamictite, Further south, in more distal sections, the Maikhan Ul Mb. con-
which is composed of limestone cobble to boulder lonestones in a tinues to thicken. On the Khongoryn Block (F701), the Maikhan Ul
shale matrix and varies from 0 to 23 m in thickness. The Khongoryn Mb. fills palaeo-canyons and varies in thickness between 160 and
diamictite is in turn overlain by micropeloidal dolostones of the Ol 283 m. One palaeo-canyon, directly west of F701, is c. 125 m deep
Mb. (Macdonald et al. 2009). In the subsequent transgression, for- and 0.6 km wide. This palaeo-canyon has an erosive base that is
merly aragonite crystal fans are developed at the dolostone– mantled with a volcanic-clast, cobble conglomerate and is filled
limestone transition. Above the post-glacial transgression, the Ol with stratified diamictite units with dropstones and striated clasts
Mb. shallows upwards from grey limestone rhythmite into c. 10 m (Macdonald et al. 2009), thin-bedded sandstone beds and a
of limestone grainstone. The overlying Ulaan Bulagyn Mb. is up c. 0.5-m-thick carbonate bed. Two massive to bedded diamictite
to 500 m thick and is composed primarily of massive weathering units lie above the canyon fill, separated by 62 m of siltstone and
dolomite; however, in more distal sections, the Ulaan Bulagyn sandstone with rare cobble lonestones and two additional
Mb. thins to less than 100 m and is composed largely of limestone. c. 0.5-m-thick carbonate beds.
The Zunne Arts Mb. begins with distinct pink-coloured colum- At Tsagaan Gol, where the member measures 304 m, again,
nar stromatolites (Boxonia grumulosa) that overly a karstic surface two diamictite units are separated by a thick sequence of flat-
with metre-scale relief (Macdonald et al. 2009). The Boxonia bio- bedded shale, siltstone and sandstone (Lindsay et al. 1996).
herms are overlain by 10 –20 m of violet and green shale that are Cobble lonestones are present in both the basal and upper metre
variably phosphatized and interbedded with lenses of dolomite of this clastic succession, between the two massive diamictites.
and microcrystalline to nodular phosphorite. This phosphatic Khomentovsky & Gibsher (1996) also reported a measured
shale is overlain by more than 100 m of blue limestone rhythmite section from Urtor Tsakhir Mountain, c. 120 km west of Tayshir,
TSAGAAN OLOOM FORMATION 333

Bayan Gol Fm. c. 1 km of where the Maikhan Ul Mb. is even thicker but still preserves this
general stratigraphic pattern of two diamictite units separated by
BG
clastic & carbonate rocks
-8 -4 0 4 8 sandstone and siltstone. In this area, mudcracks are also well devel-
oped near the top of these intervening clastic units.
In both the upper and lower diamictite units of the Maikhan Ul
Zunne Arts Member
Mb., the most common lithology comprises a matrix-dominated
diamictite with shale and sandstone encasing sub-rounded cobble
derived from the underlying Dzabkhan Formation; granite, meta-
1400 morphic and carbonate clasts of unknown origin are also present.
Also, near the base of the Maikhan Ul Mb. at Tsagaan Gol,
C clasts of deformed soft sediment have been reported (Lindsay
P et al. 1996). Clast size varies from grit to blocks .2 m across.

1200 The Khongoryn Member


OO
Ulaan Bulagyn Member

OO The Khongoryn Mb. is thickest one gully east of Tsagaan Gol


(F723); however, like the Maikhan Ul Mb., there are significant
facies changes both from north to south and from east to west
OO
(Fig. 29.4). East of Tsagaan Gol, the diamictite is 23 m thick and
composed of pebble- to boulder-sized clasts of blue-grey limestone
1000 from the underlying Tayshir member in a dark grey shale matrix
that becomes more marly and lighter coloured up-section. Striated
clasts and limestone clasts with soft sedimentary deformation are
also present. Just 6 km west, near Tsagaan Gol, the diamictite
Tsagaan Oloom Formation

is nearly absent and only 2 m of recessive shale are preserved.


The Khongoryn Mb. is also well developed on the Khongoryn
C block, south of Bayan Gol (F708), where it consists of 14.7 m of
800 Maieberg
K Ol

sub-rounded limestone pebbles, cobbles and boulders in a grey


anomaly shale matrix. Both laterally and up-section, clasts are irregularly
OO distributed, varying from clast-poor facies to boulder nests. To
the NE of the Khongoryn block, the Khongoryn Mb. is either
thin or absent.

600
Tayshir anomaly Associated carbonate rocks
Tayshir Member

The basal 10 m of the Tayshir Mb., which overlies the Maikhan


C Ul Mb., is composed of a dark grey, millimetre-laminated lime-
OO
stone. Overall, the Tayshir Mb. consists of ,650 m of limestone
OO that record three regionally extensive sequences. The base of the
400 OO first sequence is defined by a c. 10-m-thick, dark grey (weather-
ing to tan), millimetre laminated limestone that is succeeded by
c. 100 m of limestone marl and rhythmite, shoaling up-section to
c. 20 m of grainstone. The second sequence begins with c. 10 m
of limestone marl and rhythmite followed by c. 200 m of mas-
C sively bedded, blue grainstone and microbialaminite. The third
200 Rasthof sequence begins with c. 50 m of limestone rhythmite and debris
anomaly flows with numerous black chert beds and nodules, and then
shallows up-section to c. 210 m of dark, fetid limestone micro-
-8 -4 0 4 8 bialaminite and minor grainstone with giant ooids (.0.5 cm
diameter).
MU

Maikhan Ul diamictite The Ol Mb., which overlies the Khongoryn diamictite, begins
with 7–40 m of buff to pink coloured, largely recrystallized,
0 micropeloidal dolostone. Low-angle cross-stratification (Aitken
(m) vv Dzabkhan Formation
vv 1991), tubestone stromatolites (Corsetti & Grotzinger 2005),
DV

vv c. 2 km of voclanic & and giant wave ripples (Allen & Hoffman 2005) are also present
vv
volcanoclastic rocks in the Ol Mb. dolomite (Fig. 29.4). The Ol Mb. transgresses
upwards into limestone ribbonite and then rhythmite with
limestone C chert
dolomite P phosphorite
grainstone & OO giant ooids
microbialaminite Fig. 29.2. Composite carbon chemo- and lithostratigraphy of the Dzabkhan
ribbonite stromatolite basin, modified from Macdonald et al. (2009). K, Khongoryn diamictite. The
trace fossils Rasthof anomaly was first well-documented in Namibia (Yoshioka et al. 2003)
rhythmite/shale and, like the Maieberg anomaly (Halverson et al. 2005), has now been reported
debris flow globally. The Tayshir anomaly was first documented in Mongolia (Macdonald
diamictite flooding et al. 2009) and a potentially correlative mid-Cryogenian anomaly is also
siltstone/ present in the Bonahaven Fm. of the Dalradian Supergroup in Scotland and
sandstone exposure
Ireland (Prave et al. 2009).
334 F. A. MACDONALD

Tayshir Block Khongoryn Tsagaan below the Maikhan Ul Mb., indicating only a limited hiatus.
Block, F701 Block,F724 However, it is not clear if this sandstone is part of the Shargyngol
E 1 km W suite or should be included within the Maikhan Ul Mb. South of
Tsagaan Gol, the clastic units between the Dzabkhan Fm. and
the lower Maikhan Ul diamictite unit thicken to over 100 m and
lack any evidence of glacial influence on sedimentation. Conver-
sely, to the east and north of Tayshir, both the Maikhan Ul Mb.
and the Dzabkhan Fm. thin, with the diamictites of the Maikhan
250 Ul resting on an erosional contact with the Dzabkhan Fm. or the
basement rock.
Contact between the Maikhan Ul Mb. and the overlying Tayshir
Mb. is very sharp. The Tayshir Mb. rests conformably on a later-
ally persistant, c. 10-cm-thick layer of red clay that marks the
top of the Maikhan Ul Mb.
The Khongoryn diamictite typically lies above blue-grey, giant
200 ooid grainstones of the lower limestone of the Tsaagan Oloom
Fm.; however, on the Khongoryn Block (F708), there is an
additional 7.7 m of black shale and rhythmite preserved above
the ooids. The erosion of this shale likely provides the detrital
matrix for the Khongoryn diamictite. The Khongoryn diamictite
is overlain with a sharp yet conformable contact by dolostone
and limestone of the Ol Mb.

150 F718 F713


Chemostratigraphy
volcanic Strontium isotope values rise from 0.7067 to 0.7073 in the lime-
massive stones of the Tayshir Mb. In the Ulaan Bulagyn Mb. 87Sr/86Sr
diamictite values rise from 0.7073 to 0.7077, and then in the Zunne Arts
Mb. from 0.7078 to over 0.7080 (Brasier et al. 1996b; Shields
matrix composition et al. 2002).
100 Carbonate d13C values in the dark-grey laminated limestone
sandstone
above the Maikhan Ul Fm. are moderately negative with values
siltstone increasing upwards through the overlying pink marls to þ8‰
(Fig. 29.2). Values plummet abruptly at the flooding surface in
shale the middle of the Tayshir Mb., reaching a low of – 7.5‰. Mac-
limestone donald et al. (2009) refer to this sudden drop in d13C values as
clast size the Tayshir anomaly. From this nadir, d13C values increase
smoothly to þ9‰ for the upper Tayshir Mb., with values reported
50 boulder as high as þ11‰ (Brasier et al. 1996b). Shields et al. (2002) also
cobble measured d13C of organic matter in the Tayshir Mb. of the Tsagaan
gravel Oloom Fm. and found that trends roughly followed those exhibited
by the d13C in carbonate.
clast composition Overlying the upper diamictite, d13C values in the Ol Mb. begin
granite around –1‰ and follow a sigmoidal profile (Fig. 29.2). Values
clastic return to c. –1‰ at the top of the dolostone, and then decrease
again at the limestone – dolomite transition, reaching a nadir of
carbonate –6‰. Above the Ol Mb., d13C values oscillate around þ3‰
0
volcanic for most of the Ulaan Bulagyn Mb., returning to 0‰ below the
metres
sub-Zunne Arts Mb. karstic surface.
Fig. 29.3. Stratigraphy of the Maikhan Ul Mb. Locations of sections are shown Sheilds et al. (1997, 2002) reported a Ce anomaly in the Tayshir
in Figure 29.1b. Section F724 is modified from Lindsay et al. (1996). Symbols Mb. from samples collected at Tsagaan Gol. In this section, the
used that are not in the legend are in Figure 29.2. recessive strata bearing the C-isotope anomaly are not exposed,
and thus, they did not document the transgressive sequence or
the negative C-isotope values.
c. 5-cm-tall former aragonite crystal fans present at the limestone –
dolostone transition. Crystal fans are present both as individual
blades growing upwards into the sediment, and as crystal fan Palaeolatitude and palaeogeography
shrubs that are over 10 cm across.
Recent palaeomagnetic studies on the 805–770 Ma Dzabkhan Fm.
indicate that the Dzabkhan terrane was located at a latitude of
Boundary relations with overlying and underlying 478 þ 168/–128 (Levashova et al. 2010). From palaeomagnetic
non-glacial units studies on peri-Siberian terranes, including the early Cambrian
Salaany Gol Fm. on the Dzabkhan terrane, Kravchinsky et al.
The Maikhan Ul Mb. rests with an erosive base on the Shargyngol (2001) concluded that the Tuva-Mongolia belt was at low latitude,
suite and the Dzabkhan Formation (Khomentovsky & Gibsher adjacent to Siberia throughout the Ediacaran and Cambrian.
1996), and fills palaeo-topography with conglomerates lining However, this study lacked a robust confidence test (i.e. only a
palaeo-valleys (Fig. 29.3). According to Lindsay et al. (1996), at reversal test with few samples and low resolution). Moreover, an
Tsagaan Gol, soft-sedimentary deformation is present in sandstone earlier study on the Salaany Gol Fm. gave entirely different
TSAGAAN OLOOM FORMATION 335

Tayshir Block Khongoryn Block Tsagaan Block


E 18 km 3 km 10 km W E 18 km W E 15 km 6 km W
50
C C C
C C
C C
C
25
C C
(metres)

Fig. 29.4. Stratigraphy of the Khongoryn


-25 diamictite and overlying Ol Mb. Positions
F710 F716 F715 F706 F725 F708 F726 F723 F724 of measured sections are in Figure 29.1b.
micropeloidal dolomite giant wave ripples Symbols used that are not in the legend are
tubestone stromatolites crystal fans
in Figures 29.2 and 29.3.

results (Evans et al. 1996), but it was also compromised by uncer- dropstones that penetrate laminated beds. At Tsagaan Gol,
tainty in the relative ages of the folds used in the fold test and poss- cobble dropstones are also present in both the basal and upper
ible magnetic overprints. Further palaeomagnetic studies on the metre of the clastic succession, between the two massive diamictite
Dzabkhan terrane are necessary, and are in progress (Gregory units. Moreover, in more proximal settings, such as on the Khon-
et al. 2007). Nonetheless, as non-skeletal carbonate is preferen- goryn and Tayshir blocks, rare lonestones are present within the
tially produced in the warmest parts of the surface ocean (Broecker sandstone beds. These observations indicate that the deposition
& Peng 1982), and the Tsaagan Oloom Fm. is dominated by of the clastic units was influenced, at least in part, by glaciation.
shallow-water carbonates, it is likely that the Dzabkhan terrane Macdonald (2009) inferred a pro-glacial environment, including
was situated at low latitudes (less than 308) throughout the Cryo- emergent conditions and proglacial lakes, for both the diamictite
genian and Ediacaran. and the clastic units of the Maikhan Ul Mb. from the presence of
Along with other peri-Siberian terranes, it has been suggested mud cracks and 0.5-m-thick carbonate beds. A pro-glacial
that the Dzabkhan terrane occupied a Precambrian position environment is further supported by high lateral facies variability.
between Siberia and Laurentia (Gladkochub et al. 2006), and Within this context, the intervening clastic units can be interpreted
rifted away from Siberia in the late Neoproterozoic (Sengor & as a step-back of the ice-line, and the upper diamictite as an
Natal’in 1996; Kuzmichev et al. 2001; Kuzmichev et al. 2005). ice-advance.
Sengor & Natal’in (1996) further posit that throughout the late The rise in 87Sr/86Sr from 0.7067 to 0.7073 in the limestone of
Neoproterozoic and early Palaeozoic, the Dzabkhan terrane was the Tayshir Mb. is mirrored in the Rasthof Fm. in Namibia and the
attached to the Central Mongolian Block, which along with other Keele Fm. in NW Canada, suggesting that the underlying Maikhan
terranes, stretched to the present day Sea of Okhotsk. Both the Ul diamictites are early Cryogenian glacial deposits (Halverson
Tuva-Mongolia (including the Dzabkhan terrane) and Central et al. 2007). The black laminated cap carbonate above the
Mongolian Blocks host Cambrian trilobites endemic to Siberia Maikhan Ul diamictites also contains a modest negative
(Astashkin et al. 1995) and Silurian brachiopods characteristic of C-isotope anomaly similar to the Rasthof Fm. (Yoshioka et al.
the peri-Siberian realm (Hou & Boucot 1990). Alternatively, citing 2003); the extremely enriched values of the Tayshir Mb. are also
similarities in SHRIMP ages on zircons, Zhao et al. (2006) and consistent with a Cryogenian age (Hoffman & Schrag 2002;
Demoux et al. (2009) have suggested that the Baydrag and Dzab- Halverson et al. 2005). The Tayshir anomaly (Macdonald et al.
khan terranes originated from the northern margin of Gondwana. 2009) can be correlated to the moderately negative 13dC values
This reconstruction is supported by the palaeomagnetic results of obtained from the exposure-surface riddled Gruis Fm. of northern
Levashova et al. (2010), which point to Neoproterozoic connec- Namibia (Halverson et al. 2005) and the Cryogenian Bonahaven
tions with India, South China, Tarim or Australia. Dolomite of the British-Irish Caledonides (McCay et al. 2006),
or to the Trezona anomaly in Australia (McKirdy et al. 2001)
and Namibia (Halverson et al. 2005).
Geochronological constraints A glacial origin of the Khongoryn diamictite is indicated by
the presence of striated clasts and dropstones that penetrate lami-
Although the diamictites of the Dzabkhan basin have not been nated beds. The Khongoryn diamictite is thin or absent on the
directly dated radiometrically, maximum age constraints on the most proximal sections to the NE of the map area (Fig. 29.1). In
glacial deposits are provided by zircons from rhyolites within more distal sections to the SW, the diamictite is composed of
the Dzabkhan Formation of 777 + 6 Ma (Zhao et al. 2006), cobble to boulder clasts of the underlying limestone within a
803.4 + 8.0 and 773.5 + 3.6 Ma (U– Pb laser evaporation, Leva- weakly bedded shale to marl matrix. This shale matrix was
shova et al. 2010). likely derived via erosion of the shale unit in the upper portion
of the Tayshir Mb., which is only present on the Khongoryn and
Tsagaan blocks. The lack of stratigraphic architecture within the
Discussion deposit, the irregular distribution of ice-rafted debris, such as
boulder nests, and the conformable overlying contact with the Ol
A glacial origin of the Maikhan Ul diamictite units is indicated by Mb. indicate that this deposit formed as a single rainout during
the presence of faceted and striated clasts, and bullet-shaped the terminal deglaciation.
336 F. A. MACDONALD

The overlying basal dolostone of the Ol Mb. is composed of of granite geochronology in the Bayankhongor Ophiolite Zone.
fine-laminated micropeloids and contains tubestone stromatolites, Chemical Geology, 192, 23 –45.
giant wave ripples and pseudomorphosed crystal fans. These Condon, D. J., Zhu, M., Bowring, S. A., Wang, W., Yang, A. & Jin, Y.
peculiar sedimentary structures, their specific order, and the 2005. U– Pb ages from the Neoproterozoic Doushanto Formation,
distinct, sigmoidal C-isotope profile are characteristic of basal China. Science, 308, 95– 98.
Ediacaran cap carbonates globally (Hoffman et al. 2007). This Corsetti, F. A. & Grotzinger, J. P. 2005. Origin and significance of tube
suggests that the underlying Khongoryn diamictite is an structures in Neoproterozoic post-glacial cap carbonates: example
end-Cryogenian glacial deposit (Macdonald et al. 2009), with from Noonday Dolomite, Death Valley, United States. Palaios, 20,
348– 363.
the termination bracketed elsewhere by U –Pb ages of
Demoux, A., Kroener, A., Badarch, G., Jian, P., Tomurhuu, D. &
635.51 + 0.54 Ma and 635.23 + 0.57 Ma (Condon et al. 2005). Wingate, M. T. D. 2009. Zircon ages from the Baydrag block and
The phosphorites in the Zunne Arts Mb. rest above a low-angle the Bayankhongor ophiolite zone: time constraints on late Neoproter-
unconformity. Ediacaran chemostratigraphic correlations indicate ozoic to Cambrian subduction- and accretion-related magmatism in
that this surface represents a .40 Ma depositional hiatus, and as central Mongolia. Journal of Geology, 117, 377–397.
such is unrelated to the glacial deposits in the Khongoryn Dorjnamjaa, D., Bat-Ireedui, Y. A., Dashdavaa, Z. & Soelmaa, D.
member (Macdonald et al. 2009). 1993. Precambrian–Cambrian Geology of the Dzavkhan Zone.
Further sedimentological studies on the Maikhan Ul Mb. are Earth Sciences Department, Oxford.
needed to understand the significance of the interbedded sandstone Evans, D. A. D., Zhuravlev, A. Y., Budney, C. J. & Kirschvink, J. L.
and siltstone, and to determine if the Maikhan Ul diamictites 1996. Palaeomagnetism of the Bayan Gol Formation, western
represent a single episode of deglaciation or multiple ice advances Mongolia. Geological Magazine, 133, 487–496.
and retreats. Stratigraphic studies are also needed to better under- Gladkochub, D. P., Wingate, M. T. D., Pisarevsky, S. A., Donskaya,
stand the basin dynamics of the Neoproterozoic– Cambrian T. V., Mazukabzov, A. M., Ponomarchuk, V. A. & Stanevich, A.
margins of the Dzabkhan terrane, and the relationships with M. 2006. Mafic intrusions in southwestern Siberia and implications
other Mongolian terranes. The palaeogeography of the Peri- for a Neoproterozoic connection with Laurentia. Precambrian
Siberian terranes also remains speculative. It is clear, however, Research, 147, 260– 278.
that island arcs surrounded the Dzabkhan terrane for much of the Goldring, R. & Jensen, S. 1996. Trace fossils and biofabrics at the
Neoproterozoic and Cambrian, and therefore there is excellent Precambrian-Cambrian boundary interval in western Mongolia.
potential for U – Pb zircon geochronology studies in the Dzabkhan Geological Magazine, 133, 403– 415.
basin. Furthermore, the low-grade and high organic content of Gregory, L. C., Meert, J. G., Levashova, N. & Gibsher, A. S. 2007.
Paleomagnetic and geochronologic data from Central Asia: infer-
the limestone in the Tayshir Mb. is ideally suited for multi-proxy
ences for Early Paleozoic tectonic evolution and timing of worldwide
studies to better constrain the geochemical evolution of Cryogen- glacial events. American Geophysical Union, Fall Meeting,
ian oceans. GP43C –1487.
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice,
I thank field assistants B. Boldoo, E. Oyun, T. Adiya, J. Otgonhuu and U. Bold. I A. H. N. 2005. Toward a Neoproterozoic composite carbon-isotope
also thank Bayassa and A. Amaglaan for help with logistics. I thank D. Jones and record. Geological Society of America Bulletin, 117, 1181–1207.
P. Hoffman for comments and help in the field. I am grateful to P. Hoffman and Halverson, G. P., Dudás, F. O., Maloof, A. C. & Bowring, S. A. 2007.
NSF grant EAR-0417422 (to PFH) for support of this work. This represents a con- Evolution of the 87Sr/86Sr composition of Neoproterozoic Sea-
tribution of the IUGS- and UNESCO-funded IGCP (International Geoscience water. Palaeogeography, Palaeoclimatology, Palaeoecology, 256,
Programme) project #512. 103–129.
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis;
testing the limits of global change. Terra Nova, 14, 129–155.
References Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635
Aitken, J. D. 1991. The Ice Brook Formation and Post-Rapitan, Late Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Planetary
Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. Science Letters, 258, 114– 131.
Geological Survey of Canada Bulletin, 404, 1 –43. Hou, H. F. & Boucot, A. J. 1990. The Balkhash-Mongolia-Okhotsk
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves in Region of the Old World Realm. Geological Society, London,
the aftermath of a Neoproterozoic glaciation. Nature, 433, 123– 127. Memoir, 12, 297– 303.
Astashkin, V. A., Pegel, T. V. et al. 1995. The Cambrian System of the Khomentovsky, V. V. & Gibsher, A. S. 1996. The Neoproterozoic –
Foldbelts of Russia and Mongolia. International Union of Geological Lower Cambrian in northern Govi-Altai, western Mongolia: regional
Sciences, London. setting, lithostratigraphy and biostratigraphy. Geological Magazine,
Badarch, G., Byamba, J. et al. 1998. Geological Map of Mongolia. 133, 371– 390.
Mineral Resources Authority of Mongolia, Ulaan Baator. Kravchinsky, V. A., Konstantinov, K. M. & Cogne, J.-P. 2001.
Badarch, G., Cunningham, W. D. & Windley, B. 2002. A new terrane Palaeomagnetic study of Vendian and Early Cambrian of South
subdivision for Mongolia: implications for the Phanerozoic Siberia and Central Mongolia: was the Siberian platform assembled
crustal growth of Central Asia. Journal of Asian Earth Sciences, at this time? Precambrian Research, 110, 61 –92.
21, 87 –110. Kruse, P. D., Gandin, A., Debrenne, F. & Wood, R. 1996. Early Cam-
Bezzubetsev, V. V. 1986. On the Precambrian –Cambrian stratigraphy brian bioconstructions in the Zavkhan Basin of western Mongolia.
of the Dzabkhan River Basin. Materials on the Geology of MPR, Geological Magazine, 133, 429– 444.
Gostopotekhizdat, 1963, 29– 42. Kuzmichev, A., Bibikova, E. V. & Zhuravlev, D. Z. 2001. Neoproter-
Brasier, M. D., Dorjnamjaa, D. & Lindsay, J. F. 1996a. The Neoproter- ozoic (800 Ma) orogeny in the Tuva-Mongolia Massif (Siberia):
ozoic to early Cambrian in southwest Mongolia: an introduction. island arc-continent collision at the northeast Rodinia margin.
Geological Magazine, 133, 365–369. Precambrian Research, 110, 109– 126.
Brasier, M. D., Shields, G., Kuleshov, V. N. & Zhegallo, E. A. Kuzmichev, A., Kroener, A., Hegner, E., Dunyi, L. & Yusheng, W.
1996b. Integrated chemo- and biostratigraphic calibration of early 2005. The Shishkhid ophiolite, nothern Mongolia: a key to the
animal evolution: Neoproterozoic – early Cambrian of southwest reconstruction of a Neoproterozoic island-arc system in central
Mongolia. Geological Magazine, 133, 445– 485. Asia. Precambrian Research, 138, 125–150.
Broecker, W. S. & Peng, T. H. 1982. Tracers in the Sea. LDEO Press, Levashova, N. M., Kalugin, V. M., Gibsher, A. S., Yff, J., Ryabinin,
Lamont-Doherty Earth Observatory, Palisades, NY. A. B., Meert, J. & Malone, S. J. 2010. The origin of the Baydaric
Buchan, C., Pfander, J. et al. 2002. Timing of accretion and collisional microcontinent, Mongolia: constraints from paleomagnetism and
deformation in the Central Asian Orogenic Belt: implications geochronology. Tectonophysics, 485, 306–320.
TSAGAAN OLOOM FORMATION 337

Lindsay, J. F., Brasier, M., Shields, G., Khomentovsky, V. V. & Shields, G., Stille, P., Brasier, M. & Atudorei, N.-V. 1997. Stratified
Bat-Ireedui, Y. A. 1996. Glacial facies associations in a Neoproter- oceans and oxygenation of the late Precambrian environment: a post
ozoic back-arc setting, Zavkhan Basin, western Mongolia. Geologi- glacial geochemical record from the Neoproterozoic of W. Mongolia.
cal Magazine, 133, 391–402. Terra Nova, 9, 218–222.
Macdonald, F. A. 2009. Neoproterozoic stratigraphy of Alaska and Shields, G. A., Braiser, M. D., Stille, P. & Dorjnamjaa, D. 2002.
Mongolia. PhD, Harvard University. Factors contributing to high d13C values in Cryogenian limestones
Macdonald, F. A., Jones, D. S. & Schrag, D. P. 2009. Stratigraphic and of western Mongolia. Earth and Planetary Science Letters, 196,
tectonic implications of a new glacial diamictite–cap carbonate 99– 111.
couplet in southwestern Mongolia. Geology, 37, 123–126. Windley, B. F., Alexeiev, D., Xiao, W., Kroener, A. & Badarch, G.
McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial 2007. Tectonic models for accretion of the Central Asian
trinity: Neoproterozoic Earth history within the British-Irish Orogenic Belt. Journal of the Geological Society of London, 164,
Caledonides. Geology, 34, 909– 912. 31– 47.
McKirdy, D. M., Burgess, J. M. et al. 2001. A chemostratigraphic over- Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003.
view of the late Cryogenian interglacial sequence in the Adelaide Systematic variations in C, O, and Sr isotopes and elemental concen-
fold –thrust belt, South Australia. Precambrian Research, 106, trations in Neoproterozoic carbonates in Namibia: implications
149– 186. for a glacial to interglacial transition. Precambrian Research, 124,
Prave, A. R., Fallick, A. E., Thomas, C. W. & Graham, C. M. 2009. 69– 85.
A composite C-isotope profile for the Neoproterozoic Dalradian Zhao, Y., Song, B. & Zhang, S. H. 2006. The Central Mongolian micro-
Supergroup of Scotland and Ireland. Journal of the Geological continent: its Yangtze affinity and tectonic implications. In: Jahn,
Society of London, 166, 1 –13. B. M. & Chung, L. (eds) Symposium on Continental Growth and
Ruzhentsev, S. V. & Burashnikov, V. V. 1996. Tectonics of the Orogeny in Asia. Taipei, Taiwan, 135– 136.
western Mongolian Salairides. Geotectonics, 29, 379– 394. Zhegallo, L. & Zhuravelev, A. Y. 1991. Guidebook for the
Sengor, A. C. & Natal’in, B. A. 1996. Paleotectonics of Asia: fragments International Excursion to the Vendian– Cambrian Deposits
of synthesis. In: Yin, A. & Harrison, M. (eds) The Tectonic Evol- of the Dzabkhan Zone of Mongolia. Palaeontological Institute,
ution of Asia. Cambridge University Press, Cambridge, 486– 640. Moscow.
Chapter 30

The Khubsugul Group, Northern Mongolia

FRANCIS A. MACDONALD* & DAVID S. JONES


Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138, USA
*Corresponding author (e-mail: fmacdon@fas.harvard.edu)

Abstract: The Khubsugul Group of northern Mongolia contains diamictites in the Ongoluk and Khesen formations that are succeeded
by a stratiform phosphorite deposit and .2 km of early Cambrian dolomite. The stratigraphy of the Khubsugul Group, including the two
diamictites, can be correlated with that of the Dzabkhan platform in southern Mongolia. By correlation, the Ongoluk diamictite is an early
Cryogenian glacial deposit. A glaciogenic origin is inferred from the presence of striated clasts and bed-penetrating dropstones. The
younger Khesen diamictite consists predominantly of a massive carbonate-clast diamictite, but also contains bed-penetrating dropstones
in rare stratified facies, and is inferred to be end Cryogenian in age. The two diamictites are separated by as much as 250 m of allodapic
carbonate. The phosphorite in the upper Khesen Formation (Fm.) is likely latest Ediacaran to early Cambrian in age and is separated from
the glacial deposits by a major hiatus. Consequently, no links can be made between the phosphogenesis and the glacial deposits. Only
limited geochemical, geochronological and palaeomagnetic results from the Khubsugul basin have been reported to date, but work is
ongoing and there is strong potential for future studies.

Diamictites in the Ongoluk and Khesen formations of the the basal member of the Khubsugul Group; however, the diamic-
Khubsugul Group of northern Mongolia are exposed discontinu- tite is underlain by an additional several hundred metres of
ously in a c. 250 km north –south belt (Fig. 30.1). The most com- clastic rocks that he included with the underlying Arasan Fm.
plete exposures of the two diamictites and the overlying carbonate Following Osokin & Tyzhinov (1998), we include these clastic
and phosphorite occur along the Khesen and Ongoluk Gols (tr. rocks with the Ongoluk Fm. of the Khubsugul Group, and refer
Rivers), on the west side of Lake Khubsugul (Fig. 30.2; to the diamictite in the Ongoluk Fm. as the Ongoluk diamictite.
50842.50 N, 1008110 E and 50844.30 N, 100812.20 E). The upper diamictite is in the basal Khesen Fm. of the Khubsugul
Most of the studies in the Khubsugul basin have focused on the Group and is herein referred to as the Khesen diamictite. The two
phosphorite deposits and the regional tectonics, with the diamic- diamictites are separated by 100– 250 m of carbonate of the upper
tites mentioned only in passing (e.g. Ilyin 1990, 1998). Although Ongoluk Fm.
much of this work is in the Russian literature, several key
manuscripts from Litologiya i Poleznye Iskopaemye have been
translated to English in Lithology and Mineral Resources (e.g. Structural framework
Osokin & Tyzhinov 1998; Ilyin 2004). Additionally, a Russian
field guide of the Khubsugul Basin was produced for an IGCP The Khubsugul terrane is a composite Precambrian terrane, host-
excursion to the phosphorite localities (Ilyin & Byamba 1980). ing a heterogeneous Archaean and Proterozoic crystalline base-
This field trip spawned the hypothesis that ice rings orbited ment intruded by c. 800 Ma continental arc volcanism (Badarch
Precambrian Earth, that the shadow of these rings initiated et al. 2002). Based on similarities in Neoproterozoic stratigraphy,
the Neoproterozoic glaciations, and that their collapse led to radiometric ages in the underlying basement (Badarch et al. 2002)
phosphogenesis and precipitated the Cambrian radiation and the continuity of aeromagnetic anomalies associated with
(Sheldon 1984). fringing Neoproterozoic ophiolites (Buchan et al. 2002), the south-
Geological work commenced in the Khubsugul basin in the western margin of the Dzabkhan platform can be traced to the
mid-1960s with the discovery of ore-grade phosphorites (Donov western margin of the Khubsugul basin along the Tuva-Mongolia
et al. 1967). The Khubsugul Group was originally described in border (Fig. 30.1, Macdonald 2011). The eastern boundary of the
detail, including the identification of diamictites in the Ongoluk Dzabkhan terrane is obscured by Palaeozoic intrusions (Badarch
Fm., by Ilyin (1973). Osokin & Tyzhinov (1998) later differen- et al. 1998) and, consequently, pre-Ordovician connections with
tiated a second diamictite in the basal Khesen Fm. and documented the Baidrag terrane remain ambiguous. Overlap assemblages indi-
the presence of both diamictites throughout the Khubsugul basin. cate that the Dzabkhan, Khubsugul, Baidrag and Tarvagatay
The diamictites in the Khubsugul basin have not been formally terranes had amalgamated into a single continental mass by the
named, and the formations within which they occur have been Devonian (Badarch et al. 2002).
named differently in Mongolia and Siberia (see Chumakov On both the Khubsugul and Dzabkhan terranes, Palaeoprotero-
2011). The diamictite at the base of the Ongoluk Fm. in the Khub- zoic basement is overlain by thick volcanic –volcaniclastic succes-
sugul basin is likely equivalent to the diamictite in the Khushatai sions (Badarch et al. 2002). On the Khubsugul terrane, the Sarkhoi
Fm. of the Sarkhoi Group (Osokin & Tyzhinov 1998). Otherwise, volcanic rocks have been interpreted as having a continental arc
the Khubsugul Group in Mongolia is largely correlative with the affinity (Kuzmichev et al. 2001). These are unconformably over-
Boxon Group in Siberia, with the Khesen Fm. roughly equivalent lain by the rift-related volcanic rocks and clastic sediments of
to the Zabit Fm. Correlations of specific diamictites in the Khesen the Dharkhat Group (Ilyin 1990). The Khubsugul terrane trans-
and Zabit formations are complicated by multiple conglomeratic formed from a continental arc to a thermally subsiding passive
horizons within both formations (Kheraskova & Samygin 1992; margin after the c. 800 Ma Shishkhid arc accreted to its western
Osokin & Tyzhinov 1998). margin and prior to rifting along its eastern margin (Kuzmichev
Ilyin (1973, 2004) referred to the upper c. 50 m of the lower et al. 2005).
diamictite as the ‘perforated shales’ after the holes left behind Ilyin (2004) documented a deepening to the west in the
from eroded carbonate clasts, and he defined the diamictite as phosphorite-bearing strata of the Khesen Fm. and the overlying

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 339– 345. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.30
340 F. A. MACDONALD & D. S. JONES

basal Khesen diamictite (Ilyin 2004). Poorly sorted, carbonate


clast conglomerates and syn-sedimentary folding are also
common in the uppermost Khesen formation. The upper c. 2 km
of the Khubsugul Group consists of late Ediacaran and Cambrian
platformal carbonates of the Erkhelnur Fm.

Glaciogenic deposits and associated strata

The Ongoluk diamictite

Near Lake Khubsugul, the Ongoluk diamictite is composed of a


matrix-supported, stratified diamictite that ranges in thickness
from c. 200 m to 415 m, thinning to the south and west. The thick-
est and most complete section of the Ongoluk Fm. is on the ridge
north of Khesen Gol (Fig. 30.4). There, the basal contact of the dia-
mictite is gradational with gravel-sized lonestones becoming more
common upwards, both in an unsorted sandstone matrix and in a
laminated argillite matrix, and, as such, the base is difficult to
Fig. 30.1. Tectonic map of western Mongolia modified from Badarch et al.
define. Sub-angular to sub-rounded gravel clasts of dolomite and
(2002) and Windley et al. (2007). Teeth on faults indicate the inferred dip of
subduction zones. K-D Arc, Khantayshir-Dariv Arc; AZ, accretionary zone
quartzite are most common, with occasional cobbles of granite.
including arcs, metamorphic rocks and ophiolites; NP, Neoproterozoic; Ord,
Clasts become larger and more common in the upper c. 100 m of
Ordovician; Carb, Carboniferous. the diamictite (the ‘perforated shale’ member of Ilyin 1973,
2004), consisting predominantly of dolomite cobbles in an argillite
or siltstone matrix.
At Ongoluk Gol, the base of the Ongoluk diamictite is not
early Cambrian shelf carbonates, and suggested these were depos- exposed, but c. 130 m of clast poor siltstone and very coarse sand-
ited in a rift graben. Macdonald et al. (2009) alternatively posited stone are present that are very similar to the lower c. 300 m at
that most of the Neoproterozoic sediments on the Dzabkhan and Khesen Gol. The ‘perforated shale’ member is c. 50 m thick at
Khubsugul terranes were deposited on a thermally subsiding Ongoluk Gol, beginning with a 5-m-thick, massive, clast-
rifted margin, but that the late Ediacaran to early Cambrian phos- supported dolomite diamictite, and continuing upwards with a
phorites and overlying early Cambrian deposits formed in a fore- dark, laminated argillite matrix with sub-rounded to angular,
deep basin in response to the Salarian orogeny (Ruzhentsev & boulder-sized clasts of granite, quartzite, volcanic and meta-
Burashnikov 1996). morphic rocks, and gravel to boulder-sized clasts of dolomite.
Exposures of Cambrian carbonates on the west side of Lake Quartzite and volcanic clasts are commonly faceted and striated
Khubsugul are folded in tight, south-plunging synclines and are (Osokin & Tyzhinov 1998).
cut by Palaeozoic granites (Fig. 30.3), perhaps also related to an
extension of the early Cambrian Salarian orogeny (Ruzhentsev &
Burashnikov 1996) accompanying the collision between the The Khesen diamictite
Agradag arc and the Khubsugul terrane. Palaeozoic intrusions are
particularly common in the south of the Khubsugul basin where The Khesen diamictite tends to be thinner than the Ongoluk dia-
they effectively obliterate the host stratigraphy (Badarch et al. 1998). mictite, and is composed of a massive, carbonate clast-dominated
Lake Khubsugul and the Darkhat depression formed as a diamictite. The thickness of the Khesen diamictite ranges from 10
southern arm of the Neogene Baikal rifting episode. Neoprotero- to 65 m, with no systematic geographical trend. The thickest
zoic –Cambrian stratigraphy on the west side of Lake Khubsugul measured section of the Khesen diamictite in the Lake Khubsugul
was uplifted and exposed along a rift shoulder. Rifts commonly area is exposed on the ridge north of Ongoluk Gol (Figs 30.3 &
followed Precambrian structures and were accompanied by volu- 30.5) and consists predominately of a massive, carbonate-clast dia-
minous basaltic volcanism (Logatchev 1984). mictite. Clasts are composed of angular to sub-rounded dolomite
and limestone (including giant ooid and stromatolite clasts) that
are commonly imbricated, with sizes ranging from pebble to
Stratigraphy boulders. The Khesen diamictite has a yellow-weathering dolomite
matrix, with the massive, unbedded deposits broken only by thin,
The Khubsugul basin sequence begins with rift-related volcanic lenticular, rhythmically bedded marls. At Khesen Gol, quartzite
and clastic rocks of the Darkhat Group that rest unconformably and volcanic pebbles are also present, and the massive diamictite
on Precambrian basement and meta-sediments (Ilyin 1973, 1990, is interrupted with multiple fine-laminated beds that are penetrated
2004). These volcanic rocks are overlain with hundreds of by outsized clasts.
metres of clastic and carbonate rocks that are variably preserved
under the sub-Khubsugul Group unconformity (Osokin & Tyzhi-
nov 1998). The Khubsugul Group begins with as much as 100 m Boundary relations with overlying and underlying
of argillite, unsorted sandstone and minor limestone of the basal non-glacial units
Ongoluk Fm. and grades upwards into the Ongoluk diamictite,
with clasts becoming larger and more abundant upwards. The Near Lake Khubsugul the Ongoluk diamictite has a gradational
Ongoluk diamictite is overlain by 100 –250 m of allodapic dolo- basal contact with clast-size and abundance decreasing gradually
mite of the upper Ongoluk Fm. The Khesen Fm. begins with a down-section until completely disappearing from the siltstone
second diamictite that is c. 50 m thick and carbonate clast- and sandstone (Osokin & Tyzhinov 1998). At Khesen Gol, the
dominated (Osokin & Tyzhinov 1998). The Khesen diamictite is underlying clastic units consist of nearly 100 m of millimetre-
capped with a 3– 5 m dolostone, the top of which hosts laminated argillite and siltstone, with 0.5 m interbeds of unsorted
centimetre-scale barite fans. Stratiform and granular phosphorite very coarse sandstone and gravel conglomerate, and at least two
rests above a major flooding surface, less than 50 m above the beds of texture-less limestone. These units rest unconformably
THE KHUBSUGUL GROUP 341

Fig. 30.2. Geology of the western shores of


Lake Khubsugul, showing the positions of
measured sections.

on dolostone of the Arasan Group. The Ongoluk diamictite is sep- the Darkhat region of Mongolia and along the Sarkhoi River
arated from overlying dolomite by a sharp contact, although (Osokin & Tyzhinov 1998).
regionally, deposition appears to be uninterrupted. The yellow-weathering Khesen diamictite rests disconformably
Along the Bokson River in Siberia, rocks that are equivalent to on a sharp contact with the underlying light blue dolomites. The
the Ongoluk diamictite rest unconformably on the Darkhat volca- upper contact, however, is commonly broken with a couple of
nics with a 15-m-thick weathering crust at the base (Kuzmichev metres of dolomite breccia between the massive diamictite and
2001). A similar, though thinner breccia has been documented in the thin dolomite overlying the Khesen diamictite. The dolomite
342 F. A. MACDONALD & D. S. JONES

Khesen Gol, above a disconformable flooding surface (Ilyin 2004) that cuts
down into the underlying units (Fig. 30.5).
M602, M613
c. 2km Chemostratigraphy
not shown
Erkhelnur Formation C Chemostratigraphic studies have not been reported for the carbon-
ate units bounding the two diamictites, although work is in pro-
40 0 Ongoluk Gol, M607, gress. Carbon-isotope values through the phosphatic interval of
M608, M609 the Khesen Fm. at Ongoluk Gol range from – 7% to þ5% (Ilyin
& Kiperman 2000; Ilyin 2004); however, as these data have not
P yet been reproduced, their utility for correlation is questionable.
Strontium-isotope values of c. 0.7080 have also been reported
400
P from carbonate interbedded with the Khubsuglul phosphorite
C deposit (Shields et al. 2000).
300
P
P Palaeolatitude and palaeogeography
ce
r fa

Cocks & Torsvik (2007) provide a review of the palaeomagnetic and


su

Palaeozoic fauna affinity studies of Siberia and the peri-Siberian


Khubsugul Group

al

300 terranes. However, there are very few reliable palaeomagnetic


iat
dh

data on the Khubsugul terrane, particularly in the Neoproterozoic,


200 P
re

so the palaeolatitudes presented are highly speculative.


Khesen Fm.

er

From palaeomagnetic studies on peri-Siberian terranes, Krav-


inf

chinsky et al. (2001) concluded that the Tuva-Mongolia belt was


P at low latitude, adjacent to Siberia throughout the Ediacaran and
Cambrian. However, this study lacked a robust confidence test
(i.e. only a reversal test with few samples and low resolution).
200 Along with other peri-Siberian terranes, it has been suggested
100 that the Khubsugul terrane occupied a Precambrian position
between Siberia and Laurentia (Gladkochub et al. 2006), and
Ongoluk Formation

rifted away from Siberia in the late Neoproterozoic (Sengor &


Natal’in 1996; Kuzmichev et al. 2001, 2005). Sengor & Natal’in
(1996) further posit that throughout the late Neoproterozoic and
early Palaeozoic, these terranes were attached to the Central Mon-
100 golian Block, which along with other terranes, stretched to the
present day Sea of Okhotsk. However, this reconstruction is incon-
0 (m ) sistent with the presence of Ordovician accretionary zones on the
NW margins of the Baidrag and Dzabkhan terranes and a Late
Cambrian Dhizda arc on the west margin of the Khubsugul
c. 450 m terrane (Badarch et al. 2002). The Khubsugul and Tavargatay ter-
not shown ranes host Cambrian trilobites endemic to Siberia (Astashkin et al.
C 1995) and Silurian brachiopods characteristic of the peri-Siberian
0 (m) realm (Hou & Boucot 1990). Thus, although there is a paucity of
Darkhat

reliable palaeomagnetic constraints on the Khubsugul terranes,


Arasan

c. 200 m
vv several lines of evidence indicate that they were adjacent to
c. 2 km not shown
vv Siberia in the Neoproterozoic and early Palaeozoic. Pisarevsky
vv not shown et al. (2000) present a strong c. 615 Ma pole on red beds along
vv the Lena River, pinning Siberia at equatorial latitudes in the Neo-
proterozoic. Further palaeomagnetic studies on the Khubsugul
grainstone rhythmite & terrane are necessary, and are in progress (J. Meert pers. comm.).
siltstone C chert
ribbonite diamictite P phosphorite
allodapic flooding barite fans Geochronological constraints
carbonate surface
Although the diamictites of the Khubsugul Group have not been
Fig. 30.3. Stratigraphy of the Khubsugul Group along the Khesen and Ongoluk directly dated, there are at least two radiometric constraints on
Gols. The locations of measured sections are shown in Figure 30.2.
the maximum age of the deposits. In Siberia, the Bokson Group
overlies the Shishkhid arc, which contains magmatic zircons
from rhyolites with a concordant U – Pb SHRIMP age of
is only c. 1 m thick at Ongoluk Gol with centimetre-scale bladed 800 + 2 Ma (Kuzmichev et al. 2005). The Bokson Group also
barite fans at the upper dolomite –limestone transition. Approxi- overlies the Sorkhoi Group, which contains volcanic rocks that
mately 20 km north, the overlying white dolomite reaches a have been dated with whole-rock Rb –Sr at 718 + 30 Ma (Buya-
maximum thickness of 6 m, and is interspersed with bed-parallel kaite et al. 1989). In Mongolia, the volcanic rocks of the Sorkhoi
cements (Fig. 30.5). A transgression continues above the dolomite Group are stratigraphically equivalent to the Darkhat and Dzab-
into as much as 100 m of dark grey limestone rhythmite. The khan volcanics (Macdonald 2011), which have been dated at
phosphorite-bearing strata is composed largely of dolomite with 850 + 2 and 750 + 3 Ma (Pb/Pb zircon, Burashnikov 1990),
common olistostromes and mass flow deposits, and comes in and more recently, at 777 + 6 Ma (U – Pb SHRIMP zircon, Zhao
THE KHUBSUGUL GROUP 343

Khesen Gol, Ongoluk Gol, Khirbisteg Gol N. Ongoluk Gol Khesen Gol N. Khesen Ridge Bakha Gol
M611 M607, M609 M602 M615, M616 M618
M613, M614 M605, M610 P P
450 P
P C
P P P
200 100 P
P
v P P
m
P P
400 P P
150 P
P
v

50
350 v
v

100

300
50
0 (m)

250 2.5 km 3.3 km 1.0 km 14.0 km


S N
0 (m) Fig. 30.5. Detailed stratigraphy of the Khesen diamictite and overlying
basal Khesen Fm. on the western shores of Lake Khubsugul. Positions of
dolomite facies measured sections are in Figure 30.2. For legend see Figures 30.3 and 30.4.
grainstone
200 allodapic et al. 2006), and 803.4 + 8.0 and 773.5 + 3.6 Ma (laser evapor-
rhythmite ation zircon, Levashova et al. 2010). The uppermost Khesen For-
mation is thought to be latest Ediacaran in age by correlation with
diamictite facies the Zabit Fm. in Siberia, which contains Cloudina and Renalsis
(Kheraskova & Samygin 1992). The Khesen Fm. is overlain by
massive the early Cambrian, Archaeocyathid-bearing Erkhelnur Fm.
stratified matrix (Ilyin & Zhuraveleva 1968), providing a robust minimum age on
150 the diamictites.
sandstone
siltstone
Discussion
argillite
clast size Although the Dzabkhan platform of southern Mongolia may have
100 been geographically separated from the Khubsugul basin, many
boulder units in the Tsagaan Oloom Fm. (Macdonald 2011) can be corre-
cobble lated with the Khubsugul Group. Both successions are underlain
pebble by riftogenic volcanic rocks and begin with interbedded clastic
rocks and diamictites. The Maikhan Ul diamictite is an early Cryo-
clast composition genian glacial deposit (Brasier et al. 1996), and like the basal
granite Ongoluk diamictite in the Khubsugul Group, it is commonly
50 clastic over 100 m thick and dominated by siltstone and coarse sand
(Lindsay et al. 1996). The upper Ongoluk Fm. can be correlated
carbonate with the Tayshir member of the Tsagaan Oloom Fm., and the
Ongoluk Formation carbonate-rich basal Khesen diamictite can be correlated with
the Khongoryn diamictite (Macdonald et al. 2009). Chemostrati-
Arasan Formation graphy indicates that the phosphorites on both the Khubsugul
0 (m) and Dzabkhan terranes were deposited in the latest Ediacaran to
early Cambrian above a major Ediacaran hiatus (Shields et al.
Fig. 30.4. Detailed stratigraphy of the Ongoluk diamictite along the Khesen 2000; Macdonald et al. 2009).
and Ongoluk Gols. The locations of measured sections are shown in Figure 30.2. Unconformities in the Ongoluk Fm. developed near basement
m, metamorphic clasts; v, volcanic clasts. highs on rift shoulders that were active at least until the onset of
344 F. A. MACDONALD & D. S. JONES

deposition of the Ongoluk diamictite (Osokin & Tyzhinov 1998). References


While the Khubsugul basin was undergoing extension a large
thickness of diamictite accumulated near the present Lake Astashkin, V. A., Pegel, T. V. et al. 1995. The Cambrian System of the
Khubsugul. Evidence for a glaciogenic origin of the Ongoluk Foldbelts of Russia and Mongolia. International Union of Geological
diamictite includes exotic clasts with a mixed lithology, bed- Sciences, London.
truncating lonestones, and faceted and striated clasts (Osokin & Badarch, G., Byamba, J. et al. 1998. Geological Map of Mongolia.
Tyzhinov 1998). The contact with the overlying light-blue alloda- Mineral Resources Authority of Mongolia, Ulaan Baator.
pic dolostone is sharp, and as they appear regionally conformable, Badarch, G., Cunningham, W. D. & Windley, B. 2002. A new terrane
a rift –drift transition is inferred within the Ongoluk diamictite. subdivision for Mongolia: implications for the Phanerozoic crustal
Although striated clasts have not been observed in the younger growth of Central Asia. Journal of Asian Earth Sciences, 21, 87 – 110.
Brasier, M. D., Shields, G., Kuleshov, V. N. & Zhegallo, E. A. 1996.
carbonate-rich Khesen diamictite, and exotic clasts are rare, a
Integrated chemo- and biostratigraphic calibration of early animal
glacial origin of the massive deposit is inferred from the presence evolution: Neoproterozoic –early Cambrian of southwest Mongolia.
of bed-penetrating lonestones exposed along Khesen Gol, the geo- Geological Magazine, 133, 445– 485.
chemistry signature in associated carbonates (unpublished data) Buchan, C., Pfander, J. et al. 2002. Timing of accretion and collisional
and the presence of overlying barite fans, which are present deformation in the Central Asian Orogenic Belt: implications of
above basal Ediacaran cap dolostones in Australia (Kennedy granite geochronology in the Bayankhongor Ophiolite Zone. Chemi-
1996), Mauritania (Deynoux & Trompette 1976), NW Canada cal Geology, 192, 23 –45.
(Hoffman & Schrag 2002) and south China (Jiang et al. 2003). Burashnikov, V. V. 1990. Tectonics of the Urgamal Zone, Early Calido-
A basal Ediacaran age of the carbonate immediately overlying nides of Western Mongolia. Russian Academy of Sciences, Moscow.
the Khesen diamictite is inferred from stratigraphic correlation Buyakaite, M. I., Kuzmichev, A. B. & Sokolov, D. D. 1989. 718 Ma
with the Dzabkhan platform (Macdonald et al. 2009), and from Rb –Sr errorchron of the Sorkhoi Group in the East Sayan. Doklady
the occurrence of the barite fans. Akademii Nauk SSSR, 309, 150–154.
Kheraskova & Samygin (1992) rejected a glaciomarine origin Chumakov, N. M. 2011. Glacial deposits of Bokson Group, East Sayan
of diamictites in the Zabit Formation exposed on the Siberian Mountains, Buryatian Republic, Russian Federation. In: Arnaud,
side of the border, arguing that these deposits represent rift-related E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
submarine slumps and debris flows. They further suggested that the Record of Neoproterozoic Glaciations. Geological Society, London,
diamictites in the Zabit Fm. (and correlative Khesen Fm.) are latest Memoirs, 36, 285–288.
Ediacaran to early Cambrian in age, citing the presence of Cocks, L. & Torsvik, T. H. 2007. Siberia, the wandering northern terrane,
Cloudina and Renalsis in the Zabit Fm. On the Mongolian side and its changing geography through the Palaeozoic. Earth Science
of the border, there is no evidence of a glacial origin for the diamic- Reviews, 82, 29 –74.
Deynoux, M. & Trompette, R. 1976. Late Precambrian mixtite: glacial
tite at the base of the Khesen Fm., or the carbonate and carbonate
and/or non-glacial? Dealing especially with the mixtite of West
clasts conglomerate associated with the phosphorite in the upper- Africa. American Journal of Science, 276, 117–125.
most Khesen Fm., which were interpreted as olistostromes and Dobretsov, N. L. 1985. Overthrust tectonics of the East Sayans. Geotec-
slumps. It is possible that, like the Dzabkhan platform to the south tonics, 19, 26– 34.
(Macdonald et al. 2009), there is a major hiatus in the Ediacaran and Donov, N. A., Edemsky, H. B. & Ilyin, A. V. 1967. Cambrian phos-
that Kheraskova & Samygin (1992) are miscorrelating mass flows phorites of Mongolia Popular Republic. Sovetskaya Geologia, 3,
in the latest Ediacaran to early Cambrian upper Khesen Fm. with 55 – 60.
the end Cryogenian Khesen diamictite described here from the Gladkochub, D. P., Wingate, M. T. D., Pisarevsky, S. A., Donskaya,
base of the Khesen Fm. Dobretsov (1985) considered the Siberian T. V., Mazukabzov, A. M., Ponomarchuk, V. A. & Stanevich, A.
diamictites to be ‘nappe thrust olistostromes’ related to the Salarian M. 2006. Mafic intrusions in southwestern Siberia and implications
orogeny, which has been stratigraphically constrained to the early for a Neoproterozoic connection with Laurentia. Precambrian
Cambrian on the Dzabkhan terrane (Ruzhentsev & Burashnikov Research, 147, 260– 278.
1996). Again, it seems likely that there is a conflation between Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis;
the Khesen diamictite in the lower Khesen Fm., which is interpreted testing the limits of global change. Terra Nova, 14, 129–155.
here as glaciogenic, and the conglomerates interpreted as olistos- Hou, H. F. & Boucot, A. J. 1990. The Balkhash-Mongolia-Okhotsk
tromes and allodapic deposits in the upper Khesen Fm. Region of the Old World Realm. Geological Society, London,
Sheldon (1984), Osokin & Tyzhinov (1998) and Ilyin (2004) Memoir, 12, 297– 303.
have suggested a genetic relationship between the diamictite and Ilyin, A. V. 1973. Khubsugul Phosphorite-Bearing Basin. Geologiches-
phosphorite. However, Ilyin (2004) documented a disconformity kiy Institut, Akademiya Nauk SSSR, Moscow.
Ilyin, A. V. 1990. Proterozoic supercontinent, its latest Precambrian
at the base of the phosphorite (Fig. 30.5). Chemo- and lithostrati-
rifting, breakup, dispersal into smaller continents, and subsidence
graphic studies in the correlative Dzabkhan platform (Macdonald
of their margins: Evidence from Asia. Geology, 18, 1231– 1234.
et al. 2009; Macdonald 2011) suggest a major hiatus and flooding Ilyin, A. V. 1998. Rare-earth geochemistry of ‘old’ phosphorites and
surface between the Khesen diamictite and the phosphorite series, probability of syngenetic precipitation and accumulation of phos-
casting doubt on any genetic relationship. phate. Chemical Geology, 144, 243–256.
Further chemo- and lithostratigraphic studies are needed to Ilyin, A. V. 2004. The Khubsugul phosphate-bearing basin: new data and
better constrain the basin dynamics and depositional setting of concepts. Lithology and Mineral Resources, 39, 454–467.
Neoproterozoic strata in Mongolia. The palaeogeography of the Ilyin, A. V. & Byamba, J. 1980. Handbook for the Excursion ‘Phosphor-
peri-Siberian terranes also remains speculative. It is clear, ites of the Khubsugul Basin in the Mongolian Peoples’s Republic’.
however, that island arcs surrounded the Khubsugul terrane for Geologicheskiy Institut, Akademiya Nauk SSSR, Moscow.
much of the Neoproterozoic and early Cambrian, and thus, there Ilyin, A. V. & Kiperman, Y. A. 2000. Mass accumulation of biogenic
is excellent potential for geochronology in the Khubsugul basin. rocks at the Vendian/Cambrian boundary and carbon isotopic
anomalies. Soveremenny voprosy geologii (Modern Problems of
We thank our field assistants U. Bold, J. Otgonhuu and Eerie. We also thank Geology). Nauchnyi Mir, Moscow.
A. Bayasgalan and the Mongolian University of Science and Technology for Ilyin, A. V. & Zhuraveleva, I. T. 1968. On the boundary between the
making our fieldwork possible, D. Schrag for use of the Harvard University Cambrian and the Precambrian at Prikhusugulie (Mongolian PR).
Laboratory for Geochemical Oceanography, and G. Eischeid for help in the lab- Doklady Akademii Nauk SSSR, 182, 1164–1166.
oratory. We thank the National Science Foundation for funding, and P. Hoffman Jiang, G., Kennedy, M. J. & Christie-Blick, N. 2003. Stable isotopic
for inspiration and support. This represents a contribution of the IUGS- and evidence for methane seeps in Neoproterozoic postglacial cap car-
UNESCO-funded IGCP (International Geoscience Programme) project #512. bonates. Nature, 426, 822– 826.
THE KHUBSUGUL GROUP 345

Kennedy, M. J. 1996. Stratigraphy, sedimentology, and isotope geochem- (eds) The Geological Record of Neoproterozoic Glaciations.
istry of Australian Neoproterozoic postglacial cap dolostones: Geological Society, London, Memoirs, 36, 331–337.
deglaciation, d13C excursions, and carbonate precipitation. Journal Macdonald, F. A., Jones, D. S. & Schrag, D. P. 2009. Stratigraphic and
of Sedimentary Research, 66, 1050– 1064. tectonic implications of a new glacial diamictite-cap carbonate
Kherzaskova, T. N. & Samygin, S. G. 1992. Tectonic conditions in the couplet in southwestern Mongolia. Geology, 37, 123–126.
East Sayan Vendian –Middle Cambrian terrigenous carbonate associ- Osokin, P. V. & Tyzhinov, A. V. 1998. Precambrian Tilloids of the Oka-
ation. Geotectonics, 26, 445– 458. Khubsugul phosphorite-bearing basin (Eastern Sayan, Northwestern
Kravchinsky, V. A., Konstantinov, K. M. & Cogne, J.-P. 2001. Mongolia). Lithology and Mineral Resources, 33, 142– 154.
Palaeomagnetic study of Vendian and Early Cambrian of South Pisarevsky, S. A., Komissarova, R. A. & Khramov, A. N. 2000. New
Siberia and Central Mongolia: was the Siberian platform assembled palaeomagnetic result from Vendian red sediments in Cisbaikalia
at this time? Precambrian Research, 110, 61 – 92. and the problem of the relationsip of Siberia and Laurentia in the
Kuzmichev, A. 2001. Early Baikalian tectonic events in the Tuva- Vendian. Geophysics Journal International, 140, 598–610.
Mongolia Massif: arc-microcontinent collision. Geotectonics, 35, Ruzhentsev, S. V. & Burashnikov, V. V. 1996. Tectonics of the
185– 198. western Mongolian Salairides. Geotectonics, 29, 379–394.
Kuzmichev, A., Bibikova, E. V. & Zhuravlev, D. Z. 2001. Neoproter- Sengor, A. C. & Natal’in, B. A. 1996. Palaeotectonics of Asia: frag-
ozoic (800 Ma) orogeny in the Tuva-Mongolia Massif (Siberia): ments of synthesis. In: Yin, A. & Harrison, M. (eds) The Tectonic
island arc-continent collision at the northeast Rodinia margin. Evolution of Asia. Cambridge University Press, Cambridge,
Precambrian Research, 110, 109– 126. 486– 640.
Kuzmichev, A., Kroener, A., Hegner, E., Dunyi, L. & &Yusheng, W. Sheldon, R. P. 1984. Ice-ring origin of the Earth’s atmosphere and hydro-
2005. The Shishkhid ophiolite, nothern Mongolia: a key to the sphere and late Proterozoic – Cambrian phosphogenesis, Phosphorite,
reconstruction of a Neoproterozoic island-arc system in central Geological Survey of India Special Publication, 17, Udaipur,
Asia. Precambrian Research, 138, 125–150. Rajasthan, India.
Levashova, N. M., Kalugin, V. M., Gibsher, A. S., Yff, J., Ryabinin, Shields, G., Stille, P. & Brasier, M. 2000. Isotopic records across two
A. B., Meert, J. & Malone, S. J. 2010. The origin of the Baydaric phosphorite giant episodes compared: the Precambrian –Cambrian
microcontinent, Mongolia: constraints from palaeomagnetism and and the Late Cretaceous – recent. SEPM Special Publications, 66,
geochronology. Tectonophysics, 485, 306–320. 102–115.
Lindsay, J. F., Brasier, M., Shields, G., Khomentovsky, V. V. & Windley, B. F., Alexeiev, D., Xiao, W., Kroener, A. & Badarch, G.
Bat-Ireedui, Y. A. 1996. Glacial facies associations in a Neoproter- 2007. Tectonic models for accretion of the Central Asian Orogenic
ozoic back-arc setting, Zavkhan Basin, western Mongolia. Geologi- Belt. Journal of the Geological Society of London, 164, 31 –47.
cal Magazine, 133, 391–402. Zhao, Y., Song, B. & Zhang, S. H. 2006. The Central Mongolian micro-
Logatchev, N. A. 1984. The Baikal rift system. Episodes, 7, 38 –42. continent: its Yangtze affinity and tectonic implications. In: Jahn,
Macdonald, F. A. 2011. The Tsagaan Oloom Formation, southwestern B. M. & Chung, L. (eds) Symposium on Continental Growth and
Mongolia. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. Orogeny in Asia. Taipei, Taiwan, 135– 136.
Chapter 31

The Blaini Formation of the Lesser Himalaya, NW India

JAMES L. ETIENNE1,2*, PHILIP A. ALLEN3, ERWAN LE GUERROUÉ4, LARRY HEAMAN5,


SUMIT K. GHOSH6 & RAFIQUE ISLAM6
1
Present address: Neftex Petroleum Consultants Ltd, 97 Milton Park, Abingdon, Oxfordshire OX14 4RY, UK
2
Geologisches Institut, Departement Erdwissenschaften, ETH-Zentrum, Haldenbachstrasse 44, CH-8092,
Zürich, Switzerland
3
Department of Earth Sciences and Engineering, Imperial College London, South Kensington Campus, London SW7 2AZ, UK
4
Géosciences Rennes (UMR 6118 – CNRS), 263 Avenue du General Leclerc. CS 74205, Université de Rennes 1,
35042 Rennes Cedex, France
5
Department of Earth & Atmospheric Sciences, 1 – 26 Earth Sciences Building, University of Alberta, Edmonton, Alberta,
T6G 2E3, Canada
6
Wadia Institute of Himalayan Geology, 33 General Mahadeo Singh Road, Dehra Dun, Uttarakhand, India
*Corresponding author (e-mail: james.etienne@neftex.com)

Abstract: Neoproterozoic glaciogenic deposits crop out widely across the Lesser Himalaya fold and thrust belt in NW India. Underlain
by the siliciclastic Simla and Jaunsar Groups, the Blaini Formation (Fm.) includes at least two thick and regionally extensive diamictite
units, separated by siliciclastics and argillites and capped by a pink microcrystalline dolomite. A glaciogenic origin is supported by the
presence of relatively abundant striated clasts and the local preservation of polished and striated pavement on underlying Simla Group
clastics. The cap dolostone is isotopically light with respect to both 13C and 18O, which show strong covariance. The Blaini is ubiqui-
tously deformed, incorporated in regional-scale folds and thrusts, and exhibits locally intensive intra-formational deformation. Until
recently, geochronological constraints have remained poor, but new detrital zircon ages from diamictite samples provide a maximum
age limit of 692 + 18 Ma (207Pb/206Pb). Reliable palaeomagnetic data are required to constrain the position of this important passive
continental margin in palaeogeographical reconstructions.

The Lesser Himalaya fold and thrust belt in NW India comprises (1975), Bhatia & Prasad (1975), Jain & Varadaraj (1978), Bhatia
up to 10 km of variably deformed low-grade Neoproterozoic to & Prasad (1981) and Jain (1981).
Lower Palaeozoic metasediments (Schelling 1992). Originally
described as the ‘Blaini Conglomerate’ from the type section
at Baliana Nala in Himachal Pradesh (Medlicott 1864), the Structural framework
Blaini Fm. is an important stratigraphic marker in the Lesser
Himalaya and crops out across .300 km of the fold belt The Blaini Fm. crops out around the margins of a series of tectoni-
(Fig. 31.1). Some of the best exposures occur in the Mussoorie cally complex synclines and in thrust block slices between the
syncline, between Māldeota and Dhanaulti, in the Giri River Main Boundary Fault and Main Central Thrust. Resting upon a
valley near Dadahu and around Simla. A panoply of designa- succession of continental deltaic to shallow marine clastics
tions occur in the literature, including Blaini Conglomerate, (Nagthat Fm.) and more distal facies of the Simla Group (Ghosh
Blaini Group, Blaini Boulder Bed, Blaini Series, Blaini diamic- 1991), the Blaini Fm. acts as a key stratigraphic marker horizon
tites, Blaini North, Blaini South and Blaini Fm., but the latter in this part of the Himalayas.
is preferred here given the unit’s extensive mappable nature The Blaini Fm. is overlain by the Infra Krol Fm. and the Krol
(Pilgrim & West 1928; Fuchs & Sinha 1978) and use in and Tal groups, a succession that is thought to represent the
recent publications (e.g. Jiang et al. 2002, 2003a; Kaufman inner part of a north-facing passive continental margin (Brookfield
et al. 2006). 1993). Based upon palaeogeographical reconstructions (e.g.
A glaciogenic origin for diamictites of the Blaini Fm. has long Torsvik 2003) and isotopic and sequence stratigraphic similarities,
been recognized (e.g. Oldham 1887; Holland 1908; Pilgrim & this succession may have opposed the Chinese south-facing
West 1928; Auden 1946; Saxena & Pande 1969; Gaur & Dave passive margin represented by the sedimentary cover of the
1971; Bhargava & Bhattacharyya 1975; Bhatia & Prasad 1975; South China block (Jiang et al. 2003a; Kaufman et al. 2006;
Jain & Varadaraj 1978; Bhatia & Prasad 1981; Jain 1981); McFadden et al. 2008). Palaeocurrent indicators and facies
however, in the absence of good biostratigraphic data, an relationships in the underlying Jaunsar Group (Ghosh 1991)
assumed correlation with the Permo-Carboniferous Talchir tillites suggest a sediment source to the south/SE, and a basin margin
of peninsular India has hampered recognition of the true age and oriented broadly NE –SW. Sequence stratigraphic analysis of the
significance of these deposits, despite a Proterozoic age frequently Krol Group illustrates a similar trend for the Ediacaran carbonate
being posited (e.g. see Bhatia & Kanwar 1975 and references platform (Jiang et al. 2002; Kaufman et al. 2006).
therein; Brookfield 1994). Recently published biostratigraphic The Blaini Fm. is ubiquitously deformed, and faulted and folded
data and detailed work on the overlying Krol and Tal Groups on a regional scale, exhibiting penetrative cleavage and evidence
have firmly established a Proterozoic age for the Blaini Formation for pressure solution. For these reasons, gravel clast macrofabrics
(Fig. 31.2). General overviews, including textural descriptions of may locally reflect the tectonic grain rather than primary deposi-
diamictite lithofacies, are included in Bhargava & Bhattacharyya tional fabrics. In contrast, clast surface features such as glacial

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 347– 355. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.31
348 J. L. ETIENNE ET AL.

SOLAN 4 5 7 8
Siwalik Zone
3
9
1 Subathu-Dagshai
2 CHOR
6 Tal Group
10

11 Carbonate formations (Krol, Shali, Deoban etc)


12 13 15

14 Black slate formations (Infra Krol, Blaini, Shali, Sor)


MUSSOORIE
17 18 Fossiliferous Upper Palaeozoics of the
16
19 Landsdowne syncline

DEHRA DUN Blaini/Mandhali


20
21
Jaunsar Group
22 (Nagthat/Chandpur undivided)
23
24
Chail
RISHIKESH
KAUDIYALA
Basic metavolcanic rocks

Chail/Nagthat undivided
500 km
SRINAGAR
M.
B.T Simla slates
. STUDY
SHILLONG
AREA
DELHI
30 N Crystalline
basement

INDIA
CALCUTTA 20 N
MUMBAI

MADRAS NAINI TAL


0 10 20 30 40
10 N km 25
70 E 80 E 90 E

Fig. 31.1. Geological map of the Lesser Himalaya fold-and-thrust belt in NW India. Redrawn from Fuchs & Sinha (1978). Numbers indicate location of measured
sections (Table 31.1; Fig. 31.3).

striae and polish have been preserved and are distinct from tectonic facets and polish. The Blaini is capped by a pink microcrystalline
striae such as slickenlines. Brookfield (1987) noted the occurrence dolomite and is overlain by the Infra-Krol Fm. In the upper part,
of shear structures within diamictite lithofacies, which is evident the cap contains thin grey and red shale partings.
from pressure shadows and augen adjacent to rotated clasts.
Although a tectonic origin for these features needs to be con-
sidered, it is possible that the shearing reflects soft-sediment defor- Glaciogenic deposits and associated strata
mation generated during sediment transport either in the subglacial
environment, where large shear stresses can be generated beneath Massive and laminated diamictite lithofacies
overriding ice (Boulton 1996), or during downslope resedimenta-
tion of debris, which results in similar microfabrics. Diamictites occur as massive or weakly stratified clast-poor to
clast-rich units with poorly sorted silty to sandy matrices. Thick-
nesses vary across the fold belt, but are typically on the order of
Stratigraphy 10–40 m (Fig. 31.3). Although exposure is poor, sheet-like geo-
metries best explain the regional distribution of diamictite
The Blaini Formation is generally stratigraphically simple, broadly beneath the cap dolostone. Lateral continuity between stratigraphi-
characterized by a tripartite diamictite –shale –diamictite succes- cally lower occurrences of diamictite is more difficult to demon-
sion that is variable in thickness across 300 km of the fold belt strate, given the few available sections where the base of the
between Simla and Nainital. Additional diamictite intervals formation may be observed. In the eastern part of the Mussoorie
locally occur in the NW of the region and have also been reported syncline, the coarse sediment fraction of the basal diamictite
from the area around Nainital (Jiang et al. 2003a). Diamictite units fines northwards, and on the northwestern limb, the upper
are generally less than 50 m thick (Fig. 31.3) and contain relatively surface of the uppermost diamictite is characterized by a win-
abundant glacioclastic debris bearing cross-cutting glacial striae, nowed transgressive lag immediately beneath the cap dolostone.
THE BLAINI FORMATION 349

Fig. 31.2. Lithostratigraphic and biostratigraphic subdivision of the Neoproterozoic– Cambrian succession of the Lesser Himalaya, summarized from Jiang et al. (2002),
Hughes et al. (2005) and this study. Carbon–isotope data from Kaufman et al. (2006) and this study. For details on biostratigraphy, refer to Prasad et al. (1990), Tiwari
(1999), Mazumdar & Banerjee (1998), Brasier & Singh (1987), Bhatt & Mathur (1990), Kumar et al. (1983), Mathur & Srivastava (1994), Kumar et al. (1987), Joshi et al.
(1989), Jell & Hughes (1997), Mathur & Joshi (1989a, b), Tripathi et al. (1984, 1986), Singh & Rai (1983), Banerjee & Narain (1976), Bhargava et al. (1998), Mathur
et al. (1988), Rai (1987), De et al. (1994), Bhargava (1984), Joshi & Mathur (1987), Tiwari & Knoll (1994), Shanker et al. (1997), Mathur & Shanker (1989, 1990), Bhatt
(1991) and Hughes et al. (2005).

In the Solan region, at Rahed (section 3, Table 31.1, Fig. 31.3), the (1975). Some petrographic and heavy mineral data may also be
lower diamictite is erosionally based, and overlies a striated and found in Jain & Varadaraj (1978) and Bhatia & Prasad (1981).
polished surface of Simla Group argillites.
Within this lithofacies, clast types include subrounded to suban-
gular quartzite, microscale laminated siltstone, chert, vein quartz, Interpretation
gneiss, limestone, sandstone, dolomite, basalt, shale and slate.
Striated and polished clasts are relatively abundant at most localities. Numerous interpretations have been posited for the genesis of
Statistical clast shape and roundness data have previously been pre- the Blaini diamictites, emphasized by discussions in Bhatia &
sented in Bhargava & Bhattacharyya (1975) and Bhatia & Prasad Kanwar (1975, and references therein) which revolve primarily
350 J. L. ETIENNE ET AL.

Table 31.1. Locations of measured sections (Figs 31.1 and 31.3) these data support an extra-basinal provenance for some of the
material in the diamictites.
Section no. Locality Northing Easting Where diamictites directly overlie striated pavement, a primary
(Figs 1, 3) tillite interpretation is considered most likely; however, across
much of the fold belt, this association is not observed. If the intra-
1 Baliana Nala 30859.6280 77802.4660 formational shear structures observed within the diamictites are
2 Deothal from 30850.4900 7789.8540 syn-sedimentary in origin, they may reflect either subglacial pro-
to 30850.5450 7789.9200 cesses of sediment transport or sediment redistribution by debris
3 Rahed 30854.8410 77812.3260 flows, and are therefore not diagnostic palaeoenvironmental indi-
4 Giri River Valley from 308540 01.5000 778130 52.7500 cators. There is little evidence for significant glaciotectonic struc-
to 308550 00.5900 778140 49.8600 turation, stratigraphic complexity, lithofacies heterogeneity or
5 Sargaon (i) 30856.1090 77816.2460
ice-proximal processes typical of terrestrial proglacial environ-
6 Deoria from 308500 24.9300 778110 03.9700
ments or glaciomarine grounding-line fan assemblages (cf. Benn
to 308500 23.9200 778110 11.6200
& Evans 1998; Powell & Cooper 2002) and a more distal setting
7 Sargaon (ii) 30856.5550 77816.7940
8 Rajgarh-Giripool Road 30853.2230 77814.2290
from the ice margin may be more likely. Coarse-tail grading
Section locally exhibited in the basal diamictite in the Mussoorie syncline
9 Pervi River Valley 30853.6530 77814.3970 may indicate downslope reworking of glaciogenic debris distal
10 Mareog 308520 13.0800 778120 53.0000 from the ice margin, but the locally reworked tops of younger dia-
11 Dadahu-Sangar Road 30839.3110 77826.1350 mictites beneath the cap carbonate probably indicates deposition
Section within neritic water depths. The diamictite lithofacies thus encom-
12 Sangar 30841.9150 77825.2630 passes a range of different glacially influenced facies types, includ-
13 Dadahu 30837.4760 77827.8000 ing primary tillites and glaciogenic debrites, which form part of a
14 Dubrah-Mahipur 30820.6720 78808.2960 continuum of deposits interpreted by Brookfield (1987) as basal
15 Narendranagar 30809.2070 78817.9580 tills, flow tills and subaqueous meltout tills.
16 Shivpuri Bridge 308080 11.2700 788230 16.3000
17 Bhanswari 30828.6790 78810.8030
18 Dhanaulti 30825.8160 78814.2280 Sandstones, siltstones and shales
19 Dhanaulti-Raipur Road 30823.9590 78817.9550
Section Between the lower and upper diamictite units, a series of associated
20 Renuka-Sataun Road 30835.3750 77829.2690 lithofacies occur that include massive amalgamated sandstone
Section beds, flaggy siltstones and laminated silty shales. Sandstones
21 Maldeota 30820.4420 78808.1840 overlie the diamictites at the base of the Blaini Fm. and either
22 Shilla (Sataun-Shilai 30837.3040 77842.0350 pass directly upwards into shales or through stacked flaggy silt-
Road) stones into shales. Dark grey flaggy siltstones with sharp bases
23 Bhatoli (Kempte Falls– 30830.5980 78800.0360 and tops are best developed in the Dhanaulti-Raipur (section 19,
Yamuna Bridge) Table 31.1, Fig. 31.3) and Sataun-Renuka road sections. These
24 Gular Ghatti 30807.4010 78825.4380 thinly bedded (centimetre-scale) units bear no depositional
25 Nainital– Bhowali 29823.0890 79829.6610 structures, but in the Dhanaulti sections show evidence for
Section near Kaliakhan
thickening-upward, thinning-upward cycles throughout the section.
The shales achieve thicknesses in the order of at least several
tens of metres, but given an absence of key marker beds (sand-
stones rarely re-appear in section until just beneath the upper dia-
around glacial v. non-glacial processes such as turbidity currents mictite unit), and the local extent of deformation, stratigraphic
or mass failure events (see also discussions in Rupke 1968; thicknesses may locally exceed this. In the Solan region shales
Valdiya 1973). However, an enhanced understanding of modern are of limited stratigraphic thickness and sandstone beds have
glaciomarine depositional systems now recognizes the impor- thin (10 mm thick) pebbly lag bases.
tant role played by debris flows in resedimenting glaciogenic
debris in these environments and the fact that the two are not Interpretation
mutually exclusive (cf. Dowdeswell et al. 1996, 1998; Taylor
et al. 2002). The lower portion of the Blaini Fm. is characterized by a fining-
Diamictites of the Blaini Fm. are interpreted as glaciogenic in upward motif recorded by a transition from diamictite through
origin on the basis that (i) the lithofacies contains abundant glacio- sandstones and siltstones into shales. This retrogradational stacking
clastic debris (striated, faceted and polished clasts) and (ii) the pattern indicates a transition to more distal conditions of sedimen-
basal diamictite overlies striated and polished pavement (near tation, with massive sandstones and siltstones probably deposited
Rahed, section 3, Table 31.1, Fig. 31.3). below the storm-weather wave base. More proximal conditions of
Several publications have suggested that the bulk of macroclasts sedimentation are indicated up-section as the lithofacies stacking
within the Blaini diamictites are attributable to local sources pattern turns around and shales pass up into sandstones and diamic-
derived from the underlying passive margin sequence (e.g. tites. Coarser clastics and only thinly developed shales indicate
Auden 1934; Rupke 1968; Niyogi & Bhattacharya 1971; Valdiya more proximal conditions in the NW around Solan compared
1973; Jain & Varadaraj 1978), including Brookfield (1987), who with thicker developed shale successions in the SE (e.g. in the
noted a direct compositional influence on the basal diamictite Mussoorie syncline). Additional support for more proximal con-
depending on which units of the Simla Group it overlies. Petro- ditions in the NW is provided by the preservation of striated pave-
graphic and heavy mineral analyses presented in Jain and Vara- ment at Rahed indicative of grounded ice conditions.
daraj (1978) are consistent with these interpretations; however,
the observation of orthoclase, microcline, perthite, myrmekite,
zircon, tourmaline and spinel by those authors are indicative of Dolomite
provenance from acid igneous rocks, although the degree of
reworking needs consideration. Because crystalline basement Dolomite locally cements siltstones and parts of the uppermost dia-
does not crop out in this part of the fold belt, it is possible that mictite unit in the Giri Valley to the east of Solan, and has also been
THE BLAINI FORMATION 351

Fig. 31.3. Lithostratigraphic sections and d13C chemostratigraphy of the Blaini Formation in the Lesser Himalaya of NW India. Numbered sections are located in
Figure 31.1 and Table 31.1.
352 J. L. ETIENNE ET AL.

observed as discrete interbeds within equivalent units in the evidence for subglacial abrasion and the base of the formation in
Mussoorie syncline (Kaufman et al. 2006), but is most extensively this instance is clearly erosional.
developed as a regional thinly bedded, laminated pink microcrys- Recent reports of biostratigraphically significant protoconodant
talline dolomite (Fig. 31.3). Across the fold belt, the dolostone assemblages from the Gangolihat Dolomite have led Azmi & Paul
rarely exceeds 9 m in thickness. Parallel or ‘crinkly’ lamination (2004) to adopt a modified version of Valdiya’s (1995) lithostrati-
dominate, with the exception of a small patch of brecciated carbon- graphic scheme for the region, suggesting a correlation between
ate observed at Dhanaulti in the Mussoorie syncline. Like other cap the Deoban and Krol Group carbonates, and invoking a major
carbonates, the Blaini dolostones contain some enigmatic ‘tepee- regional unconformity at the base of the Blaini. This is controver-
like’ structures. In the Nigalidhar and Mussoorie synclines, these sial given the widely held view that the Deoban carbonates lie
are similar in form to ripple bedforms, with a wavelength of stratigraphically beneath the Jaunsar Group, so further data are
c. 1 m and amplitudes of the order of 0.2 m. Where observed, required to test these competing hypotheses.
these structures occur 4–4.5 m above the base of the cap and are
underlain and overlain by planar laminated dolomite. Laminae
are locally pinched out at ripple crests, but have not been observed Chemostratigraphy
to onlap against the ripple trunks. Cracks run through and penetrate
the ripple crests in the axial plane. Some millimetre-scale displace- Abundant d13C and d18O stable isotope data are available for the
ment occurs along these fractures, but no significant brecciation of Blaini cap dolostone (Kaufman et al. 2006; this study) and some
the carbonates occurs in the axial zones. strontium and magnesium values have been published (Kaufman
et al. 2006). Kaufman et al (2006) identified a negative C-isotopic
signature with up-section trends towards more negative values,
Interpretation which recover to near 0‰ in the Garhwal syncline.
We collected c. 350 samples from 10 localities across the fold belt
Most exposures of the Blaini cap carbonate are dominated by to better characterize the nature of the isotopic character and facili-
planar or crinkly laminated dolomite. The latter may indicate an tate correlation between sections (Fig. 31.3). As with most cap
algal origin, but no distinct structured mats, stromatolitic or carbonates, the Blaini dolostones are depleted in both carbon and
other biohermal forms have been observed that could provide oxygen, which show strong co-variation. In this sample set,
more definitive evidence of shallow-water sedimentation. With d13Ccarb values range from –0.44 to –6.12, and d18O values vary
the exception of very locally developed breccia observed at between –5.49 and –13.27. In Figure 31.3 we present a first
Dhanaulti, a general lack of intraclasts or sedimentary breccias attempt at a composite d13Ccarb curve throughout the cap carbonate.
suggests that deposition occurred largely below the storm-weather The curve was generated by matching absolute values and trends in
wave base. The enigmatic tepee structures, which resemble ripple values between 10 sections in the Solan region, Nigalidhar and
bedforms, are identical to examples observed elsewhere from cap Mussoorie synclines. No correction has been attempted to account
carbonates (e.g. Allen & Hoffman 2005; Alvarenga et al. 2007); for stratigraphic condensation or expansion, which may have
however, their interpretation is far from straightforward. No resulted from variation in sedimentation rates or differential com-
clear examples of laminae onlapping the ripple forms have been paction. Our data confirm the trend observed by Kaufman et al.
observed that require a sedimentary origin, and a tectonic origin (2006), with a pronounced negative excursion to –5‰ d13Ccarb fol-
has to be considered. Stromatactis-like structures associated with lowed by a recovery towards less negative values (Fig. 31.3).
similar features in the Doushantuo Fm. in South China do not Kaufman et al (2006) also report negative values for carbonate
appear to occur (which might imply linkage to methane gas or interbedded with the upper diamictites in the Mussoorie syncline.
fluid escape; cf. Jiang et al. 2006). Tepees described elsewhere
in the literature have been variably interpreted as resulting from
different syndepositional to early diagenetic processes of sedimen- Significance of isotopic analyses
tation, cementation and deformation, a discussion of which may
be found in Jiang et al. (2006). Neoproterozoic cap carbonates are renowned for being isotopi-
cally light with respect to 13C, and the use of their negative
isotope excursions as chemostratigraphic markers has become a
standard tool in both the correlation of glacially influenced strata
Boundary relations with overlying and underlying and the construction of a composite Precambrian C-isotope
non-glacial units curve (Halverson 2005).
Whether the values observed here reflect secular changes in the
Across the fold belt, the Blaini Fm. is significantly deformed and primary isotopic composition of seawater is difficult to establish. A
heavily vegetated at altitudes up to c. 2.5 km above sea level. lack of pelagic carbonate-secreting organisms makes establishing
Rheological contrasts between the different lithofacies of the for- a proxy for primary seawater composition challenging, particularly
mation and its bounding strata have often led to deformation along if the original grains have been destroyed by recrystallization. The
contact surfaces, which presents some challenges in characterizing strong covariation of d13C and d18O raises the probability of diage-
boundaries. Nevertheless, there are a number of key exposures netic alteration associated with fluids flushing through the cap
across the fold belt where stratigraphic relationships may be carbonate. Some support for fluid migration is evident given the
properly established. ubiquitous occurrence of sparry calcite-filled veins in the cap.
For example, sections at Māldeota show the diamictites to dis- If the up-section trends in values reflect variations in the primary
conformably overlie Nagthat Fm. quartzites by a slickenlined composition of seawater, then absolute values are likely to differ
faulted contact. Further north, along the road between Māldeota between localities given the transgressive nature of the carbonates
and Dhanaulti, the contact between the basal diamictite and under- deposited during post-glacial eustatic recovery. Given the poor
lying quartzites remains sharp, but with no evidence for tectonic continuity between exposures, an absence of distinct markers or
dislocation. Striated pavement has previously been reported at fauna to facilitate correlation, sections may only be compared by
this contact, but present exposures do not appear to retain glacial trends in values (Fig. 31.3). The composite curve that results is
striae. However, in the northwestern part of the fold belt, where interesting in that it broadly honours our understanding of the
the diamictites overlie softer mudrocks of the Simla Group, basin geometry, with progressive onlap (and therefore younging)
diamictites directly overlie polished striated pavement (e.g. near from NW to SE, normal to the basin margin (see earlier discussion
Rahed). Multiple sets of cross-cutting glacial striae provide and Kaufman et al. 2006, fig. 2), and little variability between more
THE BLAINI FORMATION 353

Fig. 31.4. Probability–age plot for detrital


zircons extracted from massive diamictite
sample JE-1-002 (Māldeota, Mussoorie
syncline).

closely spaced sections that occur along strike from one another. Krol Belt, the P]/] boundary is located either in the uppermost
Some perturbations to this trend may indicate palaeobathymetric part of the Krol Group (Krol E) or at the base of the Tal Group
variation resulting from deglaciation of the shelf. (Banerjee et al. 1997).
Although numerous measurements were made to characterize Crystalline basement is not exposed in the region, so maximum
the trends across the fold belt, extremely negative values have age constraints are poorly defined. However, new detrital zircon
not been observed that may implicate local methane hydrate desta- ages for samples of the diamictite at the base of the Blaini Fm.
bilization (cf. Jiang et al. 2003b; McFadden et al. 2008), although (at Māldeota in the Mussoorie syncline) provide, to our knowl-
this does not negate dissociation events as a driving mechanism for edge, the first direct radiometric data for the unit, with a youngest
negative excursions. The trend towards negative values, and recov- subpopulation of three grains giving a mean weighted-average age
ery towards more positive values is similar to that recorded of 692 + 18 Ma (207Pb/206Pb; Fig. 31.4). Four grains in the same
from the Keilberg cap on the South Congo Craton and generally sample define a slightly older subpopulation with a mean
exhibited by ‘Marinoan’ (c. 635 Ma) cap carbonates elsewhere weighted-average age of 770 + 24 Ma (207Pb/206Pb). A number
(Kennedy et al. 1998). The observation of negative values in inter- of older detrital populations are evident, which are also recorded
diamictite carbonates by Kaufman et al. (2006) differ from the in other samples. The full results of the detrital zircon analysis
general thesis of Kennedy et al. (2001), although some matrix will be published elsewhere.
samples from the Ghaub Fm. on Fransfontein Ridge exhibit nega-
tive values below the Keilberg cap (Kennedy et al. 2001, fig. 1).
Conclusions
Palaeolatitude and palaeogeography The Blaini Fm. records widespread glaciomarine conditions of
sedimentation in a passive continental margin setting. At least
Palaeomagnetic data for the Malani Igneous Suite place Rajasthan two major periods of ice advance are recorded by the widespread
along the western margin of Rodinia in mid to high latitudes at distribution of glaciogenic diamictites, which include primary sub-
c. 750 Ma (Torsvik et al. 2001). Palaeopoles for the Blaini Fm. glacial tillites deposited where the ice was grounded on the conti-
also indicate a mid- to high-latitude position for the lower diamic- nental shelf, and glaciogenic debris flows that redistributed
tite (46.58N, 1098E), but tropical palaeolatitudes for the overlying subglacial debris onto submarine fans in outer neritic to slope
cap dolostone (3.08N, 98.58E; Klootwijk 1979). Although these environments. The huge volumes of glacioclastic debris and
palaeolatitudes sit comfortably with Phanerozoic analogues of their widespread distribution point towards the existence of an
northern-hemisphere glaciation, positive fold tests are required extensive continental ice sheet, probably with polythermal charac-
to demonstrate primary remnant magnetization. Better geochrono- teristics. The limited geochronological data do not presently allow
logical control is also required to constrain the position of this the duration of these glacial events to be constrained, but both
important glaciated passive continental margin in palaeogeo- periods of ice sheet growth were probably younger than
graphic reconstructions. 692 + 18 Ma. The cap dolomite bears stable isotopic character-
istics typical of Marinoan (c. 635 Ma) post-glacial cap carbonates
elsewhere; however, further data are required to better constrain
Geochronological constraints
the age of these deposits.
A Neoproterozoic (as opposed to the formerly accepted view of
This research was funded by Schweizerischer National Fonds Grant 103502.
Permo-Carboniferous) age for the Blaini Fm. is supported by the M. Papp, M. Faccenda and R. Singh are kindly thanked for their assistance in the
occurrence of Lower Cambrian trace fossils, trilobites, small field. Stable isotope analyses were undertaken by T. Venneman at the University
shelly fauna and acanthomorphic acritarchs in basal Tal Group of Lausanne and D. Mrofka at the University of California Riverside. Laser
chert and phosphorites (e.g. Banerjee & Narain 1976; Brasier & Ablation MC ICP-MS detrital zircon analyses were undertaken at the University
Singh 1987; Bhatt 1989, 1991; Tiwari & Knoll 1994; Tiwari of Alberta, Edmonton. This study has benefited from ongoing discussions with
1999; Hughes et al. 2005; Fig. 31.2). A SHRIMP U –Pb detrital M. Kennedy (UCR), R. Rieu (Repsol YPF), G. Jiang (University of Nevada) and
zircon age of 525 + 8 Ma from Tal Group sediments at Gopichand A. Cozzi (ENI), and represents a contribution towards the IGCP 512 project.
ka Mahal in the Mussoorie syncline supports palaeontological data S. K. Ghosh and R. Islam are grateful to the Director, Wadia Institute of Himalayan
indicative of a Cambrian age, and thus reflects the youngest Geology, for providing the necessary facilities. Special thanks go to D. Banerjee
isotopic age limit for the Blaini Fm. (Myrow et al. 2003). In the and A. J. Kaufman for constructive reviews, which improved this manuscript.
354 J. L. ETIENNE ET AL.

References Brookfield, M. E. 1994. Problems in applying preservation, facies and


sequence models to Sinian (Neoproterozoic) glacial sequences in
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves in the Australia and Asia. Precambrian Research, 70, 113–143.
aftermath of a Neoproterozoic glaciation. Nature, 433, 123– 127. De, C., Das, D. P. & Andraha, P. K. 1994. Ichnostratigraphic and
Alvarenga, C. J. S. De, Dardenne, M. A. et al. 2007. Isotope strati- palaeoenvironmental significance of trace fossils from Tal Formation
graphy of Neoproterozoic cap carbonates in the Araras Group, of Nigali Dhar Syncline, Sirmur District, Himachal Pradesh, India.
Brazil. Gondwana Research, 13, 469–479. Indian Journal of Geology, 66, 77 –90.
Auden, J. B. 1934. The Geology of the Krol Belt. Records of the Geologi- Dowdeswell, J. A., Kenyon, N. H., Elverhøi, A., Laberg, J. S.,
cal Survey of India, 69, 123–167. Hollender, F.-J., Mienert, J. & Siegert, M. J. 1996. Large-scale
Auden, J. B. 1946. Blaini-Talchir. Current Science, 12, 346–348. sedimentation on the glacier-influenced Polar North Atlantic
Azmi, R. J. & Paul, S. K. 2004. Discovery of Precambrian – Cambrian margins: long-range side-scan sonar evidence. Geophysical Research
boundary protoconodants from the Gangolihat Dolomite of Inner Letters, 23, 3535– 3538.
Kumaun Lesser Himalaya: implication on age and correlation. Dowdeswell, J. A., Elverhøi, A. & Spielhagen, R. 1998. Glacimarine
Current Science, 86, 1653–1660. sedimentary processes and facies on the polar North Atlantic
Banerjee, D. M. & Narain, M. J. 1976. Trace fossils in the Lower Tal Margins. Quaternary Science Reviews, 17, 243–272.
Formation of Mussoorie and their environmental significance. Fuchs, G. & Sinha, A. K. 1978. The tectonics of the Garhwal-Kumaun
Journal of Sedimentary Petrology, 46, 235–239. Lesser Himalaya. Jahrbuch der Geologischen Bundesanstalt, 121,
Banerjee, D. M., Schidlowski, M., Siebert, F. & Brasier, M. D. 1997. 219– 241.
Geochemical changes across the Proterozoic –Cambrian transition in Gaur, G. C. S. & Dave, V. K. S. 1971. Blaini tillites near Rishikesh and
the Durmala phosphorite mine section, Mussoorie Hills, Garhwal their origin. Journal of the Geological Society of India, 12, 164– 172.
Himalaya, India. Palaeogeography, Palaeoclimatology, Palaeoecol- Ghosh, S. K. 1991. Palaeoenvironmental analysis of the Late Proterozoic
ogy, 132, 183– 194. Nagthat Formation, NW Kumaun Lesser Himalaya, India. Sedimen-
Benn, D. I. & Evans, D. J. A. 1998. Glaciers and Glaciation. Arnold, tary Geology, 71, 33 –45.
London. Halverson, G. P. 2005. A Neoproterozoic chronology. In: Xiao, S. (ed.)
Bhargava, O. N. 1984. Trace fossils from the ?Cambrian Tal Group, Neoproterozoic Geobiology. Kluwer Academic Publishers, Delft,
Sirmur District H.P. and proposed redefinition of the Tal. Journal Netherlands.
of the Palaeontological Society of India, 29, 84 –87. Holland, T. H. 1908. On the occurrence of striated boulders in the Blaini
Bhargava, O. N. & Bhattacharyya, B. K. 1975. The Blaini Formation Formation of Simla with discussion on the geological age of the beds.
of Himachal Pradesh and Uttar Pradesh. Bulletin of the Indian Records of the Geological Survey of India, 37, 129–135.
Geologists’ Association, 8, 71 – 99. Hughes, N. C., Peng, S. et al. 2005. Cambrian biostratigraphy of the Tal
Bhargava, O. N., Singh, I., Hans, S. K. & Bassi, U. K. 1998. Early Group, Lesser Himalaya, India, and early Tsanglangpuan (late early
Cambrian trace and trilobite fossils from the Nigali Dhar Syncline Cambrian) trilobites from the Nigali Dhar syncline. Geological
(Sirmaur District, Himachal Pradesh), lithostratigraphic correla- Magazine, 142, 57 –80.
tion and fossil content of the Tal Group. Himalayan Geology, 19, Jain, A. K. 1981. Stratigraphy, petrography and palaeogeography of the
89 –108. Late Paleozoic diamictites of the Lesser Himalaya. Sedimentary
Bhatia, M. R. 1981. Appendix: Late Palaeozoic diamictites of Simla Geology, 30, 43 –78.
Hills, Lesser Himalaya, India; chronostratigraphic age and geo- Jain, A. K. & Varadaraj, N. 1978. Stratigraphy and provenance of Late
chemistry. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Palaeozoic diamictites in parts of Garhwal Lesser Himalaya, India.
Pre-Pleistocene Glacial Record. Cambridge University Press, Geologische Rundschau, 67, 49 – 72.
Cambridge, 293– 294. Jell, P. A. & Hughes, N. C. 1997. Himalayan Cambrian trilobites.
Bhatia, S. B. & Kanwar, R. C. 1975. Blaini and related formations. Special Papers in Palaeontology, 58, 1– 113.
Special Issue. Bulletin of the Indian Geologists’ Association, 8, 279. Jiang, G., Christie-Blick, N., Kaufman, A. J., Banerjee, D. M. &
Bhatia, M. R. & Prasad, A. K. 1975. Some sedimentological, lithostrati- Rai, V. 2002. Sequence stratigraphy of the Neoproterozoic Infra
graphic and genetic aspects of the Blaini Formation of parts of Simla Krol Formation and Krol Group, Lesser Himalaya, India. Journal
Hills, Himachal Pradesh, India. Bulletin of the Indian Geologists’ of Sedimentary Research, 72, 524– 542.
Association, 8, 162– 185. Jiang, G., Sohl, L. E. & Christie-Blick, N. 2003a. Neoproterozoic
Bhatia, M. R. & Prasad, A. K. 1981. Evolution of Late Paleozoic glacial stratigraphic comparison of the Lesser Himalaya (India) and Yangtze
marine sedimentation in the Simla Hills, Lesser Himalaya, India. block (south China): Paleogeographic implications. Geology, 31,
Neues Jahrbuch fur Palaontologie Mh, 267– 288. 917–920.
Bhatt, D. K. 1989. Small shelly fossils, Tommotian and Meishucunian Jiang, G., Kennedy, M. J. & Christie-Blick, N. 2003b. Stable isotopic
stages and the Precambrian –Cambrian boundary – implications of evidence for methane seeps in Neoproterozoic postglacial cap
the recent studies in the Himalayan sequences. Journal of the carbonates. Nature, 426, 822–826.
Palaeontological Society of India, 34, 55 – 68. Jiang, G., Kennedy, M. J., Christie-Blick, N., Wu, H. & Zhang, S.
Bhatt, D. K. 1991. The Precambrian –Cambrian transition interval in 2006. Stratigraphy, sedimentary structures, and textures of the late
Himalaya with special reference to small shelly fossils – a review Neoproterozoic Doushantuo cap carbonate in South China. Journal
of current status of work. Journal of the Palaeontological Society of Sedimentary Research, 76, 978– 995.
of India, 36, 109–120. Joshi, A. & Mathur, V. K. 1987. Report of Cruziana type trace fossils
Bhatt, D. K. & Mathur, A. K. 1990. Small shelly fossils of the from the Arenaceous Member of Tal Formation, Mussoorie
Precambrian Cambrian boundary beds from the Krol-Tal succession Synform. Indian Minerals, 41, 61 –65.
in the Nainital Syncline, Lesser Himalaya. Current Science, 59, Joshi, A., Mathur, V. K. & Bhatt, D. K. 1989. Discovery of redlichid
218– 222. trilobites from the Arenaceous Member of the Tal Formation,
Boulton, G. S. 1996. Theory of glacial erosion, transport and deposition Garhwal Syncline, Lesser Himalaya, India. Journal of the Geological
as a consequence of subglacial sediment deformation. Journal of Society of India, 33, 538– 546.
Glaciology, 140, 43 –62. Kaufman, A. J., Jiang, G., Christie-Blick, N., Banerjee, D. M. & Rai,
Brasier, M. D. & Singh, P. 1987. Microfossils and Precambrian – V. 2006. Stable isotope record of the terminal Neoproterozoic Krol
Cambrian boundary stratigraphy at Maldeota, Lesser Himalaya. platform in the Lesser Himalayas of northern India. Precambrian
Geological Magazine, 124, 323–345. Research, 147, 156– 185.
Brookfield, M. E. 1987. Lithostratigraphic correlation of Blaini For- Kennedy, M. J., Runegar, B., Prave, A. R., Hoffman, K.-H. &
mation (late Proterozoic, Lesser Himalaya, India) with other late Pro- Arthur, M. A. 1998. Two of four Neoproterozoic glaciations?
terozoic tillite sequences. Geologische Rundschau, 76, 477–484. Geology, 26, 1059–1063.
Brookfield, M. E. 1993. The Himalayan passive margin from Precam- Kennedy, M. J., Christie-Blick, N. & Prave, A. R. 2001. Carbon isoto-
brian to Cretaceous times. Sedimentary Geology, 84, 1 – 35. pic composition of Neoproterozoic glacial carbonates as a test of
THE BLAINI FORMATION 355

paleoceanographic models for snowball Earth phenomena. Geology, In: Dowdeswell, J. A. & O. Cofaigh, C. (eds) Glacier-Influenced
29, 1135– 1138. Sedimentation on High-Latitude Continental Margins. Geological
Klootwijk, C. T. 1979. Review of palaeomagnetic data from the Indo- Society, London, Special Publications, 203, 215– 244.
Pakistani fragment of Gondwanaland. In: Farah, A. & De Long, Prasad, B., Maithy, P. K., Kumar, G. & Raina, B. K. 1990. Precam-
K. A. (eds) Geodynamics of Pakistan. Geological Survey of Pakistan, brian– Cambrian acritarchs from the Blaini-Krol-Tal sequence of
Queta, 41– 80. Mussoorie Syncline, Garhwal Lesser Himalaya, India. Memoirs of
Kumar, G., Raina, B. K., Bhatt, D. K. & Jangpangi, B. S. 1983. Lower the Geological Society of India, 16, 19 – 32.
Cambrian body- and trace-fossils from the Tal Formation, Garhwal Rai, V. 1987. Additional trace fossils from the Tal Formation (Early
Synform, Uttar Pradesh, India. Journal of the Palaeontological Cambrian) Mussoorie hills, Uttar Pradesh, India. Journal of the
Society of India, 28, 106– 11. Palaeontological Society of India, 32, 53 –59.
Kumar, G., Bhatt, D. K. & Raina, B. K. 1987. Skeletal microfauna of Rupke, J. 1968. Note on the Blaini Boulder Bed of Tehri Garhwal,
Meishucunian and Qiongzhusian (Precambrian– Cambrian bound- Kumaon Himalayas. Journal of the Geological Society of India, 9,
ary) age from the Ganga Valley, Lesser Himalaya, India. Geological 131– 133.
Magazine, 124, 167– 171. Saxena, M. N. & Pande, C. 1969. The Blaini Tillite from the type area
Mathur, V. K. & Joshi, A. 1989a. Record of inarticulate brachiopods of Simla Himalayas. Bulletin of the Indian Geological Association,
from the Arenaceous Member of the Tal Formation, Garhwal 2, 57 – 64.
Syncline, Lesser Himalaya, India. Current Science, 58, 446–448. Schelling, D. 1992. The tectonostratigraphy and structure of the eastern
Mathur, V. K. & Joshi, A. 1989b. Record of redlichiid trilobite from Nepal Himalaya. Tectonics, 11, 925– 943.
the Lower Cambrian Tal Formation, Mussoorie Syncline, Lesser Shanker, R., Mathur, V. K., Kumar, G. & Srivastava, M. C. 1997.
Himalaya, India. Journal of the Geological Society of India, 33, Additional Ediacaran biota from the Krol Group, Lesser Himalaya,
268– 270. India and their significance. Geoscience Journal, 18, 79– 94.
Mathur, V. K. & Shanker, R. 1989. First record of Ediacaran fossils Singh, I. B. & Rai, V. 1983. Fauna and biogenic structures in Krol-Tal
from the Krol Formation, Nainital syncline. Journal of the Geological succession (Vendian– Early Cambrian), Lesser Himalaya: their
Society of India, 34, 245– 254. biostratigraphic and palaeoecological significance. Journal of the
Mathur, V. K. & Shanker, R. 1990. Ediacaran medusoids from Palaeontological Society of India, 28, 67 –90.
Cambrian Tal Formation, Himachal Lesser Himalaya and the Krol Taylor, J., Dowdeswell, J. A., Kenyon, N. H. & O ´ ’Cofaigh, C. 2002.
Formation, Naini Tal syncline. Journal of the Geological Society of Late Quaternary architecture of trough-mouth fans: debris flows and
India, 36, 74 – 78. suspended sediments on the Norwegian margin. In: Dowdeswell,
Mathur, V. K. & Srivastava, M. C. 1994. Record of microgastropod J. A. & O’Cofaigh, C. (eds) Glacier-Influenced Sedimentation on
from the Arenaceous Member of the Tal Formation, Garhwal High-Latitude Continental Margins. Geological Society, London,
Syncline, Lesser Himalaya, India. Current Science, 66, 228–229. Special Publications, 203, 55 –71.
Mathur, V. K., Joshi, A. & Kumar, G. 1988. Trace fossils from Tiwari, M. 1999. Organic-walled microfossils from the Chert–
Cambrian Tal Formation Himachal Lesser Himalaya, India, and phosphorite Member, Tal Formation, Precambrian –Cambrian
their stratigraphic significance. Journal of the Geological Society of Boundary, India. Precambrian Geology, 99, 99– 113.
India, 31, 467– 475. Tiwari, M. & Knoll, A. H. 1994. Large acanthomorphic Acritarchs from
Mazumdar, A. & Banerjee, D. M. 1998. Siliceous sponge spicules in the Infrakrol Formation of the Lesser Himalaya and their stratigraphic
the Early Cambrian Chert-phosphate Member of the Lower Tal significance. Journal of Himalayan Geology, 5, 193– 201.
Formation, Krol belt, Lesser Himalaya. Geology, 26, 899– 902. Torsvik, T. H. 2003. The Rodinia jigsaw puzzle. Science, 300,
McFadden, K. A., Huang, J. et al. 2008. Pulsed oxidation and biological 1379–1381.
evolution in the Ediacaran Doushantuo Formation. Proceedings of the Torsvik, T. H., Carter, L. M., Ashwal, L. D., Bhushan, S. K., Pandit,
National Academy of Sciences of the United States of America, 15, M. K. & Jamtveit, B. 2001. Rodinia refined or obscured: palaeomag-
3197– 3202. netism of the Malani igneous suite (NW India). Precambrian
Medlicott, H. B. 1864. On the geological structure and relation of the Research, 108, 319–333.
southern portion of the Himalayan ranges between the rivers Tripathi, C., Jangpangi, B. S., Bhatt, D. K., Kumar, G. & Raina, B. K.
Ganges and Ravee. Memoirs of the Geological Survey of India, 3, 1984. Early Cambrian brachiopods from ‘Upper Tal’, Mussoorie
1 – 212. syncline, DehraDun District, Uttar Pradesh, India. Geophytology,
Myrow, P. M., Hughes, N. C. et al. 2003. Integrated tectonostrati- 14, 221– 227.
graphic analysis of the Himalaya and implications for its tectonic Tripathi, C., Kumar, G., Mehra, S., Bhatt, D. K., Mathur, V. K. &
reconstruction. Earth and Planetary Science Letters, 212, 433–441. Joshi, A. 1986. Additional Early Cambrian (Botomian) brachiopod
Niyogi, D. & Bhattacharya, S. C. 1971. A note on the Blaini boulder fossil localities in Tal Formation, Lesser Himalaya, India, and their
beds of the Lower Himalaya. Himalayan Geology, 1, 111– 122. significance. Current Science, 55, 585–588.
Oldham, R. D. 1887. Notes on some points in Himalayan Geology. Valdiya, K. S. 1973. Blaini conglomerates of Himachal Pradesh and
Records of the Geological Survey of India, 20, 155–161. Garhwal. Recent Researchs in Geology, 1, Hindustan Publishing
Pilgrim, G. E. & West, W. D. 1928. The structure and correlation of Corporation, Delhi.
Simla rocks. Memoirs of the Geological Survey of India, 53, 1– 140. Valdiya, K. S. 1995. Proterozoic sedimentation and Pan-African
Powell, R. D. & Cooper, J. M. 2002. A glacial sequence strati- geodynamic development in the Himalaya. Precambrian Research,
graphic model for temperate, glaciated continental shelves. 74, 35– 55.
Chapter 32

Neoproterozoic glacial records in the Yangtze Region, China

QI-RUI ZHANG*, XUE-LEI CHU & LIAN-JUN FENG


State Key Laboratory of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy of Sciences,
PO Box 9825, Beijing 100029, China
*Corresponding author (e-mail: qrzhang@mail.iggcas.ac.cn)

Abstract: The Nanhua rift occurred at c. 820 Ma. The last rifting stage was likely associated with the Neoproterozoic glaciations. The
glacially influenced sequence, in ascending order, comprises the Jiangkou Group (c. 720– 663 Ma), the interglacial Datangpo Formation
(Fm.) (c. 663 Ma) and the Nantuo Fm. (,656– 635 Ma). In deep-water environments the glacially influenced units exhibit conformable
contacts with a transitional lower boundary and a sharp upper boundary. In shallow marine environments, the lower boundary is always
unconformable and erosive and the upper boundary is sharp but conformable.
The Jiangkou glaciogenic deposits likely comprise the waxing, the maximum and the waning stages of glaciation. The lower part of the
Chang’an Fm. corresponds to the waxing stage. It initially inherited the pre-glacial deep-water environment, and later, the thickening and
coarsening upward trend likely indicates a lowering of sea level and ice-sheet advance. The predominant, massive diamictite in the upper
Chang’an Fm. probably highlights the maximum glacial stage, and likely resulted from an amalgamation of debris flow deposits, suspen-
sion and ice-rafted debris. The Fulu Fm. records the waning stage; its lower member is characterized by arkosic sandstone and greywacke
indicative of an interstadial between glacial events, while the upper member represents the final glacial event of the Jiangkou glaciation.
The carbonaceous black shale and Mn carbonate at the base of the interglacial Datangpo Fm. were likely indicative of sedimentary star-
vation associated with Jiangkou deglaciation. The Nantuo glaciation is characterized by an abrupt onset and termination with negligible
waxing and waning stages. The sediments likely resulted from subglacial deposition, sedimentary gravity flows and ice-rafted debris with
reworking by currents.
The onset of Jiangkou glaciation took place around c. 720 Ma and terminated at c. 663 Ma. The Nantuo glaciation started at
,c. 656 Ma and terminated at 635 Ma. The palaeolatitude during the Jiangkou glaciation is unconstrained and was 37 + 78 during
the Nantuo glaciation. The palaeogeographic reconstruction for the Yangtze Block is controversial: it was either located between
Australia and Laurentia or to the NW of Australia and to the north of India.
Values of d34Spyrite from the top of the Fulu and the base of Datangpo formations exhibit extraordinarily high values up to þ60‰. Values
of d13Ccarb from the base of the Datangpo Fm. range from –7.4‰ to –13.0‰ and from the cap carbonate of the Doushantuo Fm. are gen-
erally lower than –5‰; however, methanogenic values as low as –48‰ have been reported. From the Datangpo Fm., dinosterane has been
detected, and phytane, triterpanes and hopane-type pentacyclic triterpanes are also abundant. Biomarker data from glacially influenced rocks
indicate that the photosynthetic process never ceased and that the marine environment was probably normal during glaciations.
The banded iron formation (BIF) at the base of the Fulu Fm. is either basaltic, tuffaceous, siliciclastic or carbonate-hosted. Rhodochrosite
at the base of the Datangpo Fm. was probably associated with submarine volcanic and hydrothermal activity, but a detrital oxide origin
cannot be excluded.
The difference between the Jiangkou and the Nantuo glaciations is significant. Tectonics and associated topography are speculated to be
the main controlling factors on the Neoproterozoic glaciations in the Yangtze region.

The Neoproterozoic glacially influenced (GI) rocks in China Cryogenian Period (850 –650 Ma) (Remane 2000). However, this
were first recognized by Willis et al. (1907). Wang et al. (1981) paper redefines its range as from c. 720 Ma to 635 Ma on the basis
summarized the Neoproterozoic GI record in China and Lu et al. of the most recent geochronological results.
(1985) edited the first memoir of the Neoproterozoic GI rocks in
China. By the 1990s, a series of geological memoirs on the
regional geology in each province reported the GI successions in The Jiangkou Group and the Jiangkou glaciation
detail. This paper focuses on the Neoproterozoic GI records in
the Yangtze region of South China. The Jiangkou Group, formerly the Jiangkou Fm. (Yang 1985), is a
The stratigraphic classification scheme of Neoproterozoic GI term for the lower GI deposits and encompasses the Chang’an and
deposits used in this paper is shown in Figure 32.1b. The strati- Fulu formations (BGMR-Hunan 1997; Zhang et al. 2003).
graphic nomenclature changes considerably, so brief explanations Before the 1980s, the term ‘Chang’an glaciation’ (Wang et al.
are necessary. 1981) was used to represent the early glaciation, and Lu et al.
(1985) proposed the term ‘Gucheng glaciation’ to replace it.
The ‘Sinian’ System and Period However, neither is appropriate, so the term Jiangkou glaciation
is preferred (Zhang et al. 2003).
The term ‘Sinian’ first appeared in the nineteenth century
(Richthofen 1882), and since 1983 it has been restricted to the The Fulu Fm.
Neoproterozoic in South China (Liu et al. 1991). More recently,
however, it was redefined as the stratigraphic unit between the The term ‘Fulu Fm.’ has been redefined at least twice. It was first
Nantuo Fm. and the Cambrian (NCS 2001), and coincides with defined as an interglacial unit between the Nantuo and the
the Ediacaran Period (Knoll et al. 2004). Chang’an formations (Fig. 32.2b, A) (BGMR-Guangxi 1985).
The upper boundary of the Fulu Fm. was then lowered to the
The Nanhuan System and Period base of the later established Datangpo Fm. (Lu et al. 1985).
Recently, the upper boundary of the Fulu Fm. was again lowered
The Nanhuan ‘period’ ranges between 800 Ma and 680 Ma (NCS to the base of the ‘Gucheng Fm.’ (Fig. 32.2b, A), which was
2001), and so almost coincides with the chronometrically defined emphasized to be the record of a single glaciation (BGMR-Hunan

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 357– 366. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.32
358 Q. R. ZHANG ET AL.

Fig. 32.1. (a) Outcrops of Neoproterozoic glacial deposits in the Yangtze region (adapted from Liu et al. 1991): (i) outline of the mainland of China and (ii) the inset
rectangle in (i). The star in (ii) is the location of Zhaoxing section and the pentagon the location of Shixian section. (b) A comprehensive Neoproterozoic stratigraphic
succession in southeastern Guizhou (adapted from BGMR-Guizhou 1987). 1, Condon et al. 2005; 2, Zhou et al. 2004; 3, Zhang et al. 2008a; 4, Zhou et al. 2007a; 5, Wang
et al. 2003a; *, Hao & Zhai 2004.

1997; Zhu et al. 2007). This paper adopts the definition given by uplift and erosion in the interior of the region and the accommo-
Lu et al. (1985). dation created in the rift basin. Syn-sedimentary faulting is
By 1985, based on the occurrence of BIF and rhodochrosite, the evident from abrupt changes in thickness of the GI units. The
Fulu Fm. was considered to be an interglacial unit sandwiched manganese deposits of the Datangpo Fm. were likely related to
between the ‘Chang’an glaciation’ and the Nantuo glaciation. hydrothermal and submarine volcanic activity (Liu 1994; Yang
After the establishment of the interglacial Datangpo Fm., the et al. 1997).
former ‘interglacial’ nature of the Fulu Fm. became invalid and The great thickness (.2000 m, Fig. 32.3) of the Nantuo Fm. in
in fact it records the waning stage of the Jiangkou glaciation. the SW of Hunan (BGMR-Hunan 1988) and SE of Guizhou
(BGMR-Guizhou 1987) probably signifies the final filling of the
rift basin. The ‘Nd isotopic shift’ (c. 765 Ma to 635 Ma) revealed
Structural framework by Li & McCulloch (1996) also reflects the final filling stage of
the Nanhua Rift.
At c. 1140 –900 Ma, the Yangtze and the Cathaysian plates col-
lided along the southeastern margin of the Yangtze region (Li
et al. 2008). The Nanhua rift started at c. 820 Ma (Fig. 32.1b) at Stratigraphy
the marginal area of the plate (Wang & Li 2003). At c. 830–
795 Ma and c. 780 –745 Ma, two pulses of bimodal anorogenic A comprehensive stratigraphic column of the Neoproterozoic is
magmatic activity occurred (Li et al. 2003). The early pulse given in Figure 32.1b. The GI units (c. 720– 635 Ma) bracketed
began prior to, and peaked at the onset of the rift; the later pulse by the Ediacaran and the ‘Yangzian’ (Hao & Zhai 2004) are
of volcanism accompanied the main rifting stage. The first three defined as Nanhuan (Cryogenian) in this paper. There are two GI
rifting phases were likely pre-glacial, and the fourth phase associ- units: the Jiangkou Group and the Nantuo Fm. (Fig. 32.1b).
ated with glaciation (Wang & Li 2003). The Jiangkou Group comprises two formations: the Chang’an
The surface below the Neoproterozoic GI deposits is important and the Fulu. The Chang’an Fm. occurs only in the western parts
and mostly represents a characteristic stratigraphic break. It prob- of the SE margin (Figs 32.2 & 32.3). The Fulu Fm. is composed
ably resulted from the uplift and erosion of the rift shoulder. The of two members: the Liangjiehe and the Gucheng (sections A to
pre-glacial units immediately below the erosive surface become D in Fig. 32.2); the contact between them is generally conformable
older towards the north (Fig. 32.2b), while units above the (section A & B in Fig. 32.2). The Nantuo Fm. (Fig. 32.1b) is the
erosive surface become younger (Fig. 32.2b), probably indicating most widespread Neoproterozoic GI unit (Figs 32.2 and 32.3).
that rift-induced subsidence persisted throughout the Nanhuan Sandwiched between the two GI units is the interglacial Datangpo
period. The Doushantuo cap dolostone likely marks the end of Fm., which is thinner and finer grained. In the east of the Yangtze
rifting and the onset of thermal subsidence. region, this tripartite structure is also typical, but the two lower
Magmatism during the Nanhuan was relatively weak and units are much thinner. Generally, the Nanhuan in the Yangtze
probably represents the final episode of magmatism and region is better preserved because of the tectonic simplicity and
rifting. The rifting may reconcile the great contrast between the low degree of metamorphism.
THE YANGTZE REGION 359

Fig. 32.2. Neoproterozoic lithostratigraphy in the Yangtze region. (a) Variation along a north–south direction (locations of sections are given as stars in Fig. 32.3). The
distance between the two end sections is c. 607 km. Distances between adjacent sections are also given. Vertical scale is the same except for section A. (b) Schematic
stratigraphic break structure. Section A is adapted from BGMR-Guangxi (1985). Sections E, F and G are adapted from Sha et al. (1963). Section G shows no
Neoproterozoic glacial deposits, and section A is complete.

Glaciogenic deposits and associated strata with cross-stratification and loading structures, and lenses of con-
glomerate and breccia (Fig. 32.4a). Usually it is grey to greenish in
The sedimentological description and facies analysis of the GI colour with 60 to 80% clay minerals. Clast content is less than 5%,
deposits are updated regularly, with the most recent and detailed with sizes ranging from 0.2 cm to 5 cm and maximum 80 cm, sub-
version being provided by Dobrzinski & Bahlburg (2007). angular to sub-rounded. Rock types are quartz, sandy mudstone,
Figure 32.3 shows the distribution of the Nantuo Fm., in which sandstone, quartzite-like sandstone, minor volcanic rocks and,
the dashed line outlines the distribution of the Jiangkou Group. rarely, striated clasts.

The Chang’an Fm. The Fulu Fm.

In deep-water environments, the contact between the Chang’an The Fulu Fm. comprises two members (Fig. 32.4b): the Liangjiehe
and the pre-glacial Gongdong formations is generally conformable and the Gucheng (Zhang et al. 2003).
(BGMR-Guangxi 1985; BGMR-Guizhou 1987). The transitional The Liangjiehe Member (583 m) comprises mainly massive,
base of the Chang’an Fm. is distinguished from the pre-glacial fine-grained arkosic arenite or greywacke with lonestones, and
grey to greenish mudstone by sandy mudstone with clasts. occasionally with beds and lenses of granular conglomerate and
The Chang’an Fm. can be divided into two parts. The majority diamictite. Cross-stratification and ripple marks, as well as
of the lower part (619 m thick) is mudstone (Fig. 32.4a), with graded bedding, are present. Locally at the base there are layers
cycles showing 0.1– 0.8 mm stratification. There are also coarse of purple to reddish banded ironstone with graded bedding and
to fine-grained, massive to graded sandstones, with few current cross-stratification. Beds or lenses of micritic dolostone are
ripples and sole flutes, stratified and massive diamictites and also reported.
minor granular conglomerate at the base. Clast size varies from From the sedimentary characteristics of the Liangjiehe
0.2 to 5 cm, clast shape from sub-rounded to angular, and the Member in the Zhaoxing section (Fig. 32.4b), it is hard to deter-
clast types are mainly quartz, sandstone and slate, with minor mine its climatic character. However, there are sections where
fragments of basic to acidic igneous rocks. Locally, there are the Liangjiehe Member is typified by lonestone and beds of
lonestones (likely dropstones), striated clasts and mudstone massive diamictite. Therefore, it is reasonable to suggest that
breccia. Reverse grading is occasionally seen in coarse- it is interstadial.
grained intervals. The Gucheng Member in the type section in Changyang County
The upper part (1279 m) comprises massive diamictite (92% of Hubei Province is only 5.6 m thick, and comprises beds of
of whole), with subordinate stratified mudstone, sandstone diamictite with about 30– 40 vol% clasts, and lenses of sandy
360 Q. R. ZHANG ET AL.

Fig. 32.3. Distributions of Jiangkou Group (outlined by the black dashed line; adapted from BGMR-Guangxi 1985 and Wang & Li 2003) and Nantuo Fm. (grey areas;
adapted from Wang & Li 2003). The contours show thickness of the Nantuo Fm. only.

conglomerate, sandy mudstone with lonestones (interpreted as striated and bullet-shape clasts (Fig. 32.4c) are frequently observed
dropstones), and striated clasts (Ma & Wang 1983). In the deep- (BGMR-Guangxi 1985). Because more than 90% of its thickness is
water Zhaoxing section the Gucheng Member is about 217 m massive diamictite, it looks similar to the upper part of the
thick (Fig. 32.4b). Chang’an Fm.; however, its clasts are larger, more common and
The Fulu Fm. is also characterized by its calcareous composition lithologically more variable (e.g. granite, diabase, volcanic rocks,
(Dobrzinski & Bahlburg 2007) and dolomite mounds or lenses carbonate etc.), especially in its upper part. Allochthonous sedi-
(Zhou et al. 2007b). mentary blocks, in cases up to a maximum thickness of 10 m, fre-
quently occur on shelf and upper slope environments. Diamictites
with sharp base and deformed stratification, and occasionally
The Datangpo Fm. gravel lag deposits, can be recognized in fresh outcrops.

This is named after a section in northeastern Guizhou Province,


with more than 500 m (Fig. 32.5) of fine-grained muddy The cap carbonate of the Doushantuo Fm.
sandstone, muddy siltstone, mudstone and about 5 m of black
carbonaceous manganese shale at the base (Ma & Wang The cap carbonate represents the first member of the Doushantuo
1983). A range of sedimentary structures (gas holes, diapirs, Fm., typically 3– 6 m thick and extending basin wide with the
seeping tubes, soft sediment deformation and mud volcanoes) Nantuo Fm. (Fig. 32.3). It can be classified into three intervals:
are reported from layers and lenses of rhodochrosite associated C1, the strongly disrupted and cemented basal layer; C2, the
with dolomite lenses (Zhou et al. 2007b). The upward coarsen- middle laminated layer with local tepee-like structures; and
ing and thickening trend is obvious. To the SE of the Yangtze C3, the thinly laminated silty and shaly limestone and dolomite
region, it thins to a thickness of c. 12 m of dark carbonaceous (Jiang et al. 2006). More than 90% of the rock minerals are
shale (Fig. 32.4c) (BGMR-Guangxi 1985). However, there are micritic and microcrystalline dolomite with little quartz, feldspar
sections without black carbonaceous shale but only grey-green and muscovite; calcite occurs mainly as filling materials in
massive mudstone. fissures (Wang et al. 2005). On shelf environments the cap
carbonate is thicker, less deformed, with cemented breccias,
tepee-like structures, stromatactis-like cavities, sheet cracks
The Nantuo Fm. and barite fans. On the outer shelf or upper slopes it is thinly
laminated and deformed, and in the basal part (C1 of Jiang
The maximum thickness of the Nantuo Fm. is more than 2000 m et al. 2006). The centimetre-scale dolostone layers frequently
(Fig. 32.3); it is composed predominantly of diamictite with intercalate with lenses or layers of silicate. The cap carbonate
scarce mudstone, sandstone and conglomerate. Dropstones, and is missing in the lower slope and basin environments.
THE YANGTZE REGION 361

Fig. 32.4. Stratigraphic columns of the (a) Chang’an and (b) Fulu formations along the Zhaoxing section (the star in Fig. 32.1a,ii; adapted from BGMR-Guizhou 1987)
and (c) the Datangpo and Nantuo formations along the Shixian section (the pentagon in Fig. 32.1a,ii; adapted from BGMR-Guangxi 1985).

Boundary relations with overlying and underlying The upper boundary of the Jiangkou Group
non-glacial units
The boundary is conformable but lithologically sharp. Below the
The boundaries of the GI units are conformable in slope to surface is diamictite (gravel bearing sandy mudstone) of the
basin environments; in shelf environments the upper boundary Gucheng Member and above it is usually the carbonaceous black
remains conformable but the lower boundaries are always shale and locally grey-greenish mudstone of the Datangpo Fm.
unconformable.

The lower boundary of the Nantuo Fm.


The lower boundary of the Jiangkou Group
On shelf and upper slope environments, the contact between
This is mostly unconformable and in contact with different units; Nantuo Fm. and the Datangpo Fm. is lithologically sharp and
however, it is conformable in the SW deep-water environment erosive. In deep-water environments, however, it is locally transi-
(Fig. 32.2b), for example, in the Zhaoxing section (the star in tional from mudstone to inversely graded diamictite beds as seen in
Fig. 32.1a(ii)) it is transitional. The top part of the Gongdong the Pancao section of the Guzhang County in Hunan Province and
Fm. (Fig. 32.1b) comprises thinly bedded, grey sandy mudstone in the Dacao section of the Ningguo County in Southern Anhui
in which a single layer (c. 4 cm thick) is usually composed of a Province.
parallel laminated lower part and a structureless upper part; this
is likely to correspond to the Td and Te of the Bouma divisions,
respectively. The base of the Chang’an Fm. is, however, char- The upper boundary of the Nantuo Fm.
acterized by a thickly bedded greywacke, c. 20 m thick, with
sporadic pebbles (,5 mm), on top of which is laminated sandy The contact between the Nantuo and Doushantuo formations is
mudstone, where dropstones and clast clusters are occasionally sharp, conformable and occasionally pyrite-bearing (Jiang et al.
found. 2006). The top of the Nantuo Fm. usually comprises diamictite
362 Q. R. ZHANG ET AL.

Fig. 32.5. The type section of Datangpo


Fm. in Songtao County in the NE of
Guizhou Province (adapted from Yang
et al. 2002).

in shelf environments; however, in slope and basin environments Datangpo Fm. At the base of the Datangpo Fm., values of d13Ccarb
there could exist a thin interval of sandy mudstone or of rhodochrosite range from –7.4‰ to –13.0‰ (Tang & Liu 1999;
muddy siltstone. Yang et al. 2002), so values of d13Ccarb from the base of Datangpo
Fm. are probably lower than –5‰ (Chu et al. 2003).
Isotope stratigraphy Doushantuo Fm. Carbon-isotope data from the cap carbonate of
34 Doushantuo Fm. are abundant. Jiang et al. (2003a) and Wang
d Spyrite et al. (2008a) reported very negative C-isotope values as low
as –48‰ in the Yangtze Gorges area and in Changyang
Fulu Fm. This is about 30 m thick in the Minle section of Hunan County, interpreted as evidence for methanogenesis. Generally
Province. d34Spyrite values of þ24.1‰ and þ23.5‰ at c. 1 m there is a negative excursion in the cap carbonate of the
and 10 m from the base and þ41.8‰ to þ58.4‰ at c. 23 m and Doushantuo Fm. (Jiang et al. 2007; Zhou & Xiao 2007; Zhu
29 m are reported by Li et al. (1999). Zhou et al. (2007b) speculate et al. 2007).
that these high positive d34Spyrite values were likely associated with
cold methane seepage.
Nd depleted mantle model ages (tDM)
Datangpo Fm. Super heavy values of d34Spyrite (40 –60‰) are
reported from the vicinity of rhodochrosite mineralization (Li
The tDM ages of sediments from the Neoproterozoic to lower
et al. 1999, 2006; Tang & Liu 1999) and are attributed alternatively
Danzhou Group (c. 820– 765 Ma, Fig. 32.1b) are stable at
to the anoxic conditions, low level of dissolved sulphate S (Liu
c. 1.8 Ga (Li & McCulloch 1996). In the upper Danzhou Group
et al. 2006) or restricted basin (Li et al. 1999). However, Wang
(c. 765–720 Ma, Fig. 32.1b) the tDM ages drop to 1.3–1.4 Ga,
et al. (2005) suggest that these extraordinarily high d34S values
and during the Nanhuan (c. 720–635 Ma, Fig. 32.1b) return to
were probably associated with cold methane seepage, as they are
c. 1.8 Ga. From the Ediacaran onwards (,635 Ma), the tDM ages
similar to the values of c. 47.8– 67.1‰ from ODP Site 1146 in
stabilize at c. 1.8 Ga again. This ‘Nd isotopic shift’ (Li &
the South China Sea.
McCulloch 1996, fig. 3) probably reveals the final filling of the
Nanhua rift by predominant juvenile crustal materials.
Cap carbonate of the Doushantuo Fm. Values of d34Spyrite ranging
from 36.35‰ to 50.62‰ are reported by Wang et al. (2005),
with the high values interpreted as recording sulphate reduction
at cold methane seeps. Other characteristics

Banded iron formation (BIF)


d13Ccarb
BIF appears at the base of the Fulu Fm. (Fig. 32.4), and can be
Fulu Fm. In the upper part of the Fulu Fm., values of d13Ccarb from traced from the northern Guangxi Region in the west to the
dolomite and pebbly dolomitic sandstone are higher than –2.86‰ middle of the Jiangxi Province in the east of the Yangtze region.
and were proposed to be associated with cold spring carbonate Tang et al. (1987) classify the deposits into four types: basaltic,
(Zhou et al. 2007b). tuffaceous, siliciclastic and carbonate interbedded.
THE YANGTZE REGION 363

Manganese deposits Geochronological constraints

The black shale-hosted manganese deposits of the Datangpo Fm. Onset of the Jiangkou glaciation
are important in China, although the deposits are relatively small
(Roy 2006). They are mostly associated with faults and formed A sample from the top unit of the Banxi Group (Fig. 32.2b, B) pro-
in local rift or graben basins. The content of volcanic materials vides a SHRIMP U –Pb zircon age of 725 + 10 Ma (Zhang et al.
may reach 25–50% (Yang & Pang 2006). Submarine volcanic 2008a). Because the Danzhou Group (Fig. 32.2b, A) is transitional
and hydrothermal activities are considered to be the major factor upward into the Jiangkou Group, this age becomes a new
in the formation of the deposits (Yang et al. 1997; Qin et al. maximum age constraint on the onset of the Jiangkou glaciation,
2005; Yang & Pang 2006). They are also closely associated with and the age of the onset is reasonably estimated to be c. 720 Ma.
deglacial transgression after the Jiangkou glaciation, so Liu et al.
(2006) speculate that they resulted from diagenetic remobilization
of lateritic soil residues deposited in restricted basins by low- Termination of the Jiangkou glaciation
latitude glaciation. These two factors are not mutually exclusive,
and a combined model is proposed (Tang & Liu 1999). The U –Pb zircon age of 663 + 4 Ma from a sample near the base
of the Datangpo Fm. (Zhou et al. 2004) approximates to the ter-
mination of the Jiangkou glaciation (Fig. 32.1b).
Biomarkers
Onset of the Nantuo glaciation
Wang et al. (2003b, 2008b) studied biomarkers from samples of
the Nanhuan System. The total organic content (TOC) of diamic- A recent U– Pb SHRIMP zircon age of 656 + 3 Ma from the top of
tites and cap carbonate range from 0.01 to 0.13%, 1–2 orders of the Dantangpo Fm. in west Hunan indicates that the onset of the
magnitude lower than those of the non-GI sedimentary rocks. Nantuo glaciation should be younger than c. 656 Ma (Zhang
The positive correlation trend between TOC and the concen- et al. 2008b). Considering the shelf setting in the sampled
trations of phytane and/or pristane implies that photosynthetic section and significant erosion at the Nantuo/Datangpo boundary,
autotrophs were probably the primary organisms contributing to the Nantuo glaciation could be significantly younger than 656 Ma.
the sedimentary organic matter. The unusually low concentrations
of phytane plus pristane (0.005 –0.076 ng g21) in diamictites
reveal a very weak photosynthetic process, and likely indicate Termination of the Nantuo glaciation
that the photosynthetic process never ceased, allowing the survival
of photosynthetic eukaryotes and other organisms. The GI sedi- After the dating of 635.2 + 0.6 Ma from samples near the base of
mentary environment was probably normal marine (Wang et al. the cap carbonate on top of the Nantuo Fm. (Condon et al. 2005),
2003b, 2008b). the age of the termination of Nantuo glaciation is constrained to
Dinosterane has been detected from black shale of the Datangpo c. 635 Ma (Ogg & Gradstein 2008).
Fm. (Meng et al. 2003), indicating that dinoflagellates and other
eukaryotic algae had evolved by the mid-Neoproterozoic. Xie Discussion
et al. (1999) report n-alkane distribution curves for manganese
deposits of the Datangpo Fm. Phytane is predominant, while the There has been a prolonged debate on the correlation of the
Pr– Ph ratios (0.71 –0.83) indicate a reducing environment. Triter- Liantuo Fm. with various GI units (Yin et al. 2003 among
panes and hopane-type pentacyclic triterpanes are abundant, and others), which has brought great difficulty to the investigation of
the organic matter originated from algae and microbes was the evolution of Neoproterozoic glaciations in South China.
likely associated with methane seeps (Zhou et al. 2007b). However, a newly achieved age of 725 + 10 Ma (Zhang et al.
2008a) from the top of the Banxi Group, the recalculated age of
766 + 18 Ma of the former 748 + 12 Ma of the Liantuo Fm.
Palaeolatitude and palaeogeography (Zheng 2003) and 765 + 14 Ma from the Danzhou Group (Zhou
et al. 2007a) seem to have solved the problem and confirm that
Palaeopoles the Liantuo Fm. is not a GI unit and can only be correlated with
the middle of the Banxi (Danzhou) Group.
Li et al. (1996) report a palaeolatitude of 38 + 88 for the Liantuo Based on the above evidence, it is almost certain that in the
Fm. (766 + 18 Ma) in Hubei Province. Zhang & Piper (1997) Yangtze region there were two Neoproterozoic glaciations: the
achieve a similar palaeolatitude of 37 + 78 for the Nantuo Fm. Jiangkou (c. 720 to 663 + 4 Ma) and the Nantuo (,656+3 to
(635 Ma) in Yunnan Province. 635.2 + 0.6 Ma). The salient decrease in tDM ages from c.
The proximity of the palaeopoles of Liantuo and Nantuo 765 Ma to c. 720 Ma probably indicates that the uplifted rift
formations indicates that either the palaeopoles remained almost shoulder was unroofed, weathered and eroded, and supports the
unchanged from the Liantuo stage to the Nantuo stage, or proposition that the atmospheric content of CO2 was significantly
that there was a loop in the apparent polar wander path from drawn down, triggering glaciations (Goddéris et al. 2007). The ‘Nd
c. 766 –635 Ma. isotopic shift’ reveals that a great amount of juvenile mantle-
derived material was incorporated into the sedimentary prove-
nance (Li & McCulloch 1996). The peculiar hiatus architecture
Palaeogeographical position of the GI stratigraphy in the Yangtze region (Fig. 32.2b) testifies
independently to the uplifting, erosion and filling of the rift
The relative position of the Yangtze region is controversial. Li basin. It is therefore the tectonics and greatly evolved topography
et al. (2008) places the Yangtze region (South China Block) that finally controlled the geological records of the glaciations.
between Australia and Laurentia during the assembly of the
supercontinent Rodinia at c. 900–780 Ma and proposes that it
moved to the NE of Australia during the break-up of Rodinia at The depositional environment
c. 750–630 Ma. Kirschvink (1992), Zhang & Piper (1997), Jiang
et al. (2003b) place the South China Block to the NW (present It was the Nanhua rift that provided the accommodation necessary
coordinate) of Australia and to the north of India. for GI deposition. The northwardly expanding distribution of
364 Q. R. ZHANG ET AL.

younger GI units (Fig. 32.2b) likely signifies the protracted The reasons for these differences are controversial. Dobrzinski
subsidence of the basin. Except for a few outcrops in the west of & Bahlburg (2007) emphasize the different reaches of the ice
the region, the Neoproterozoic GI deposits are believed to sheets and ascribe the differences to the palaeolatitude, intensity
be glaciomarine. and extension of the two glaciations. In fact, the role of the tectonic
control implied in Figure 32.2b was likely crucial, and the
The Jiangkou glaciations. The maximum thickness is 2698 m difference of the evolving topography in different rift stages was
(BGMR-Guizhou 1987) and it can be divided into three stages. probably a secondary controlling factor. The Jiangkou glaciation
The first or waxing stage corresponds to the lower part of the happened within a background of contrasting relief, while the
Chang’an Fm. (Fig. 32.4a). The laminated sandy mudstones with Nantuo glaciation occurred within a glacially levelled topography.
dropstones and clast-clusters at the base indicate an inherited Sedimentary structures (Liu et al. 1999), redox proxies
deep-water environment with the addition of glacially derived (Dobrzinski et al. 2004), and biomarkers (Wang et al. 2003b,
sediments. The thickening and coarsening upward trend probably 2008b) probably suggest that hydrological cycle was not shut
signifies the lowering of the sea level and advance of ice sheets. down during glaciations in the Yangtze region.
The predominant massive diamictite with graded sandstones and
mudstone in the upper part of the Chang’an Fm. (Fig. 32.4a) prob- This paper was supported by project nos 40532012 and 40373011 of the Natural
ably records the amalgamation of debris-flow deposits and rainout Science Foundation of China (NSFC) and project KZCX3-SW-141 of the Chinese
in an ice-distal environment during the second or main glacial Academy of Sciences (to X.-L. Chu). We thank G. A. Shields and E. Arnaud for
stage. The shallow ice-proximal records of these two stages were their encouragement and help in writing this paper. Detailed and insightful
likely deposited originally, but were not preserved because of reviews by G.-Q. Jiang, C.-M. Zhou, H. Bahlburg and G. A. Shields-Zhou are
erosion by the later glacial events. In the waning or third stage, gratefully acknowledged. This is a contribution to International Geological
Correlation Program Project 512.
BIF and massive and cross-stratified greywacke are probably
indicative of an interstadial, where the frequent occurrence of
lonestones indicates the existence of icebergs and ice sheets. The References
diamictite in the Gucheng Member was pro-glacial, as described
by Dobrzinski & Bahlburg (2007), and likely subglacial or ice- BGMR-GUANGXI (BUREAU OF GEOLOGY AND MINERAL
proximal too in certain sections. RESOURCES OF GUANGXI ZHUANG AUTONOMOUS
REGION) 1985. Regional Geology of Guangxi Zhuang Autonomous
The Datangpo interglacial. The thin starved sediments of the Region. Geological Publishing House, Beijing, 853 (in Chinese with
Datangpo Fm. in the Zhaoxing-Shixian area (the star and pentagon English abstract).
in Fig. 32.1a(ii)) are probably indicative of a deep-water environ- BGMR-GUIZHOU (BUREAU OF GEOLOGY AND MINERAL
ment. With few exceptions, nothing coarser than muddy fine- RESOURCES OF GUIZHOU PROVINCE) 1987. Regional
grained sandstone is reported from the unit in this region, so it Geology of Guizhou Province. Geological Publishing House,
probably represents a distal marine environment. The basal black Beijing, 698 (in Chinese with English abstract).
shale and Mn dolomite were probably deposited in a restricted BGMR-HUNAN (BUREAU OF GEOLOGY AND MINERAL
RESOURCES OF HUNAN PROVINCE) 1988. Regional Geology
shelf environment at the early transgression induced by the degla-
of Hunan Province. Geological Publishing House, Beijing, 719 (in
ciation of the Jiangkou glaciation. Chinese with English abstract).
BGMR-HUNAN (BUREAU OF GEOLOGY AND MINERAL
The Nantuo glaciation. The Nantuo Fm. in the Yangtze Gorges area RESOURCES OF HUNAN PROVINCE) 1997. Multiple Classifi-
was originally believed to be terrestrial (Wang et al. 1981, among cation of the Stratigraphy of China (43): Stratigraphy (Lithostratic)
others) and was known as the Nantuo Tillite for more than half a of Hunan Province. China University of Geosciences Press,
century. In fact, the Nantuo Fm. in the Yangtze Gorges area Wuhan, 292 (in Chinese).
(Zhang 1995), as well as in the Nanhua rift, was glaciomarine. Chu, X.-L., Zhang, Q.-R., Zhang, T.-G. & Feng, L.-J. 2003. Sulfur and
The predominant diamictite was likely subglacial and ice-proximal carbon isotopic variations in Neoproterozoic sedimentary rocks from
in an inner shelf environment, including sediment gravity flow and southern China. Progress in Natural Science, 13, 875–880.
rainout. In ice-distal environments, amalgamation of debris flow Condon, D., Zhu, M.-Y., Bowring, S., Wang, W., Yang, A. & Jin, Y.
deposits, rainout diamictite and disturbed deposition are dominant. 2005. U– Pb ages from the Neoproterozoic Doushantuo Formation,
The thin layer of lag deposits (Liu et al. 1999) likely indicates the China. Science, 308, 95– 98.
existence of a winnowing current. The inversely graded pebbly Dobrzinski, N. & Bahlburg, H. 2007. Sedimentology and environ-
mudstone at the transitional lower boundary was likely hemipela- mental significance of the Cryogenian successions of the Yangtze
gic, and the sandstone or sandy mudstone at the top was turbiditic Platform, South China block. Palaeogeography, Palaeoclimatology,
in origin; however, the waxing and waning stages are not Palaeoecology, 254, 100–122.
significant. Dobrzinski, N., Bahlburg, H., Strauss, H. & Zhang, Q.-R. 2004. Geo-
chemical climate proxies applied to the Neoproterozoic glacial suc-
Jiang et al. (2006) consider that the cap carbonate of the Doush-
cession on the Yangtze platform, south China. In: Jinkin, G. S.,
antuo Fm. shares morphological and petrographic attributes with McMenamin, M. A. S., McKay, C. P. & Sohl, L. (eds) The
modern and ancient methane seeps, as a result of post-glacial Extreme Proterozoic: Geology, Geochemistry and Climate. Ameri-
warming of the ocean, and is indicative of relatively deep-water can Geophysical Union, Washington DC, Geophysical Monograph,
deposition, most likely below storm wave base. 146, 13– 32.
Goddéris, Y., Donnadieu, Y. et al. 2007. Coupled modeling of global
carbon cycle and climate in the Neoproterozoic: links between
Comparison between the Jiangkou and the Nantuo glaciations Rodinia breakup and major glaciations. Comptes Rendus Geoscience,
339, 212– 222.
The differences between the two glaciations are obvious. The most Hao, J. & Zhai, M. 2004. Jinning movement and Sinian System in China:
conspicuous difference is in stratigraphic complexity. The Jiangkou their relationship with Rodinia Supercontinent. Chinese Journal of
glaciation can be divided into three stages, the waxing, the Geology, 39, 139–152 (in Chinese with English abstract).
maximum and the waning. The preservation of the Jiangkou Jiang, G.-Q., Kennedy, M. J. & Christie-Blick, N. 2003a. Stable
glacial units is highly limited in the southeastern marginal region isotopic evidence for methane seeps in Neoproterozoic postglacial
of the Yangtze block, because the complete succession is restricted cap carbonates. Nature, 426, 18– 25.
to the deep-water environment. The Nantuo glaciation was relatively Jiang, G. Q., Sohl, L. E. & Christie-Blick, N. 2003b. Neoproterozoic
simple, usually with negligible or no waxing and waning records. stratigraphic comparison of the Lesser Himalaya (India) and
THE YANGTZE REGION 365

Yangtze block (south China): Paleogeographic implications. Richthofen, F. von. 1882. China, 2. Ergebnisse Eigener Reisen und
Geology, 31, 917–920. Darauf Gegrundeter Studien, Berlin.
Jiang, G. Q., Kennedy, M. J., Christie-Blick, N., Wu, H.-C. & Zhang, Remane, J. 2000. International Stratigraphic Chart, with Explanatory
S.-H. 2006. Stratigraphy, sedimentary structures, and textures of the Note. Sponsored by ICS, IUGS and UNESCO (distributed at the
late Neoproterozoic Doushantuo cap carbonate in South China. 31st International Geological Congress, Rio de Janeiro 2000), 16.
Journal of Sedimentary Research, 76, 978–995. Roy, S. 2006. Sedimentary manganese metallogenesis in response to the
Jiang, G. Q., Kaufman, A. J., Christie-Blick, N., Zhang, S.-H. & evolution of the Earth system. Earth-Science Reviews, 77, 273– 305.
Wu, H.-C. 2007. Carbon isotope variability across the Ediacaran Sha, Q.-A., Liu, H.-Y., Zhang, S.-S. & Chen, M.-E. 1963. Nantuo tillite
Yangtze platform in South China: implications for a large in the eastern Yangtze Gorges area. Scientia Geologica Sinica, 3,
surface-to-deep ocean d13C gradient. Earth and Planetary Science 139– 148 (in Chinese).
Letters, 261, 303–320. Tang, J.-F., Fu, H.-Q. & Yu, Z.-Q. 1987. The stratigraphic horizon, type,
Kirschvink, J. L. 1992. A paleogeographic model for Vendian and and formation condition of Precambrian siliceous iron formation in
Cambrian time. In: Schopf, J. W. & Klein, C. (eds) The Proterozoic south China. Ore Geology, 6, 1 – 10 (in Chinese).
Biosphere: A Multidisciplinary Study. Cambridge University Press, Tang, S.-Y. & Liu, T.-B. 1999. Origin of the early Sinian Minle manga-
Cambridge, 569–581. nese deposit, Hunan Province, China. Ore Geology Reviews, 15,
Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blick, N. 71– 78 (in Chinese).
2004. A new period for the geologic time scale. Science, 305, 621–622. Wang, J. & Li, Z.-X. 2003. History of Neoproterozoic rift basins in South
Li, X.-H. & McCulloch, M. T. 1996. Secular variation in the Nd isotopic China: implications for Rodinia break-up. Precambrian Research,
composition of Neoproterozoic sediments from the southern margin 122, 141–158.
of the Yangtze Block: evidence for a Proterozoic continental collision Wang, Y.-L., Lu, S.-N., Gao, Z.-J., Lin, W.-X. & Ma, G.-G. 1981. Sinian
in southeast China. Precambrian Research, 76, 67– 76. tillites of China. In: Hambrey, M. J. & Harland, W. B. (eds)
Li, Z.-X., Zhang, L. & Powell, C. Mc. A. 1996. Positions of the east Earth’s Pre-Pleistocene Glacial Record. Cambridge University
Asian cratonic blocks in the Neoproterozoic supercontinent Press, 386–401.
Rodinia. Australian Journal of Earth Science, 43, 593–604. Wang, J., Li, X.-H., Duan, T., Liu, D., Song, B., Li, Z. & Gao, Y. 2003a.
Li, R.-W., Chen, J., Zhang, S., Lei, J., Shen, Y. & Chen, X. 1999. Zircon SHRIMP U– Pb dating for the Cangshuipu volcanic rocks and
Spatial and temporal variations in carbon and sulfur isotopic compo- its implications for the lower boundary age of the Nanhua strata in
sitions of Sinian sedimentary rocks in the Yangtze platform, South South China. Chinese Science Bulletin, 48, 1663–1669.
China. Precambrian Research, 97, 59– 75. Wang, T.-G., Wang, C.-J., Zhang, W.-B., Shi, Q., Zhu, L. & Chen,
Li, Z.-X., Li, X.-H., Kinny, P. D., Wang, J., Zhang, S. & Zhou, H. 2003. J.-Y. 2003b. Initial organic geochemical investigation on Late
Geochronology of Neoproterozoic syn-rift magmatism in the Yangtze Neoproterozoic– Early Cambrian sediments in the Yangtze region,
Craton, South China and correlations with other continents: evidence China. Progress in Natural Science, 13, 936– 941.
for a mantle superplume that broke up Rodinia. Precambrian Wang, J.-S., Gan, H.-Y., Wei, Q., Hu, G.-W. & Ge, Q. 2005. Stable
Research, 122, 85– 109. isotopes of carbon and sulfur of cap dolomite in the Three Gorges
Li, Z.-X., Bogdanova, S. V. et al. 2008. Assembly, configuration, and and its mechanism discussion. Geoscience, 19, 14 –20 (in Chinese
break-up history of Rodinia: a synthesis. Precambrian Research, with English abstract).
160, 179– 210. Wang, J.-S., Jiang, G.-Q., Xiao, S.-H., Li, Q. & Wei, Q. 2008a. Carbon
Liu, J.-S. 1994. Early Sinian sedimentary formations, turbidite and isotope evidence for widespread methane seeps in the c. 635 Ma
manganese ore genesis in Hunan Province. Geotectonica et Metallo- Doushantuo cap carbonate in south China. Geology, 36, 347– 350.
genia, 18, 173–182 (in Chinese). Wang, T.-G., Li, M.-J., Wang, C.-J., Wang, G.-L., Zhang, W.-B., Shi,
Liu, H.-Y., Dong, R.-S. et al. 1991. The Sinian System of China. Scien- Q. & Zhu, L. 2008b. Organic molecular evidence in the Late
tific Press, Beijing (in Chinese). Neoproterozoic Tillites for a palaeo-oceanic environment during
Liu, H.-Y., Hao, J. & Li, Y.-J. 1999. The Late Precambrian Stratigraphy the snowball Earth era in the Yangtze region, southern China.
and Geological Evolution of Middle-Eastern China. Science Press, Precambrian Research, 162, 317– 326.
Beijing. Willis, B., Blackwelder, E. & Sargent, R. H. 1907. Research in
Liu, T.-B., Maynard, J. B. & Alten, J. 2006. Superheavy S isotopes China. Vol. 1 part 1. The Carnegie Institution of Washington,
from glacier-associated sediments of the Neoproterozoic of south Washington, DC.
China: oceanic anoxia or sulfate limitation? In: Kesler, S. E. & Xie, Q.-L., Chen, D.-F. & Chen, X.-P. 1999. Characteristics of sedimen-
Ohmoto, H. (eds) Evolution of Early Earth’s Atmosphere, Hydro- tary organic matter in Songtao manganese deposits, Guizhou. Acta
sphere – Constraints from Ore Deposits. Geological Society of Sedimentologica Sinica, 17, 280– 284.
America Memoir, 198, 205–222. Yang, Y.-J. 1985. The glacigenous strata of lower Sinian in Hunan
Lu, S.-N., Ma, G.-G., Gao, Z. J. & Lin, W.-X. 1985. Primary research on Province. In: PRECAMBRIAN GEOLOGY EDITORIAL COMMIT-
glacigenous rocks of Late Precambrian in China. In: PRECAM- TEE (ed.) Precambrian Geology No. 1: The Collected Works of Late
BRIAN GEOLOGY EDITORIAL COMMITTEE (ed.) Precambrian Precambrian Glacigenous Rocks in China. Geological Publishing
Geology No. 1, The Collected Works of Late Precambrian Glacigen- House, Beijing, 224–241 (in Chinese with English abstract).
ous Rocks in China. Geological Publishing House, Beijing, 1 –86 (in Yang, S.-X. & Pang, K.-T. 2006. Mineralization model for the manganese
Chinese with English abstract). deposits in northwestern Hunan – an example from Minle manganese
Ma, G.-G. & Wang, Y.-G. 1983. Discussion on the glacial geology of the deposit in Huayuan, Hunan. Sedimentary Geology and Tethyan
early Sinian in the area adjacent to Hubei, Hunan, Sichuan and Geology, 26, 72 –80.
Guizhou Provinces of China. Bulletin of Yichang Institute of Yang, J.-D., Xue, Y.-S., Sun, W.-G., Tao, X.-C., Wang, Z.-Z. & Zhou,
Geology and Mineral Resources, Chinese Academy of Geological C.-M. 1997. The origin and age of the manganese ore of Nantuo stage.
Sciences, 7, 43– 51 (in Chinese with English abstract). Chinese Science Bulletin, 42, 1538– 1541.
Meng, F.-W., Yuan, X. L., Zhou, C.-M. & Chen, Z.-L. 2003. Dinoster- Yang, R.-D., Ouyang, Z.-Y., Zhu, L.-J., Wang, S.-J., Jiang, L.-J.,
ane from the Neoproterozoic Datangpo black shales and its biological Zhang, W.-H. & Gao, H. 2002. A new understanding of manganese
implications. Acta Micropalaeontologica Sinica, 20, 97 –102. carbonate deposits in early Sinian Datangpo stage. Acta Mineralogica
NCS (NATIONAL COMMITTEE ON STRATIGRAPHY) 2001. Strati- Sinica, 22, 329– 336.
graphic Guide in China and Explanation (revised). Geological Pub- Yin, C.-Y., Liu, D.-Y., Gao, L.-Z., Wang, Z.-Q., Xing, Y.-S., Jian, P. &
lishing House, Beijing, 59 (in Chinese). Shi, Y.-R. 2003. Lower boundary age of the Nanhua System and the
Ogg, J. G. & Gradstein, F. M. 2008. The Concise Geologic Time Scale. Gucheng glacial stage: evidence from SHRIMP dating. Chinese
Cambridge University Press, Cambridge. Science Bulletin, 48, 1657– 1662.
Qin, Y., Zhou, Q. & Zhang, S. 2005. Elementary properties of manga- Zhang, Q.-R. 1995. The origin of the Sinian Nantuo Formation in
nese of Nanhuan in the Northeastern Guizhou. Guizhou Geology, Yichang County, Hubei Province. Scientia Geologica Sinica, 30,
122, 246– 251. 147– 152 (in Chinese with English abstract).
366 Q. R. ZHANG ET AL.

Zhang, Q. R. & Piper, J. D. A. 1997. Palaeomagnetic study of Zhou, C.-M. & Xiao, S.-H. 2007. Ediacaran d13C chemostratigraphy of
Neoproterozoic glacial rocks of the Yangzi Block: palaeolatitude South China. Chemical Geology, 237, 89– 108.
and configuration of South China in the late Proterozoic Superconti- Zhou, C.-M., Tucker, R., Xiao, S.-H., Peng, Z.-X., Yuan, X.-L. &
nent. Precambrian Research, 85, 173– 199. Chen, Z. 2004. New constraints on the ages of Neoproterozoic
Zhang, Q.-R., Chu, X.-L., Bahlburg, H., Feng, L.-J., Dobrzinski, N. glaciations in south China. Geology, 32, 437– 440.
& Zhang, T.-G. 2003. Stratigraphic architecture of the Neoproter- Zhou, J.-B., Li, X.-H., Ge, W.-C. & Li, Z.-X. 2007a. Age and origin of
ozoic glacial rocks in the ‘Xiang-Qian-Gui’ region of the central middle Neoproterozoic mafic magmatism in southern Yangtze
Yangtze Block, South China. Progress in Natural Science, 13, Block and relevance to the break-up of Rodinia. Gondwana Research,
783– 787. 12, 184– 197.
Zhang, Q.-R., Li, X.-H., Feng, L.-J., Huang, J. & Song, B. 2008a. A Zhou, Q., Du, Y.-S., Yan, J.-X., Zhang, M.-Q. & Yin, S.-L. 2007b.
new age constraint on the onset of the Neoproterozoic glaciations Geological and geochemical characteristics of the cold seep
in the Yangtze Platform, South China. Journal of Geology, 116, carbonates in the Early Nanhua System in Datangpo, Songtao,
423– 429. Guizhou Province. Earth Science – Journal of China University of
Zhang, S.-H., Jiang, G.-Q. & Han, Y.-G. 2008b. The age of the Nantuo Geosciences, 32, 845–852 (in Chinese with English abstract).
Formation and Nantuo glaciation. Terra Nova, 20, 289– 294. Zhu, M.-Y., Zhang, J.-M. & Yang, A.-H. 2007. Integrated Ediacaran
Zheng, Y.-F. 2003. Neoproterozoic magmatic activities and global (Sinian) chronostratigraphy of South China. Palaeogeography,
evolution. Kexue Tongbao, 48, 1705–1720 (in Chinese). Palaeoclimatology, Palaeoecology, 254, 7 – 61.
Chapter 33

Neoproterozoic glaciogenic diamictites of the Tarim Block, NW China

MAOYAN ZHU* & HAIFENG WANG


State Key Laboratory of Palaeobiology and Stratigraphy, Nanjing Institute of Geology and Palaeontology,
Chinese Academy of Sciences, Nanjing, 210008, China
*Corresponding author (e-mail: myzhu@nigpas.ac.cn)

Abstract: Neoproterozoic glaciogenic outcrops are mostly limited to the southwestern and northwestern margins of the Tarim basin and
the Tianshan Mountains. Only two glaciogenic diamictites are recorded in the Neoproterozoic successions from the Tielikeli and Aksu-
Wusi areas of the Tarim basin. The lower diamictite from the Polong Formation in the Tielikeli area is c. 800 m thick. The upper dia-
mictite in the area is very thin (Yutang diamictite, 10 m; Yulmeinak diamictite, 51 m). In contrast to the Tarim basin, the Neoproterozoic
successions along the northern margin of the Tarim Block in the Tianshan Mountains record three or four glaciogenic diamictites, which
are well developed and crop out in the Qurugtagh area. Although the glaciogenic nature of the oldest Beiyixi diamictite remains ques-
tionable, all other diamictites (Altungol, Tereeken and Hankalchough) exhibit distinct glaciogenic sedimentary features. SHRIMP zircon
U –Pb dating demonstrate that the Bayisi diamictite was deposited between 740+7 Ma and 725+10 Ma, the Altungol and Tereeken
diamictites between 725+10 Ma and 615+6 Ma, respectively, and the Hankalchough diamictite between 615+6 Ma and c. 542 Ma.
Carbonates atop the Altungol, Tereeken and Hankalchough diamictites show distinct C-isotopic values that are typical for those recorded
in ‘cap carbonates’ on other continents. The striking similarity between the cap carbonate of the Tereeken and Marinoan-age Nantuo
diamictites further support a 635 Ma ‘Marinoan’ age for the Tereeken glaciation. Unlike the 582 Ma Gaskiers glaciations of Newfound-
land, the Hankalchough diamictite is possibly ,551 Ma, as suggested by C-isotope chemostratigraphy and biostratigraphy, indicating
that post-Marinoan glaciations on different continents may be diachronous.

Supplementary material: Data are available at: http://www.geolsoc.org.uk/SUP18469.

Neoproterozoic glaciogenic diamictites on the Tarim Block were unconformably overly the Sukulok Group. Interbedded within
first recognized by a Swedish geologist, E. Norin, following a the massive conglomerate unit are lenticular or bedded feldspathic
joint Chinese –Swedish expedition 1927–1935 in the Qurugtagh sandstones and thinly bedded silty mudstone. The conglomerate is
area (Norin 1937). Although several detailed studies have been dominated by sub-rounded or rounded, well-sorted clasts of chert
carried out since and three Neoproterozoic glaciogenic diamictites and porphyry.
had already been reported in the early 1980s (see Gao et al. 1985a The Yalaguz Fm. is overlain by the Qakmaklik Group which
and references therein), these diamictites only recently received consists in ascending order of the Polong, Kilix and Yutang for-
wide attention (Xiao et al. 2004). Various research groups have mations. The Polong Fm. is dominated by purplish-brown
now investigated the Neoproterozoic successions of the Tarim massive diamictite of c. 880 m thick. The diamictite is underlain
basin margin, particularly in the Qurugtagh region, focusing on by c. 200 m of grey laminated cherts and silicified shale. The
the study of Neoproterozoic tectonics, magmatism, lithologies, Kilix Fm. consists of interbeds of purplish-green laminated mud-
geochronology, chemostratigraphy and palaeontology. stone and siltstone with thin beds of silty micritic limestone (c.
The present chapter aims to review both old and recent studies, 312 m) in the lower part, purple quartz sandstone (42.9 m) in the
focusing on the stratigraphy of the Neoproterozoic diamictites in middle, and purple interbeds of feldspathic sandstone and con-
four palaeogeographical areas located on or close to the Tarim glomerate (86.8 m) at the top. The feldspathic sandstone exhibits
Block (Fig. 33.1). Using detailed descriptions and recent geochro- large-scale cross-stratification and horizontal stratification. The
nological and chemostratigraphic data, this chapter discusses cor- Yutang Fm. is subdivided into three parts. The lower part (c.
relations and ages of the Neoproterozoic diamictites of the Tarim 50 m) consists of grey-greenish laminated muddy cherts and mud-
Block. Published palaeomagnetic data and their palaeogeographic stone with a conglomerate 0.1–0.5 m thick at the base. The middle
implications are also reviewed. part (c. 10 m) consists of purplish brown diamictite. The upper part
(121.9 m) of the Yutang Fm. consists of reddish and brownish
feldspathic sandstone and conglomerate showing large-scale
Glaciogenic diamictites in the Tielikeli area, cross-stratification and wave ripples.
SW Tarim Basin Overlying the Qakmaklik Group is the Kurkak Fm., which is
composed of four parts. A 10-m-thick laminated dolomite marks
The Tielikeli area is located at the southwestern margin of the the contact between the Qakmaklik Group and the Kurkak Fm.
Tarim basin, where Neoproterozoic outcrops are well exposed The lower part of the Kurkak Fm. is c. 110 m in thickness and
along the northern fringes of the western Kunlun Mountains consists of black shale. This is succeeded by 100 m of grey
(Fig. 33.1(1) & Fig. 33.2) (Ma et al. 1991; Wang et al. 2004). quartz sandstone and conglomerate in the middle, and dark grey
The Neoproterozoic strata and tillites of the area, which had orig- shale with interbeds of siltstone and rare phosphorite nodules/
inally been mapped as part of the Devonian Qizlafu Group, were concretions or lamina in the upper part of this unit.
first recognized by the local geologist Ma and his colleagues in The Kzisuhum Fm. is composed of four parts in ascending
1978 and described in detail a decade later (Ma et al. 1989). order. The lower interval is 55.6 m thick and consists of dark
green quartz sandstone with thin interbeds of feldspathic sandstone
and pyrite-bearing siltstone in the lower part, and dark grey silt-
Stratigraphy stone, greyish-green sandstone and grey coarse-grain quartz sand-
stone with three phosphorite beds in the upper part. The overlying
The Cryogenian and Ediacaran successions in the area (Fig. 33.3) interval comprises 86.1 m of yellowish brown sandy dolostone
start with c. 230 m of conglomerate (Yalaguz Fm.), which with interbeds of purple siltstone and minor sandy siderite

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 367– 378. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.33
368 M. ZHU & H. WANG

Fig. 33.1. Geological map of the Tarim Block and adjacent areas (modified from Lu et al. 2008). Key: PH, Phanerozoic rocks; NP, Neoproterozoic rocks; MP,
Mesoproterozoic rocks; PP, Palaeoproterozoic rocks; AR, Archaean rocks; RNP, early Neoproterozoic granitoids; F, Faults; IF, Inferred faults; Diamictites, Cryogenian
and Ediacaran diamictites; Q, Quaternary desert and sedimentary deposits; Square: 1, Tielikeli area, SW Tarim Basin; 2, Aksu-Wusi area, NW Tarim Basin; 3, Quruqtagh
area, southern margin of Eastern Tianshan Mts.; 4, Guozigou-Keguqing Mt. area, western range of Northern Tianshan Mountains.

layers. Approximately 52.2 m quartz sandstone with large-scale dolostone, granite, gneisses, diabase and so on. Clasts are domi-
planar cross-stratifications overlies the dolostone interval. The nated by pebbles of 2 –10 cm in diameter, with a maximum diam-
upper part ofthe Kzisuhum Fm. is dominated by dolostone eter of 100 cm. Glaciogenic features include striated marks, snicks
(177.4 m). The contact to the overlying Devonian strata is sharp and notches on the surface of cobble and pebble size clasts, and
and unconformable. parallel cracks of clasts. The diamictite at the base and top of the
interval is characterized by 30–50% poorly sorted clasts with
traces of bedding. Laminated mudstone and siltstone with drop-
Description of the diamictites stones are observed in the middle and top of the interval.
The upper diamictite interval of the Yutang Fm. is only 10 m
There are two diamictite intervals in the Tielikeli area. Both are thick, showing similar sedimentary features and a similar clast
interpreted as glaciogenic diamictites or tillites. The lower diamic- composition as the lower diamictite. Similar dropstone-bearing
tite interval of the Polong Fm. is dominated by massive diamictite laminated mudstone also occurs at the top of the upper diamictite.
with 5 –30% randomly organized, angular, sub-angular or It is interesting to note that both diamictite intervals are
sub-rounded clasts of porphyry, quartz porphyry, chert, quartzite, underlain by an interval consisting of grey laminated muddy

Fig. 33.2. Geological map showing


outcrop of the Tielikeli area, SW Tarim
Basin (modified from Ma et al. 1989).
THE TARIM BLOCK, NW CHINA 369

Stratigraphy

The Neoproterozoic succession in the area (Fig. 33.5) consists of


the Qiaoenblaq and Wushinanshan groups; the former overlies
the metamorphic rocks of the Aksu Group unconformably. The
Aksu Group is older than 962 + 12 Ma (whole-rock Rb –Sr
dating) and is dominated by blueschist (Liou et al. 1989;
Gao et al. 1993). The Qiaoenblaq Group, overlying the Aksu
blueschists, consists of (in ascending order) the Xifangshan, Dong-
qiaoenblaq, Muyangtan and Dongwu formations. The Xifangshan
Fm. is c. 1715 m thick and is characterized by rhythmic beds of
grey-green feldspathic sandstone, feldspathic quartz sandstone
and siltstone. The rhythmites exhibit sandstone with grading fea-
tures in the lower part, and siltstone with horizontal and cross-
stratification in the upper part. The rhythmic beds range from
1 cm to 4 m in thickness, but are predominantly 50– 100 cm.
They have been interpreted as gravity flow deposits or proximal
turbidites (Gao et al. 1993).
The Dongqiaoenblaq Fm. consists of 15 –20-m-thick grey-green
conglomerate with interbeds of sandstone in the lower and massive
diamictite with a maximum thickness of 70 m in the upper part.
Overlying the diamictite is the Muyangtan Fm., which is
c. 107 m thick and is composed of grey-green calcareous siltstone
and quartz sandstone with parallel and cross-stratification. The
overlying Dongwu Fm., c. 176 m thick, is composed of grey-green
and grey-brownish coarse-grained feldspathic sandstone. There is
a 1– 2-m-thick conglomerate bed at the base of the Dongwu Fm.
The conglomerate is dominated by sub-rounded or rounded
cobbles and pebbles of granite.
The Wushinanshan Group consists of the Yulmeinak, Sugetblaq
and Qigeblaq formations. The Yulmeinak Fm. overlies sandstone
of the Dongwu Fm. above an angular unconformable contact and
is composed of purplish massive diamictite with a maximum thick-
ness of 70 m, but the diamictite gradually decreases and eventually
pinches out towards the east. The overlying Sugetblaq Fm. consists
of two members. The lower member is c. 728 m thick is and com-
posed of brownish coarse-grained feldspathic sandstone with inter-
beds of yellow-green siltstone and sandstone in the lower part, and
red, grey and brownish ferrous quartz sandstone with thin interbeds
of hematite and a 15-m-thick unit of peridotite basalt in the upper
part. The uppermost part of the lower member of the Sugetblaq
Fm. is characterized by 79 m of dark green and brownish amygda-
loidal olivine basalt. The upper member of the Sugetblaq Fm. is
composed of brownish and greyish thin-bedded sandy limestone
with edgewise clasts and interbedded thin siltstone in the lower
part and yellow to green, thin-bedded, glauconitic feldspathic
sandstone in the upper part. The overlying Qigeblaq Fm. is a
characteristic interval of thick-bedded dolostone with a thickness
of 186 m. This unit is overlain by early Cambrian cherts and phos-
phorite, which contain small shelly fossils (Gao et al. 1981; Qian &
Xiao 1984; Yao et al. 2005).

Fig. 33.3. Composite stratigraphic log of the Cryogenian and Ediacaran in the Description of the diamictites
Tielikeli area, SW Tarim Basin (modified from Ma et al. 1989). Key as in
Figure 33.10. As in the Tielikeli area, there are two diamictite intervals inter-
preted as glaciogenic in the Aksu-Wushi area. The lower diamic-
cherts and silicified shale, demonstrating a similar depositional tite interval of the Dongqiaoenblaq Fm. is composed of 20–80%
sequence. randomly organized sub-rounded clasts of felsite, dacite, granite,
jasper, quartzite, sandstone, andesite, diabase and so on. Clasts
are dominated by c. 10-cm-sized pebbles with a maximum diam-
Glaciogenic diamictites in the Aksu-Wusi area, eter of 60 cm. The matrix is composed of feldspathic sandstone
NW Tarim Basin and siltstone with a mud component, showing soft deformation
features and contorted bedding in some beds. Although clasts are
Neoproterozoic outcrops exposed in the Aksu-Wusi area, NW randomly organized, and striated marks have been observed on
Tarim basin, are easily accessible (Figs 33.1, 33.2 & 33.4). Glacio- the surface of some clasts from the diamictite of the Dongqiaoen-
genic diamictite in the area was first documented in local geologi- blaq Fm., the high degree of sorting, contorted bedding and sub-
cal reports during the 1950s and has been well studied by Gao & rounded shapes of the clasts, which are similar in lithology to
Qian (1985) and Gao & Chen (1993). the clasts from the underlying conglomerate, led Gao et al.
370 M. ZHU & H. WANG

Fig. 33.4. Geological map showing the


Cryogenian and Ediacaran outcrop of the
Aksu-Wusi area, NW Tarim Basin
(modified from Gao et al. 1985b).

(1993) to reinterpret the diamictite as having been rapidly depos- c. 800-m-thick carbonate unit, the Paergangtagh Fm., with an
ited from submarine debris flows. unconformable contact (Zhao et al. 1985; Gao et al. 1993). The
The upper diamictite interval of the Yulmeinak Fm. is a discon- Quruqtagh Group is subdivided (in ascending order) into nine
tinuous interval with an unconformable basal surface. The diamic- formations: the Bayixi, Zhaobishan, Altungol, Huangyanggou,
tite shows distinctly glaciogenic characteristics, including a large Tereeken, Zhamoketi, Yukkengol, Shuiquan and Hankalchough
abraded surface at its base with notching, scouring and striations, formations (Gao & Zhu 1984; Xiao et al. 2004; Kou et al. 2008).
which is a clear indication of ice movement or glacial pavement. The best continuous outcrops are located in the Middle
The diamictite is massive and matrix-supported with 20– 30% Quruqtagh area.
clasts of various lithologies including granites and volcanic rocks,
blueschists, sandstone and siltstone originally derived from the
underlying strata. The clasts vary in size, with the largest ones at The Bayixi Fm. This consists predominantly of shallow-water
the base (maximum diameter of 6 m), and show a poor degree of marine siliciclastic and volcanic rocks with several intervals of
sorting. Dominant clasts are subangular with a muddy or calcar- diamictite. The overall thickness of the formation ranges from
eous envelope, but characteristic selliform (saddle-shaped) and 640 to 1670 m. It starts with a basal conglomeratic interval with
other irregularly shaped clasts are also observed. Striation marks, predominant clasts of dolostone and limestone from the underlying
nicks and notches are common on the surface of clasts. Dropstones Paergangtagh Fm. The clasts are angular to subangular in shape,
of various lithologies occur in laminated siltstone in the middle with the largest clasts being 30 cm in diameter. The lower part
of the diamictite interval. of the Bayixi Fm. consists of volcanic rocks, slates and sandstones
associated with diamictite. The middle part of the formation con-
sists of sandstone, and siltstone interbedded with diamictite and
Glaciogenic diamictites in the Quruqtagh area, southern slates. The upper part of the formation consists of two diamictite
margin of the eastern Tianshan Mountains intervals of c. 200 m thickness each, separated by interbedded
slates and siltstones.
The Quruqtagh area is located at the northeastern margin of the
Tarim basin and belongs to the southern margin of the Eastern
Tianshan, covering an area of about 30 000 km2 (Figs 33.1– The Zhaobishan Fm. This is a characteristic interval with thick-
33.3). The Quruqtagh area is customarily subdivided into three dis- nesses ranging from 358 to 570 m, consisting predominantly of
tricts. The area west of Xindi is called the East Quruqtagh, the area sandstones and calcareous siltstones in the lower part and slates
between Xindi and Xinger is the Middle Quruqtagh, while the area with interbedded thin calcareous siltstones and sandstones in the
east of Xinger is the West Quruqtagh. The Neoproterozoic out- upper part. The sandstones in the lower part of the formation are
crops and glaciogenic diamictites are well exposed in all three dis- lithic, coarser-grained (0.2 –2 mm), better rounded and sorted
tricts (Fig. 33.6) and were first reported by a Swedish geologist, and show large cross-bedding, whereas sandstones in the upper
E. Norin, after a joint 1927 –1935 Chinese – Swedish expedition part are more fine-grained (0.1 –1 mm) and show a lower degree
in the area (Norin 1937). The Neoproterozoic stratigraphy of the of textural maturity. The overall sequence of the Zhaobishan
area has been well established by local geological surveys during Fm. represents a deepening-upward sequence.
the 1950s and has been documented in detail by Gao & Zhu
(1984). Since then, the latest Neoproterozoic successions have The Altungol Fm. This is characterized by three massive diamictite
been investigated using various methods (e.g. Zhong & Hao intervals associated with sandstones, siltstones and shales. The
1990; Xiao et al. 2004; Duan et al. 2005; Huang et al. 2005; He thickness of the formation increases from East Quruqtagh
et al. 2007a; Liu et al. 2007; Zhang et al. 2007; Shen et al. (764 m) to West Quruqtagh (2395 m). In the South Quruqtagh
2008; Kou et al. 2008; Zhu et al. 2008; Xu et al. 2009). area, the formation is only 8 m thick. The basal and upper
massive diamictites mark the basal and top boundaries of the
Stratigraphy Altungol Fm. The thickness of the three diamictite intervals
ranges from 50 to 100 m. Associated sandstones, siltstones and
The latest Neoproterozoic successions in the Quruqtagh area shales exhibit rhythmic beddings with average bed thicknesses
(Fig. 33.7) are known as the Quruqtagh Group, and overlie a of 40–60 cm.
THE TARIM BLOCK, NW CHINA 371

The Huangyanggou Fm. This is a newly erected formation repre-


senting an interval between the uppermost diamictite of the Altun-
gol Fm. and the diamictite of the Tereeken Fm. (Cao 1991; Kou
et al. 2008). The interval without diamictite was originally the
upper part of the Altungol Fm., and was interpreted to represent
an interglacial interval between the Altungol and Tereeken glacia-
tions (Kou et al. 2008). The base of the Huangyanggou Fm. is
marked by c. 2 m of grey-purplish massive micritic dolostone,
which represents a cap carbonate above the Altungol diamictite.
The lower part of the formation consists of meta-andesite, clinker-
ing breccia and silty tuff, while the upper part consists of silty mud-
stone and muddy siltstones with thin interbeds or lenses of
limestone. A 1–3-m-thick limestone unit marks the top boundary
of the Huangyanggou Fm., underlying the base of the Tereeken
Fm. A c. 50-m-thick dolostone interval reported from the Altungol
Fm. by Xiao et al. (2004) at the Xishankou section may correspond
to the Huangyanggou Fm.

The Tereeken Fm. This is characterized by grey massive diamictites


with thicknesses ranging from 689 to 1845 m. The diamictite is
separated into several units by thin interbeds of finely laminated
silty mudstones and carbonate.

The Zhamoketi Fm. This is a siliciclastic interval consisting of


metre-scale rhythmites of sandstone, siltstone and mudstone,
with thicknesses ranging from 297 m to 793 m. The base of the for-
mation is marked by a 2 –10-m-thick unit of dolostone with char-
acteristics of basal Ediacaran cap carbonates, overlying the
Tereeken diamictites. The top of this formation is marked by a c.
20– 80-m-thick diabase sill. The rhythmites have been interpreted
as turbidites, showing distinct structures of the Bouma sequences
deposited on a continental slope (Li & Dong 1991; Kou et al.
2008). The rhythmites exhibit grading at their base with overlying
laminated and wavy or cross-bedded intervals; convoluted lami-
nations or soft-deformation have been observed within some
beds. Sole marks such as flute casts, groove casts and load casts
are common at the base of the rhythmites.

The Yukkengol Fm. This consists of green-greyish finely laminated


siltstones and shales with thicknesses ranging from 80 to 583 m,
representing condensed sequences of deeper- water facies.

The Shuiquan Fm. This is characterized by laminated carbonates


and silty shales. Its variable thicknesses between 22 and 307 m
was interpreted to result from erosion at the top of the formation
(Gao & Zhu 1984). Centimetre-scale ribbonites, stromatolites
and microbially laminated dolostone were reported in the Shui-
quan Fm., suggesting sedimentation in the euphotic zone (Xiao
et al. 2004). The wormiform carbonaceous fossils – filaments c.
2 mm wide and up to 40 mm long – from the Shuiquan Fm.
reported by Gao et al. (1985) were reinterpreted as Vendotaenid
fossils (Xiao et al. 2004).

The Hankalchough Fm. This consists of light grey diamictite. Its


thickness decreases eastward from 467 m to less than 10 m. A
dolostone unit (1 –6 m thick) at the base of the Xishanblaq Fm.
marks the top of the Hankalchough Fm.
The Xishanblaq Fm. consists of black cherts and cherty phos-
phorites associated with volcanic rocks. Cambrian microfossils
have been reported from the phosphorites (Yao et al. 2005).

Description of the diamictites

The four major intervals of diamictite in the Quruqtagh Group


have been assigned to three glaciations (Gao & Zhu 1984).
Below is a summary of these four diamictites based on detailed
Fig. 33.5. Composite stratigraphic log of the Cryogenian and Ediacaran in the descriptions by Gao & Zhu (1984), Xiao et al. (2004) and Kou
Aksu-Wusi area, NW Tarim Basin (after Gao et al. 2005b). Key as in
et al. (2008).
Figure 33.10.
372 M. ZHU & H. WANG

Fig. 33.6. Geological map showing the


Cryogenian and Ediacaran outcrop of the
Quruqtagh area, southern margin of the
Eastern Tianshan Mountains (modified
from Gao & Zhu 1984).

The Bayixi diamictite. The Bayixi diamictites are distributed across The Tereeken diamictite. The Tereeken diamictite is the most con-
only a limited area largely near the Middle Quruqtagh. The thick- tinuous and distinct interval in the Quruqtagh area, showing unam-
ness of the diamictite units in the Bayixi Fm. varies from a few biguous glaciogenic features, striated clasts and dropstones
metres to more than 100 m. Poorly sorted and angular clasts (5 – throughout. Clasts in the diamictites are dominated by carbonate,
20%) occur in the silty and muddy matrix of the diamictites (see granite and other igneous rocks; marble and siliciclastic clasts
Supplementary Material, Fig. 1S-a). Clasts are mostly derived are more abundant in comparison to the Bayixi and Altungol dia-
from older metamorphic and igneous rocks, including gneiss, mictites (see Supplementary Material, Fig. 1S-c). Striated clasts
granite, diabase, dolostones or marbles, and quartz; siliciclastic are common (Fig. 1S-d). Three to eight individual diamictite
clasts are rare. Clast sizes are predominantly ,6 cm in diameter. units can be recognized within the Tereeken Fm. (Gao & Zhu
Some boulders are perpendicularly or obliquely oriented relative 1984). These units are intercalated with non-glaciogenic, lami-
to bedding plane. The Bayixi diamictites were interpreted as gla- nated silty marine mudstones and carbonates. Several thin carbon-
ciomarine in origin by Gao & Zhu (1984), supported by the vari- ate laminae (,10 mm thick and laterally continuous for more than
ety of lithic compositions, the angular and irregular shapes of the 100 m) occur within the silty mudstone in the lower Tereeken Fm.
clasts, scratch marks and non-unidirectional striations on the at the Heishan –Zhaobishan section (Xiao et al. 2004). These
clast surfaces. However, glacial striations were not confirmed by laminae consist of vertically oriented, upward-growing calcite
Xiao et al. (2004). In addition, no cap carbonate has been reported. crystals, and are typically inundated by overlying siltstone. In
addition, a bedded carbonate unit occurs in the lower Tereeken
The Altungol diamictites. Based on Kou et al. (2008), there are three Fm. between two massive diamictite units, at the same section.
diamictite units within the newly defined Altungol Fm. However, As suggested by Xiao et al. (2004), the existence of laminated silt-
as documented by Gao & Zhu (1984), the Altungol diamictite units stones and bedded dolostones between Tereeken diamictite units
appear unevenly distributed because only one unit has been indicates that there was an active hydrological system during the
observed in some sections. The Altungol diamictites are massive, Tereeken ice age.
have a matrix of low maturity and include clasts of quartzite, A 2–10-m-thick carbonate unit above the Tereeken diamictites
granite, cherts and quartz. The clasts are generally .1 cm with represents a distinct lithostratigraphic marker, and was recognized
the largest being c. 15 cm in diameter (see Supplementary as the ‘Yukkengol limestone’, defining the boundary between the
Material, Fig. 1S-b). Clasts with stepped fractures and glaciogenic Tereeken diamictite and overlying strata (Norin 1937). The same
striations have been observed (Kou et al. 2008). Gao & Zhu (1984) boundary was subsequently adapted to define the boundary
claimed that the Altungol diamictites have a closer affinity with the between the Lower and Upper Sinian Systems (Gao & Zhu
overlying Tereeken diamictites and argued that the Altungol and 1984). The carbonate unit in the Yukkengol area has been
Tereeken diamictites jointly represent a mid-Quruqtagh ice age. described in detail as a cap carbonate atop the Tereeken diamictite
However, Xiao et al. (2004) suggest the alternative interpretation by Xiao et al. (2004). According to observations from a section
that the Altungol diamictites resemble more closely the Bayixi dia- between Xidashan and Moheshan, similar to the Yukkengol area,
mictites in clast composition, degree of metamorphism and thick- the carbonate unit consists of three subunits (see Supplementary
ness of individual diamictite beds. Xiao et al. (2004) argued that if Material, Fig. 1S-e, f ). The lower subunit (c. 1.5 m) is a massive
all the diamictites in the Bayixi and Altungol formations were dolostone with angular allochthonous clasts of various compo-
found to be glaciogenic, it would be more likely that the Altungol sitions. The middle subunit (c. 2.5 m) consists of thinly bedded,
and Bayixi diamictites represent pulses of a single ice age. Xiao finely laminated dolostone with thin interbeds of shales. The
et al.’s (2004) hypothesis is further supported by the observations upper subunit (c. 1 m thick) consists of medium-bedded dolostone.
of Kou et al. (2008) who confirmed the existence of an interglacial It is interesting to note that the uppermost part of the Tereeken
interval, the Huangyanggou Fm., separating the Altungol diamic- diamictite underlying the cap carbonate shows well-defined
tites from the Tereeken ice age. A 2 –4-m-thick, grey-purplish car- bedding and represents deposits of a deglaciation interval.
bonate unit with negative values of d13C at the base of the
Huangyanggou Fm. was considered to be a cap carbonate above The Hankalchough diamictite. The Hankalchough diamictite is
the Altungol diamictites (Kou et al. 2008). However, in opposition composed of light grey and greenish-grey massive diamictite
to Xiao et al.’s (2004) hypothesis, Kou et al. (2008) assigned the (see Supplementary Material, Fig. 2S-a, b, c). The clast compo-
Altungol diamictites to a distinctly separate ice age. sition differs from that of the underlying three diamictite intervals,
THE TARIM BLOCK, NW CHINA 373

consisting of angular clasts dominated by carbonate and granite.


Siliciclastic clasts from underlying strata are abundant. Some car-
bonate boulders contain overturned stromatolite, probably derived
from older stromatolitic dolostones in the Paergangtagh Group,
which underlies the Quruqtagh Group (Zhao et al. 1985). The Han-
kalchough diamictite was originally considered to be terrestrial in
origin (Gao & Zhu 1984), but has been reinterpreted as glaciomar-
ine by Xiao et al. (2004) due to the occurrence of dropstones.
A distinct 1– 5-m-thick grey-greenish or light grey, finely lami-
nated muddy-silty dolostone or calcareous mudstone unit above
the Hankalchough diamictite (see Supplementary Material,
Fig. 2S-d) contains common lithic fragments and small clasts
(3 –5%) with disseminated pyrite granules (5 –10 mm). It was orig-
inally interpreted as glaciolacustrine in origin (Gao & Zhu 1984)
and reinterpreted as an atypical type of cap dolostone (Xiao
et al. 2004).

Glaciogenic diamictites in the Guozigou-Keguqingshan


area, western range of North Tianshan Mountains

The Guozigou-Keguqingshan area belongs to the western range of


the North Tianshan Mts. (Figs 33.1– 33.4). The well-exposed late
Precambrian successions in the area (Fig. 33.8) is composed of
three intervals of diamictite, which were first recognized by local
geologists in 1976 and subsequently documented in detail by
Wang et al. (1983) and Gao et al. (1985b) based on fieldwork in
1978 –1979. Since then, no more new information has been
published.

Stratigraphy

The Cryogenian and Ediacaran Kailaketik Group in the area


(Fig. 33.9) is c. 700–1000 m thick and is subdivided into six
formations (in ascending order): the Kulutieliekti, Tulasu,
Biexibastao, Keyindi, Tarqat and Talisayi formations. The
outcrop quality and thicknesses of the succession in the area
vary from west to east. The lower part of the succession is well
exposed in the Kuguqingshan area and the upper part of the succes-
sion is well developed in the Guozigou area. The succession is
overlain disconformably by the Lingkuanggou Fm., which con-
tains early Cambrian phosphatized small shelly fossils, and is
underlain by the carbonate of the Kusongmqiek Group with
unconformable contact.

The Kulutieliekti Fm. This is c. 115 m thick. A basal conglomerate


interval of 7.2 m is composed predominantly of limestone and
dolostone clasts derived from the underlying Kusongmqiek
Group. The well-sorted clasts have rounded or sub-rounded shapes
with diameters ranging from 1 to 25 cm. The basal conglomerate is
overlain by a 9-m-thick greenish tuff bed. The tuffaceous bed con-
tains c. 30% volcanic clasts (0.03 –0.3 cm in diameter) and carbon-
ate clasts (0.5 –1.5 cm in diameter). The subsequent strata are
composed of dark-greyish, massive amygdaloidal basalt with a
thickness of c. 90 m; the basalt occasionally shows subangular
carbonate and volcanic clasts (2 –5 cm in diameter). Overlying
the basalt is a 9.4-m-thick greyish, massive diamictite.

The Tulasu Fm. This is composed of laminated, black, carbon-


aceous, silicified mudstone with intercalations of black siltstone
with a thickness of 105 m. The siltstone layers increase up-section
and form rhythmic thin mudstone –siltstone couplets. Organic
microfossils are reported from this interval.

Fig. 33.7. Composite stratigraphic log of the Cryogenian and Ediacaran in The Biexibastao Fm. This consists of a lower member (25 m) of silt-
the Quruqtagh area, southern margin of the Eastern Tianshan Mountains stone and sandstone and an upper member of diamictite (30 m).
(after Gao & Zhu 1984; He et al. 2007; Kou et al. 2008). Legends as in Limestone beds occur within the well-bedded siltstone and
Figure 33.10. sandstone interval.
374 M. ZHU & H. WANG

Fig. 33.8. Geological map showing the


Cryogenian and Ediacaran outcrop of the
Guozigou-Keguqingshan area, western
range of Northern Tianshan Mountains
(modified from Wang et al. 1983).

The Keyindi Fm. This is a thin interval in the Kuguqing Mountains an average diameter of 0.2–5 cm. The large clasts (up to 13 cm in
area and becomes thicker in the Guozigou area, where it is more diameter) exhibit characteristics of dropstone within the finely
than 115 m thick, but the basal part is unexposed. The formation laminated mudstone, providing solid evidence of glaciogenic pro-
is composed of laminated, dark-greyish muddy siltstone and venance of these diamictite, dolostone and mudstone units.
black silty mudstone.
The Talisayi diamictite. This contains 10% clasts with an average
The Tarqat Fm. This is also composed predominantly of laminated, diameter ranging from 0.2 –15 cm. Diameters of the largest
muddy siltstone but with grey-greenish or purplish colour and clasts range up to 100 cm. The clasts consist of predominantly car-
cross-stratification in some layers. Thin lenticular limestone beds bonate, chert and quartz, and rare clasts of granite and other volca-
(0.7– 3 cm) occur in some sections. nic or metamorphic rocks. The shapes of the clasts vary from round
to angular. Some clasts show striation marks, rolling cracks or
The Talisayi Fm. This is a distinct interval consisting of greyish or lacunules on the surface.
purplish, massive diamictite with a thickness ranging from 79 to Based on the sedimentary features described above, the Biexi-
424 m. Several dark-greyish, laminated siltstone layers are recog- bastao and Talisayi diamictites have been considered by Wang
nized within the diamictite. The diamictite is overlain by Cambrian et al. (1983) to be marine glaciogenic diamictites.
phosphorite of the Linkuanggou Fm., which contains early small
shelly fossils and trilobites. A conglomerate bed at the base of
the Linkuanggou Fm. and an irregular contact marks the unconfor- Correlation
mity between the Talisayi and the Linkuanggou formations.
Based on the stratigraphic descriptions of the four type areas,
the successions in the Tielikeli and Aksu-Wusi areas in the
Description of the diamictites Tarim Block show significant similarities. Similar to the well-
studied Cryogenian and Ediacaran successions of the Yangtze
The three diamictite intervals in the area were first interpreted by Block, only two glaciation intervals are developed, and the
Wang et al. (1983) as representing three glaciations. These inter- uppermost part of the succession is composed of thick carbon-
vals were used for correlation with the Quruqtagh area. ate. The C-isotope variation of the upper Sugetblaq and Qige-
blaq formations in the Aksu-Wusi area is similar to that at the
The Kulutieliekti diamictite. This diamictite consists of 40– 50% top Doushantuo and Dengying formations in the Yangtze
clasts of dolostone, limestone, marbles, cherts and basalt. The Block (He et al. 2007a; Zhan et al. 2007). In particular, a
diameter of these clasts varies from 1 to 25 cm, and the largest major negative excursion in the lower part of the Upper
diameter reaches 35 cm. However, the diamictite lacks convincing Member of the Sugetblaq Fm. can be correlated with the
glaciogenic features, and its glaciogenic provenance remains ‘DOUNCE’ excursion at the top of the Doushantuo Fm. (Zhu
uncertain. et al. 2007). The chemostratigraphy supports the correlation of
the Yulmeinak glaciation with the Nantuo glaciation (635 Ma),
The Biexibastao diamictite. This is not a continuous interval but even though there is no distinct cap carbonate recorded atop
often changes laterally to silty dolostone. The diamictite shows the Yulmeinak diamictite.
distinct glaciogenic features, such as striated clasts. The diamic- While the successions in the Quruqtagh and Guozigou-
tite contains 5 –20% clasts with average diameters ranging from Keguqingshan areas in the Tianshan Mountains are similar to
0.2 to 15 cm. The diameters of the largest clasts reach up to each other, they show significant differences from those in the
100 cm. The clasts are generally sub-angular and sub-rounded, Tielikeli and Aksu-Wusi areas. At least three glaciogenic diamic-
dominated by carbonate, chert and volcanic rocks. The laterally tites with possible cap carbonates are recorded in the area. The cap
equivalent silty dolostone also contains 5% angular or subangular carbonates atop the Altungol and Tereeken diamictites show
clasts with average diameters of 0.3–3 cm. The diameters of the characteristic C-isotopic features (Xiao et al. 2004; Kou et al.
largest clasts within the dolostone range up to 20 cm. Similar to 2008; Shen et al. 2008). However, C-isotope values from the cap
the laterally equivalent diamictite, the clast composition is also carbonate atop the Hankalchough diamictite vary dramatically
dominated by carbonate. The associated mudstone, which occurs from section to section (Xiao et al. 2004), demonstrating great
either below the diamictite or is laterally equivalent, contains differences when compared with the cap carbonates atop the
5–25% sand grains and 5% angular or sub-angular clasts with Altungol and Tereeken diamictites. The reason for the spatial
THE TARIM BLOCK, NW CHINA 375

variation of C-isotope values of the cap carbonate atop the Hankal-


chough diamictite remains problematic. Nevertheless, the distinct
association of glaciogenic diamictites and cap carbonates provides
strong evidence for correlation of these diamictites. In particular,
the cap carbonate above the Tereeken diamictite (see Supplemen-
tary Material, Fig. 1S) shows similar sedimentary features to that
of the Nantuo-Marinoan diamictite in other continents, supporting
their correlation. Furthermore, C-isotope chemostratigraphy also
supports a post-Nantuo or post-Cryogenian age of the Hankal-
chough diamictite, because the Dounce-Shuram-Wonaka excur-
sion, which is regarded as a distinct global mid-Ediacaran event,
has been recorded from the Shuiquan Fm. (Xiao et al. 2004).
Detailed integrated correlations of the Cryogenian and Edia-
caran successions of these four areas are given in Figure 33.10.
It should be noted that biostratigraphic correlations using organic-
walled microfossils have been intensively adapted for correlation
of the Neoproterozoic successions in the literature. Because taxo-
nomic problems and restricted occurrences resulted from poor
preservation, biostratigraphic correlation of organic-walled micro-
fossils remains problematic (Gao et al. 1985a; Zhong & Hao
1990). Magnetostratigraphy may support an improved correlation
of the Neoproterozoic successions; however, because of complex
and long-lasting tectonic processes, the original magnetic polarity
can only rarely be discerned. Correlations based on previous
palaeomagnetic studies (see Gao & Zhu 1984; Gao et al. 1985a,
1993) are inconsistent with other correlations and are thus not
yet considered reliable.

Geochronological constraints on the diamictites

Volcanic rocks and tuffaceous beds are well developed within the
Cryogenian and Ediacaran sequences in the Tarim Block. Three
major volcanic episodes are recorded particularly in the Quruqtagh
area, and have been termed (in ascending order) the Beiyixi, Zha-
moketi and Shuiquan episodes by Gao et al. (1985a). These volca-
nic rocks and tuffaceous beds, including basalt, diabase, andesite,
rhyolite and quartz porphyry, provide potential samples for radio-
metric dating. Numerous radiometric ages have been published in
past decades (Lu et al. 1985; Gao et al. 1993; Xu et al. 2005, 2009
and references therein). However, as these ages were based on
different radiometric methods, it is difficult to adapt them for stra-
tigraphic correlation. The most reliable SHRIMP zircon U –Pb
ages from the Qurugtagh area published by Xu et al. (2009)
provide constraints on the four diamictites in the area, suggesting
that the Bayixi diamictite was deposited between 740+7 Ma and
725+10 Ma, the Altungol and Tereeken diamictites between
725+10 Ma and 615+6 Ma, and the Hankalchough diamictite
between 615+6 Ma and c. 542 Ma.
These ages indicate that the oldest (Beiyixi) glaciation in the
Tarim Block is younger than 740 Ma. Although we are not sure
whether the Beiyixi glaciation represents the oldest Neoproterozoic
glaciation on a global scale, however, geochronological data from
the magmatic rocks underlying all the lowest glaciogenic diamic-
tites in the Tarim Block suggest that the oldest glaciation should
be younger than 780 Ma. These ages were reported by Zhang
et al. (2009), including a U–Pb zircon age of 773+3 Ma for the
mafic dyke swarms intruding the c. 820 Ma granite in the Quruqtagh
area, a U–Pb zircon age of 759+7 Ma for the mafic dyke swarms
intruding the blueschist of the Aksu Group of the Aksu-Wusi
region, and a U–Pb zircon age of 815+57 Ma for a mantle-sourced
gneissic granite in the Tielikeli area. Nevertheless, no sedimentary
evidence of glaciation overlying these 760–780 Ma magmatic
rocks and underlying the Beiyixi diamictite rules out glaciations(s)
between 740 Ma and 780 Ma in the Tarim Block.
The reported ages also support the correlation of the Tereeken
Fig. 33.9. Composite stratigraphic log of the Cryogenian and Ediacaran in the glaciation with the c. 635 Ma Nantuo glaciation, and that the
Guozigou-Keguqingshan area, western range of Northern Tianshan Mountains Hankalchough diamictite represents a younger Ediacaran gla-
(after Wang et al. 1983; Gao et al. 1985a). Legends as in Figure 33.10. ciation. However, whether the Hankalchough and Talisayi
376 M. ZHU & H. WANG

Fig. 33.10. Stratigraphic correlation of the


Cryogenian and Ediacaran successions in
the Tarim area. Key: 1, dolostone; 2,
limestone; 3, conglomerate; 4,
coarse-grained sandstone; 5, sandstone;
6, siltstone; 7, diamictite; 9, bedded
diamictite; 8, shale and mudstone;
10, silicates or cherts; 11, phosphorite;
12, blueschist; 13, volcanic rocks;
14, dropstones.

diamictites are time-correlative to the 582 Ma Gaskiers glaciation Palaeolatitudes and palaeogeography
remains poorly constrained. Based on the C-isotope data from the
Shuiquan Fm. in the Qurugtagh area (Xiao et al. 2004), it is pro- Systematic palaeomagnetic analyses of the samples from Cryogen-
posed here the Hankalchough diamictite is younger than 551 Ma, ian and Ediacaran strata of the Aksu-Wusi, Quruqtagh and
because the negative C-isotopic data from the Shuiquan Fm. Guozigou-Keguqinshan areas indicate that the palaeolatitudes of
exhibit distinct values of the DOUNCE excursion at the top of the three areas are ,278 (Table 33.1), which supported the low-
the Doushantuo Fm. of the Yangtze Block. Strata at the top of latitude glaciations during the late Neoproterozoic put forward
this excursion have a high-resolution zircon U – Pb age of by the Snowball Earth hypothesis. These palaeomagnetic data,
551 Ma (Condon et al. 2005). There is no doubt that Hankal- combined with new data from the mafic dyke swarms (c. 810–
chough and Talisayi diamictites represent an end-Ediacaran gla- 820 Ma) intruding the strata below the Cryogenian successions
ciation rather than a Cambrian one, because Cambrian small in the Tielikeli (328N; Zhang et al. 2004) and Aksu-Wusi
shelly fossils have been reported from the strata overlying the (43 + 68N; Chen et al. 2004) areas, collectively suggest a close
Hankalchough and Talisayi diamictites (Wang et al. 1983; Qian palaeogeographic affinity of the Tarim and Yangtze plates with
1999; Yao et al. 2005). Australia and Antarctica during the late Neoproterozoic,
THE TARIM BLOCK, NW CHINA 377

Table 33.1. Palaeolatitude data from four areas of the Tarim Block (Gao et al. Kou, X. W., Wang, Y., Wei, W., He, J. Y. & Xu, B. 2008. The Neo-
1993; Huang et al. 2005; Zhan et al. 2007) proterozoic Altungol and Huangyanggou formations in Tarim
plate: recognized newly glaciation and interglaciation? Acta Petrolo-
Guozigou-Keguqinshan Quruqtagh Aksu-Wusi gica Sinica, 24, 2863– 2868 (in Chinese with English abstract).
Li, H. & Dong, Y. 1991. Sedimentary features of the Sinian Zhamoketi
Tarqat Fm. 8.18 Shuiquan Fm. 21.78 Sugetblaq Fm. 24.38 Formation in the Middle Quruqtagh area of Xinjiang. Xinjiang
6.68 Zhamoketi Fm. 21.98 c. 278 Geology, 9, 340–351.
Kulutieliekti Fm. 10.38 Tereeken Fm. 19.98 Yulmeinak Fm. 24.18 Liou, J. G., Graham, S. A. et al. 1989. Proterozoic blueschist belt in
Beiyixi Fm. 0.98 western China: best documented Precambrian blueschists in the
world. Geology, 17, 1127–1131.
Liu, B., Xu, B., Meng, X. Y., Kou, X. W., He, J. Y., Wei, W. & Mi, H.
2007. Study on the chemical index of alteration of Neoproterozoic
strata in the Tarim plate and its implications. Acta Petrologica
supporting the supercontinental configuration and subsequent Sinica, 23, 1664–1670 (in Chinese with English abstract).
rifting of Rodinia. Lu, S., Ma, G., Gao, Z. & Lin, W. 1985. Sinian ice ages and glacial
sedimentary facies – areas in China. Precambrian Research, 29,
This work benefited from discussions with many colleagues, including C. Zhou, 53– 63.
G. Li, B. Xu, S. Xiao and G. Shields. Fieldwork was jointly carried out with Lu, S., Li, H., Zhang, C. & Niu, G. 2008. Geological and geochronolo-
G. Li, F. Zhao, X. Zhao and Z. Yin. This project is supported by the National gical evidence for the Precambrian evolution of the Tarim craton
Natural Science Foundation of China (nos 40715005 and 40930211). This rep- and surrounding continental fragments. Precambrian Research,
resents a contribution of the IUGS- and UNESCO-funded IGCP (International 160, 94– 107.
Geoscience Programme) project #512. Ma, S., Wang, Y. & Fang, X. 1989. The Sinian at north slope of western
Kunlun Mountains. Xinjiang Geology, 7, 68– 79 (in Chinese with
English abstract).
References Ma, S., Wang, Y. & Fang, X. 1991. Basic characteristics of Proterozoic
Eonothem as a table cover on northern slope of western Kunlun
Cao, R. 1991. New observation of the Sinian System in the southern Mountain. Xinjiang Geology, 9, 59– 71 (in Chinese with English
Yardang Mountains, Xinjiang. Regional Geology of China, 1991, abstract).
30 – 34 (in Chinese with English abstract). Norin, E. 1937. Reports from the Scientific Expedition to the Northwes-
Chen, Y., Xu, B., Zhan, S. & Li, Y. 2004. First mid-Neoproterozoic tern Provinces of China under the Leadership of Dr. Sven Hedin, III.
palaeomagnetic results from the Tarim Basin (NW China) and their Geology, 1. Geology of Western Quruqtagh, Eastern Tien-Shan.
geodynamic implications. Precambrian Research, 133, 271– 281. Bokförlags Aktiebolaget Thule, Stockholm.
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005. Qian, Y. (ed.) 1999. Taxonomy and Biostratigraphy of Small Shelly
U –Pb ages from the Neoproterozoic Doushantuo Fm., China. Fossils in China. Science Press, Beijing (in Chinese with English
Science, 308, 95 – 98. summary).
Duan, J. Y., Xia, D. X. & An, S. L. 2005. Deep-water sedimentation and Qian, J. & Xiao, B. 1984. An Early Cambrian small shelly fauna from
tectono-palaeogeography of the Neoproterozoic – early Palaeozoic Aksu-Wushi region, Xinjiang. In: Professional Papers of Strati-
aulacogen in Kuruktag, Xingjiang, China. Acta Geologica Sinica, graphy and Palaeontology, No. 13. Geological Publishing House,
79, 7 – 14 (in Chinese with English abstract). Beijing, 65– 90 (in Chinese with English abstract).
Gao, Z. & Chen, K. 2003. The Nanhua System of Xinjiang and some Shen, B., Xiao, S., Kaufman, A. J., Bao, H., Zhou, C. & Wang, H.
geological issue of Nanhua System in China. Geological Survey 2008. Stratification and mixing of a post-glacial Neoproterozoic
and Research, 26, 8 –14 (in Chinese with English abstract). ocean: evidence from carbon and sulfur isotopes in a cap dolostone
Gao, Z. & Qian, J. 1985. Sinian glacial deposits in Xinjiang, Northwest from northwest China. Earth and Planetary Science Letters, 265,
China. Precambrian Research, 29, 143– 147. 209– 228.
Gao, Z. & Zhu, S. 1984. Precambrian Geology in Xinjiang, China. Xin- Wang, J., Cheng, S., Bai, W. & Wang, L. 1983. Glacigenous strata in
jiang People’s Publishing House, Urumuqi, China (in Chinese with the western part of the northern Tianshan Mountains. In: Precam-
English summary). brian Geology, No.1 of the Collected Works of Late Precambrian
Gao, Z., Wu, S., Li, Y. & Qian, J. 1981. Sinian –Cambrian stratigraphy of Glacigenous Rocks in China. Geological Publishing House,
Aksu-Keping area, Xinjiang. Chinese Science Bulletin, 12, 741–743 Beijing, 105– 118 (in Chinese with English summary).
(in Chinese). Wang, A. G., Zhang, C. L. & Guo, K. Y. 2004. Depositional types and its
Gao, Z., Wang, W. et al. 1985a. The Sinian System of Xinjiang. Xinjiang tectonic significance of lower member of Nanhuan System in North
People’s Publishing House, Urumuqi, China (in Chinese with English margin of West Kunlun. Journal of Stratigraphy, 28, 248– 256 (in
abstract). Chinese with English abstract).
Gao, Z., Wang, W. et al. 1985b. The Sinian System on Aksu-Wushi Xiao, S., Bao, H. et al. 2004. The Neoproterozoic Quruqtagh Group in
Region, Xinjiang, China. Xinjiang People’s Publishing House, eastern Chinese Tianshan: evidence for a post-Marinoan glaciation.
Urumuqi, China (in Chinese with English summary). Precambrian Research, 130, 1– 26.
Gao, Z., Chen, J., Lu, S., Peng, C. & Qin, Z. 1993. The Precambrian Xu, B., Jian, P., Zheng, H., Zou, H., Zhang, L. & Liu, D. 2005. U– Pb
Geology in Northern Xinjiang (Precambrian Geology, No. 6). Geo- zircon geochronology and geochemistry of Neoproterozoic volcanic
logical Publishing House, Beijing (in Chinese with English rocks in the Tarim Block of northwest China: implications for the
summary). breakup of Rodinia supercontinent and Neoproterozoic glaciations.
He, J. Y., Xu, B., Meng, X. Y., Kou, X. W., Liu, B., Wang, Y. & Mi, H. Precambrian Research, 136, 107– 123.
2007a. Neoproterozoic sequence stratigraphy and correlation in Xu, B., Xiao, S. et al. 2009. SHRIMP zircon U– Pb age constraints on
Kurugtagh area, Xinjiang. Acta Petrologica Sinica, 23, 1645–1654 Neoproterozoic Quruqtagh diamictites in NW China. Precambrian
(in Chinese with English abstract). Research, 168, 247–258.
He, X. B., Xu, B. & Yuan, Z. Y. 2007b. C-isotope composition and Yao, J., Xiao, S., Yin, L., Li, G. & Yuan, X. 2005. Basal Cambrian
correlation of the Upper Neoproterozoic in Keping area, Xingjiang. microfossils from the Yurtus and Xishanblaq formations (Tarim,
Chinese Science Bulletin, 52, 504– 511. north-west China): systematic revision and biostratigraphic corre-
Huang, B. C., Xu, B., Zhang, C. X., Li, Y. A. & Zhu, R. X. 2005. Palaeo- lation of Micrhystridium-like acritarchs from China. Palaeontology,
magnetism of the Baiyisi volcanic rocks (c. 740 Ma) of Tarim, North- 48, 687– 708.
west China: a continental fragment of Neoproterozoic Western Zhan, S., Chen, Y., Xu, B., Wang, B. & Faure, M. 2007. Late Neopro-
Australia? Precambrian Research, 142, 83 –92. terozoic palaeomagnetic results from the Sugetbrak Formation of the
378 M. ZHU & H. WANG

Aksu area, Tarim basin (NW China) and their implications to palaeo- Zhao, W., Li, C., Gao, Z. & Miao, C. 1985. Stromatolites from the Paer-
geographic reconstructions and the snowball Earth hypothesis. gangtag Group at the southern slope of Mt. Kuluketag, Xinjiang. Acta
Precambrian Research, 154, 143– 158. Palaeontologica Sinica, 24, 71 –82.
Zhang, C. L., Shen, J. L. & Guo, K. Y. 2004. Geochemistry of the Neo- Zhong, D. & Hao, Y. (eds) 1990. Sinian to Permian Stratigraphy
proterozoic mafic dyke swarm and basalt in south of Tarim Plate and and Palaeontology of the Tarim Basin, Xinjiang, (I) Kuruktag
its tectonic significance. Acta Petrologica Sinica, 20, 473– 482 (in Region. Nanjing University Press, Nanjing (in Chinese with
Chinese with English abstract). English abstract).
Zhang, C. L., Li, X. H., Li, Z. X., Lu, S. N., Ye, H. M. & Li, H. M. 2007. Zhu, M., Strauss, H. & Shields, G. A. 2007. From Snowball Earth to the
Neoproterozoic ultramafic –mafic– carbonatite complex and grani- Cambrian bioradiation: calibration of Ediacaran– Cambrian Earth
toids in Quruqtagh of northeastern Tarim Block, western China: geo- history in South China. Palaeogeography, Palaeoclimatology,
chronology, geochemistry and tectonic implications. Precambrian Palaeoecology, 254, 1 –6.
Research, 152, 149– 169. Zhu, W., Zhang, Z., Shu, L., Lu, H., Su, J. & Yang, W. 2008. SHRIMP
Zhang, C. L., Li, Z. X., Li, X. H. & Ye, H. M. 2009. Neoproterozoic mafic U –Pb zircon geochronology of Neoproterozoic Korla mafic dykes in
dyke swarms at the north margin of the Tarim Block, NW China: age, the northern Tarim Block, NW China: implications for the long
geochemistry, petrogenesis and tectonic implications. Journal of Asia lasting breakup process of Rodinia. Journal of the Geological
Earth Sciences, 35, 167– 179. Society, London, 165, 887–890.
Chapter 34

The Hula Hula Diamictite and Katakturuk Dolomite, Arctic Alaska

FRANCIS A. MACDONALD
Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138, USA (e-mail: fmacdon@fas.harvard.edu)

Abstract: The Katakturuk Dolomite is a c. 2-km-thick Neoproterozoic carbonate succession (units K1– K4) exposed in the NE Brooks
Range of Alaska. These strata were deposited on a south-facing (present coordinates), rifted passive margin on the North Slope subterrane
(NSST) of the Arctic Alaska-Chukotka Plate (AACP). The glaciogenic Hula Hula diamictite rests below the Katakturuk Dolomite and
consists of 2– 50 m of diamictite that interfingers with the underlying Mt. Copleston volcanic rocks. Unit K1 of the Katakturuk Dolomite
begins with less than 10 m of dark grey, finely laminated limestone with ‘roll-up’ structures, and continues upwards with nearly 500 m of
recrystallized, ooid-dominated grainstone. The Nularvik dolomite (unit K2 of the Kataktruk Dolomite) rests on unit K1 with a knife-sharp
contact on a heavily silicified surface. The Nularvik dolomite is composed predominantly of laminated micro-peloids hosting tubestone
stromatolites and giant wave ripples, followed by decametres of dolomatized, pseudomorphosed former aragonite crystal fans.
Carbon-isotope chemostratigraphy suggests that the Hula Hula diamictite is an early Cryogenian glacial deposit, and that, despite the
absence of directly underlying glacial deposits, the Nularvik dolomite is a basal Ediacaran cap carbonate. These correlations are sup-
ported by the characteristic sedimentological features in both the carbonate capping the Hula Hula diamictite and the Nularvik dolomite.
Detrital zircon and Palaeozoic fauna provenance studies support the inference that much of the AACP is exotic to Laurentia; however, the
pre-Mississipian relationship between the NSST and the rest of the AACP remains uncertain. Previous palaeomagnetic surveys have been
hampered by pervasive Late Cretaceous overprints. Additional geological mapping, sequence stratigraphy and geochronological data are
needed to correlate Neoproterozoic and Palaeozoic units across the AACP, and constrain relationships between subterranes in the AACP.

The Neoproterozoic Katakturuk Dolomite (units K1 –K4) and the type section of the Nularvik dolomite (referred to as the Nularvik
overlying Cambrian to Ordovician Nanook Limestone form the cap carbonate in Macdonald et al. 2009b) is on the west side of the
backbone of the northernmost ranges of the Arctic National Wild- Nularvik River (Figs 34.2 and 34.3b, section F601; 69837.5880 N,
life Reserve of Alaska (for a field guide and an introduction to the 145805.4510 W). The Nularvik dolomite is succeeded by a major
general geology of the region, see Molenaar et al. 1987). The pre- transgression marked by shale and allodapic carbonate, and then
Mississippian stratigraphy of the NE Brooks Range has been an additional c. 1200 m of Ediacaran shallow-water dolomite
described through geological mapping (Leffingwell 1919; Reed (units K3 –K4).
1968; Reiser 1971; Sable 1977; Robinson et al. 1989), palaeonto- Exposure of the Katakturuk Dolomite is limited to the North
logical reconnaissance (Dutro 1970) and sequence analysis (Clough Slope subterrane (NSST) of the Arctic Alaska-Chukotka Plate
& Goldhammer 2000). Recently, Macdonald et al. (2009b) con- (AACP). The NSST is tied to the AACP by the Carboniferous
ducted integrated chemo- and lithostratigraphic studies, identified Lisburne Group; however, the pre-Mississippian relationship
two putative glacial horizons and suggested a late Neoproterozoic between the NSST, the AACP and Laurentia remain uncertain.
age for the Katakturuk Dolomite. Dark-coloured limestones, potentially correlative to the lower-
The Katakturuk Dolomite was named by Dutro (1970) for its most unit of the Katakturuk Dolomite (K1), are present in the
exposure in the Katakturuk River canyon in the Sadlerochit Moun- Third and Fourth Ranges of the NSST (Fig. 34.1), c. 10 km
tains. Katakturuk is an English derivation of the Inupiaq word and 15 km south of the Shublik Mountains, respectively (Reiser
Qattaqtuuraq, which translates to ‘a wide open place’ (Clough 1971). However, the thickness of this unit is unknown as it is
1989). The Katakturuk Dolomite has only been positively ident- poorly exposed in the Third Range and structurally duplicated
ified in three localities: the Sadlerochit Mountains, the Shublik in the Fourth Range (Macdonald et al. 2009b). Precambrian dolo-
Mountains and on both sides of the Hula Hula River near mites dominated by stromatolites and coated grains, possibly cor-
Kikitak Mountain (Fig. 34.1). The succession attains its greatest relative with the Katakturuk Dolomite, have been described on
thickness in the Sadlerochit Mountains, where it is composed of Seward Peninsula and near Snowden Mountain on the southern
c. 2080 m of shallow-water dolomite, and progressively thins to subterranes of the AACP (Dumoulin 1988; Dumoulin & Harris
the south in the Shublik Mountains and at Kikitak Mountain. 1994), and in the Farewell Terrane (Babcock et al. 1994);
Exposures in the NE Brooks Range capture the shelf – slope tran- however, no diamictites have been described at these localities.
sition, facing to the south (present coordinates). Terminal Neoproterozoic diamictites occur in the Upper Tindir
Macdonald et al. (2009b) identified the Hula Hula diamictite at Group of East-Central Alaska, but these were definitively depos-
the base of the Katakturuk Dolomite in the eastern Sadlerochit ited on Laurentia (Allison et al. 1981; Young 1982; Macdonald
Mountains and along the Hula Hula River near Kikitak Mountain & Cohen 2011).
(Fig. 34.1). The type section of the Hula Hula diamictite is on the
east side of the Hula Hula River along Eustik Creek (Figs 34.2 &
34.3a, section F621; 69825.4910 N, 144823.2160 W). At this locality, Structural framework
the diamictite is overlain by a 10-m-thick, dark, variably dolomi-
tized limestone (unit K1 of the Katakturuk Dolomite) that contains Present exposures of the Katakturuk Dolomite in the Sadlerochit,
microbial ‘roll-up’ structures (Hoffman et al. 1998; Pruss et al. Shublik and Kikitak Mountains (Fig. 34.1) are the product of
2010). This is one of only two limestone horizons within the Palaeogene north-vergent thrusting (Wallace & Hanks 1990).
Katakturuk Dolomite. These structures formed during a late-stage reactivation of the
In the Sadlerochit Mountains, nearly 500 m higher in the mostly Mesozoic Brookian orogeny (Moore et al. 1997), and are
sequence, a micro-peloidal dolostone containing tubestone stroma- related to the piecemeal accretion of southern Alaska (Fuis et al.
tolites, giant wave ripples and decametres of pseudomorphosed 2008). Pre-Mississippian strata in the NE Brooks Range also pre-
former aragonite crystal fans rests on a silicified surface. The serves Early to Middle Devonian SE-vergent structures associated

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 379– 387. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.34
380 F. A. MACDONALD

Fig. 34.1. Geological map of the Shublik, Sadlerochit and Kikitak Mountains, and the Fourth Range, with a simplified tectonic and location map inset. Post-Ordovician
geology modified from Robinson et al. (1989) and Bader & Bird (1986). Tectonic map modified and simplified from Johnston (2001) and Colpron et al. (2007).
Abbreviations for terranes: AACP, Arctic Alaska-Chukotka Plate; FR, Farewell-Ruby-Hammond Terranes; NAm, Ancestral North America; LA, Late Accreted
Terranes; ?, poorly exposed and of uncertain affinity.

with the Romanzof orogeny (Oldow et al. 1987). These structures Clough & Goldhammer (2000) inferred a NE– SW palaeo-
are thought to be distinct from those related to the Late Devonian strandline (present coordinates) from palaeo-current data measured
Ellesmerian orogeny in the Yukon and Canadian Arctic Islands in tabular cross-bedded grainstones, the orientation of elongated
(Lane 2007). stromatolites, and from the thickening of outer-ramp to slope
Metamorphic grade generally increases from north to south facies to the south. While readily measurable elongated stromato-
with exposures of the Katakturuk Dolomite in the Shublik and lites are not present in the lower kilometre of the Katakturuk Dolo-
Sadlerochit Mountains displaying little folding and simple mite, facies changes are apparent between the Sadlerochit, Shublik
block faults. In the NE Sadlerochit Mountains, the underlying and Kikitak Mountains, deepening from north to south (Macdonald
Neoproterozoic rocks are tightly folded and chloritized; et al. 2009b); deepening is not demonstrable in facies changes
however, this apparent difference in deformation may be due from west to east along the ranges. Consequently, for the lower
entirely to lithology and rheology. The overlying Nanook Lime- kilometre of the Katakturuk Dolomite (the Hula Hula Diamictite
stone contains conodonts with a Conodont Alteration Index and units K1 –K3), Macdonald et al. (2009b) assumed an east –
(CAI) of 3.5 suggesting that at least these strata have experienced west palaeo-strandline. Balanced cross-sections from seismic data
less than 300 8C (Harris et al. 1990). The Katakturuk Dolomite is reveal a c. 30% Palaeogene north – south shortening in the NE
heavily dolomitized and recrystallized; however, the age of the Brooks Range (Molenaar et al. 1987; Moore et al. 1997). This
dolomitzation is unknown and, typically, primary sedimentary results in a restored distance between the Sadlerochit and Shublik
features are preserved. In the Kikitak Mountain area, small-scale Mountains of c. 11 km, and an additional 10 km north –south and
folding is more common, and the basalts are pervasively chlori- 50 km east –west between the Shublik and Kikitak Mountains.
tized. These modifications, coupled with incomplete exposures, However, there are few constraints on the pre-Palaeogene tectonic
make an exact determination of the thickness of the Hula Hula movement between the two ranges, inhibiting confidence in three-
diamictite impossible. dimensional basin reconstructions.
HULA HULA DIAMICTITE AND KATAKTURUK DOLOMITE 381

Fig. 34.2. Composite carbon chemo- and


lithostratigraphy of the Katakturuk
Dolomite (K1– K4) and the Nanook
Limestone (N1, N2) in the NE Brooks
Range. Sadlerochit data from measured
sections F601, F602, F607, F501 and F505;
Shublik data from measured sections F513,
F514, F517, F613 and F614; Kikitak data
from measured sections F619 and F624 (see
Figs 34.1 & 34.3 for locations). All
carbonate carbon measurements in ‰
notation. Depositional distance assumes
c. 30% Palaeogene shortening (Molenaar
et al. 1987; Moore et al. 1997). Time scale
is inferred from the detrital zircon ages in
the ‘O.G.’ (Macdonald et al. 2009b),
Cambrian trilobites in the Nanook
Limestone and the similarity of the
chemostratigraphic profile and sedimentary
structures in the intervening carbonates to
other Neoproterozoic age successions (see
text for discussion). HH, Hula Hula
diamictite; V, Mt. Copleston
volcanic rocks.

Stratigraphy and the Katakturuk Dolomite are separated by the Hula Hula
diamictite, which ranges in thickness from c. 2 m to 50 m.
As much as 3 km of Neoproterozoic strata are exposed in the NE
Brooks Range of Arctic Alaska. The oldest strata in the region are
in map unit ‘O. G.’, which was previously mapped as the Neurokpuk Katakturuk Dolomite
Formation (Fm.) and consists of tightly folded, mixed siliciclastic
and carbonate rocks (Reiser et al. 1980; Robinson et al. 1989; In the Sadlerochit Mountains, the Katakturuk Dolomite is over
Macdonald et al. 2009b). The thickness of this unit is unknown 2000 m thick, whereas in the Shublik Mountains, it is only c.
and its base is not exposed. Map unit O. G. is succeeded by the Mt. 1000 m thick. This is due both to internal thinning and a basal trun-
Copleston volcanic rocks. Although this contact is structural in the cation (Macdonald et al. 2009b). Generally, the Katakturuk Dolo-
Sadlerochit Mountains (Macdonald et al. 2009b), regionally it mite is composed of massive, light-grey, shallow-water dolomite
has been described as an unconformity (Reiser et al. 1980; Robin- with common ooids and cement. However, further to the SE,
son et al. 1989). The Mt. Copleston volcanic rocks underlie and along the Hula Hula River, the Katakturuk Dolomite is composed
interfinger with the Hula Hula diamictite, which is in turn suc- predominantly of allodapic carbonate of unit K1 and measures
ceeded by the c. 2-km-thick Neoproterozoic Katakturuk Dolomite. only 400 m thick, with the upper units of the Katakturuk Dolomite
Subsidence analysis suggests that these strata were accommodated truncated under the sub-Mississippian unconformity (Figs 34.2 &
by extension on the southern margin of the North Slope subterrane 34.3b). To highlight the major unconformities and disconformities,
with units K1 – K3 of the Katakturuk Dolomite deposited on a ther- and departing from Robinson et al.’s (1989) lithostratigraphic sub-
mally subsiding passive margin (Macdonald et al. 2009b). division, Macdonald et al. (2009b) divided the Katakturuk Dolo-
mite into four informal units (Fig. 34.2): the Cryogenian map
unit K1, the basal Ediacaran Nularvik dolomite (map unit K2),
Mt. Copleston volcanic rocks the early –middle Ediacaran map unit K3 and the late Ediacaran
map unit K4. The uppermost portion of the Katakturuk Dolomite
The Mt. Copleston volcanic rocks are rusty weathering, dark in the Sadlerochit Mountains, as described by Clough & Goldham-
maroon to black and green tholeitic basalts with 5 mm chlorite, mer (2000) has been included with the lowermost Nanook Lime-
calcite, zeolite amygdales, and common native copper. In the stone (Macdonald et al. 2009b).
western Shublik Mountains, the basalt is up to 450 m thick with
metre-scale individual flows. Along the Hula Hula River, the Mt.
Copleston volcanic rocks are c. 500 m thick with a true thickness Nanook Limestone
difficult to determine due to structural complexities. The volcanic
rocks are often greenstone, and dominated by volcaniclastic units The Katakturuk Dolomite is overlain by the Cambrian to Ordovi-
in the upper c. 100 m. On the west flank of Kikitak Mountain, the cian Nanook Limestone. Previous studies suggested an unconfor-
basalts are over 100 m thick and metamorphosed to greenstone. In mity at this level (Clough & Goldhammer 2000; Macdonald
the eastern Sadlerochit Mountains, the basalts appear relatively et al. 2009b); however, this interpretation may have been compli-
low grade with minimal chlorite, 5 –10 mm long plagioclase cated by cave breccias from a karstic surface higher in the succes-
lathes, and intact, spherical amygdales. Exposures measure up to sion (personal observations). The lower c. 250 m of the Nanook
105 m thick with an unconformable, often faulted basal contact Limestone host bed-parallel ichnogenera and C-isotopic profiles
with the underlying unit O. G. The Mt. Copleston volcanic rocks that are consistent with an Early to Middle Cambrian age
382 F. A. MACDONALD

Fig. 34.3. (a) Geological map of the Kikitak Mountain area near the Hula Hula River. (b) Geological map of the eastern Sadlerochit Mountains along Nularvik Creek. See
Figure 34.1 for locations and keys for both maps.

(Macdonald et al. 2009b). The upper Nanook Limestone contains Creek, where the diamictite is only 2 m thick, and is composed
Late Cambrian trilobites (Blodgett et al. 1986), and terminates of angular cobbles of dolomite, quartzite and basalt in a coarse,
with c. 160 m of Middle to Late Ordovician strata with the upper- arkose grit (Fig. 34.4).
most beds containing a diverse gastropod assemblage along with
other molluscs, ostracods and brachiopods with a definitive Katakturuk Dolomite
Late Ordovician age (Blodgett et al. 1986) and a Siberian affinity
(Blodgett et al. 2002). Near Kikitak Mountain and on both sides of the Hula Hula River,
the basal unit of the Katakturuk Dolomite (K1) consists of c. 10 m
Glaciogenic deposits and associated strata of dark limestone, and an additional c. 75 m of rhythmite and allo-
dopic carbonate. Near the Hula Hula River, on the north side of
Hula Hula diamictite Eustik Creek, the basal 10 m limestone overlying the Hula Hula
diamictite exhibits ‘roll-up’ microbial structures, reminiscent of
Reiser et al. (1970) described the orange weathering diamictite the basal Rasthof Formation of northern Namibia (Hoffman
along the Hula Hula River as carbonate debris flows with clasts et al. 1998; Pruss et al. 2010), whereas these are not present in
of basalt, and included these deposits with the Katakturuk Dolo- the Sadlerochit Mountains. These shallow upwards to ,500 m
mite. Macdonald (2009) separated the Hula Hula diamictite from of massively bedded, often silicified, grainstone and packstone
the Katakturuk Dolomite and suggested that sedimentation with poorly defined parasequences that are littered with giant
occurred under a glacial influence. On the east side of the Hula ooids (c. 5 mm in diameter). Unit K1 culminates with c. 30 m of
Hula River, along Eustik Creek, the Hula Hula diamictite is c. resistant, silicified grainstone and broken beds of recrystallized
50 m thick. Although incomplete exposure and structural rep- black chert that weather to a distinct black and white (the lower
etition compromise the measurement of exact thicknesses, discrete zebra dolomite of Robinson et al. 1989).
thrust panels allow for confidence in the general stratigraphic
relations. The lower 12 m of the diamictite is composed of cobble- Nularvik dolomite
sized clasts of orange dolomite and green to black basalt, together
with quartzite pebbles, in a green to tan siltstone, and is interfin- The Nularvik dolomite (K2) is equivalent to the upper portion of
gered with at least four basaltic flows that range in thickness the zebra dolomite (Robinson et al. 1989; Clough & Goldhammer
from 0.2 to 2 m. This lower diamictite is overlain with c. 30 m 2000). In the Sadlerochit Mountains, the Nularvik dolomite (K2)
of poorly exposed, fine millimetre-laminated siltstone with rare, consists of 20– 45 m of white to buff-coloured, recrystallized,
gravel- and cobble-sized bedding-piercing outsized clasts and mul- finely laminated, micro-peloidal dolomite, overlain by tens of
tiple orange allodapic carbonate beds. The upper 5 m of the Hula metres of former aragonite crystal fans (Macdonald et al.
Hula diamictite is a massive, clast-supported diamict with boulders 2009b). Funnel-shaped calcite and silica cements are common in
of dolomite and cobbles of basalt in a calcareous silt matrix. No the lower 15 m of unit K2. In cross-section, the funnels are less
striated clasts have been observed. than 5 cm tall, and up to 2 cm wide, taper downward, and are com-
On the west side of the Hula Hula River near Kikitak Mountain monly linked at the top along bed parallel cements. The funnels are
(Figs 34.2 & 34.3b), the basal Hula Hula diamictite is composed filled with isopacous, void-filling cements. In the western Shublik
of c. 10 m of diamictite with cobbles of dolomite and basalt in a Mountains, these funnel-shaped cements are laterally equivalent
siltstone matrix, and an additional c. 40 m of fine millimetre- with tubestone stromatolite bioherms (Corsetti & Grotzinger
laminated siltstone with rare outsized clasts and multiple beds of 2005). The peculiar tubestone stromatolites are distinguished by
orange allodapic carbonate. The Hula Hula diamictite is also evenly distributed, c. 1-cm-diameter, cement-filled cylindrical
exposed in the eastern Sadlerochit Mountains along the Nularvik tubes, and span as much as 8 m of stratigraphy.
HULA HULA DIAMICTITE AND KATAKTURUK DOLOMITE 383

Fig. 34.4. C-isotope chemo- and


lithostratigraphy of the Mt. Copleston
volcanic rocks, the Hula Hula diamictite,
and unit K1 along the Nularvik Creek in the
eastern Sadlerochit Mountains (F607), and
along the east side of the Hula Hula River
(F619).

In the central and western Sadlerochit Mountains, giant wave deposit is a debris flow or a glacial diamictite formed of repro-
ripples (Allen & Hoffman 2005) are also present in the Nularvik cessed volcanic rocks. Macdonald et al. (2009b) included this
dolomite, at the top of the micro-peloidal dolomite. These are suc- unit with the Mt. Copleston volcanic rocks rather than the Hula
ceeded by breccia, and tens of metres of pseudomorphosed arago- Hula diamictite because it lacks foreign clasts, shows no obvious
nite crystal fans, with individual fans measuring as tall as 60 cm evidence for a glacial origin, and because volcanic breccias
(Clough & Goldhammer 2000; Macdonald et al. 2009b). The occur at other horizons within the Mt. Copleston volcanic rocks.
strata hosting the fans are dominated by grainstone and cement Along the Nularvik Creek in the Sadlerochit Mountains the
and are often broken, brecciated and recrystallized. Crystal fans basal contact of the Katakturuk Dolomite (unit K1) is poorly
are not present at this horizon in the Shublik Mountains. exposed, whereas near Kikitak Mountain and on both sides of
In the Sadlerochit Mountains, unit K2 is succeeded by 2 m of the Hula Hula River, the Hula Hula diamictite is overlain with a
shale and laterally discontinuous allodapic carbonate beds. In the knife-sharp contact by c. 10 m of dark limestone. The Nularvik
Shublik Mountains, these deeper water facies expand to over dolomite (K2) rests above unit K1 on a heavily silicified surface
100 m of shale, rhythmite and allodapic carbonate. These are in with a knife-sharp contact (Fig. 34.5).
turn overlain by an additional c. 1200 m of dolomite, primarily in
grainstone, biolaminate and stromatolitic facies (units K3 and K4).
Chemostratigraphy
Boundary relations with overlying and underlying High-resolution, carbonate C- and O-isotope chemostratigraphy
non-glacial units through the Katakturuk Dolomite and Nanook Limestone were
reported by Macdonald et al. (2009b). In the Sadlerochit Moun-
In the eastern Sadlerochit Mountains, along the Nularvik Creek, tains, above the Hula Hula diamictite, in the basal 20 m of K1,
the Hula Hula diamictite rests disconformably on pillow basalt C-isotope values rise from – 2‰ to þ6‰ (Fig. 34.4) where they
of the Mt. Copleston volcanic rocks, with a basally erosive hover between þ3‰ and þ6‰. In the Kikitak Mountain area,
contact. In the Kikitak Mountain area, the Hula Hula diamictite C-isotope values rise from þ1‰ to þ8‰, and also oscillate
also rests disconformably on the Mt. Copleston volcanic rocks; around þ5‰ with slightly more variability. In the last para-
yet, the upper 50 m of the Mt. Copleston volcanic rocks consists sequence of K1, values drop to 0‰.
of a volcaniclastic diamictite of outsized volcanic gravel and C-isotope profiles of the Nularvik dolomite display an inverted
cobbles in matrix-supported volcanic grit. It is unclear if this S-shaped profile with a nadir at –2‰ (Fig. 34.5). In the Sadlerochit
384 F. A. MACDONALD

Fig. 34.5. Chemo- and lithostratigraphy of the Nularvik dolomite. See Figure 34.1 for locations of measured sections. All carbonate carbon (filled) and oxygen (hollow)
data are in ‰ notation.

Mountains, above the Nularvik dolomite, C-isotope values are palaeobiogeographic affinities of the AACP do not necessarily
highly variable through the cement-dominated crystal fans. In contradict the rotation model for the opening of the Arctic
the Shublik Mountains, no crystal fans are present, and instead Ocean, but they do indicate that the pre-Devonian AACP was
the transgressive sequence progresses from grainstone, to ribbo- exotic to Laurentia (Macdonald et al. 2009b). These models are
nite, to variably dolomitized limestone rhythmite and shale. difficult to test directly because most palaeomagnetic studies in
C-isotope values bottom out at – 3‰ in these rhythmites, then northern Alaska have been compromised by a pervasive Late
jump to þ3‰ above a sharp surface below the overlying allodapic Cretaceous overprint (Plumley et al. 1989; Stone 1989).
carbonate (Fig. 34.5).
Geochronological constraints
Palaeolatitude and palaeogeography
A minimum age constraint on the Katakturuk Dolomite is provided
The most popular model for the opening of the Arctic Ocean by Late Cambrian trilobites in the upper portion of the overlying
involves a c. 668 counterclockwise rotation of the AACP away Nanook Limestone (Blodgett et al. 1986). A lower age constraint
from the Canadian Arctic islands about a pole in the Mackenzie is provided by map unit O. G., which is stratigraphically below
Delta region (Carey 1955, 1958; Hamilton 1970; Grantz et al. the Mt. Copleston volcanic rocks and the Hula Hula diamictite,
1979). Barring any earlier movement relative to Laurentia, this and contains c. 760 Ma (206Pb/207Pb LA-ICPMS) detrital zircon
model would place the Neoproterozoic exposures in the NE grains (Macdonald et al. 2009b). A coarse, diabase sill within
Brooks Range offshore of what is now Banks Island. Lane map unit O. G., previously assumed to be coeval with the Mt.
(1997) pointed out multiple geological inconsistencies with the Copleston volcanic rocks, yielded a whole rock Rb –Sr isochron
rotation model, including ages of deformation and deposition age of 801 + 20 Ma (Moore 1987; Clough & Goldhammer
and proposed a model pinning Arctic Alaska to near its present 2000); however, recent U –Pb dates of badellyite in these sills
position since Palaeozoic times. However, his ‘fixed’ Alaska suggest they are Cretaceous in age (Macdonald 2009).
model does not account for growing palaeontological evidence
of Siberian and Baltican Palaeozoic fauna in Alaskan terranes
(Blodgett et al. 2002; Dumoulin et al. 2002). Discussion
There are several modified versions of the rotation model that
include differential motion within the AACP (Miller et al. 2006), A glacial origin of the Hula Hula diamictite is indicated by
and pre-rotation displacement relative to North America the presence of bed-penetrating outsized clasts interpreted as
(Sweeney 1982). The Neoproterozoic stratigraphy and Palaeozoic dropstones and the association with a geochemically and
HULA HULA DIAMICTITE AND KATAKTURUK DOLOMITE 385

sedimentologically distinct overlying dark-coloured limestone. and sedimentary textures in a particular order that are both charac-
Although the bulk of the Mt. Copleston volcanic rocks are below teristic of basal Ediacaran cap carbonates globally (Allen &
the Hula Hula diamictite with a few small flows inter-fingering Hoffman 2005). The lack of glacial deposits can be attributed to
with the lowermost Hula Hula diamictite, it is not certain that poor preservation potential as the glacio-eustatic sea-level drop
the Mt. Copleston volcanic rocks mark the onset of glaciation, as left the carbonate platform exposed or possibly covered with
there could have been glacial activity prior to the emplacement grounded ice until the post-glacial transgression. Glacial diamic-
of the basalts that failed to leave a record. Nonetheless, a potential tites are also rare under the Keilberg cap carbonate on the Otavi
interpretation of the stratigraphy of the Hula Hula diamictite along platform in Northern Namibia (Hoffman & Halverson 2008). In
Eustik Creek is that the basal c. 10 m of massive diamictite with the Sadlerochit and Shublik Mountains, a late Cryogenian glacia-
interfingering basalts were deposited as glaciomarine deposits tion may be contained in a silicified surface, with the overlying
during the encroachment of sea ice; the middle c. 30 m of Nularvik dolomite representing the basal Ediacaran cap carbonate.
millimetre-laminated silts with occasional debris flows and rare The Katakturuk Dolomite has previously been correlated with
dropstones formed under total ice cover; and the upper c. 5 m of the early Neoproterozoic deposits of NW Laurentia (Rainbird
massive diamictite represents the ice-retreat phase. et al. 1996), such as the Lower Tindir Group of the Yukon-Alaska
Above the Hula Hula diamictite, C-isotope values rise from border area (Young 1982; Macdonald et al. 2010a), the
–2‰ to þ6‰ in the Sadlerochit Mountains and from þ1‰ to pre-717.4 Ma Fifteenmile Group in the Ogilvie Mountains (Mac-
þ8‰ near Kikitak Mountain (Fig. 34.4). Although it is not clear donald et al. 2010b), the Little Dal Group in the Mackenzie Moun-
why values are more enriched in the deeper-water sections, the tains (Aitken 1981) and the pre-716.3 Ma Shaler Supergroup of
positive, concave trend, and the extremely enriched values are Victoria Island (Young 1981; Macdonald et al. 2010b). This corre-
typical of Cryogenian post-glacial carbonates (Halverson et al. lation was based in large part on palaeogeographical reconstruc-
2005), and the basal negative anomaly is similar in magnitude to tions that tie the AACP to northwestern Laurentia. However, the
the basal Rasthof Formation (Yoshioka et al. 2003). An early Katakturk Dolomite is younger than c. 760 Ma and likely Cryogen-
Cryogenian age of the Hula Hula diamictite is also suggested by ian to Ediacaran in age (Macdonald et al. 2009b). Using the
the presence of ‘roll-up’ microbial structures in the overlying 801 + 20 Ma Rb –Sr date, Clough & Goldhammer (2000)
dark limestones, which are reminiscent of the basal Rasthof cap argued that the Mt. Copleston volcanic rocks represent a rifting
carbonate of Northern Namibia (Hoffman et al. 1998). episode coeval with the c. 780 Ma Gunbarrel Event (Park et al.
C-isotope profiles of the Nularvik dolomite (K2) display an 1995; Harlan et al. 2003), and that the Katakturuk Dolomite was
inverted S-shaped profile with a nadir at –2‰ (Fig. 34.5). Normal- deposited during the thermal subsidence stage on a passive carbon-
ized for thickness (and excluding isotopic values of cements), this ate ramp. Instead, the Mt. Copleston basalts could be correlative
isotopic profile is similar to that of the Ediacaran basal Doushantuo with the Franklin igneous event. However, NW Laurentia lacks
in South China (Jiang et al. 2003; Zhou & Xiao 2007), which has potentially correlative Cryogenian to Ediacaran successions domi-
been dated at 635.2 + 0.6 Ma (Condon et al. 2005). The C-isotope nated by platformal carbonate rocks. At present, the origin and
profile is also reminiscent of slope sections of the Keilberg Fm. in Neoproterozoic position of the NSST and AACP are unknown;
northern Namibia, where underlying glacial deposits have been however, the Neoproterozoic and Palaeozoic strata of the AACP
dated at 635 + 0.5 Ma (Hoffmann et al. 2004), and shelf sections potentially have features in common with peri-Siberian and Balti-
are 3–4‰ lighter than foreslope sections (Hoffman et al. 2007). can terranes (Colpron & Nelson 2009; Macdonald et al. 2009b).
Assuming that carbon is well-mixed in the oceans and on plat- Further work is necessary to test these correlations.
forms, the relatively enriched values of the Nularvik dolomite
suggest it was deposited early compared to shelf sections in I thank my field assistants B. Black, P. Kreycik and W. Macdonald for pushing
northern Namibia, and then truncated by exposure surfaces through inclement weather and difficult terrain. I thank D. Schrag and G. Eischied
before seawater reached extremely negative values. for use of and help in Harvard’s Paleooceanography Laboratory. U. Bold,
Funnel-shaped calcite and silica cements in unit K2 are similar to K. Knudson, W. Macdonald and K. Wecht are thanked for helping to prepare
those in the basal Ediacaran Ol cap carbonate of Mongolia (Macdo- samples. I am grateful to P. Hoffman and the NSF Arctic program for providing
nald et al. 2009a), in the Keilberg Fm. Of Northern Namibia, and in financial support, and VECO polar resources for providing logistical support. I
the basal Doushantuo in South China (Macdonald, unpublished also thank the GSA for a student research grant. I would like to thank
P. Hoffman, D. Jones and J. Clough for helpful discussions and comments
data). As these are laterally equivalent with tubestone stromatolite
throughout this work. Finally, I thank B. McClelland and E. Arnaud for their
bioherms and are reminiscent in plan view, they may be a related
helpful comments that greatly improved the manuscript. This represents a con-
facies that is characteristic of basal Ediacaran cap carbonates.
tribution of the IUGS- and UNESCO-funded IGCP (International Geoscience
The crystal fans in the Nularvik dolomite are formed in grain- Programme) project #512.
stone with multiple exposure surfaces, pervasive cements, and
broken and brecciated beds (Clough & Goldhammer 2000; Macdo-
nald et al. 2009b). Occasionally, individual fans are tipped over on References
their side from the buckling of teepees. C-isotope values are highly
variable through this interval. Together, these data suggest the fans Aitken, J. D. 1981. Stratigraphy and sedimentology of the Upper Proter-
were formed in a restricted, lagoonal setting with multiple ozoic Little Dal Group, Mackenzie Mountains, Northwest Territories.
exposure surfaces (Macdonald et al. 2009b). This is a very differ- In: Campbell, F. H. A. (ed.) Proterozoic Basins of Canada. Geologi-
ent depositional environment than the settings for sea-floor pre- cal Survey of Canada Paper 81-10.
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves
cipitate crystal fan development in other basal Ediacaran cap
in the aftermath of a Neoproterozoic glaciation. Nature, 433,
carbonates (Peryt et al. 1990; James et al. 2001; Hoffman &
123– 127.
Halverson 2008). Isotopic scatter, evidence of exposure and perva- Allison, C. W. A., Young, G. M., Yeo, G. M. & Delaney, G. D. 1981.
sive cementing is also a common feature in other basal Ediacaran Glaciogenic rocks of the Upper Tindir Group, east-central Alaska. In:
cap carbonates that were deposited in basins lacking active stretch- Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene
ing, such as the upper portion of the Doushantuo Fm. in South Glacial Record. Cambridge University Press, Cambridge.
China (Jiang et al. 2003) and the Jbeliat dolostone in Mauritania Babcock, L. E., Blodgett, R. B. & St. John, J. 1994. New Late(?)
(Hoffman & Schrag 2002; Shields et al. 2006). This pre- Proterozoic-age formations in the vicinity of Lone Mountain,
transgression shoaling could be a product of isostatic rebound out- McGrath Quadrangle, West-Central Alaska. In: Till, A. B. &
pacing passive subsidence and post-glacial eustatic sea-level rise. Moore, T. E. (eds) Geologic Studies in Alaska by the U.S. Geological
Although no late Cryogenian glacial diamictites have been Survey, 1993. United States Government Printing Office,
identified, the Nularvik dolostone contains an isotopic profile Washington.
386 F. A. MACDONALD

Bader, J. W. & Bird, K. J. 1986. Geologic map of the Demarcation Point, Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
Mt. Michelson, Fluxman Island quadrangles, northeastern Alaska. Rice, A. H. N. 2005. Toward a Neoproterozoic composite
U.S. Geological Survey Miscellaneous Investigations 1791, 1 sheet, carbon-isotope record. Geological Society of America Bulletin, 117,
scale 1:250,000. 1181– 1207.
Blodgett, R. B., Clough, J. G., Dutro, J. T., Ormiston, A. R., Palmer, Hamilton, W. 1970. The Uralides and the motion of the Russian and
A. R. & Taylor, M. E. 1986. Age revisions of the Nanook Limestone Siberian Platforms. Geological Society of America Bulletin, 81,
and Katakturuk Dolomite, northestern Brooks Range, Alaska. In: 2553– 2576.
Bartsch-Winkler, S. & Reed, K. M. (eds) Geological Studies in Harlan, S. S., Heaman, L. M., LeCheminant, A. N. & Premo, W. R.
Alaska by the Geological Survey during 1985. US Geological 2003. Gunbarrel mafic magmatic event: a key 780 Ma time marker
Survey Circular 978. for Rodinia plate reconstructions. Geology, 31, 1053– 1056.
Blodgett, R. B., Rohr, D. M. & Boucot, A. J. 2002. Palaeozoic links Harris, A. G., Lane, R. H. & Tailleur, I. L. 1990. Conodont thermal
among some Alaskan accreted terranes and Siberia based on mega- maturation patterns in Palaeozoic and Triassic rocks, Northern
fossils. In: Miller, E. L., Grantz, A. & Klemperer, S. L. (eds) Alaska – geological and exploration implications. In: Grantz, A.,
Tectonic Evolution of the Bering Shelf– Chukchi Sea –Arctic Johnshon, L. & Sweeney, J. F. (eds) The Arctic Ocean Region,
Margin and Adjacent Landmasses. Geological Society of America DNAG Series L. Geological Society of America, Boulder, CO.
Special Paper 360, Boulder, Colorado. Hoffman, P. F. & Halverson, G. P. 2008. Otavi Group of the western
Carey, S. W. 1955. The orocline concept in geotectonics. Royal Society of Northern Platform, the Eastern Kaoko Zone and the western Northern
Tasmania Proceedings, 89, 255– 288. Margin Zone. In: Miller, R. M. (ed.) The Geology of Namibia, vol. 2.
Carey, S. W. 1958. Continental drift. In: Carey, S. W. (ed.) Continental Handbook of the Geological Survey of Namibia, Windhoek.
Drift, a Symposium. University of Tasmania, Hobart, Australia. Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis;
Clough, J. G. 1989. General stratigraphy of the Katakturuk Dolomite in testing the limits of global change. Terra Nova, 14, 129–155.
the Sadlerochit and Shublik Mountains, Arctic National Wildlife Hoffman, P. F., Kaufman, A. J. & Halverson, G. P. 1998. Comings and
Refuge, northeastern Alaska. Alaska Division of Geological & goings of global glaciations on a Neoproterozoic tropical platform in
Geophysical Surveys, Public Data File 89-4a, 1– 11. Namibia. GSA Today, 8, 1– 9.
Clough, J. G. & Goldhammer, R. K. 2000. Evolution of the Neoproter- Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
ozoic Katakturuk dolomite ramp complex, northeastern Brooks Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran
Range, Alaska. In: Grotzinger, J. P. & James, N. P. (eds) Carbonate (635 Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Pla-
Sedimentation and Diagenesis in the Evolving Precambrian World. netary Science Letters, 258, 114–131.
SEPM Special Publication, No. 67. Society of Sedimentary Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
Geology, Tulsa, Oklahoma. 2004. U– Pb zircon date from the Neoproterozoic Ghaub
Colpron, M. & Nelson, J. L. 2009. A Palaeozoic Northwest Passage: Formation, Namibia: constraints on Marinoan glaciation. Geology,
incursion of Caledonian, Baltican and Siberian terranes into eastern 32, 817– 820.
Panthalassa, and the early evolution of the North American Cordil- James, N. P., Narbonne, G. M. & Kyser, T. K. 2001. Late Neoprotero-
lera. In: Cawood, P. & Kroner, A. (eds) Accretionary Orogens zoic cap carbonates; Mackenzie Mountains, northwestern Canada;
through Space and Time. Geological Society, London, Special precipitation and global glacial meltdown. Canadian Journal of
Publications, 318, 273–307. Earth Sciences, 38, 1229–1262.
Colpron, M., Nelson, J. L. & Murphy, D. C. 2007. Northern Cordilleran Jiang, G., Kennedy, M. J. & Christie-Blick, N. 2003. Stable isotopic
terranes and their interactions through time. GSA Today, 17, 1 –7. evidence for methane seeps in Neoproterozoic postglacial cap car-
Condon, D. J., Zhu, M., Bowring, S. A., Wang, W., Yang, A. & Jin, Y. bonates. Nature, 426, 822– 826.
2005. U– Pb ages from the Neoproterozoic Doushanto Formation, Johnston, S. T. 2001. The Great Alaskan Terrane Wreck: reconciliation
China. Science, 308, 95 – 98. of palaeomagnetic and geological data in the northern Cordillera.
Corsetti, F. A. & Grotzinger, J. P. 2005. Origin and significance of tube Earth and Planetary Science Letters, 193, 259–272.
structures in Neoproterozoic post-glacial cap carbonates: example Lane, L. S. 1997. Canada Basin, Arctic Ocean: evidence against a
from Noonday Dolomite, Death Valley, United States. Palaios, 20, rotational origin. Tectonics, 16, 363–387.
348– 363. Lane, L. S. 2007. Devonian –Carboniferous palaeogeography and oroge-
Dumoulin, J. A. 1988. Stromatolite- and coated-grain-bearing carbonate esis, northern Yukon and adjacent Arctic Alaska. Canadian Journal
rocks of the western Brooks Range. In: Galloway, J. P. & Hamil- of Earth Sciences, 44, 679– 694.
ton, T. D. (eds) Geologic Studies in Alaska by the U.S. Geological Leffingwell, E. 1919. The Canning River region, northern Alaska. US
Survey during 1987. United States Government Printing Office, Geological Survey Professional Paper 109.
Washington. Macdonald, F. A. 2009. Neoproterozoic stratigraphy of Alaska and
Dumoulin, J. A. & Harris, A. G. 1994. Depositional framework and Mongolia. PhD thesis, Harvard University.
regional correlation of pre-Carboniferous metacarbonate rocks of Macdonald, F. A. & Cohen, P. A. 2011. The Tatonduk inlier, Alaska-
the Snowden Mountain area, Central Brooks Range, Northern Yukon border. In: Arnaud, E., Halverson, G. P. & Shields, G.
Alaska. US Geological Survey Professional Paper 1545, 1 – 55. (eds) The Geological Record of Neoproterozoic Glaciations.
Dumoulin, J. A., Harris, A. G., Gagiev, M., Bradley, D. C. & Geological Society, London, Memoirs, 36, 389–396.
Repetski, J. E. 2002. Lithostratigraphic, conodont, and other faunal Macdonald, F. A., Jones, D. S. & Schrag, D. P. 2009a. Stratigraphic
links between lower Palaeozoic strata in northern and central and tectonic implications of a new glacial diamictite-cap carbonate
Alaska and northeastern Russia. In: Miller, E. L., Grantz, A. & couplet in southwestern Mongolia. Geology, 37, 123–126.
Klemperer, S. L. (eds) Tectonic Evolution of the Bering Shelf – Macdonald, F. A., McClelland, W. C., Schrag, D. P. & Macdonald,
Chukchi Sea– Arctic Margin and Adjacent Landmasses. Geological W. P. 2009b. Neoproterozoic glaciation on a carbonate platform
Survey of America Special Paper 360, Boulder, Colorado. margin in Arctic Alaska and the origin of the North Slope subterrane.
Dutro, J. T. 1970. Pre-Carboniferous carbonate rocks, northeastern Geological Society of America Bulletin, 121, 448–473.
Alaska. In: Adkison, W. L. & Brosge, M. M. (eds) Proceed- Macdonald, F. A., Cohen, P. A., Dudás, F. O. & Schrag, D. P. 2010a.
ings of the Geological Seminar on the North Slope of Alaska, Early Neoproterozoic scale microfossils in the Lower Tindir Group
American Association of Petroleum Geologists, Pacific Section. of Alaska and the Yukon Territory. Geology, 38, 143–146.
Los Angeles, CA. Macdonald, F. A., Schmitz, M. D. et al. 2010b. Calibrating the Cryo-
Fuis, G. S., Moore, T. E. et al. 2008. Trans-Alaska Crustal Transect and genian. Science, 327, 1241–1243.
continental evolution involving subduction underplating and syn- Miller, E. L., Toro, J. et al. 2006. New insights into Arctic palaeo-
chronous foreland thrusting. Geology, 36, 267–270. geography and tectonics from U –Pb detrital zircon geochronology.
Grantz, A., Eittreim, S. & Dinter, D. A. 1979. Geology and tectonic Tectonics, 25, 1 –19.
development of the continental margin north of Alaska. Tectonophy- Molenaar, C. M., Mull, C. G. & Swauger, D. A. 1987. Geologic fea-
sics, 59, 263– 291. tures of Ignek Valley and adjacent mountains, northeastern Alaska.
HULA HULA DIAMICTITE AND KATAKTURUK DOLOMITE 387

Centennial Field Guide Volume 1: Cordilleran Section of the Geo- and Shublik Mountains, Mt. Michelson C-1, C-2, C-3, C-4 Quadran-
logical Society of America. gles, Alaska. US Geological Survey Open-file report, 70-273, scale
Moore, T. E. 1987. Geochemistry and tectonic setting of some volcanic 1:63,360, 5 sheets.
rocks of the Franklinian assemblage, central and eastern Brooks Reiser, H. N., Brosge, W. P., Dutro, J. T. & Detterman, R. L. 1980.
Range. In: Tailleur, I. & Weimer, P. (eds) Alaskan North Slope Geologic map of the Demarcation Point quadrangle, Alaska,
Geology. SEPM, Pacific Section, California and Alaska Geological 1:250,000, map I-1133. Miscellaneous Investigation Series. U.S.
Society, Alaska. Geological Survey.
Moore, T. E., Wallace, W. K., Mull, C. G., Adams, K. E., Plafker, G. Robinson, M. S., Decker, J., Clough, J. G., Reifenstuhl, R. R.,
& Nokleberg, W. J. 1997. Crustal implications of bedrock geology Dillon, J. T., Combellick, R. A. & Rawlinson, S. E. 1989.
along the Trans-Alaska Crustal Transect (TACT) in the Brooks Geology of the Sadlerochit and Shublik Mountains, Alaska National
Range, northern Alaska. Tectonics, 102, 20 645–20 684. Wildlife Reserve, northeast Alaska, Professional Report 100,
Oldow, J. S., Lallemant, A., Julian, F. E. & Seidensticker, C. M. 1:10,000 scale. State of Alaska, Department of Natural Resources,
1987. Ellesmerian (?) and Brookian deformation in the Franklin Division of Geological and Geophysical Survey.
Mountains, northeastern Brooks Range, Alaska, and its bearing on Sable, E. G. 1977. Geology of the western Romanzof Mountains,
the origin of the Canada Basin. Geology, 15, 37 – 41. Brooks Range, northeastern Alaska: a comprehensive study of
Park, J. K., Buchan, K. L. & Harlan, S. S. 1995. A proposed giant plutonic, sedimentary, and metamorphic rocks in an eastern Brooks
radiating dyke swarm fragmented by the separation of Laurentia Range area. Geological Survey, Washington, D.C., Professional
and Australia based on palaeomagnetism of ca. 780 Ma mafic intru- Paper 897.
sions in western North America. Earth and Planetary Science Shields, G., Deynoux, M., Strauss, H., Paquet, H. & Nahon, D.
Letters, 132, 129– 139. 2006. Barite-bearing cap dolostones of the Toudeni Basin, north-
Peryt, T. M., Hoppe, A., Bechstadt, T., Koster, J., Pierre, C. & west Africa: sedimentary and isotopic evidence for methane
Richter, D. K. 1990. Late Proterozoic aragonitic cement crusts, seepage after a Neoproterozoic glaciation. Precambrian Research,
Bambui Group, Minas Gerais, Brazil. Sedimentology, 37, 279–286. 153, 209–235.
Plumley, P. W., Vance, M. S. & Milazzo, G. 1989. Structural and Stone, D. B. 1989. Palaeogeography and rotations of Arctic Alaska – an
palaeomagnetic evidence for Tertiary bending of the Eastern unresolved problem. In: Kissel, C. & Laj, C. (eds) Palaeomagnetic
Brooks Range Flexure, Alaska. In: Hillhouse, J. W. (ed.) Deep Rotations, Continental Deformation. Kluwer Academic, Norwell,
Structure and Past Kinematics of Accreted Terranes, Geophysical MA.
Monograph 50. American Geophysical Union, Washington, DC. Sweeney, J. F. 1982. Mid-Palaeozoic travels of Arctic Alaska. Nature,
Pruss, S. B., Bosak, T., Macdonald, F. A., McLane, M. & Hoffman, 298, 647–649.
P. F. 2010. Microbial facies in a Sturtian cap carbonate, the Rasthof Wallace, W. K. & Hanks, C. R. 1990. Structural provinces of the North-
Formation, Otavi Group, northern Namibia. Precambrian Research, eastern Brooks Range, Arctic National Wildlife Refuge, Alaska.
181, 187– 198. AAPG Bulletin, 74, 1100– 1118.
Rainbird, R. H., Jefferson, C. W. & Young, G. M. 1996. The early Yoshioka, H., Asahara, Y., Tojo, B. & Kawakami, S. 2003. Systematic
Neoproterozoic sedimentary Succession B of Northwestern Lauren- variations in C, O and Sr isotopes and elemental concentrations in
tia: correlations and palaeogeographic significance. Geological Neoproterozoic carbonates in Namibia: implications for a glacial to
Society of America Bulletin, 108, 454– 470. interglacial transition. Precambrian Research, 124, 69 – 85.
Reed, B. L. 1968. Geology of the Lake Peters area, northeastern Brooks Young, G. M. 1981. The Amundsen Embayment, Northwest Territories:
range, Alaska. US Geological Survey Bulletin 1236, 1 –132. relevance to the Upper Proterozoic evolution of North America. In:
Reiser, H. N. 1971. Northeastern Brooks Range – a surface expression of Campbell, F. H. A. (ed.) Proterozoic Basins of Canada. Geological
the Prudhoe Bay Section. In: Adkison, W. L. & Brosge, M. M. (eds) Survey of Canada Paper 81-10.
Proceedings of the Geological Seminar on the North Slope of Alaska. Young, G. M. 1982. The late Proterozoic Tindir Group, east-central
American Association of Petroleum Geologists, Pacific Section, Alaska; evolution of a continental margin. Geological Society of
Los Angeles, CA. America Bulletin, 93, 759–783.
Reiser, H. N., Dutro, J. T., Brosge, W. P., Armstrong, A. K. & Zhou, C. & Xiao, S. 2007. Ediacaran d13C chemostatigraphy of South
Detterman, R. L. 1970. Progress map, geology of the Sadlerochit China. Chemical Geology, 237, 107–126.
Chapter 35

The Tatonduk inlier, Alaska –Yukon border

FRANCIS A. MACDONALD* & PHOEBE A. COHEN


Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138, USA
*Corresponding author (e-mail: fmacdon@fas.harvard.edu)

Abstract: Glaciogenic deposits of the Rapitan and Hay Creek Groups are exposed in the Tatonduk inlier of east-central Alaska and the
western Yukon. The Rapitan Group ranges in thickness from c. 50 to 700 m with Fe-formation common in the upper 10 m. In the most
distal settings, the Rapitan Group is separated from the diamictite of the Hay Creek Group by over 100 m of sandstone and siltstone;
however, the Hay Creek Group contains large erosive surfaces and cannibalizing breccia, and rarely preserves strata between the two
glaciogenic deposits. The diamictite of the Hay Creek Group is capped by a white- to buff-coloured dolostone with pseudo-teepee struc-
tures, bed-parallel, isopachous sheet-crack cements, and a depleted C-isotope signature. Late Neoproterozoic glacial deposits in the
Tatonduk inlier were formerly assigned to the Tindir Group. To simplify the nomenclature in the northwestern Canadian Cordillera,
the Tindir Group was abandoned and replaced with nomenclature consistent with that of the Windermere Supergroup in the Mackenzie
Mountains. The mixed lithology and anchizone-grade metamorphism distinguish the Rapitan and Hay Creek Groups in the Tatonduk
inlier as attractive future targets for integrated micropalaeontology, geochemistry, palaeomagnetism and geochronology.

Cairnes (1914) referred to Precambrian stratigraphy exposed along followed Mertie (1933) in designating the exposures along the
the international border between the United States and Canada Tatonduk River as the type locality of the diamictite-bearing
as the Tindir Group. Mertie (1930, 1933) later described the strati- units. This choice is due in part to the relative ease of access, but
graphy (and natural history of the region) in remarkable detail and is problematic because of structural complications and the pres-
divided the Tindir Group into seven units. Brabb & Churkin (1969) ence of a major disconformity under the diamictite in the Hay
produced an excellent map of the geology on the Alaskan side of Creek Group. Owing to lack of exposure, the critical features of
the border, while the Tindir Group was mapped on the Yukon both diamictites are not present at one single locality. Along
side of the border by Norris (1982); however, exact correlations Pass Creek (Fig. 35.3), Fe-formation of the Rapitan Group and por-
of specific units of the Tindir Group across the border remained tions of both diamictites are exposed, but the overlying teepee
ambiguous until more recent mapping and compilation (Dover dolomite is not present. Along Hard Luck Creek (Fig. 35.3), the
1992; Van Kooten et al. 1997) and integrated litho- and chemostra- contact between the Fe-formation and underlying diamictite can
tigraphy (Macdonald et al. 2010a, b, 2011; Macdonald & Roots be observed, and the teepee dolomite is present, although the dia-
2010). To improve the consistency of geological maps of the mictite of the Hay Creek Group is not exposed. This lack of
Yukon and promote the synthesis of geological data, Macdonald exposure coupled with the presence of multiple disconformities
et al. (2011) reassigned the Lower Tindir Group to the Pinguicula creates a significant challenge (and many uncertainties) in regional
and Fifteenmile Groups of the Mackenzie Mountains Supergroup stratigraphic correlations.
and the Upper Tindir Group to the Rapitan, Hay Creek and
‘upper’ groups of the Windermere Supergroup (Figs 35.1 & 2).
The Upper and Lower Tindir Groups were first distinguished by Structural framework
Payne & Allison (1981), despite the fact that no unconformable
contacts have been observed between the two (Young 1982). Neoproterozoic diamictites of the Windermere Supergroup are
The Lower Tinder Group consists primarily of dolomite and exposed through an erosional window, referred to as the Tatonduk
shale that are commonly cut with north-trending mafic dykes inlier, which extends across the Canadian border into Alaska
(Van Kooten et al. 1997). Unlike the Lower Tindir Group, the (Fig. 35.2). Present exposures are the product of a mid-Cretaceous
units of the Upper Tindir Group are not cut by mafic dykes (Van to Palaeogene foreland fold-and-thrust belt (Norris 1972) associ-
Kooten et al. 1997). Young (1982) separated the Upper Tindir ated with the underplating of the Yakutat Block (Fuis et al.
Group into five units in ascending stratigraphic order from unit 1 2008). The Tatonduk inlier protrudes from the elbow of the
to unit 5: unit 1 consists primarily of mafic volcanic rocks; unit Ogilvie deflection, where the orientation of folds and thrusts
2, purple mudstones and diamictite, including Fe-formation; unit transition from north-vergent to west-vergent due to Cenozoic
3, an additional massive diamictite; unit 4, platfomal dolomites activity on the Kaltag fault and associated dextral transpressional
and green to light grey shales; unit 5, grey to black shales and lime- structures (Norris 1972). Anchizone-grade metamorphism is
stones. In the updated stratigraphic framework (Macdonald et al. suggested by the presence of economic oil deposits in the overlying
2011), unit 1 is renamed the Pleasant Creek volcanic rocks, unit Phanerozoic strata sourced in part from Palaeozoic shales (Van
2 is the Rapitan Group, unit 3 and the ‘teepee’ dolomite at the Kooten et al. 1997).
base of unit 4 constitute the Hay Creek Group, and the rest of Late Neoproterozoic strata in the Tatonduk inlier are a north-
unit 4 and unit 5 are assigned to the ‘upper’ group. The informal western continuation of the Windermere Supergroup in the Mack-
upper group is a provisional name that encompasses the Sheepbed, enzie (Aitken 1989), Wernecke (Pyle et al. 2004) and Ogilvie
Gametrail, Blueflower and Risky formations in the Mackenzie Mountains (Mustard & Roots 1997), for which deposition com-
Mountains (e.g. Aitken 1989; Dalrymple & Narbonne 1996; Mac- menced with rifting of the northwestern margin of Laurentia
Naughton et al. 2000, 2008) and equivalent strata in the Ogilvie (Stewart 1975). Neoproterozoic strata in the Tatonduk inlier
Mountains (Macdonald et al. 2011). have also been correlated with the Katakturuk Dolomite in the
Exposures of the Windermere Supergroup in the Tatonduk inlier northeastern Brooks Range of Arctic Alaska (Rainbird et al.
are typically limited to the walls of creek beds. Allison et al. (1981) 1996); however, the Katakturuk Dolomite was likely deposited

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 389– 396. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.35
390 F. A. MACDONALD & P. A. COHEN

Ogilvie Mountains surfaces and breccias. The Hard Luck Creek Fault (HLF) marks
an abrupt expansion of the stratigraphy, and can be extended to
Tatonduk Inlier Coal Creek Inlier
Macdonald the SE into the Ogilvie Mountains where it has been named the
et al., 2011

Macdonald &
Young, 1992

et al., 2010b;
Van Kooten
et al., 2010a

Roots, 1997
Roots, 2010
Thompson
Mt. Harper Fault, and roughly marks a former rift shoulder of

Macdonald

et al. 1994;
et al., 1997

Macdonald

Mustard &
the Laurentian margin (Mustard & Roots 1997). The HLF and
other Precambrian faults in the region were inverted during mid-
Cretaceous to Palaeogene shortening (Brabb & Churkin 1969;
Jones Ridge & Funnel Ck., Bouvette Bouvette Van Kooten et al. 1997).
Bouvette Fms., Palaeozoic Adams, Jones Ridge Formation Formation
Cambrian undiff. Hillard, Jones Formation
Undiff. Ridge Fms. PH5 PH5

Stratigraphy
Windermere Supergroup

pCtl

Upper Harper Group


5 5 PH4 PH4

Upper Tindir Group


Upper Tindir Group
Upper Tindir Group

“upper” The stratigraphy of the diamictite-bearing units in the Tatonduk


group
4b inlier was reviewed by Allison et al. (1981) and described in
4 PH3 PH3 detail by Young (1982). Additional measured sections are pre-
4a pCtss sented in Macdonald et al. (2010a, supplementary material,
Hay Creek 3b
Group 3 PH2 2010b, supplementary material, 2011). Diamictite of the Rapitan
3a
Group is underlain by the Pleasant Creek volcanic rocks, which
Rapitan Gp. 2 pCtr 2 PH1 PH1/PH2 are up to 200 m thick and consist chiefly of amygdaloidal pillow
2
Pleasant Creek/Mt. 1 pCtbs 1 MHVC 1
MHVC basalt and cherty hyaloclastic breccia, with minor tuff, shale and
Harper volcanic rocks
conglomerate. The Rapitan Group is composed primarily of fine-
1 717.43 ± 0.14 Ma 2 716.47 ± 0.24 Ma unconformity laminated purple and red mudstone and siltstone that are sprinkled
with dolomite and basaltic lonestones. The upper c. 15 m of the
Fig. 35.1. Nomenclature chart depicting names used by different authors to
Rapitan Group commonly hosts a massive diamictite and
describe Neoproterozoic stratigraphy in the Tatonduk and Coal Creek inliers.
Fe-formation. The Rapitan Group is overlain by the Hay Creek
Dates are U –Pb CA-IDTIMS zircon ages (Macdonald et al. 2010b).
Group, which is composed of up to 150 m of planar laminated
siltstone and sandstone with minor dolomite marl, massive diamic-
tite, and dolomite breccia capped by a white to buff-coloured
on a separate margin, as it is doubtful that the pre-Mississippian dolostone with bed-parallel cements. The uppermost dolostone
Arctic Alaska-Chukotka Plate was part of Laurentia (McClelland, of the Hay Creek Group is less than 5 m thick and rests discon-
1997; Blodgett et al. 2002; Dumoulin et al. 2002). formably on all of the underlying units of the Windermere Super-
Although unconformities have not been observed in the Wind- group and the Pleasant Creek volcanic rocks (Fig. 35.4). The upper
ermere Supergroup of the Tatonduk inlier, multiple disconformi- group is composed largely of black shale with minor allodapic
ties are present in the Hay Creek Group marked by erosional carbonate.

Fig. 35.2. Location map of the NW


Cordillera with Cryogenian and Ediacaran
strata in grey. AACP, Arctic Alaska-
Chukotka Plate. Inset shows the
distribution of the Mackenzie Mountains
and Windermere Supergroups in the
Tatonduk Inlier. Teeth on thrust faults.
THE TATONDUK INLIER, ALASKA– YUKON BORDER 391

Fig. 35.3. Geological map of exposures


along the Tatonduk River and Hard Luck
Creek.

Glaciogenic deposits and associated strata Calcareous concretions and jasper lenses are present in the
Fe-formation, while foreign clasts are rare. Chemical analyses
Rapitan Group of the Fe-formation yield FeO concentrations up to 50%, high
SiO2/Al2O3, and depletions in most other elements, suggesting
The Rapitan Group (formerly Upper Tindir unit 2) is exposed near low clastic deposition and a chemical silica contribution
the international border along Pleasant Creek and in outcrops close (Young 1982).
to the Tatonduk River (Macdonald et al. 2010b). These strata are
chiefly composed of fine-laminated purple and red mudstone and
siltstone speckled with dolomite gravel lonestones. Scattered Hay Creek Group ( formerly units 3a, 3b and 4a of the Upper
throughout the siltstone are thin diamictite beds (,1 m) dominated Tindir Group)
by clasts of dolomite, clastic grit and volcanic cobbles. The
Rapitan Group also contains faceted clasts, boulders with stria- In the most western sections along Pass Creek (T710), the Rapitan
tions, slumped beds, flame structures, groove casts and flute struc- Group is overlain by c. 140 m of planar bedded, siltstone, sand-
tures (Young 1982). The upper c. 15 m of the Rapitan Group stone and marl. These beds lack any evidence of tidal influence,
(Fig. 35.5a; section T709) hosts multiple c. 10-cm-thick beds of lack lonestones, are commonly stacked in fining-upward Bouma
Fe-formation, which are interbedded with a laminated diamictite sequences, and as such, are interpreted as deep-water turbidites.
with bed-penetrating, outsized clasts. The majority of clasts in The turbidite beds are succeeded by an additional diamictite unit,
the Rapitan Group consist of dolomite derived from the underlying but the contact between the two is not exposed. In section T710 the
Fifteenmile Group. Where the upper contact of the Fifteenmile diamictite of the Hay Creek Group is 22 m thick and consists of
Group is exposed, it is overlain by a well-sorted, imbricated car- 10 m of massive diamictite, 10 m of laminated centimetre-beds
bonate matrix conglomerate with abundant dolostone clasts of of siltstone and an additional 2 m of stratified diamictite (Fig.
variable size, interpreted as a debris flow, which is followed by 35.5b). The massive diamictite is clast-dominated with boulders
the parallel-bedded siltstone and sandstone of the Hay Creek of dolomite, and cobbles of Fe-formation, siltstone, conglomerate
Group. The thickness of the Rapitan Group varies greatly from and basalt in a marly pink matrix. The bedded diamictite has the
,50 m near Pleasant Creek to .700 m c. 20 km to the NW same clast composition as the rock below it, but clasts are slightly
(Young 1982). Palaeocurrent measurements in the interbedded smaller with no boulders, and a matrix of purple silt. Approxi-
siltstone suggest a west-facing margin (present coordinates) mately 20 km to the NW of Pass Creek (section Y.S. 9), still
(Young 1982). west of the HLF, the diamictite of the Hay Creek Group is poorly
392 F. A. MACDONALD & P. A. COHEN

Fig. 35.4. Chemo- and lithostratigraphy of the ‘upper’ group in east-central Alaska. Y.S. indicates section from Young (1982). Sections have been projected along the
Hard Luck Fault (i.e. to the NW) to an east– west line, with distances between sections estimated from the projection. Note the change in scale for two sections SW of the
Hard Luck fault, which are approximately four times as thick as sections NE of the fault. See Figure 35.3 for section locations.

exposed, but lonestones c. 2 m in diameter are present, and the dia- several basal Ediacaran cap carbonates globally (Hoffman &
mictite is capped by a buff-coloured dolostone (Young 1982). Macdonald 2010).
Along the Tatonduk River, Young (1982) described a massive,
over 250-m-thick, crudely stratified, purple diamictite interbedded The ‘Upper’ Group
with minor amounts of contorted purple mudstone and lenses
of chert. We assign this diamictite to the Hay Creek Group. The Hay Creek Group is overlain by the ‘upper’ group (formerly
Clasts are up to 0.6 m across, and consist predominantly of units 4b and 5 of the Upper Tindir Group), which consists of as
dolomite from the Fifteenmile Group, with minor limestone, much as 50 m of planar laminated siltstone, sandstone and dolomi-
basalt and chert. Faceted and striated clasts have been described tic marl, and an additional sequence of black shale interbedded
at this locality (Allison et al. 1981), but have not been observed with minor organic-rich limestone (Fig. 35.4; section T707).
elsewhere in the Tatonduk inlier. The diamictite of the Hay Like the underlying units, the ‘upper’ group displays a major stra-
Creek Group is also very thick c. 20 km due north, NE of the tigraphic expansion to the SW ranging from 40–75 m thick in the
HLF (Van Kooten et al. 1997), where it cuts down into the Yukon to c. 700 m thick along the Tatonduk River in Alaska (Mac-
underlying stratigraphy. donald et al. 2010a). Both Hay Creek and the ‘upper’ group strata
Northeast of the HLF, the diamictite of the Hay Creek Group are consistent with a SW-facing margin (present coordinates).
is either absent, or represented by a dolomite (matrix and clast) These ‘upper’ group units form the final clastic-carbonate cycle
breccia. No foreign clasts have been identified with the exception prior to deposition of the Cambrian sandstone (Backbone Ranges
of some clasts of the Pleasant Creek volcanic rocks near the base. Formation) and the commencement of the miogeocline after
This breccia cuts out the underlying units of the Windermere break-up of Rodinia. This group completes Windermere Super-
Group and rests on the Pleasant Creek volcanic rocks (Fig. 35.4). group sedimentation and episodic extension.
The overlying dolomite, ,5 m thick, rests disconformably on
underlying units of the Hay Creek Group as well as the Pleasant
Creek volcanic rocks and Rapitan Group strata. It is white to Boundary relations with overlying and underlying
buff-coloured dolostone with bed-parallel cements (pseudo- non-glacial units
teepee structures of Young 1982). These pseudo-teepees do not
show a polygonal plan-form or a concentration of cements The basal contact of the Rapitan Group was not seen; however,
along the broken pieces, as is typical of teepees that are of a sub- volcanic fragments similar in composition to the underlying Plea-
aerial exposure origin (Kendall & Warren 1987). Instead, sant Creek volcanic rocks are common in the lower half of the
cements are isopachous and bed-parallel, and beds are contorted massive diamictite. The uppermost exposures of the Rapitan
and irregularly buckled, suggesting intraformational detachment Group commonly consist of Fe-formation. At some localities the
during deposition. Similar ‘sheet-crack’ cements are present in upper contact is exposed, and overlain by a well-sorted, dolomite
THE TATONDUK INLIER, ALASKA– YUKON BORDER 393

(a) (b) erosional disconformity. These interpretations are consistent


with the erosional disconformity and breccia in the Hay Creek
Upper Unit 2 Unit 3b Group NE of the HLF where much of the underlying stratigraphy
Hard Luck Creek Pass Creek is missing. Although the top contact is not exposed SW of the HLF,
to the NE the breccia is sharply overlain by the buff-coloured
T709 T710 dolostone that defines the top of the Hay Creek Group.

v
Chemostratigraphy
10 20
Carbon-isotope values of the dolomite at the top of the Hay Creek
Group are extremely depleted, although somewhat variable
(Fig. 35.4; Macdonald et al. 2010a, supplementary data). Near
8 16 the international border (section T708), values are consistently
between –3‰ and –3.5‰, whereas near Hard Luck Creek (sec-
tions T701 and T709), values are more scattered, ranging from
–6‰ to þ2‰. In the ‘upper’ group, C-isotope values tend to be
(metres)

6 12 enriched (Fig. 35.4).


Carbon and Sr isotopes were reported from unit 5 of the Upper
Tindir Group in Canada (Kaufman et al. 1992); however, these
sections are actually correlative with the Fifteenmile Group
v
(Macdonald et al. 2010a, b, 2011). The lowest Sr-isotope values
4 8 from these sections are near 0.7064 (Macdonald et al. 2010a), a
value that is typical for pre-Sturtian carbonate rocks (Halverson
et al. 2007).

2 4 Palaeolatitude and palaeogeography

v No palaeomagnetic studies have been reported from the Tatonduk


inlier. Although the lithologies of Pleasant Creek volcanic rocks
0 0 and the Rapitan Group (basalts and Fe-rich clastic sediments,
respectively) are ideal targets, most palaeomagnetic studies in
northern Alaska are compromised by a pervasive Late Cretaceous
overprint (Plumley et al. 1989). However, palaeopoles on the
diamictite clast size Rapitan Group in the Mackenzie Mountains yield a palaeolatitude
lithologies of 6 + 48 (Park 1997). This pole is consistent with the grand mean
massive boulder pole on the contemporaneous Franklin LIP (Denyszyn et al. 2009),
unsorted and demonstrates that Laurentia straddled the equator during depo-
cobble sition of the Rapitan Group (Macdonald et al. 2010b). Although
stratified Laurentia remained at low latitudes until 615 Ma according to
gravel the controversial Long Range Dyke pole (Murthy et al. 1992;
imbricated Hodych et al. 2004), Laurentia appears to have migrated to high
conglomerate latitudes by 590 Ma (Murthy 1971). McCausland et al. (2007)
clast provide an excellent review of the Ediacaran palaeomagnetic
matrix compostion data from Laurentia.
composition v volcanic
iron- Geochronological constraints
formation clastic
dolomite Radiometric data
marl carbonate
The NNW-trending mafic dykes that intrude the Fifteenmile Group
siltstone iron- in the Tatonduk inlier have yielded a wide range of K –Ar biotite
formation
and whole-rock ages: 532 + 11 Ma, 572 + 16 Ma, 588 + 14 Ma
and 644 + 18 Ma (Van Kooten et al. 1997). These dykes have
Fig. 35.5. (a) Stratigraphy of the Rapitan Group exposed along Hard Luck not been observed to intrude any of the Windermere Supergroup
Creek, showing diamictite and its relationship to the Fe-formation. (b) units and are possibly feeder dykes for the Pleasant Creek
Stratigraphy of the Hay Creek diamictite along Pass Creek, including diamictite volcanic rocks. Because alteration can lead to argon loss (Westcott
and associated lithofacies. See Figure 35.3 for section locations. 1966), these dates are minimum age constraints for the Fifteen-
mile Group.
Along strike to the east in the Coal Creek inlier (Fig. 35.3),
clast, dolomite matrix diamictite, interpreted as a debris flow, Rapitan Group correlatives rest above the Mt. Harper volcanic
which is followed by the parallel-bedded siltstone and sandstone complex (MHVC). Zircons extracted from rhyolite in member D
with Bouma sequences of the Hay Creek Group. of the MHVC were dated at 717.43 + 0.14 Ma, and zircons
Young (1982) interpreted the diamictite of the Hay Creek Group from a tuff within the Rapitan Group correlatives were dated at
along the Tatonduk River as allochthonous in the sedimentary 716.47 + 0.24 Ma (U – Pb ID-TIMS; Macdonald et al. 2010b).
sense, having been derived from a more proximal setting to the In the Rapitan Group of the Mackenzie Mountains, a clast of leu-
east of the HLF in a massive slope failure, with an unconformable cogranite in the Rapitan Group has a U –Pb TIMS bulk zircon age
basal contact. Allison et al. (1981) interpreted this contact as an of 755 + 18 Ma (Ross & Villeneuve 1997).
394 F. A. MACDONALD & P. A. COHEN

Micropalaeontology geochronology of correlative rocks in the Coal Creek inlier


(Macdonald et al. 2010b, 2011).
Fossils interpreted as microscopic flatworm impressions by
Allison (1975), but possibly sponge spicules (Andy Knoll
pers. comm.), have been described in the shale interbedded Discussion
with basalt in the lower portion of the Windermere Supergroup.
It is not clear if these samples were collected from the Pleasant Depositional setting
Creek volcanic rocks, Rapitan Group or the Hay Creek Group, as
they are reported in the ‘basalt and red beds’ unit of Brabb & The Windermere Supergroup contains two Cryogenian glaciogenic
Churkin (1969), which incorporates all three. Recent biomarker deposits (Rapitan Group and the diamictite of the Hay Creek
work suggests that the presence of sponge spicules in the Wind- Group) separated by c. 140 m of non-glacial strata (Fig. 35.4). A gla-
ermere Supergroup is not inconsistent with a Cryogenian age ciomarine depositional setting for the Rapitan Group is suggested
(Love et al. 2009). Microfossils have also previously been by the presence of faceted and striated clasts, bed-penetrating
described in chert of Fifteenmile Group (in strata previously mis- dropstones, and common outsized and exotic clasts, along with evi-
mapped as unit 5 of the Upper Tindir Group; Macdonald et al. dence for subaqueous slumping in the form of graded grain flows
2010a), including cyanobacterial coccoids, acritarchs such as and debris flows. Young (1988) ascribed the Fe-formations of the
Trachyhystrichosphaera, vase-shaped microfossils and unique, Rapitan Group, and equivalent strata in the Tatonduk inlier, to
enigmatic siliceous scales (Allison 1980; Allison & Hilgert rift-related hydrothermal activity. However, these Fe-formations
1986; Allison & Awramik 1989). More recent studies have are intimately associated with well-developed dropstones. Presum-
extracted these microfossils from the surrounding carbonate ing these diamictites were deposited during the terminal ice retreat,
rock, and have demonstrated that they are composed of phos- and presuming an increased solubility of iron in the ocean due
phate rather than silica, suggesting a green algae taxonomic affi- to low oxygen levels under long-lived sea ice (Martin 1965),
nity (Cohen et al. 2011). A pre-717 Ma for these fossils is Fe-formation can be attributed to an influx of oxygenated fresh
supported by C- and Sr-isotope correlations (Macdonald et al. water concentrated at the termini of ice streams (Kirschvink 1992).
2010a), the presence of dykes cutting the sections that are The parallel bedded sandstones and siltstones in the lower
co-magmatic with the Pleasant Creek volcanic rocks, and portion of the Hay Creek Group are interpreted as turbidites as

Fig. 35.6. Neoproterozoic stratigraphy in


the Tatonduk inlier and correlations with
the Windermere Supergroup exposed c.
500 km east in the Mackenzie Mountains of
Canada. The schematic stratigraphy of the
Windermere Supergroup is modified from
Halverson et al. (2008). IRD mudstones are
laminated fine-grained sediments with
lonestones interpreted as ice rafted debris.
THE TATONDUK INLIER, ALASKA– YUKON BORDER 395

they lack any evidence of wave action or traction currents, and Allison, C. W. A. 1975. Primitive fossil flatworm from Alaska: new
beds are stacked in fining-upward Bouma cycles. These green-grey evidence bearing on ancestry of the Metazoa. Geology, 3, 649– 653.
turbidites are distinguished from the red and purple grain flows in Allison, C. W. A. 1980. Siliceous microfossils from the lower Cambrian
the Rapitan Group by the lack of lonestones. of northwest Canada: possible source for biogenic chert. Science,
Young (1982) interpreted the diamictite of the Hay Creek Group 211, 53– 55.
SW of the HLF as the product of a massive slope failure, and ident- Allison, C. W. A. & Hilgert, J. W. 1986. Scale micro-fossils from the
ified an unconformity at the base. Nonetheless, there is evidence of Early Cambrian of northwest Canada. Journal of Paleontology, 60,
a glaciomarine influence on deposition with the presence of 973– 1015.
Allison, C. W. A. & Awramik, S. M. 1989. Organic-walled microfossils
striated clasts (Allison et al. 1981), and a well-bedded diamictite
from earliest Cambrian or latest Proterozoic Tindir Group rocks,
with dropstones near the top of Hay Creek Group along Pass northwest Canada. Precambrian Research, 43, 253–294.
Creek (Fig. 35.5b). A depositional setting along the upper slope, Allison, C. W. A., Young, G. M., Yeo, G. M. & Delaney, G. D. 1981.
with diamictite being the product of syn-glacial slope failure, is Glaciogenic rocks of the Upper Tindir Group, east-central Alaska. In:
consistent with the interpretation that the expansion of the strati- Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene
graphy of all of the Windermere Supergroup across the HLF Glacial Record. Cambridge University Press, Cambridge, 720–723.
roughly approximates the slope – shelf transition. Blodgett, R. B., Rohr, D. M. & Boucot, A. J. 2002. Paleozoic links
From outcrops along the Tatonduk River, Allison et al. (1981) among some Alaskan accreted terranes and Siberia based on mega-
cited deformation in Rapitan Group sediments near the overlying fossils. In: Miller, E. L., Grantz, A. & Klemperer, S. L. (eds) Tec-
disconformity to suggest that the underlying stratigraphy was tonic Evolution of the Bering Shelf – Chukchi Sea –Arctic Margin and
unconsolidated at the time of deposition of the upper diamictite. Adjacent Landmasses. Geological Society of America Special Paper
However, slumps and syn-sedimentary folds are common through- 360, Boulder, Colorado, 273– 290.
out the Rapitan Group (Young 1982), so this deformation in the Brabb, E. E. & Churkin, M. J. 1969. Geologic Map of the Charley River
Rapitan Group may be syn-sedimentary or the product of grounded Quadrangle, East-Central Alaska. U.S. Geological Survey Map
ice during the Rapitan glaciation. I-573.
Cairnes, D. D. 1914. The Yukon-Alaska International Boundary Between
Regional correlations Porcupine and Yukon Rivers. Geological Survey of Canada Memoir
67, Map 140-A.
Another line of evidence that the upper diamictite represents a Cohen, P. A., Schopf, J. W., Butterfield, N. J., Kudryaytsev, A. &
second, later glacial event is that it is overlain with a buff-coloured Macdonald, F. A. 2011. Phosphate biomineralization in mid-
Neoproterozoic protists. Geology, 39, 539– 542.
dolostone. The latter is considered a basal Ediacaran cap carbonate
Dalrymple, R. W. & Narbonne, G. M. 1996. Continental slope sedi-
as it contains bed-parallel, isopachous sheet-crack cements mentation in the Sheepbed Formation (Neoproterozoic, Windermere
(Hoffman & Macdonald 2010) and a C-isotope profile similar to Supergroup), Mackenzie Mountains, N.W.T. Canadian Journal of
that of thin basal Ediacaran cap carbonates in China (Zhou & Earth Sciences, 33, 848–862.
Xiao 2007), Australia (Kennedy 1996) and Namibia (Halverson Denyszyn, S. W., Halls, H. C., Davis, D. W. & Evans, D. A. D. 2009.
et al. 2005). Paleomagnetism and U– Pb geochronology of Franklin dykes in High
Lithologically, the Rapitan Group in the Tatonduk inlier is very Arctic Canada and Greenland: a revised age and paleomagnetic pole
similar to the clast poor siltstone and Fe-formation of the Sayunei for constraining block rotations in the Nares Strait region. Canadian
Formation in the Mackenzie Mountains (Young 1976; Yeo 1984). Journal of Earth Sciences, 46, 689– 705.
The diamictite of the Hay Creek Group can be correlated with the Dover, J. H. 1992. Geologic map and fold and thrust belt interpretation of
Ice Brook Formation (Fig. 35.6). This correlation is particularly the south eastern part of the Charley River Quadrangle, East Central
attractive because the carbonate at the top of the Hay Creek Alaska. US Geological Survey Miscellaneous Investigations, Map
Group shares sedimentological and isotopic characteristics with I-1942, scale 1:100,000, 2 sheets.
the Ravensthroat cap dolomite in the Mackenzie Mountains Dumoulin, J. A., Harris, A. G., Gagiev, M., Bradley, D. C. &
(Aitken 1991; James et al. 1999). Repetski, J. E. 2002. Lithostratigraphic, conodont, and other
Regional correlations of Neoproterozoic strata in NW Canada faunal links between lower Paleozoic strata in northern and central
are particularly important in light of the recent geochronological Alaska and northeastern Russia. In: Miller, E. L., Grantz, A. &
constraints provided by volcanic tuffs interbedded with the Fif- Klemperer, S. L. (eds) Tectonic Evolution of the Bering Shelf –
teenmile and Rapitan Groups in the Ogilvie Mountains (Macdo- Chukchi Sea –Arctic Margin and Adjacent Landmasses. Geological
nald et al. 2010b). Coupled with the robust palaeomagnetic poles Survey of America Special Paper 360, Boulder, Colorado, 291– 312.
Evans, D. A. D. 2000. Stratigraphic, geochronological and paleomagnetic
in NW Canada (Evans 2000; Denyszyn et al. 2009) and the rich
constraints upon the Neoproterozoic climatic paradox. American
micropalaeontological record in the Tatonduk inlier (Allison & Journal of Science, 300, 347–433.
Arwimik 1989; Macdonald et al. 2010b), the inliers of the Fuis, G. S., Moore, T. E. et al. 2008. Trans-Alaska Crustal Transect
Ogilvie Mountains hold great promise of a calibrated record of and continental evolution involving subduction underplating and
tectonics, climate, chemistry and life in Cryogenian oceans. synchronous foreland thrusting. Geology, 36, 267–270.
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
We are grateful to the Yukon Geological Survey, the Department of Earth and Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
Planetary Sciences, Harvard University, P. Hoffman, A. Knoll and the NSF for isotope record. Geological Society of America Bulletin, 117,
support. We thank R. B. Blodgett and D. Jones for helpful discussions. We also 1181–1207.
thank the YUCH National Park office in Eagle for allowing access to the Halverson, G. P., Dudás, F. O., Maloof, A. C. & Bowring, S. A. 2007.
field area. This represents a contribution of the IUGS- and UNESCO-funded Evolution of the 87Sr/86Sr composition of Neoproterozoic Seawater.
IGCP (International Geoscience Programme) project #512. Palaeogeography, Palaeoclimatology, Palaeoecology, 256, 103–129.
Hodych, J. P., Cox, R. A. & Kosler, J. 2004. An equatorial Laurentia at
References 550 Ma confirmed by Grenvillian inherited zircons dated by LAM
ICP-MS in the Skinner Cove volcanics of western Newfoundland:
Aitken, J. D. 1989. Uppermost Proterozoic formations in central Macken- implications for inertial interchange true polar wander. Precambrian
zie Mountains, Northwest Territories. Geological Survey of Canada Research, 129, 93– 113.
Bulletin, 368, 1 –26. Hoffman, P. F. & Macdonald, F. A. 2010. Sheet-crack cements and
Aitken, J. D. 1991. The Ice Brook Formation and Post-Rapitan, Late early regression in Marinoan (635 Ma) cap dolostones: regional
Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. benchmarks of vanishing ice-sheets? Earth and Planetary Science
Geological Survey of Canada Bulletin, 404, 1 – 43. Letters, 300, 374– 384.
396 F. A. MACDONALD & P. A. COHEN

James, N. P., Narbonne, G. M. & Kyser, K. T. 1999. Neoproterozoic Cap Murthy, G. S. 1971. The paleomagnetism of diabase dykes from the
Carbonate Facies; Mackenzie Mountains, NW Canada; Abiotic Pre- Grenville Province. Canadian Journal of Earth Sciences, 8,
cipitation and Global Glacial Meltdown. Geological Society of 802– 812.
America Abstracts with Programs, 31, 487. Murthy, G. S., Gower, C. F., Tubrett, M. & Patzold, R. 1992.
Kaufman, A. J., Knoll, A. H. & Awramik, S. M. 1992. Biostratigraphic Paleomagnetism of Eocambrian Long Range dykes and Double
and chemostratigraphic correlation of Neoproterozoic sedimentary Mer Mormation from Labrador, Canada. Canadian Journal of
successions: Upper Tindir Group, northwestern Canada, as a test Earth Sciences, 29, 1224–1234.
case. Geology, 20, 181– 185. Mustard, P. S. & Roots, C. F. 1997. Rift-related volcanism, sedimen-
Kendall, C. G. S. C. & Warren, J. 1987. A review of the origin and tation and tectonic setting of the Mount Harper Group, Ogilvie
setting of teepees and their associated fabrics. Sedimentology, 34, Mountains, Yukon Territory. Geological Survey of Canada Bulletin,
1007– 1027. 492.
Kennedy, M. J. 1996. Stratigraphy, sedimentology and isotope geo- Norris, D. K. 1972. En echelon folding in the northern Coldillera of
chemistry of Australian Neoproterozoic postglacial cap dolostones: Canada. Bulletin of Canadian Petroleum Geology, 20, 634– 642.
deglaciation, d13C excursions and carbonate precipitation. Journal Norris, D. K. 1982. Geology, Ogilvie River, Yukon Territory, Geological
of Sedimentary Research, 66, 1050–1064. Survey of Canada, Map 1526A, 1:250,000 scale.
Kirschvink, J. L. 1992. Late Proterozoic low-latitude global glaciation: Park, J. K. 1997. Paleomagnetic evidence for low-latitude glaciation
the snowball earth. In: Schopf, J. W. & Klein, C. (eds) The Proter- during deposition of the Neoproterozoic Rapitan Group, Mackenzie
ozoic Biosphere. Cambridge University Press, Cambridge, 51 –52. Mountains, N.W.T., Canada. Canadian Journal of Earth Sciences,
Love, G. D., Fike, D. A. et al. 2009. Fossil steroids record the appearance 34, 34 – 49.
of Demospongiae during the Cryogenian period. Nature, 457, Payne, M. W. & Allison, C. W. A. 1981. Paleozoic continental-margin
718– 722. sedimentation in east-central Alaska. Geology, 9, 274–279.
Macdonald, F. A. & Roots, C. F. 2010. Upper Fifteenmile Group in the Plumley, P. W., Vance, M. S. & Milazzo, G. 1989. Structural and
Ogilvie Mountains and correlations of early Neoproterozoic strata in paleomagnetic evidence for Tertiary bending of the Eastern
the northern Cordillera. In: MacFarlane, K. E., Weston, L. H. & Brooks Range Flexure, Alaska. In: Hillhouse, J. W. (ed.) Deep
Blackburn, L. R. (eds) Yukon Exploration and Geology 2009. Structure and Past Kinematics of Accreted Terranes, Geophysical
Yukon Geological Survey, Whitehorse, YT, 237–252. Monograph 50. American Geophysical Union, Washington, DC,
Macdonald, F. A., Cohen, P. A., Dudás, F. O. & Schrag, D. P. 2010a. 127– 150.
Early Neoproterozoic scale microfossils in the Lower Tindir Group Pyle, L. J., Narbonne, G. M., James, N. P., Dalrymple, R. W. &
of Alaska and the Yukon Territory. Geology, 38, 143– 146. Kaufman, A. J. 2004. Integrated Ediacaran chronostratigraphy,
Macdonald, F. A., Schmitz, M. D. et al. 2010b. Calibrating the Wernecke Mountains, northwestern Canada. Precambrian Research,
Cryogenian. Science, 327, 1241– 1243. 132, 1– 27.
Macdonald, F. A., Smith, E. F., Strauss, J. V., Cox, G. M., Halverson, Rainbird, R. H., Jefferson, C. W. & Young, G. M. 1996. The early Neo-
G. P. & Roots, C. F. 2011. Neoproterozoic and early Paleozoic cor- proterozoic sedimentary Succession B of Northwestern Laurentia:
relations in the western Ogilvie Mountains, Yukon. In: MacFar- correlations and paleogeographic significance. Geological Society
lane, K. E., Weston, L. H. & Blackburn, L. R. (eds) Yukon of America Bulletin, 108, 454– 470.
Exploration and Geology 2010. Yukon Geological Survey, White- Ross, G. M. & Villeneuve, M. E. 1997. U –Pb geochronology of stranger
horse, 161– 182. stones in Neoproterozoic diamictites, Canadian Cordillera: impli-
MacNaughton, R. B., Narbonne, G. M. & Dalrymple, R. W. 2000. cations for provenance and ages of deposition. Radiogenic age and
Neoproterozoic slope deposits, Mackenzie Mountains, northwestern isotopic studies, Report 10, Geological Survey of Canada, Current
Canada: implications for passive-margin development and Ediacaran Research 1997-F, 141– 155.
faunal ecology. Canadian Journal of Earth Sciences, 37, 997–1020. Stewart, J. H. 1975. Initial deposits in the Cordilleran geosyncline:
MacNaughton, R. B., Roots, C. F. & Martel, E. 2008. Neoprotero- evidence of a late Precambrian (,850 m.y.) continental seperation.
zoic– (?)Cambrian lithostratigraphy, northeast Sekwi Mountain map Geological Society of America Bulletin, 83, 1345–1360.
area, Mackenzie Mountains, Northwest Territories: new data from Van Kooten, G. K., Watts, A. B. et al. 1997. Alaska Division of Geo-
measured sections. Geological Society of Canada, Current Research logical and Geophysical Surveys, Report of Investigations 96-6A, 3
2008, 16, 1– 17. sheets, scale 1:125,000, Fairbanks.
Martin, H. 1965. Beobachtungen zum Problem der jung-präkambrischen Westcott, M. R. 1966. Loss of argon from biotite in a thermal meta-
Glazialen Ablagerungen in Südwestafrika. (Observations concerning morphism. Nature, 210, 83– 84.
the problem of the late Precambrian glacial deposits in South West Yeo, G. M. 1984. The Rapitan group: relevance to the global association
Africa.) Geologische Rundschau, 54, 115– 127. of Late Proterozoic glaciation and iron-formation. PhD thesis,
McCausland, P. J. A., Van der Voo, R. & Hall, C. M. 2007. University of Western Ontario, London, Ontario.
Circum-Iapetus paleogeography of the Precambrian – Cambrian Young, G. M. 1976. Iron-formation and glaciogenic rocks of the Rapitan
transition with a new paleomagnetic constraint from Laurentia. Group, Northwest Territories, Canada. Precambrian Research, 3,
Precambrian Research, 156, 125– 152. 137– 158.
McClelland, W. C. 1997. Detrital zircon studies of the Proterozoic Young, G. M. 1982. The late Proterozoic Tindir Group, east-central
Neruokpuk Formation, Sadlerochit and Franklin Mountains, northern Alaska; evolution of a continental margin. Geological Society of
Alaska. Geological Society of America Abstracts with Programs, 25, 28. America Bulletin, 93, 759–783.
Meertie, J. B. 1930. Geology of the Eagle-Circle district, Alaska. US Young, G. M. 1988. Proterozoic plate tectonics, glaciation and
Geological Survey Bulletin, 816, 121–122. iron-formations. Sedimentary Geology, 58, 127– 144.
Meertie, J. B. 1933. The Tatonduk-Nation district, Alaska. US Geologi- Zhou, C. & Xiao, S. 2007. Ediacaran d13C chemostatigraphy of South
cal Survey Bulletin, 836-E, 345–454. China. Chemical Geology, 237, 107– 126.
Chapter 36

Neoproterozoic glacial record in the Mackenzie Mountains, northern Canadian Cordillera

P. F. HOFFMAN1,2* & G. P. HALVERSON3,4


1
Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138, USA
2
School of Earth and Ocean Sciences, University of Victoria, Victoria, BC V8W 2Y2, Canada
3
School of Earth and Environmental Sciences, The University of Adelaide, North Terrace, Adelaide, SA 5005, Australia
4
Present address: Department of Earth and Plantary Sciences, McGill University, Montreal, QC H3A 2A7, Canada
*Corresponding author (e-mail: paulfhoffman@yahoo.com)

Abstract: In the Mackenzie Mountains, an arcuate foreland thrust-fold belt of Late Cretaceous–Paleocene age in the northern Canadian
Cordillera, two discrete glacial– periglacial sequences of Cryogenian age (the Rapitan Group and the Stelfox Member of the Ice Brook
Fm.) are separated by c. 1.0 km of non-glacial strata. The older Rapitan diamictite occurs in an amagmatic rift basin; the younger Stelfox
diamictite occurs on a passive-margin continental slope.
The Rapitan Group consists of three formations. The lower Mount Berg Fm. is a complex of diamictites and conglomerates of limited
extent. The middle Sayunei Fm. is a thick sequence of maroon-coloured mudrocks hosting innumerable graded layers of silt- and fine-
grained sandstone. It lacks wave- or traction current-generated bedforms, and is lightly sprinkled with granule aggregates (‘till pellets’)
and lonestones of dolostone and rare extrabasinal granitoids. It is capped by a hematitic Fe-formation that was reworked into the discon-
formably overlying Shezal diamictite. The Shezal Fm. is a complex of olive-green coloured boulder diamictites with subordinate, dark-
grey shales, siltstones and parallel-sided sandstones. Some of the boulders are faceted and striated, and include dolostone, quartzite, silt-
stone and gabbro in declining order of abundance. Diamictite terminates abruptly at the top of the Shezal Fm., which is sharply overlain
by dark shales or by ,52 m of fetid, dark-grey, 13C-depleted limestone with graded bedding.
The Stelfox Member is dominated by non-stratified, carbonate-clast diamictite with faceted and striated clasts, locally associated with sub-
ordinate, well-laminated shales containing till pellets and ice-rafted dropstones. It is thin or absent on the palaeocontinental shelf, but thickens
seaward (southwestward) on the palaeocontinental slope. A thin clay drape separates it from a laterally continuous post-glacial ‘cap’ dolo-
stone, which is a very pale coloured, micro- to macropeloidal dolostone with low-angle cross-laminae, giant wave ripples and local bioherms
of corrugated stomatolites. In the NW, the dolostone is followed by reddish and greenish marls, followed by black shale of the Sheepbed Fm.
In the SE, the dolostone is overlain by pink or grey limestones with well-developed sea-floor cements pseudomorphic after aragonite. In this
area, the top of the dolostone is ferruginous and contains digitate rosettes of sea-floor barite cement, variably calcitized. The dolostone–
limestone contact is perfectly conformable, and synclinal structures previously intepreted as karst features are tectonic in origin.
The grand mean palaeomagnetic pole for the well-studied Franklin Large Igneous Province (c. 718 Ma) of Arctic Laurentia, coeval with
the basal Rapitan Group in the Mount Harper area, Yukon Territory, places the Mackenzie Mountains firmly in the tropics, at 18 + 38N
palaeolatitude, at the onset of the Rapitan glaciation.
Carbon (d13C), oxygen (d18O) and strontium (87Sr/86Sr) isotopes have been measured in carbonates bracketing the Rapitan and Stelfox
diamictites. Sulphur isotope data (d34S) have been obtained from carbonate-associated sulphate and barite above the younger diamictite, and
calcium isotope data (d44Ca) from the younger carbonate itself. The results are broadly consistent with data from other areas. Iron isotope
(d57Fe) and cerium anomaly (Ce/Ce*) values increase systematically upwards through the Sayunei Fe-formation, supporting an interpret-
ation that deposition occurred within a redox chemocline through which the basin floor descended as a consequence of isostatic loading by the
advancing Shezal ice sheet.

Supplementary material: Data are available at http://www.geolsoc.org.uk/SUP18470.

Glacial marine diamictites (‘tillites’) were recognized in the A second, thinner, glaciogenic unit was discovered .1000 m
northwestern Mackenzie Mountains (Fig. 36.1) and to the west stratigraphically above the Rapitan Group by Aitken (1991a, b).
in the Yukon Territory by Shell Oil Company geologists in He defined it as the Stelfox Member (,272 m) of the Ice Brook
1958 (Ziegler 1959), during the course of economic assessment Fm., the type section of which is located in the Sayunei Range at
of associated Fe-formations. At the time, they were thought to 648080 0700 N, 1298000 3800 W (Aitken 1991b). It is overlain by a dis-
be early Palaeozoic in age (Ziegler 1959). They were referred to tinctive carbonate unit, the ‘Tepee dolostone’ of Eisbacher (1978,
as the Rapitan Group by Green & Godwin (1963). Correlative 1981a) and Aitken (1991b), subdivided into the Ravensthroat and
strata were studied in the Hayhook Lake area (Fig. 36.1), 300 km Hayhook formations by James et al. (2001). The dolomite features
to the SE, by Upitis (1966), Gabrielse et al. (1973), Young unusual sedimentary structures, including sheet-crack cements,
(1976) and Eisbacher (1978), where they were recognized as reverse-graded peloids, giant wave ripples, corrugated stromato-
being of Neoproterozoic (Hadrynian) age. Gabrielse et al. (1973) lites and sea-floor cements (crystal fans) originally composed of
erected a type section for the Rapitan Group near Hayhook aragonite and barite (Eisbacher 1981a; Aitken 1991b; James
Lake at 63834’03”N, 127802’41”W. It includes the Sayunei, et al. 2001; Hoffman & Schrag 2002; Allen & Hoffman 2005;
Shezal and Twitya formations (Fig. 36.2). Eisbacher (1978) Hoffman & Macdonald 2010). It has many similarities with post-
enlarged the Rapitan Group to include the Keele Fm. because glacial cap dolostones on other palaeocontinents (Hoffman et al.
of its gradational relationship with the Twitya Fm. However, 2011), which allow the base of the Ediacaran Period to be recog-
Yeo (1981) and most subsequent workers follow the original nized globally (Knoll et al. 2006).
definition (Green & Godwin 1963) and limit the Rapitan Correlatives of the Mackenzie Mountains and Windermere super-
Group to glaciogenic strata beneath the Twitya Fm. Additional groups, including both glacial horizons, are discontinuously exposed
stratigraphic and sedimentological studies of the Rapitan in Wernecke and Ogilvie in central Yukon Territory (Macdonald &
Group along the 400 km arc of the Mackenzie Mountains were Roots 2009; Macdonald et al. 2010b), and in the Tatonduk inlier
carried out by Eisbacher (1981a, b, 1985) and Yeo (1981, 1984, of east-central Alaska (Young 1982; Macdonald et al. 2010a;
1986). Macdonald & Cohen 2011).

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 397– 411. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.36
398 P. F. HOFFMAN & G. P. HALVERSON

134 oW 132 oW 128oW 126o W


130oW
M Norman
platfor a
ck Wells
Iron SNAKE RIVER
Ck
m en
mou zi
Cranswick R.
ntain e
Ri
v 65oN

e
Corn Gayna

r
Creek R. P
la

fr
Arctic te
Red R. a u

on
Stone Knife R. t

Th
Shale Lake

ru
st
Twitya R. 64oN
Boomerang L.
Moose Horn R.

Stelfox Mtn Hayhook


June Lake Lake
REDSTONE
Ravens-
This Sekwi throat R.
figure Brook
63oN

PLATEAU Coates L.
Windermere Supergroup Thunder-
cloud R. Fig. 36.1. Outcrop distribution of the
Sheepbed Rapitan Group and the Windermere
Rapitan Group Supergroup in the Mackenzie Mountains,
Creek
measured sections showing locations of measured sections
forming the basis of this report.

This chapter is a synopsis of previous work, augmented by River (648490 2300 N, 1308270 1100 W: 22 m), Stone Knife River
seven new sections of the Rapitan Group, measured by P.F.H. (648410 5300 N, 1298530 1900 W: 114 m), Shale Lake (648320 4100 N,
in the Iron Creek (658030 1500 N, 1338070 3100 W: .775 m), Gayna 1298220 3700 W: 995 m), Boomerang Lake (638460 4600 N, 1278280 0500 :
.236 m), Hayhook Lake (638340 3800 N, 1278050 3400 W: 997 m)
and Ravensthroat River (638130 5700 N, 1278030 1800 W: 230 m)
Group Formation Member Reference areas; and 13 sections of the Stelfox diamictite (all ,35 m)
Gabrielse and its post-glacial carbonate sequence at Cranswick River
Sheepbed (658050 5400 N, 1328260 1700 W), Arctic Red River (648560 0500 N,
et al. (1973) 1318030 3000 W), Gayna River (648490 2200 N, 1308280 2700 W), Stone-
Hayhook James et al. knife River (648400 3600 N, 1298530 4800 W), Shale Lake (648310 1200 N,
(2001) 1298290 0900 W), Twitya River (648130 0700 N, 1288380 1100 W),
Ravensthroat
Moose Horn River (638570 1300 N, 1278300 3900 W), Hayhook
Stelfox Lake (638340 1700 N, 1278110 4400 W), Stelfox Mountain (638360 0100 N,
Ice Brook Aitken 1278500 4700 W; 638350 4600 N, 1278530 0400 W; 638350 6000 N,
Delthore
(1991b) 1278530 5700 W; 638350 3700 N, 1278550 1100 W) and Ravensthroat
Durkan River (638170 0500 N, 1278080 4000 W). The named locations are
Keele indicated in Figure 36.1. The coordinates refer to the base of the
section.
Twitya
Eisbacher
Shezal
(1978)
Rapitan Sayunei Structural framework
Mt Berg
The Mackenzie Mountains are the physiographic expression of an
Copper Cap Gabrielse arcuate, NE-vergent, foreland thrust –fold belt of Late Cretac-
Coates Lake Redstone River et al. (1973) eous –Paleocene age in the northern Canadian Cordillera (Aitken
& Long 1978; Aitken 1982; Narbonne & Aitken 1995). Late
Thundercloud Jefferson (1983) Neoproterozoic strata of the Windermere Supergroup (Fig. 36.3)
Little Dal Aitken (1981) are principally exposed in the hanging wall of the Plateau Thrust
system (Fig. 36.1). Their subhorizontal (Redstone Plateau) or
Fig. 36.2. Stratigraphic nomenclature and defining references for units SW-dipping attitude depends on their position above a flat or
associated with Cryogenian glaciogenic strata (black with inverted triangles) in ramp in the thrust plane. Outcrop-scale strain is heterogeneous;
the Mackenzie Mountains. structurally intact, low-strain sections can be located by
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 399

(a) (b) (c) (d)


Mackenzie Mountains composite Hayhook Lake (type sec.) Hayhook Lake (ref. sec.) Iron Creek
63o34’03”N, 127o02’41”W 63o34’38”N, 127o05’34”W 65o03’15”N, 133o07’31”W
0 0.0
Sekwi (km)
CAMBRIAN

(m) (m)
(km) -0.1

Shezal
Vampire
-50 -50
Backbone -0.2
Ranges

Keele
Ingta -0.3
Risky -100 -100
-2 -0.4
EDIACARAN

Blueflower
-0.5 fs BIF
-150 -150
WINDEREMERE SUPERGROUP

Gametrail

olive green (Little Dal-Katherine derivation)


-0.6

Sayunei
fs
Sheepbed -0.7 -200 -200
fs
Ravensthroat BIF
-4 Ice Stelfox -0.8 fs
Brook BIF
Twitya

Keele -0.9 -250 -250

Shezal
fs
CRYOGENIAN

-1.0 BIF

maroon
Twitya -300 -300
-1.1

Mount Berg
unnamed cap
Shezal -1.2
-6 Rapitan -350 -350
Group Sayunei -1.3 fs

Mt Berg
-1.4
Shezal

Coates Lake -400 -400

Coates?
Group -1.5

-1.6 -450 -450


-8 BIF
MACKENZIE MTNS SUPERGROUP

Little -1.7
LEGEND
Dal
-1.8 -500 mud-matrix
Group R diamictite
BIF
Sayunei

R -1.9
wacke-matrix
-2.0 -550 diamictite
TONIAN(?)

-10 Katherine
maroon

Group -2.1 scaly-matrix


Sayunei

-600 diamictite
-2.2

Tsezotene 780 Ma mudstone with


(U-Pb) -2.3
silty turbidites,
Coates

-650
-2.4 debrites, sparse
'Unit H1' ice-rafted debris
-12
R fs

carbonate reefal sulphate mudstone siltstone sandstone basalt- haematitic haematite- flooding
carbonate evaporite gabbro mudstone jaspilite surface
Fig. 36.3. (a) Composite columnar section of Neoproterozoic strata in the Mackenzie Mountains (modified after Narbonne & Aitken 1995). (b) Type section of the
Rapitan Group (Gabrielse et al. 1973) in the Hayhook Lake area (see Fig. 36.1 for location) as remeasured by P.F.H. (c) Detailed section of the upper Sayunei and Shezal
formations 1.6 km NW of the Rapitan Group type section. Note Fe-formation (‘BIF’) at the top of the Sayunei Formation, representing ‘basin-facies’ Fe-formation
(Eisbacher 1985). (d) Section of the lower Rapitan Group near Iron Creek, representative of ‘transitional-facies’ Fe-formation (Eisbacher 1985).

mapping. Long slopes are unattractive because of scree. Published An unconformity at the base of the Backbone Ranges Fm. (Early
1:250 000-scale map coverage is incomplete and stratigraphically Cambrian) cuts progressively down-section from SW to NE, with
inconsistent. the result that the Windermere Supergroup is missing to the NE of
400 P. F. HOFFMAN & G. P. HALVERSON

the Plateau Thrust system. To the SW, the glaciogenic units are Keele Fm. (Day et al. 2004). A thin, jasper-pebble conglomerate
buried by younger strata. Although the outcrop belt of Windermere occurs widely near its base. Glendonites, interpreted as pseudo-
strata is less than 30 km wide, it fortuitously preserves the outer morphs after ikaite (CaCO3.6H2O), occur close to the same
shelf-edge and upper slope of a continental terrace developed in horizon (James et al. 2005).
late Cryogenian (Keele Fm.) to early Ediacaran (Sheepbed Fm.) The outer edge of the Keele platform is exposed near Shale Lake
time (Ross 1991; Narbonne & Aitken 1995; Dalrymple & Nar- and on Stelfox Mountain (Fig. 36.1), where a SW-dipping ‘break-
bonne 1996; Day et al. 2004). Continental rifting, leading to the away scarp’ (Aitken 1991b, p. 26) separates shallow-marine Keele
formation of the continental terrace, controlled sedimentation strata on the footwall from hotel-size megaclast breccia (Durkan
during the Coates Lake and Rapitan groups (Eisbacher 1981a, Member), turbiditic siltstone (Delthore Member) and glaciogenic
1985; Jefferson 1983), as well as during the preceding Mackenzie diamictite (Stelfox Member) of the Ice Brook Fm. on the
Mountains Supergroup (Turner & Long 2008). The Coates Lake hanging wall (Aitken 1991b; Shen et al. 2008). The Stelfox dia-
and Rapitan basins are interpreted by Jefferson & Ruelle (1986) mictite is buttressed against the palaeoscarp and its feather-edge
and Yeo (1981), respectively, as rhombochasms (‘pull-apart’ steps across the trace of the palaeoscarp on Stelfox Mountain (at
basins) associated with hypothetical strike –slip systems, systems 638350 5800 N, 1278520 0200 W) onto the Keele Formation of the foot-
that are not mutually compatible in orientation. wall (Fig. 36.4b). The Ravensthroat Fm. passes across the fault line
without displacement, proving that movement on the palaeoscarp
ended before the glacial termination. Eisbacher (1981a) inferred
Stratigraphy that the Durkan megabreccia was triggered by glacioeustatic fall,
but this interpretation was rejected by Aitken (1991b), who
Mackenzie Mountains Supergroup argued that pore-fluid overpressures would not develop in uncon-
solidated sediments unless the base-level fall was catastrophically
The Mackenzie Mountains Supergroup is a broadly conformable rapid. He (Aitken 1991b) insisted that only the Stelfox Member
succession composed of sandstone, siltstone, carbonate and eva- was glacial in origin. P.F.H. therefore prefers the name Stelfox gla-
porite of mostly shallow-marine origin (Fig. 36.3a). Its upper ciation over Ice Brook glaciation, despite the latter’s priority in the
part, the Little Dal Group, is 2 km thick in its type section (Gab- literature (Kaufman et al. 1997).
rielse et al. 1973; Aitken 1981; Halverson 2006) and is dominated
by carbonate with a recessive interval of gypsiferous siltstone. The
carbonates are platformal in the SE and basinal in the NW, with Ravensthroat, Hayhook and Sheepbed formations
reefal build-ups in the lower part (Aitken 1981; Turner et al.
1997). The carbonate is conformably overlain by and locally inter- As with the Rapitan glaciation, major marine flooding followed by
stratified with pillow basalt, the Little Dal lavas of Aitken (1982). Stelfox diamictite deposition, providing accommodation on the
These are possibly the extrusive equivalents of the Tsezotene Keele platform for the formation of a transgressive, post-glacial
dykes and sills (Aitken 1982), which belong to the 780 Ma Gun- carbonate (Raventhroat/Hayhook formations) and 446 –562 m of
barrel large igneous province (Harlan et al. 2003). black shale and minor siltstone of the Sheepbed Fm. (Pyle et al.
2004; Shen et al. 2008). On the continental slope SW of the
Keele platform, the Sheepbed Fm. reaches 1050 m in thickness
Coates Lake and Rapitan groups at Sekwi Brook (Fig. 36.1) and includes significantly more silt-
stone, deposited by both turbidity flows and unidirectional
The Coates Lake Group is an assemblage of sandstone, carbonate- (NW-directed) contour currents (Dalrymple & Narbonne 1996).
clast conglomerate, gypsiferous siltstone and carbonate, including The contourites are estimated to have formed in water depths of
fetid basinal limestone with turbidites and ‘debrites’ (i.e. coarse- 1.0– 1.5 km and imply that the slope was open to the world
grained mass-flow deposits) (Ruelle 1982; Jefferson & Ruelle ocean in a gulf .100 km wide (Dalrymple & Narbonne 1996).
1986). The Rapitan Group (Eisbacher 1978, 1981a) includes the
Sayunei (sigh-YOU-knee) Fm., composed of maroon coloured,
subaqueously deposited, fine-grained clastics with subordinate
debrites and lonestones, and the overlying Shezal (shiz-ALL) Glaciogenic and associated strata
Fm., a stack of mostly olive-coloured, polymictic ‘diamictites’
(i.e. massive, foliated or bedded wackestone with randomly dis- Rapitan Group
persed, matrix-supported pebbles and boulders, characteristically
faceted and striated) with thin interbeds of dark shale and sand- The remeasured type section of the Rapitan Group (Gabrielse
stone. Locally, an older diamictite complex (Mount Berg Fm.) et al. 1973) is shown in Figure 36.3b. The Sayunei Fm.
occurs below the Sayunei Fm. (Yeo 1981). The Rapitan Group (487 m) consists of maroon-coloured mudstone with innumerable
onlaps tilted Coates Lake and Little Dal group strata unconform- graded beds of siltstone and/or fine-grained sandstone, some
ably (Eisbacher 1978, 1981a, 1985). Basinal facies of the Coates with outsized clasts of dolostone. Granule- to small pebble-sized
Lake Group are black, pyritic and organic-rich; those of the lonestones are sprinkled throughout the mudstone. The clasts are
Rapitan Group are maroon, hematitic and organic-poor. commonly rimmed by chlorite, formed by reaction between the
dolostone and the host mudstone. Some clasts are aggregates of
sand or granules, interpreted as ‘till pellets’ (Ovenshine 1970) by
Twitya, Keele and Ice Brook formations Young (1976) and Eisbacher (1981a). Bedforms attributable to
wave action are absent, as is evidence of bottom traction cur-
A major marine transgression followed the Rapitan glaciation, pro- rents. Greenish mudstone beds occur, but none are dark grey
viding accommodation for 330–765 m of dark grey shale, siltstone or black, in stark contrast to the basinal facies of the underlying
and fine-grained sandstone of the Twitya Fm. (Eisbacher 1978, Coates Lake Group.
1981a; Aitken 1982). Resting sharply upon boulder diamictite of In the Hayhook Lake area (Fig. 36.1), the top of the Sayunei
the upper Shezal Fm., at the base of the Twitya Fm., are 0– 52 m Fm. is marked by 10–16 m of brick-red, hematitic mudstone and
of dark-grey, thin-bedded limestone (Eisbacher 1978, 1981a; hematite jaspilite (Fig. 36.3c). This interval is parallel-laminated
Aitken 1982), which are absent in the Rapitan Group type and lacks evidence for bottom traction currents. Within the
section (Hayhook Lake) and thickest at Shale Lake (Fig. 36.1). Fe-rich formation are lonestones and thin dolostone-clast debrites
Above the Twitya Fm. lie 220 –600 m of cyclic, shallow-marine (Young 1976). The former include rounded cobbles and small
sandstone, siltstone, limestone and dolomitized limestone of the boulders of porphyritic quartz monzonite (i.e. two-feldspar
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 401

(a)
o
155-210

NW SE
50 50

Sheepbed Fm
40 40
G Sheepbed Formation
o G
200-220
R
30 G 30

Hayhook fm
G
o R
198-210 R
20 20
o 190-200o
215
G

Ravensthroat fm
240o
10 190o
R R 190-210o 10
195o
190o 210-230o
230o 195o
0 0

Ice Brook Fm (Stelfox Mb)


-10 -10
aeolian(?)
-20 -20
Keele Formation

-30 -30

Keele Fm
o
150-225 -40
-40 o
200

o
200
-50 -50
Cranswick R. Arctic Red R. Gayna R. Stone Knife R. Shale Lake Twitya R. Moose Horn R. Ravensthroat R.
65o05’54”N, 64o56’05”N, 64o49’22”N, 64o40’36”N, 64o31’12”N, 64o13’07”N, 63o57’13”N, 63o17’05”N,
132o26’17”W 131o03’30”W 130o28’27”W 129o53’48”W 129o29’09”W 128o38’11”W 127o30’39”W 127o08’40”W

(b)
W RG red / green marly shale
LEGEND E
50 black shale 50
allodapic limestone current ripples (flow azimuth)
peloidal dolostone (primary) sea-floor aragonite cement
40 carbonate-clast diamictite sea-floor barite cement 40
limestone giant wave ripples (crestline)
dolostone (secondary) corrugate stromatolite
30 sheet-crack cement
30
quartz sandstone
quartz-sericite siltstone crossbedding (flow azimuth) Hayhook
formation
20 20

Sheepbed Formation
10 Ravensthroat 10
o o 295-300o
formation
170-205 275-280 235o
0 0

-10 -10
en, 1991b)

Stelfox Member Fig. 36.4. Stratigraphic relations based on


Ice Brook Formation
columnar sections spanning the Stelfox
-20 -20 Member and its post-glacial carbonate
scarp” (Aitk

sequence (Raventhroat and Hayhook


Keele Formation
-30 -30 formations) in lines oriented parallel (a) and
transverse (b) to the depositional strike of the
“breakaway

Neoproterozoic passive margin. The


-40 Delthore Member -40 ‘breakaway scarp’ (Aitken 1991b) marks the
Ice Brook Formation
outer edge of the Keele Formation shelf.
Megaclast breccia of the Durkan Member
-50 -50 (not shown) conformably underlies the
Stelfox Mt. Stelfox Mt. Stelfox Mt. Stelfox Mt. Stelfox Mt. Moose Horn R. Hayhook Lake
63o36’08”N, 63o35’37”N, 63o36’00”N, 63o35’46”N, 63o36’01”N, 63o57’13”N, 63o34’17”N, Delthore Member and both abut the
127o57’43”W 127o55’11”W 127o53’57”W 127o53’04”W 127o50’47”W 127o30’39”W 127o11’44”W ‘breakaway scarp’ at a buttress unconformity.

granitoid with 5–15% quartz). The Fe-formation is overlain dis- In the Iron Creek area (Fig. 36.3d), an aggregate thickness of 100–
conformably by maroon-coloured diamictite (basal Shezal Fm.), 120 m of hematite jaspilite with pea-sized jasper nodules (mini-
which contains rounded and angular clasts from the lower unit. concretions) is sandwiched between units of poorly stratified,
In some sections, the Fe-formation is missing and was probably boulder diamictite with subordinate beds and lenses of tabular cross-
removed by erosion at the base of the Shezal Fm. bedded, medium-grained, well-rounded, monocrystalline, quartz
402 P. F. HOFFMAN & G. P. HALVERSON

sandstone (Yeo 1981, 1986; Klein & Beukes 1993). Outsized clasts has a terrigenous matrix and the carbonate clasts, freed in float,
and debrites occur within the jaspilite intervals, as do thin seams and are visibly faceted and striated. Ice-rafted dropstones and ‘till
lenses of Fe-carbonate (siderite). The Hayhook Lake area and Iron pellets’ (Ovenshine 1970) are well-developed in laminated facies
Creek sections represent respectively the ‘basinal’ and ‘transitional’ (Aitken 1991a, b).
(more proximal) facies of Fe-formation distinguished in the Rapitan
Group by Eisbacher (1985).
The Shezal Fm. diamictite is mostly olive-green or grey in Ravensthroat– Hayhook post-glacial carbonate couplet
colour, except for the basal part, which is maroon and clearly
derived from the underlying Sayunei Fm. (Fig. 36.3c). The clasts The Stelfox diamictite or its equivalent erosion surface is every-
consist of dolomite (Little Dal Group), commonly stromatolitic, where sharply overlain by a transgressive, 10–15-m-thick, cap
quartzite (Katherine Group), siltstone (Coates Lake Group and dolostone (Fig. 36.4), previously known informally as the ‘Tepee
Tsezotene Fm.), basalt and gabbro. Extrabasinal clasts are rare. dolostone’ (Eisbacher 1981a, 1985; Aitken 1991b). The pale, incan-
Faceted and striated clasts, particularly siltstone, are abundant descent, yellowish-grey scree profusely generated by the informally
and unequivocally glacial in origin (Young 1976; Eisbacher defined Ravensthroat Fm. (James et al. 2001) could not be more
1981a). Diamictite occurs in crudely tabular, poorly stratified, unlike the jet-black feathers of its avian namesake. East of 1308W
matrix-supported bodies up to 90 m thick. The matrix consists longitude, the dolostone is overlain by up to 15 m of ledge-forming
variably of wackestone, structureless mudstone, or mudstone grey or pink limestone (Fig. 36.4), informally defined as the
with anastomosing (‘scaly‘) foliation (Fig. 36.3c) that is subparal- Hayhook Fm. (James et al. 2001). West of 1308W longitude, it is
lel to bedding but unrelated to the axis of maximum tectonic com- overlain by an ascending sequence of red, green and grey shales
pression. The Shezal Fm. (320 m thick) is unusually well-exposed with thin beds of current-rippled limestone (Fig. 36.4). This tri-
in a river canyon 1.6 km NW of the type area (Fig. 36.3c), where coloured sequence is remarkably similar to that overlying the
diamictite units are separated by subordinate, recessive intervals of homologous postglacial cap dolostones (Hambrey & Spencer
well-stratified siltstone, graded sandstone and dark-grey shale. A 1987; Halverson et al. 2004) in East Greenland (e.g. Kap Weber,
6-m-thick duplex structure with thrusting directed toward the NE Andrée Land) and East Svalbard (e.g. Ditlovtoppen, Ny Friesland).
occurs in the reference section between –320 and –330 m The Ravensthroat Fm. consists of micro- and macropeloidal
(Fig. 36.3c). In this section, the tops of the diamictite bodies dolostone with ubiquitous, low-angle, cross-lamination defined
tend to be more sharply defined than their bases. by normal and reverse-graded laminae (James et al. 2001). The
No post-glacial carbonate is present in the type section of the peloids are structureless, sub-spherical aggregates of dolomicrite
Rapitan Group (Fig. 36.3b), but elsewhere the basal Twitya Fm. up to 3 mm in diameter. They commonly rest on abraded facets.
consists of thin graded beds of dark-grey, allodapic limestone (Eis- Giant wave ripples (cf. Allen & Hoffman 2005) are well developed
bacher 1978, 1981a; Aitken 1982). The carbonate reaches a thick- in many sections (Fig. 36.4). They are trochoidal in profile, with
ness of 52.5 m at Shale Lake. At Stone Knife River (Fig. 36.1), sharp crests and lobate troughs, and individual ripple trains
where 40 m of limestone is exposed, its lower part is hummocky aggrade for up to 1.4 m through bidirectional accretion of peloidal
cross-stratified. Near Corn Creek (Fig. 36.1), bordering the Wer- laminae. Their synoptic relief, crest to trough, is up to 40 cm. The
necke Mountains, up to 300 m of massive dolomite (Mount structures are different in form and origin from peritidal ‘tepees’
Profeit Fm.) is laterally equivalent to the lower Twitya post-glacial (Assereto & Kendall 1977; Kendall & Warren 1987); they are
limestone (Eisbacher 1981a). not brecciated, lack syndepositional cements, and their crestlines
are linear and parallel, not polygonal, in plan view (Eisbacher
1981a; Aitken 1991b; James et al. 2001). Their crestlines generally
Stelfox Member (Ice Brook Fm.) trend NE, at a high angle to the Keele shelf edge (Fig. 36.5).
Decametric bioherms composed of corrugated stromatolites
Whereas the Rapitan Group is up to 1500 m thick (Yeo 1981), the (James et al. 2001) oriented subparallel to the crests of giant
Stelfox Member is less than 40 m thick in most sections (Fig. 36.4) wave ripples occur in some sections (Fig. 36.4). The corrugated
and the thickest, 308 m (Atiken 1991a), is also the deepest-water
section of Stelfox diamictite in the Mackenzie Mountains. Basin-
ward thickening could account for the modest thickness of the
Stelfox diamictites, which are only preserved in shelf and upper-
slope settings, in contrast to the deep rift-basin setting of the
1 130o 128o
Rapitan Group. 65o
At Shale Lake (Fig. 36.4a), the Stelfox diamictite is directly 3
underlain by 5–9 m of well-sorted, well-rounded, medium-
grained, quartz sandstone of possible aeolian origin. Large-scale
1 Keele
tabular cross-bedding indicates transport to the SSW (c. 2058
azimuth, n ¼ 2), consistent with easterly palaeowinds (see
4
below). The presence of aeolianite beneath the diamictite would
suggest that glacioeustasy preceded or outranked glacioisostasy.
shelf
The Stelfox Member is typically recessive and masked by scree
slope 5
from the overlying Ravensthroat Fm., which along with its modest o
thickness accounts for its delayed recognition. The dominant 64
lithology is massive to weakly stratified diamictite, in which
rounded clasts of tan dolostone and grey limestone are supported
by a matrix of brownish-tan carbonate-rich mudrock with dis- 50 km 5
persed grains of quartz sand and granules. The clasts are derived
from the underlying Keele Fm. and the absence of quartzite 1
clasts suggests that carbonates were the only components of the 132o 130o
Keele Fm. that were strongly lithified at the time of glaciation.
Beds and lenses of quartz-chert sandstone, micaceous wacke and Fig. 36.5. Mean orientations of crestlines (double bars) of giant wave ripples in
polymictic debrite are subordinate components of the Stelfox the Ravensthroat Fm. at different locations. Numerals are the numbers of
Member. At Cranswick River section (Fig. 36.4a), the diamictite independent ripple trains measured at each location.
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 403

stromatolites are closely similar to those that are intergradational Little Dal groups (Gabrielse et al. 1973; Eisbacher 1978, 1981a,
with ‘tubestone’ stromatolites in the homologous Keilberg cap 1985; Aitken 1982; Jefferson & Ruelle 1986). In the Hayhook
dolostone in northern Namibia. Evidence of subaerial exposure Lake area (Fig. 36.1), the Coates Lake Group is onlapped by the
is absent, but there can be no doubt that the Ravensthroat Fm. Sayunei Fm. along the unconformity (Eisbacher 1978, 1981a).
was deposited above wave-base and within the euphotic zone. The presence of submarine carbonate-clast talus breccias in the
In virtually all sections from Shale Lake to Ravensthroat River lower Rapitan Group (beneath and within the lower Sayunei
(Fig. 36.1), a distance of nearly 200 km, the uppermost 4– 10 cm of Fm.) on the northeastern margin of the outcrop belt near Shale
the Ravensthroat Fm. contains a remarkable digitate crust com- Lake (648320 4100 N, 1298220 3700 W) suggests a master normal fault
posed of sea-floor barite (BaSO4) cement, variably pseudomor- on that side of the basin. It may have been structurally inverted
phosed by calcite. At Shale Lake, the barite crust is developed during orogenic contraction to form a part of the Plateau Thrust
directly on a train of giant wave ripples. Because of pseudo- system. The maroon colour of the Sayunei stems in part from its
morphic replacement in some sections, the barite cement was for- derivation from the Coates Lake Group, which is dominated by
merly misinterpreted as an aragonite cement (Aitken 1991b, fig. 8) reddish siltstones of the Redstone River Fm. Similarly, the olive-
or as microdigitate stromatolites (James et al. 2001, fig. 10). The green colouration of the overlying Shezal diamictites reflects its
digits consist of rosettes or bundles of bladed crystals whose mor- derivation from the grey carbonates and greenish basalts of the
phology is quite distinct from the prismatic (pseudohexagonal) Little Dal Group.
habit of the former aragonite cements in the overlying Hayhook A knife-sharp, marine, flooding surface separates the Shezal Fm.
Fm. (James et al. 2001). The digits preserve growth lamellae, 1– from fetid, dark-grey, flaggy limestone and shale of the lower
2 mm thick, and typically tilt to the SW due to preferential Twitya Fm. (Eisbacher 1978, 1981a). Where the Rapitan Group
growth in the seaward direction. The spaces between the digits is absent, the Twitya transgresses older units disconformably (Nar-
are filled by laminated, peloidal, ferroan dolomite, the dark bonne & Aitken 1995).
brown colour of which contrasts with the pale colouration of the
rest of the Ravensthroat Fm. The borders of the barite digits are
ragged due to the growth of vertically standing barite crystals Stelfox Member
that lap out onto the upper surfaces of individual laminae of infill-
ing peloidal dolomite. This provides textural proof that the barite A low-angle unconformity (Fig. 36.4a) separates the Stelfox dia-
digits formed simultaneously with the peloidal infills. This mictite from shelf carbonates and clastics of the underlying
interpretation agrees with that of barite cements at the homologous Keele Fm. (Day et al. 2004). West of the palaeoscarp at Stelfox
post-glacial dolostone– limestone transition in central Australia Mountain (Fig. 36.4b), a sharp disconformity separates the
(Kennedy 1996). Stelfox diamictite from parallel-laminated, greenish-grey siltstone
The tips of the barite digits, marking the Ravensthroat – of the underlying Delthore Member (Aitken 1991b). As the
Hayhook transition, are typically ‘colonized’ by slender prismatic Delthore Member was deposited below storm wave-base, glacial
fans of pseudomorphosed aragonite, while at the same level the erosion may be required to account for the sharp disconformity.
peloidal ferroan dolostone changes to calcimicrite, initially with A major marine flooding surface separates the Stelfox Member,
some reworked dolostone. Locally, the barite crust was dissolved or the Keele Fm. where the Stelfox is absent, from the overlying
away instead of pseudomorphically replaced, and the barite Ravensthroat Fm. Where the basal contact of the dolostone is
cement layer is represented by a collapse microbreccia of not covered by its own scree (e.g. Cranswick River, Arctic Red
ferroan dolostone clasts, overlain by calcimicrite of the basal River, Stoneknife River), 10–20 cm of clay separate it from the
Hayhook Fm.. The local occurrence of this microbreccia contribu- underlying diamictite.
ted to the misinterpretation of the contact as a karstic unconformity The top of the Hayhook limestone is a conformable marine
(James et al. 2001, fig. 9B). flooding surface overlain by organic-rich black shale of the
Numerous mesoscopic folds at the Ravensthroat –Hayhook Sheepbed Fm. (Pyle et al. 2004; Shen et al. 2008).
contact, previously interpreted as karst channels (James et al.
2001, fig. 9A), are consistently oriented parallel to tectonic clea-
vage and regional folds. They are not karst channels because the Chemostratigraphy
primary stratification in both units is everywhere parallel to the
contact. We interpret them as lobate-cuspate folds (Ramsay & Carbon isotopes
Huber 1987) of tectonic origin and Late Cretaceous –Paleocene
age, related to bedding-parallel shortening at the rheological inter- Carbon-isotopic records have been published for the Little Dal and
face between the stiff Ravensthroat dolostone and the weak Coates Lake groups (Halverson 2006), the Twitya post-glacial
Hayhook limestone. Contrary to James et al. (2001), we find no limestone and Keele Fm. (Narbonne et al. 1994; Kaufman et al.
evidence of karstic unconformity, subaerial exposure or isostatic 1997; Hoffman & Schrag 2002), and the Ravensthroat –Hayhook
adjustment at the Ravensthroat –Hayhook contact. post-glacial carbonate (James et al. 2001). The Little Dal Group
The Hayhook Fm. is 0 –15 m thick (Fig. 36.4) and consists of is enriched in 13C relative to the PDB standard, except for a
flaggy micrite with pseudomorphosed crystal fans of aragonite depleted interval in the Upper Carbonate Fm. The overlying
cement (James et al. 2001). The cement fraction generally Coates Lake Group begins with a deep negative excursion to
increases upward and reaches nearly 100% in some sections, –8‰ in d13C, followed by a positive excursion to þ8‰ before
notably Shale Lake (Fig. 36.1). The detailed interplay between settling back to modestly enriched values.
cement growth, burial by micrite and renewed cement develop- Values of d13C in the Twitya post-glacial limestone increase
ment indicates precipitation at a free-face on the sea-floor. up-section from – 2.5‰ at the base. Carbonates of the Keele Fm.
are strongly enriched,  þ 8‰, but plunge to – 8‰ in the last
30 m of section beneath the Stelfox diamictite.
Boundary relations with overlying and underlying The Ravensthroat Fm. is also depleted in 13C (Fig. 36.6) and values
non-glacial units decline up-section. The Ravensthroat–Hayhook contact coincides
with a –2‰ step function. This is broadly consistent with low-
Rapitan Group temperature equilibrium fractionation between dolomite and calcite
(Friedman & O’Neil 1977). Within the Hayhook Fm., d13C rises
A low-angle unconformity separates the glaciogenic Rapitan by c. 1‰ up-section, with no systematic difference in values
Group from various units of the underlying Coates Lake and between micro-drilled micrite and sea-floor cement (Fig. 36.6).
404 P. F. HOFFMAN & G. P. HALVERSON

(a) (per mil V-PDB) An alternative, less consequential, interpretation of the d18O
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2 0 25 data is that the micrite was preferentially altered by diagenetic
fluids on account of its finer grain size compared with the
20 cements, and therefore its greater surface-to-volume ratio. The
d18O data are more depleted in 18O and display more scatter in
15
the second section (Fig. 36.6b), at Ravensthroat River. This
section illustrates the ease with which d18O, but not d13C, was
diagenetically altered.
10
Sheepbed Fm
Hayhook fm 5 Strontium isotopes
Ravensthroat fm
Stelfox Mb 0 New and previously published (Kaufman et al. 1993, 1997)
87
(m) Sr/86Sr data from Sr-rich (300 –3000 ppm), low-Mn (Mn/
Sr , 0.1) limestones bracketing the Rapitan and Stelfox glacia-
(b) tions in the Mackenzie Mountains are given in Halverson et al.
30 (2007). Non-radiogenic values of 0.70550 –0.70622 (n ¼ 10) are
observed in the Little Dal Group, compared with 0.70644–
25 0.70669 (n ¼ 5) in the Coates Lake Group. No data are currently
available for the post-Rapitan limestone (basal Twitya Fm.), but
20
limestones directly beneath the Stelfox diamictite (uppermost
Keele Fm.) have ratios of 0.70718–0.70720 (n ¼ 3). Sea-floor
δ O δ C
18 13
cements from the Hayhook post-glacial limestone are marginally
15
calcite (micrite) less radiogenic at 0.70714– 0.70716 (n ¼ 4) than the Coates Lake
calcite (cement) Group, whereas the coexisting micrites are more radiogenic, with
dolopelmicrite 10 minimum values of 0.70751–0.70792 (n ¼ 5), which is consistent
James et al. (2001) with preferential alteration. All reported Hayhook values are from
5 samples with Sr concentrations .600 ppm and Mn/Sr , 0.1.

0
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2 0 (m) Sulphur isotopes and iron speciation
(per mil V-PDB)
Sulphur isotopes have been measured from carbonate-associated
Fig. 36.6. C- and O-isotope data from dolomite of the Ravensthroat Fm. sulphate in the Ravensthroat and Hayhook formations, as well as
and from coexisting micrite and sea-floor cement of the Hayhook formation from the barite sea-floor cement at the top of the Ravensthroat
at (a) Shale Lake and (b) Ravensthroat River. Fm. (M.T. Hurtgen, pers. comm. 2008). The d34SCAS and
d34Sbarite values are mutually consistent and agree with values
Oxygen isotopes from correlative strata in Namibia (Hurtgen et al. 2006), but the
data have yet to be published.
In order to test the possibility of low-temperature dolomite – calcite Sulphide S isotopes and Fe-speciation (ratio of highly reactive
equilibrium fractionation, which would imply a sea-floor origin Fe to total Fe) have been studied in organic-rich black shales of
for the Ravensthroat dolomite, and also for the possibility that the Sheepbed Fm. (Fig. 36.3) in order to infer changes in the oxy-
the Hayhook sea-floor cements and associated micrites might genation of the atmosphere and the deep ocean, respectively, in
retain a record of bottom and surface waters, respectively, we the aftermath of the Stelfox glaciation (Shen et al. 2008). Large
plot unpublished O-isotope data for two sections that we variability (.35‰) in d34Ssulphide is observed from bottom to
sampled in detail (Fig. 36.6). Exceptional petrographic preser- top, indicating that sulphate was not limiting (i.e. .2 mM). This
vation of the sea-floor cements motivated this effort. At Shale implies higher rates of oxidative weathering (the major source
Lake, the Ravensthroat-Hayhook contact coincides with a –6‰ of marine sulphate) than before the glaciation (Hurtgen et al.
step function in d18O, taking the micritic component to be repre- 2002). There is a dramatic shift in the Fe-speciation ratio
sentative of the surface waters, from which the Ravensthroat was (FeHR/FeT), however, between 140 and 160 m above the base
also derived (Fig. 36.6a). The magnitude of the step function in of the lithologically homogeneous black shale sequence that is
d18O, roughly 3 that in d13C, is compatible with a low- 420 m thick in total. The average value below 140 m is 0.43
temperature equilibrium fractionation between dolomite and (n ¼ 32), indicating bottom water anoxia, whereas above 160 m
calcite (Friedman & O’Neil 1977). it is 0.11 (n ¼ 27), which is consistent with oxygenated bottom
The sea-floor cements are consistently more enriched in 18O waters (Raiswell & Canfield 1998; Lyons & Severmann 2006;
than the micritic component of the Hayhook (Fig. 36.6a), consist- Canfield et al. 2007).
ent with lower temperatures and/or higher salinities of the bottom
waters, from which the cements were precipitated, relative to the
surface waters. The cements become more enriched in 18O Calcium isotopes
up-section (Fig. 36.6a), which could indicate an increase in the
temperature and/or salinity gradient, or alternatively an increase Calcium isotopes (d44Ca) were measured in limestone of the lower
in water depth, with time. The cements exhibit considerably Twitya Fm., and in dolostone and limestone of the Ravensthroat
more scatter in d18O than does the micrite, which could reflect and Hayhook formations, respectively (Silva-Tamayo et al. 2010).
the difficulty in micro-drilling cement without contamination The data from the first are consistent with mid-Cryogenian post-
from the micrite that fills in between the slender prisms of glacial carbonates in Brazil, and the latter with basal Ediacaran
former aragonite and even their originally hollow interiors. post-glacial sequences in Brazil and Namibia (Silva-Tamayo
Accordingly, the reduction in scatter towards the top of the et al. 2010). Compared with the younger sequences, the older
Hayhook reflects the more massive and continuous nature of the d44Ca profiles appear basally truncated, similar to their sequence
uppermost cements. stratigraphies and d13C profiles (Hoffman & Schrag 2002). The
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 405

younger sequences exhibit paired negative and positive d44Ca d57Fe values increase systematically upsection from – 0.7‰ to
excursions that span twice the entire Phanerozoic range of variabil- 1.2‰. The largest rise coincides with the transition from hematitic
ity (Farkas et al. 2007). The paired anomaly in the lower Hayhook mudstone to hematite-jaspilite, which also corresponds to the
limestone is preceeded (Ravensthroat Fm.) and followed (upper highest Fe concentration (c. 58 wt% Fe2O3). The increase in
Hayhook Fm.) by relatively stable values lying well within the d57Fe is accompanied by a rise in Ce/Ce* (Fig. 36.7).
Phanerozoic range (Silva-Tamayo et al. 2010).
Equilibrium isotope fractionation during carbonate precipitation
maintains seawater roughly 1.0‰ heavier than the sources and Palaeontology
sinks of Ca in the ocean. Thus, imbalance between Ca input and
output will cause deviations in the d44Ca of seawater, and therefore Neoproterozoic palaeontology of the Mackenzie Mountains is
in marine carbonate. The magnitude of the equilibrium fraction- summarized in Narbonne & Aitken (1995). When combined
ation is expected to vary with temperature and carbonate precipi- with contemporary work in the Wernecke Mountains (Narbonne
tation rate, but relatively modest variability in d44Ca (n ¼ 10) & Hofmann 1987) and ongoing stratigraphic and palaeontological
within the Ravensthroat Fm. (Silva-Tamayo et al. 2010), deposited studies in correlative strata of the Tatonduk inlier, a pattern of
during global deglaciation when changes in temperature and pre- change over time emerges that is broadly consistent with other
cipitation rate may have been large, implies that those dependen- areas (Macdonald et al. 2010a, b; Macdonald & Cohen 2011). A
cies were either small or counteracting at that time. The paired number of eukaryotic crown groups, both algal and protistan,
negative and positive excursions of 1.0‰ apiece in the lower had evolved and diversified before the Rapitan glaciation. The
Hayhook limestone and correlatives (n ¼ 31) represents first, a record of these groups is cryptic from shortly before the Rapitan
large excess of Ca input over output, followed by the reverse, a glaciation until long after the Stelfox glaciation.
large excess of Ca output over input (Silva-Tamayo et al. 2010). Soft-bodied, Ediacara-type macrofossils, representing benthic
polypoid and frond-like organisms, occur in the upper Ediacaran
Blueflower Fm. (Fig. 36.1), where they coexist with infaunal
Synglacial hematite jaspilite (Fe-formation), Rapitan Group burrows (simple, meandering and subordinate patterned) that
extend into the overlying Risky Fm. (Narbonne & Aitken 1990;
REE geochemistry of the Sayunei jaspilite (Klein & Buekes, 1993) Narbonne 1994). In the same area, near Sekwi Brook (Fig. 36.1),
is described in Hoffman et al. (2011). Planavsky et al. (2010) simple radially symmetric body fossils and rare trace fossils
present elemental data (P, Fe, Mn and Al) for jaspilite samples occur in the middle Sheepbed Fm. (Narbonne & Aitken 1990;
from the Sayunei Fm. (n ¼ 41) and correlative strata (Upper Narbonne 1994). The Stelfox glaciation is not recognized at
Tindir Group) in the Tatonduk inlier. In common with other Sekwi Brook, but at Shale Lake (Fig. 36.1), Fe-speciation data
Cryogenian glaciogenic jaspilites, the samples have much higher indicate that bottom waters became oxygenated just below the
phosphate contents (average P/Fe ¼ 1.44%, n ¼ 44) than Palaeo- middle Sheepbed Fm. (Shen et al. 2008).
proterozoic or Archaean Fe-formations (Planavsky et al. 2010). An assemblage of simple, centimetric annuli and discs, thought
Iron isotopes (d57Fe) and accompanying rare-earth elements to be biogenic, occur in the upper Twitya Fm. (Fig. 36.3a) of the
(REE), including redox-sensitive cerium anomalies (Ce/Ce*), Sayunei Range (Hofmann et al. 1990), south of Twitya River
have been measured through the 16.4-m-thick interval of hematitic (Fig. 36.1). As possible sponge-grade metazoan fossils pre-dating
mudstone and hematite-jaspilite (banded Fe-formation) at the top the terminal Cryogenian glaciation, the ‘Twitya-discs’ have
of the Sayunei Fm. in the Hayhook Lake area (Fig. 36.7). The recently been joined by calcareous structures at different horizons

Fig. 36.7. Fe isotopes (d57Fe), wt% Fe2O3


and Ce anomaly data (Ce/Ce*) from the
Sayunei Fe-formation in the Hayhook Lake
reference section. Data tabulated in
Halverson et al. (2011).
406 P. F. HOFFMAN & G. P. HALVERSON

in South Australia (Wallace & Woon 2008; Maloof et al. 2010), diamictite, correlative with the lower Rapitan Group, was depos-
and a sponge biomarker in the Oman Salt Basin (Love et al. 2009). ited during the waning stages of bimodal volcanism in an active
rift basin (Mustard & Roots 1997; Macdonald et al. 2010b). The
presence of faceted and striated clasts, bed-penetrating
Other characteristics dropstones, common outsized and exotic clasts, and glacial push
structures (Hart River inlier) within a marine succession support
The Crest deposit at Iron Creek (Fig. 36.3d), near Snake River, is a a glacial marine origin for the diamictite at an ice-sheet grounding-
hematite –jaspilite interval 120 m thick within glaciogenic dia- line. A quartz-phyric rhyolite within the volcanic pile, stratigraphi-
mictites of the Rapitan Group (Yeo 1986; Klein & Beukes cally beneath the diamictite, and a felsic tuff within diamictite,
1993). The deposit contains 5.6 billion tonnes of ‘ore’ averaging c. 60 m above the base of the glacial –periglacial sequence, give
47.2% Fe and regional reserves are estimated to exceed 18.6 U –Pb zircon dates of 717.43 + 0.14 and 716.47 + 0.24 Ma
billion tonnes (Yeo 1986). Because of remoteness and difficult (2s), respectively (Macdonald et al. 2010b). These dates are indis-
terrain, the deposit has never been economic. tinguishable in age from mafic dykes and sills of the Franklin
Large Igneous Province (Heaman et al. 1992; Denyszyn et al.
2009; Macdonald et al. 2010b), which straddled the palaeoequator
Palaeolatitude and palaeogeography at the time of their emplacement. The beginning of glacial marine
sedimentation in the Mount Harper rift basin at 717 Ma is the best
Palaeomagnetic data from the Mackenzie Mountains relating to indicator of the onset of Rapitan glaciation in the northern
the Rapitan and Stelfox glaciations have been reviewed (Evans Canadian Cordillera (Macdonald et al. 2010b). These data are
2000; Evans & Raub 2011). Morris (1977) recognized three consistent with maximum age constraints obtained from the
natural remanent magnetic components in the Rapitan Group, of Mackenzie Mountains and Windermere Supergroup, including
which the one inferred to be the youngest (Cretaceous?) gives a a weighted mean 207Pb/206Pb date of 779 + 2.3 Ma (Harlan
high palaeolatitude and the other two give low palaeolatitudes. A et al. 2003) obtained for two baddeleyite analyses from a
larger data set obtained by Park (1997) gives broadly similar gabbro sill (Carcajou Canyon Gabbro) intruding the Tsezotene
results, but only after anomalous declinations are corrected for Fm. (Fig. 36.3), and a weighted average 207Pb/206Pb date of
large vertical-axis rotations of presumed structural origin. The 755 + 18 Ma obtained for two concordant but low-Pb zircon
high-inclination poles were acquired during tectonic folding and grains from a single leucogranite ‘dropstone’ within turbidites
the low-inclination poles before that time, although Morris close to the base of the Sayunei Fm. near Shale Lake (Ross &
(1977) and Park (1997) disagreed on which of the two clusters Villeneuve 1997).
of low-inclination poles was older. The palaeolatitudes of their The Stelfox glaciation has no independent radiometric age con-
inferred oldest components are 08 + 28 (Morris 1977) and straint. Mafic tuffs in the upper Keele Fm. (Fig. 36.3a) contain only
06 + 78 (Park 1997). Neither result is definitively syndepositional, xenocrystic zircons. The Stelfox glaciation is assumed, largely on
but the near-equatorial palaeolatitudes agree with more recent the basis of its post-glacial carbonate sequence, to be correlative
palaeomagnetic results from mid-Neoproterozoic (800 –740 Ma) with the terminal Cryogenian glaciation (Knoll et al. 2006),
sedimentary rocks in western Laurentia (Weil et al. 2004, 2006). which terminated in 635 Ma.
Shallow mafic igneous rocks provide more reliable palaeomag-
netic results than sedimentary rocks because they are strongly
magnetized, resist low-temperature remagnetization, have not Discussion
been compacted (no flattening of palaeomagnetic inclination),
and provide baked-contact tests of the primary age of remanent Palaeoenvironmental interpretation of the Rapitan Group
magnetization, which can be directly dated as the crystallization
age of the rock. The Franklin Large Igneous Province (LIP) com- The Rapitan Group is widely interpreted as a glacial marine suc-
prises comagmatic mafic dykes, sills and lavas extending collec- cession because of the occurrence of faceted and striated clasts,
tively for 2500 km across Arctic Laurentia. The grand mean sand aggregates (‘till pellets‘), erratic dropstones with impact-
palaeopole (8.48N, 163.88E, A95 ¼ 2.88, n ¼ 78 sites) for the related structures, and soft-sediment deformation horizons (e.g.
Franklin LIP (Denyszyn et al. 2009) places the Mount Harper push structures, scaly matrix in particular diamictite sheets) of
area and the Mackenzie Mountains, respectively, at 21+3 and apparent glacitectonic origin (Young 1976; Eisbacher 1981a,
18 + 38N palaeolatitude at the time of magmatism, which 1985; Yeo 1981; Aitken 1982; Macdonald et al. 2010b). Eisbacher
spanned the onset of the Rapitan glaciation (see below). These (1985) presents a comprehensive facies model ranging from fluc-
results for the Rapitan glaciation are more reliable than those tuating grounding-line diamictites (Shezal Fm.) through transi-
from the Rapitan Group itself. tional facies to distal proglacial turbidites and fine-grained
There are no direct palaeomagnetic constraints on the Stelfox suspension fallout accumulated in structurally controlled, deep-
glaciation. McMechan (2000b) cites Park (1994) as giving a water basins (Sayunei Fm.). The distal nature of the Sayunei turbi-
palaeolatitude of c. 358 for the Stelfox glaciation, but this was dites, characterized by Bouma C-D sequences, is emphasized (Yeo
based on palaeopoles from the Risky Fm., which is c. 2 km strati- 1981; Eisbacher 1985) and the source of suspended sediment is
graphically above the Stelfox Member (Fig. 36.3a). Park (1994) inferred to be turbid meltwater plumes emanating from discharge
estimated the Risky Fm. to be younger than the 615 Ma Long sites at ice grounding-lines at the basin margin (Young 1976; Yeo
Range Dykes of Newfoundland and Labrador (Kamo et al. 1989; 1981; Klein & Beukes 1993). Eisbacher (1985, p. 242, fig. 8)
Kamo & Gower 1994). Although more results are needed for suggests that the paucity of dropstones and diamictite, respect-
better statistics, the preliminary palaeopole for the Long Range ively, in the Sayunei Fm. reflect a suppression of icebergs and a
Dykes (19.08N, 355.38E, A95 ¼ 17.48, n ¼ 5 dykes) places the buttressing of ice grounding-lines by a ‘stable sea-ice cover’
Mackenzie Mountains near the palaeoequator at 615 Ma (McCaus- (Dowdeswell et al. 2000). Such a permanent ice cover would
land et al. 2007). account for the low organic content and deep maroon colour of
the Sayunei mudrocks, in contrast to the basinal shale and lime-
stone of the underlying non-glacial Copper Cap Fm., which are
Geochronological constraints mostly fetid and black in colour. Removal of the permanent ice
cover by climate warming would have allowed icebergs and ice
In the Mount Harper area of the Ogilvie Mountains, 310 km west grounding-lines to advance basinward at the time of the
of Corn Creek (Fig. 36.1), up to 120 m of massive and bedded Sayunei-Shezal transition (Eisbacher 1985).
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 407

Palaeomagnetic data place the Rapitan rift basins in palaeolati- would have quickly oxygenated the surface waters, leading to
tudes comparable to the present southern Red Sea Rift basin abiotic Fe(III) precipitation.
(Evans & Raub 2011). At the Last Glacial Maximum (LGM, After the Sayunei basin became ice-free, the water vapour
c. 20 ka), arguably as severe a glaciation as any in the Phanerozoic content of air blowing across the basin would have increased due
eon, the lowest nearby glacial moraines, in the Ethiopian Highlands, to evaporation. This would likely have caused a positive change
are 3750 m above sea level (Umer et al. 2004). Air temperatures at in the mass-balance of downwind ice sheets, due to increased pre-
that altitude are on average 26 8C colder than the same latitude at sea cipitation and decreased melting on account of fog and clouds. On
level, reached by Rapitan glaciers. This gives some sense of the con- a ‘modern Snowball Earth’ (i.e. present geography with complete
trast between the Rapitan glaciation and the LGM. sea-ice cover, Voigt & Marotzke 2009), the Red Sea Rift would be
isolated from the Indian Ocean because of the roughly 1.0 km fall
in global mean sea level (glacioeustasy). As tropical sea-surface
Origin of ‘transitional-facies’ Fe-formation temperatures rose to melting point, sea ice on the Red Sea basin
would disappear, while the Indian Ocean would remain ice-bound
A local hydrothermal source of Fe for transitional-facies because of sea – glacial flow from the south (Warren et al. 2002;
Fe-formation (Fig. 36.3d) was favoured by Yeo (1981) and Goodman & Pierrehumbert 2003). Accordingly, the Shezal dia-
Young (1988, 2002). However, when compared with Neoarchaean mictite would represent the time interval between the opening of
and Palaeoproterozoic Fe-formations, the rare-earth element the Sayunei basin ‘oasis’ and overall glacial termination. Once
(REE) chemistry of the Rapitan Fe-formation is ‘much less dis- illuminated, primary productivity in the basin, now crowded
tinctly influenced by hydrothermal input’, suggesting that such with icebergs, would be stimulated.
input was ‘highly diluted by ocean waters at Rapitan time’ Fe-isotope and Ce-anomaly data (Fig. 36.7) shed additional light
(Klein & Beukes 1993). Moreover, where submarine volcanism on the basin-facies Fe-formation near Hayhook Lake (Fig. 36.3c).
was demonstratively active during the Rapitan glaciation (Mount Controls on Fe-isotope variation in natural systems are not well
Harper area), Fe-formation is absent. Where Fe-formation is known, and data from ancient rocks are still quite limited. Labora-
present, in the Tatonduk inlier (Macdonald & Cohen 2011) and tory experiments indicate that broadly similar equilibrium and
the Mackenzie Mountains, interbedded volcanic rocks are non- kinetic isotope fractionations are associated with both abiotic
existent (Macdonald et al. 2010b). and biological Fe-oxidation pathways (Johnson & Beard 2006).
Global Fe-speciation data suggest that Cryogenian and Edia- Consider the d57Fe trend in the Fe-formation (Fig. 36.7) not in
caran deep waters were ferruginous, rather than euxinic (Canfield terms of a secular change in seawater Fe-isotope composition,
et al. 2008). This raises the question why Fe-formation (hematite but as a change in net Fe-isotope fractionation (D57Fehem-Fe2þ)
jaspilite) was only deposited during glacial times, and possibly between hematite (or ferric-oxyhydroxide precursor) and dis-
only during the ‘Sturtian’ glaciation (Hoffman et al. 2011). If the solved Fe(II) across a redox chemocline in the water column
Rapitan rift-basins were covered by thick permanent marine ice (Halverson et al. 2011). This could come about either because
during fine-grained sedimentation of the Sayunei Fm. (Eisbacher the Fe-isotope fractionation varied as a function of the Fe(II)
1985), primary productivity would have crashed, limiting bacterial concentration, assumed to increase with depth across the chemo-
sulphate reduction and allowing high concentrations of Fe in acidic cline, or because of progressive oxidation of Fe(II) as it upwelled
waters in the absence of hydrothermal sources (Mikucki et al. across the chemocline. Progressive oxidation would drive the iso-
2009). The Crest Fe deposit at Iron Creek (Fig. 36.3d) is situated topic composition of the dissolved Fe(II) to more 57Fe-depleted
within a sequence of transitional-facies diamictites (Eisbacher values, lowering the d57Fe of the hematite produced. In either
1985), where local oxidizing power capable of titrating dissolved case, the d57Fe of hematite should decrease from the base of the
Fe would have been provided by subglacial meltwater discharges chemocline to the top. Accordingly, the upward increase in
at basin-margin ice grounding-lines, assuming air bubbles in the d57Fe observed within the Fe-formation (Fig. 36.7) suggests an
meteoric ice (compressed snow) contained oxygen. increase in water depth with time, which is consistent with more
Low primary productivity related to the permanent ice cover as reducing conditions up-section inferred from the Ce/Ce* data
well as low seawater pH would account for elevated P/Fe ratios in and with transgression inferred on sedimentological grounds
the Fe-formation at Iron Creek and perhaps other synglacial Cryo- over the same stratigraphic interval (Klein & Beukes 1993). As
genian Fe-formations (Planavsky et al. 2010). The explanation for the Sayunei Fe-formation is directly overlain by the Shezal diamic-
the high P/Fe ratio favoured by Planavsky et al. (2010) involves tite (Fig. 36.3), the rise in relative sea level is logically attributed to
‘unprecedented continental P fluxes during postglacial and inter- lithospheric downwarping under the load of the advancing Shezal
glacial time periods, given the extraordinary extent and duration ice sheet (Halverson et al. 2011).
of Cryogenian ice cover and the high levels of P delivery expected
from glaciated catchments’. The problem with their explanation is
that six of the seven Cryogenian Fe-formations in their database Palaeoenvironmental interpretation of the Stelfox (Ice Brook)
were deposited before the termination of the first major glaciation
Cryogenian glaciation.
The Stelfox Member of the Ice Brook Fm. is a subaqueous unit
interpreted as glaciomarine because of the presence of faceted
Origin and stratigraphic localization of ‘basin-facies’ and striated clasts in diamictite, sand and granule aggregates
Fe-formation (‘till pellets’), dropstones with impact-related structures in finely
laminated mudrocks, extremely angular quartz grains derived
Why does Fe-formation in the basin facies (Eisbacher 1985) occur from well-rounded quartz arenites of the Keele Fm., and diamictite
at the top of the Sayunei Fm., disconformably beneath the Shezal in a shallow-shelf setting where large-scale mass-flows are less
diamictite (Fig. 36.3b,c)? If we accept Eisbacher’s (1985) expla- likely to occur (Aitken 1991a, b).
nation that the advance of the Shezal grounding-line was triggered In discussing the Stelfox-equivalent Vreeland Diamictites in
by the removal of permanent sea ice over the basin, two factors northeastern British Columbia, McMechan (2000a, b) described
would have contributed to the deposition of Fe-formation. First, mud-rich diamictites and related fine-grained terrigenous facies
as the ice cover thinned and finally disappeared, anoxic and oxy- as resembling deposits associated with temperate, wet-based gla-
genic photosynthesis could have precipitated Fe2O3-precursor ciers in southern Alaska, in contrast to those of cold-based glaciers
from anoxic Fe(II)-rich basin waters. Second, after the ice cover in Antarctica. She argued that they are inconsistent with the
was removed, air –sea gas exchange and wind-driven mixing existence of a Cryogenian ‘Snowball’ Earth, assuming that its
408 P. F. HOFFMAN & G. P. HALVERSON

mean-annual tropical surface temperature was close to that of carbonate from dissolution of carbonate rock ‘flour’, exposed by
present-day Antarctica. She noted the prevalence of biogenic deglaciation (Fairchild 1993). Once the ocean was mixed, the post-
facies (e.g. diatomaceous ooze) beneath Antarctic ice shelves, on glacial excess of Ca output over input drove the positive d44Ca excur-
the Antarctic continental shelf and adjacent to the outlets of sion. Silva-Tamayo et al. (2010), who assume that the stratigraphic
fast-flowing ice streams, implying a near-absence of fine-grained expression of the negative d44Ca excursion corresponds to its time
terrigenous input from meltwater plumes (Domack 1988; Ander- of formation, attribute the negative excursion to Ca influx from
son et al. 1991). Citing the same authors, she posited that signifi- weathering immediately after deglaciation. Their model requires
cant meltwater flow requires surface melting and therefore that Ca input greatly exceeded output, which appears unlikely at a
summer air temperatures above 0 8C, which do not occur on Ant- time of ocean warming (raising saturation) and mixing (facilitating
arctica or on a ‘Snowball’ Earth (McMechan 2000b). degassing), when the surface ocean is known to have been critically
Much has been learned about the Antarctic glacial regime since oversaturated from the widespread occurrence of syndeglacial cap
the studies cited by McMechan (2000a, b). More than 150 subgla- dolostones (Hoffman et al. 2007).
cial lakes have been found, some at the origins of fast-flowing ice
streams, and they are not stagnant but are subject to active hydro-
dynamic exchange (Siegert et al. 2005; Wingham et al. 2006; Bell Correlation and palaeogeography
et al. 2007). Modelling suggests that basal melting occurs under
most of the interior of the East Antarctic Ice Sheet and all of its The post-glacial carbonate sequences following the older and
outlet ice streams (Pattyn 2010). Recent collapse of the Larsen B younger Cryogenian glaciations are stratigraphically, lithologi-
ice shelf revealed the former existence of both terrigenous and bio- cally and isotopically distinct (Kennedy et al. 1998; Hoffman &
genic sedimentation beneath the ice shelf (Domack et al. 2005; Schrag 2002). They provide a surprisingly successful basis for dis-
Damiani & Giorgetti 2008). Springtime diatom blooms on the tinguishing the glaciations globally (Knoll et al. 2006; Hoffman
East Antarctic margin may be triggered, not inhibited, by Fe-rich et al. 2011). As cosmopolitans, however, they are less well
sub-glacial meltwater disharges, from which terrigenous mud sub- suited for palaeogeographic reconstruction. In a wide orogen like
sequently settles (Leventer et al. 2006). There is no doubt that the North American Cordillera, the distinction between Neoproter-
accumulation rates for muds are lower in polar than in temperate ozoic successions that are indigenous to Laurentia and those orig-
settings, but sedimentation rates in the Cryogenian are as yet inating elsewhere is fundamental (Johnston 2008; Hildebrand
unknown in the absence of biogenic facies, varve chronologies, 2009). All exposed Neoproterozoic rocks in the Mackenzie Moun-
or data on meteoritic dust content. We agree that ‘any model for tains (Fig. 36.1) were thrust northeastward relative to cratonic
Neoproterozoic glaciation should include a significant period of Laurentia in the Upper Cretaceous –Paleocene. Sequence and
more moderate glacial conditions prior to the end of glaciation’ chemostratigraphic comparison of successions deposited before,
(McMechan 2000b), but we maintain that even the most extreme between and after the Cryogenian glaciations implies that the
glaciation must experience more moderate conditions at its termin- Tatonduk, Coal Creek, Hart River and Wernecke inliers of the
ation, especially in tropical marine environments, and that this is Yukon Territory expose a Neoproterozoic succession continuous
fully consistent with the hysteresis loop predicted by the Snowball with that of the Mackenzie Mountains (Macdonald & Roots
Earth hypothesis (Hoffman 2009). 2009; Macdonald et al. 2010a, b; Macdonald & Cohen 2011).
The co-occurrence of 718– 716 Ma mafic magmatism in the
Mount Harper area (Coal Creek inlier) and on cratonic Laurentia
Calcium isotope record of the Stelfox glaciation and (Franklin LIP) therefore implies that the Neoproterozoic succes-
its aftermath sion of the Ogilvie and Mackenzie Mountains formed on the
margin of Laurentia.
The Snowball Earth hypothesis (Hoffman & Schrag 2002) pro- In contrast, the lithostratigraphically alien Katakturuk Dolomite
vides a potential explanation for the paired negative and positive (Ediacaran) in the northeastern Brooks Range, Arctic Alaska, is
d44Ca excursions and for their stratigraphic location in the lower unlikely to have had a Laurentian affinity before Siluro-Devonian
Hayhook Fm. (Silva-Tamayo et al. 2010). During a snowball gla- time, at the earliest (Macdonald et al. 2009). Accordingly, the stra-
ciation, a steady rise in CO2 drives carbonate dissolution in the tigraphically underlying Hula Hula diamictite represents a Cryo-
ocean (a source of Ca) and precludes carbonate deposition genian glaciation of Chukotka-Arctic Alaska, not Laurentia, and
(Hoffman & Schrag 2002, fig. 10). This is consistent with the the Mount Copleston volcanics within and beneath the diamictite
absence of primary carbonate in the relevant synglacial strata. are predicted to be asynchronous with the Franklin LIP.
The d44Ca of dissolved Ca in snowball brine should therefore What about Neoproterozoic successions in the rest of the North
fall towards the value of the Ca input. The first carbonate that pre- American Cordillera? They host named glacial horizons in British
cipitates from the brine will be depleted in 44Ca by c. 1.0‰ com- Columbia (Toby and Vreeland), Idaho (Scout Mountain), Utah
pared with normal carbonates. In the snowball aftermath, Ca input (Dutch Peak and Mineral Fork), California (Surprise and Wild-
to the ocean from erosion is high, but rapid ocean warming and the rose) and Sonora (Mina el Mezquite). Conventionally, all have
more leisurely drawdown of CO2, through silicate weathering, been assumed to originate at the rifted margin of Laurentia
drive excess Ca output (carbonate deposition) over input, raising (Stewart 1972). In the Cordilleran collisional model (Johnston
d44Ca. During deglaciation, however, ocean surface waters are 2008; Hildebrand 2009), all except possibly the Mineral Fork ori-
flooded by glacial meltwater, marine and terrestrial, and are thus ginated elsewhere. In that model, they collided with the Laurentian
isolated from the deep brine by an exceptionally stable density margin in the Late Cretaceous as part of a composite ribbon con-
stratification (Shields 2005). Consequently, the 44Ca-depleted tinent named ‘Rubia’ (Hildebrand 2009) or ‘Saybia’ (Johnston
snowball signature is not observed in the Ravensthroat Fm., 2008). Targeted detrital zircon dating of the various glacial hor-
which was deposited in surface waters during the snowball melt- izons, and their host successions, should be an effective means
down (James et al. 2001). As melting neared completion, the of testing the Rubia/Saybia hypothesis. From the perspective of
flux of meltwater waned and ocean mixing by winds and tides the Mackenzie Mountains, the sequence stratigraphy and lithology
resumed. The negative d44Ca excursion in the lower Hayhook Fm. of Neoproterozoic successions assigned to the Windermere Super-
(Silva-Tamayo et al. 2010) records the mixing of 44Ca-depleted group in British Columbia (Ross et al. 1989) and to the South are
snowball brine and the meltwater lid. It does not record the gener- not similar. The differences may reflect facies changes, as conven-
ation of the isotope anomaly in dissolved Ca of the brine, but tionally assumed, but it seems fair to conclude that Neoproterozoic
rather its delayed expression as precipitated carbonate when the strata of the North American Cordillera provide no grounds at
brine was mixed into the meltwater lid, which was saturated with present for rejecting the Rubia/Saybia hypothesis. To date, all
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 409

that we do know is that palaeomagnetic data from the Franklin LIP Day, E. S., James, N. P., Narbonne, G. M. & Dalrymple, R. W.
place the then northward-facing rifted margin of Laurentia in the 2004. A sedimentary prelude to Marinoan glaciation, Cryogenian
Mackenzie Mountains at 18 + 38N latitude at the onset of the (Middle Neoproterozoic) Keele Formation, Mackenzie
Rapitan glaciation, and possibly closer to the palaeo-equator Mountains, northwestern Canada. Precambrian Research, 133,
during the Stelfox glaciation. 223– 247.
Denyszyn, S. W., Halls, H. C., Davis, D. W. & Evans, D. A. D. 2009.
Fieldwork in the Mackenzie Mountains was licensed by the Aurora Research Paleomagnetism and U– Pb geochronology of Franklin dykes in High
Institute and supported by a grant from the Astrobiology Institute of the US Arctic Canada and Greenland: a revised age and paleomagnetic pole
National Aeronautics and Space Administration (NASA), and by grants constraining block rotations in the Nares Strait region. Canadian
EAR-9905495 and EAR-0417422 from the US National Science Foundation Journal of Earth Sciences, 46, 689– 705.
(NSF) to P.F.H. C. A. Ferguson, M. T. Hurtgen, F. A. Macdonald, A. C. Domack, E. W. 1988. Biogenic facies in the Antarctic glacimarine
Maloof, Y. Shen and A. V. Turchyn participated in the fieldwork. C and O iso- environment: basis for a polar glacimarine summary. Palaeogeogra-
topes were analysed at the Laboratory for Geochemical Oceanography at phy, Palaeoclimatology, Palaeoecology, 63, 357– 372.
Harvard University, directed by D. P. Schrag. Sr isotope data were obtained at Domack, E. W., Duran, D. et al. 2005. Stability of the Larsen B ice shelf
the Geochronology Laboratory at the Massachusetts Institute of Technology, on the Antarctic Peninsula during the Holocene epoch. Nature, 436,
directed by S. A. Bowring. Fe isotopes were measured at the Laboratoire des 681– 685.
Mécanismes et Transferts en Géologie (LMTG) at the Université P. Sabatier, Dowdeswell, J. A., Whittington, J. A., Jennings, A. E., Andrews, J.
Toulouse, France, directed by F. Poitrasson. P.F.H. acknowledges discussions T., Mackensen, A. & Marienfield, P. 2000. An origin for lami-
with J. D. Aitken, R. W. Dalrymple, E. W. Domack, G. H. Eisbacher, nated glacimarine sediments through sea-ice build-up and suppressed
H. Gabrielse, L. M. Heaman, H. J. Hofmann, N. P. James, C. W. Jefferson, iceberg rafting. Sedimentology, 47, 557– 576.
A. J. Kaufman, T. Kurt Kyser, M. E. McMechan, W. A. Morris, D. C. Murphy, Eisbacher, G. H. 1978. Re-definition and Subdivision of the Rapitan
G. M. Narbonne, J. K. Park, G. M. Ross, G. M. Yeo and G. M. Young. Comments Group, Mackenzie Mountains. Geological Survey of Canada Paper
by E. Arnaud, G. Narbonne, E. C. Turner and G. Yeo substantially improved 77– 35.
the manuscript. This represents a contribution of the IUGS- and UNESCO- Eisbacher, G. H. 1981a. Sedimentary tectonics and glacial record in the
funded IGCP (International Geoscience Programme) project #512. Windermere Supergroup, Mackenzie Mountains, northwestern
Canada. Geological Survey of Canada Paper, 80– 27.
Eisbacher, G. H. 1981b. Late Precambrian tillites of the northern Yukon-
References Northwest Territories region, Canada. In: Hambrey, M. J. &
Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial Record.
Aitken, J. D. 1981. Stratigraphy and sedimentology of the Upper Proter- Cambridge University Press, Cambridge, 724–727.
ozoic Little Dal Group, Mackenzie Mountains, Northwest Territories. Eisbacher, G. H. 1985. Late Proterozoic rifting, glacial sedimentation and
In: Campbell, F. H. A. (ed.) Proterozoic Basins of Canada. Geologi- sedimentary cycles in the light of Windermere deposition, western
cal Survey of Canada Paper, 81 – 10, 47– 71. Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 51,
Aitken, J. D. 1982. Precambrian of the Mackenzie fold belt — a strati- 231– 254.
graphic and tectonic overview. In: Hutchinson, R. W., Spence, C. Evans, D. A. D. 2000. Stratigraphic, geochronological and Paleomagnetic
D. & Franklin, J. M. (eds) Precambrian Sulfide Deposits. Geologi- constraints upon the Neoproterozoic climatic paradox. American
cal Association of Canada Special Paper, 25, 149– 161. Journal of Science, 300, 347–433.
Aitken, J. D. 1991a. Two late Proterozoic glaciations, Mackenzie Moun- Evans, D. A. D. & Raub, T. D. 2011. Neoproterozoic glacial Palaeolati-
tains, northwestern Canada. Geology, 19, 445– 448. tudes: a global update. In: Arnaud, E., Halverson, G. P. & Shields,
Aitken, J. D. 1991b. The Ice Brook Formation and post-Raptian, Late G. (eds) The Geological Record of Neoproterozoic Glaciations. Geo-
Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. logical Society, London, Memoirs, 36, 93 –112.
Geological Survey of Canada Bulletin, 404. Fairchild, I. J. 1993. Balmy shores and ice wastes: the paradox of carbon-
Aitken, J. D. & Long, D. G. F. 1978. Mackenzie tectonic arc – reflection ates associated with glacial deposits in Neoproterozoic times. Sedi-
of early basin configuration? Geology, 6, 626–629. mentology Review, 1, 1 –16.
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves in the Farkas, J., Böhm, F. et al. 2007. Calcium isotope record of Phanerozoic
aftermath of a Neoproterozoic glaciation. Nature, 433, 123–127. oceans: implications for chemical evolution of seawater and its
Anderson, J. B., Kennedy, D. S., Smith, M. J. & Domack, E. W. 1991. causative mechanisms. Geochimica et Cosmochimica Acta, 71,
Sedimentary facies associated with Antarctica’s floating ice masses. 5117–5134.
In: Anderson, J. B. & Ashley, G. M. (eds) Glacial Marine Sedimen- Friedman, I. & O’Neil, J. R. 1977. Chapter KK. Compilation of stable
tation: Paleoclimatic Significance. Geological Society of America isotope fractionation factors of geochemical interest. In: Fleischer,
Special Paper, 261, 1– 25. M. (ed.) Data of Geochemistry 6th edn. United States Geological
Assereto, R. L. A. M. & Kendall, C. G. St. C. 1977. Nature, origin and Survey Professional Paper 440– KK, Washington, DC.
classification of peritidal tepee structures and related breccas. Sedi- Gabrielse, H., Blusson, S. L. & Roddick, J. A. 1973. Geology of the
mentology, 24, 153– 210. Flat River, Glacier Lake and Wrigley Lake map-areas, District of
Bell, R. E., Studinger, M., Shuman, C. A., Fahnestock, M. A. & Mackenzie and Yukon Territory. Geological Survey of Canada
Joughin, I. 2007. Large subglacial lakes in East Antarctica at the Memoir 366.
onset of fast-flowing ice streams. Nature, 445, 904– 907. Goodman, J. & Pierrehumbert, R. T. 2003. Glacial flow of floating
Canfield, D. E., Poulton, S. W. & Narbonne, G. M. 2007. Late- marine ice in ‘Snowball Earth’. Journal of Geophysical Research,
Neoproterozoic deep-ocean oxygenation and the rise of animal life. 108, 3308, doi: 10.1029/2002JC001471.
Science, 315, 92 – 95. Green, L. H. & Godwin, C. I. 1963. Snake River area: mineral industry
Canfield, D. E., Poulton, S. W., Knoll, A. H., Narbonne, G. M., of Yukon Territory and southwestern District of Mackenzie, 1962.
Ross, G., Goldberg, T. & Strauss, H. 2008. Ferruginous conditions Geological Survey of Canada Paper 63 –38, 15 – 18.
dominated later Neoproterozoic deep-water chemistry. Science, 321, Halverson, G. P. 2006. A Neoproterozoic chronology. In: Xiao, S. &
949– 952. Kaufman, A. J. (eds) Neoproterozoic Geobiology and Paleobiology.
Dalrymple, R. W. & Narbonne, G. M. 1996. Continental slope sedi- Springer, Dordrecht, 231– 271.
mentation in the Sheepbed Formation (Neoproterozoic, Windermere Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
Supergroup), Mackenzie Mountains, N.W.T. Canadian Journal of glaciation (Neoproterozoic) in northeast Svalbard. Basin Research,
Earth Sciences, 33, 848–862. 16, 297– 324.
Damiani, D. & Giorgetti, G. 2008. Provenance of glacial-marine sedi- Halverson, G. P., Dudás, F. Ö., Maloof, A. C. & Bowring, S. A. 2007.
ments under the McMurdo/Ross Ice Shelf (Windless Bight, Antarc- Evolution of the 87Sr/86Sr composition of Neoproterozoic seawater.
tica): Heavy minerals and geochemical data. Palaeogeography, Palaeogeography, Palaeoclimatology, Palaeoecology, 256,
Palaeoclimatology, Palaeoecology, 260, 262– 283. 103– 129.
410 P. F. HOFFMAN & G. P. HALVERSON

Halverson, G. P., Poitrasson, F., Hoffman, P. F., Nedelec, A., Kaufman, A. J., Jacobsen, S. B. & Knoll, A. H. 1993. The Vendian
Montel, J.-M. & Kirby, J. 2011. Fe isotope and trace element geo- record of Sr and C isotopic variations in seawater: implications for
chemistry of the Neoproterozoic syn-glacial Raptian iron formation. tectonics and paleoclimate. Earth and Planetary Science Letters,
Earth and Planetary Science Letters, doi: 10.1016/j.epsl.2011.06.21. 120, 409– 430.
Hambrey, M. J. & Spencer, A. M. 1987. Late Precambrian Glaciation of Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes,
Central East Greenland. Meddelelser om Grønland, Geoscience 19. ice ages and terminal Proterozoic earth history. Proceedings of
Harlan, S. S., Heaman, L., LeCheminant, A. N. & Premo, W. R. 2003. the National Academy of Sciences (USA), 94, 6600– 6605.
Gunbarrel mafic magmatic event: a key 780 Ma time marker for Kendall, C. G. St. C. & Warren, J. 1987. A review of the origin and
Rodinia plate reconstructions. Geology, 31, 1053–1056. setting of tepees and their associated fabrics. Sedimentology, 34,
Heaman, L. M., LeCheminant, A. N. & Rainbird, R. H. 1992. Nature 1007– 1027.
and timing of Franklin igneous events, Canada: implications for a Kennedy, M. J. 1996. Stratigraphy, sedimentology and isotopic geochem-
Late Proterozoic mantle plume and the break-up of Laurentia. istry of Australian Neoproterozoic postglacial cap dolostones: degla-
Earth and Planetary Science Letters, 109, 117–131. ciation, d13C excursions and carbonate precipitation. Journal of
Hildebrand, R. S. 2009. Did westward subduction cause Cretaceous- Sedimentary Research, 66, 1050– 1064.
Tertiary orogeny in the North American Cordillera? Geological Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K.-H. &
Society of America, Special Paper 457. Arthur, M. A. 1998. Two or four Neoproterozoic glaciations?
Hoffman, P. F. 2009. Pan-glacial – a third state in the climate system. Geology, 26, 1059–1063.
Geology Today, 25, 107– 114. Klein, C. & Beukes, N. J. 1993. Sedimentology and geochemistry of the
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: glaciogenic Late Proterozoic Rapitan Fe-formation in Canada.
testing the limits of global change. Terra Nova, 14, 129–155. Economic Geology, 88, 542–565.
Hoffman, P. F. & Macdonald, F. A. 2010. Sheet-crack cements and Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blick, N.
early regression in Marinoan (635 Ma) cap dolostones: regional 2006. The Ediacaran Period: a new addition to the geologic time
benchmarks of vanishing ice-sheets? Earth and Planetary Science scale. Lethaia, 39, 13 – 30.
Letters, 300, 374–384. Leventer, A., Domack, E. et al. 2006. Marine sediment record from
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., the East Antarctic margin reveals dynamics of ice sheet recession.
Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran GSA Today, 16, 4– 10.
(635 Ma) post-glacial ‘cap dolostones’ diachronous? Earth and Pla- Love, G. D., Grosjean, E. et al. 2009. Fossil steroids record the appear-
netary Science Letters, 258, 114–131. ance of Demospongiae during the Cryogenian period. Nature, 457,
Hoffman, P. F., Macdonald, F. A. & Halverson, G. P. 2011. Chemical 718– 721, doi: 10.1038/nature07673.
sediments associated with Neoproterozoic glaciation: iron formation, Lyons, T. W. & Severmann, S. 2006. A critical look at iron paleoredox
cap carbonate, barite and phosphorite. In: Arnaud, E., Halverson, proxies based on new insights from modern euxinic marine basins.
G. P. & Shields, G. (eds) The Geological Record of Neoproterozoic Geochimica et Cosmochimica Acta, 70, 5698– 5722.
Glaciations. Geological Society, London, Memoirs, 36, 67 –80. Macdonald, F. A. & Roots, C. F. 2009. Upper Fifteenmile Group in the
Hofmann, H. J., Narbonne, G. M. & Aitken, J. D. 1990. Ediacaran Ogilvie Mountains and correlations of early Neoproterozoic strata in
remains from intertillite beds in northwestern Canada. Geology, 18, the northern Cordillera. In: MacFarlane, K. E., Weston, L. H. &
1199– 1202. Blackburn, L. R. (eds) Yukon Exploration and Geology 2009,
Hurtgen, M. T., Arthur, M. A., Suits, N. S. & Kaufman, A. J. 2002. Yukon Geological Survey, Whitehorse, 237–252.
The sulfur isotopic composition of Neoproterozoic seawater sulfate: Macdonald, F. A. & Cohen, P. A. 2011. The Tatonduk inlier,
implications for a snowball Earth? Earth and Planetary Science Alaska-Yukon border. In: Arnaud, E., Halverson, G. P. &
Letters, 203, 413–429. Shields, G. (eds) The Geological Record of Neoproterozoic
Hurtgen, M. T., Halverson, G. P., Arthur, M. A. & Hoffman, P. F. Glaciations. Geological Society, London, Memoirs, 36, 389–396.
2006. Sulfur cycling in the aftermath of a 635-Ma snowball Macdonald, F. A., Mcclelland, W. C., Schrag, D. P. & Macdonald,
glaciation: evidence for a syn-glacial sulfidic deep ocean. Earth W. P. 2009. Neoproterozoic glaciation on a carbonate platform
and Planetary Science Letters, 245, 551–570. margin in Arctic Alaska and the origin of the North Slope subterrane.
James, N. P., Narbonne, G. M. & Kyser, T. K. 2001. Late Neoprotero- Geological Society of America Bulletin, 121, 448–473.
zoic cap carbonates: Mackenzie Mountains, northwestern Canada: Macdonald, F. A., Cohen, P. A., Dudás, F. Ö. & Schrag, D. P. 2010a.
precipitation and global glacial meltdown. Canadian Journal of Early Neoproterozoic siliceous scale microfossils in the Lower Tindir
Earth Science, 38, 1229– 1262. Group of Alaska and the Yukon Territory. Geology, 38, 143– 146.
James, N. P., Narbonne, G. M., Dalrymple, R. W. & Kyser, T. K. 2005. Macdonald, F. A., Schmitz, M. D. et al. 2010b. Calibrating the
Glendonites in Neoproterozoic low-latitude, interglacial, sedimentary Cryogenian. Science, 327, 1241– 1243.
rocks, northwest Canada: insights into Cryogenian ocean and Maloof, A. C., Rose, C. V. et al. 2010. Possible animal-body fossils in
Precambrian cold-water carbonates. Geology, 33, 9 – 12. pre-Marinoan limestones from South Australia. Nature Geoscience,
Jefferson, C. W. 1983. The Upper Proterozoic Redstone Copper Belt, 3, 653– 659.
Mackenzie Mountains, Northwest Territories. PhD thesis, University McCausland, P. J. A., Van der Voo, R. & Hall, C. M. 2007.
of Western Ontario, London, Ontario, 445. Circum-Iapetus paleogeography of the Precambrian– Cambrian
Jefferson, C. W. & Ruelle, J. C. L. 1986. The Late Proterozoic transition with a new paleomagnetic constraint from Laurentia.
Redstone Copper Belt, Mackenzie Mountains, Northwest Territories. Precambrian Research, 156, 125– 152.
In: Morin, J. A. (ed.) Mineral Deposits of Northern Cordillera. McMechan, M. E. 2000a. Vreeland Diamictites – Neoproterozoic
Canadian Institute of Mining and Metallurgy Special Volume, 37, glaciogenic slope deposits, Rocky Mountains, northeast British
154– 168. Columbia. Bulletin of Canadian Petroleum Geology, 48, 246– 261.
Johnson, C. M. & Beard, B. L. 2006. Fe isotopes: an emerging technique McMechan, M. E. 2000b. Reply to discussion. Vreeland Diamictites –
for understanding modern and ancient biogeochemical cycles. GSA Neoproterozoic glaciogenic slope deposits, Rocky Mountains, north-
Today, 16, 4– 10. east British Columbia. Bulletin of Canadian Petroleum Geology, 48,
Johnston, S. T. 2008. The Cordilleran ribbon continent of North America. 364– 366.
Annual Reviews of Earth and Planetary Sciences, 36, 495– 530. Mikucki, J. A., Pearson, A. et al. 2009. A contemporary microbially
Kamo, S. L. & Gower, C. F. 1994. Note: U– Pb baddeleyite maintained subglacial ferrous ‘ocean’. Science, 324, 397– 400.
dating clarifies age of characteristic paleomagnetic remanence of Morris, W. A. 1977. Paleolatitude of upper Precambrian glaciogenic
Long Range dykes, southeastern Labrador. Atlantic Geology, 30, Rapitan Group and the use of tillites as chronostratigraphic marker
259– 262. horizons. Geology, 5, 85 – 88.
Kamo, S. L., Gower, C. F. & Krogh, T. E. 1989. Birthplace for the Mustard, P. S. & Roots, C. F. 1997. Rift-related volcanism, sedimen-
Iapetus Ocean? A precise U –Pb zircon and baddeleyite age for the tation and tectonic setting of the Mount Harper Group, Ogilvie Moun-
Long Range dikes, southeast Labrador. Geology, 17, 602–605. tains, Yukon Territory. Geological Survey of Canada Bulletin, 492.
THE MACKENZIE MOUNTAINS, NORTHERN CANADIAN CORDILLERA 411

Narbonne, G. M. 1994. New Ediacaran fossils from the Mackenzie Silva-Tamayo, J. C., Nägler, T. F. et al. 2010. Global Ca isotope
Mountains, northwestern Canada. Journal of Palaeontology, 68, variations in c. 0.7 Ga old post-glacial carbonate successions. Terra
411– 416. Nova, 22, 188– 194.
Narbonne, G. M. & Hofmann, H. J. 1987. Ediacaran biota of the Stewart, J. H. 1972. Initial deposits in the Cordilleran Geosyncline: evi-
Wernecke Mountains, Yukon, Canada. Palaeontology, 30, 647– 676. dence of a Late Precambrian (,850 m.y.) continental separation.
Narbonne, G. M. & Aitken, J. D. 1990. Ediacaran fossils from the Sekwi Geological Society of America Bulletin, 83, 1345– 1360.
Brook area, Mackenzie Mountains, northwestern Canada. Palaeon- Turner, E. C. & Long, D. G. F. 2008. Basin architecture and syndeposi-
tology, 33, 945– 980. tional fault activity during deposition of the Neoproterozoic Macken-
Narbonne, G. M. & Aitken, J. D. 1995. Neoproterozoic of the Macken- zie Mountains supergroup, Northwestern Territories, Canada.
zie Mountains, northwestern Canada. Precambrian Research, 73, Canadian Journal of Earth Sciences, 45, 1159– 1184.
101– 121. Turner, E. C., James, N. P. & Narbonne, G. M. 1997. Growth
Narbonne, G. M., Kaufman, A. J. & Knoll, A. H. 1994. Integrated che- dynamics of Neoproterozoic calcimicrobial reefs, Mackenzie
mostratigraphy and biostratigraphy of the Windermere Supergroup, Mountains, northwest Canada. Journal of Sedimentary Research,
northwestern Canada: implications for Neoproterozoic correlations 67, 437– 450.
and early evolution of animals. Geological Society of America Bulle- Umer, M., Kebede, S. & Osmaston, H. 2004. Quaternary glacial activity
tin, 106, 1281– 1292. on the Ethiopian mountains. In: Ehlers, J. & Gibbard, P. L. (eds)
Ovenshine, T. A. 1970. Observations of ice rafting in Glacier Bay, Quaternary Glaciations – Extent and Chronology, Part III: South
Alaska, and the identification of ancient ice-rafted deposits. Geologi- America, Asia, Africa, Australia, Antarctica. Elsevier, Amsterdam,
cal Society of America Bulletin, 81, 891–894. 171–174.
Park, J. K. 1994. Palaeomagnetic constraints on the position of Laurentia Upitis, U. 1966. The Rapitan Group, southwestern Mackenzie Mountains,
from middle Neoproterozoic to Early Cambrian times. Precambrian Northwest Territories. MSc thesis, McGill University, Montreal.
Research, 69, 95 – 112. Voigt, A. & Marotzke, J. 2009. The transition from the present-day
Park, J. K. 1997. Paleomagnetic evidence for low-latitude glaciation climate to a modern Snowball Earth. Climate Dynamics, doi:
during deposition of the Neoproterozoic Rapitan Group, Mackenzie 10.1007/s00382-009-0633-5.
Mountains, N.W.T., Canada. Canadian Journal of Earth Sciences, Wallace, M. W. & Woon, E. 2008. Giant Cryopgenian reefs as windows
34, 34 – 49. into pre-Ediacaran life. In: Gallagher, S. J. & Wallace, M. W.
Pattyn, F. 2010. Antarctic subglacial conditions inferred from a hybrid (eds) Selwyn Symposium 2008: Neoproterozoic Extreme Climates
ice sheet/ice stream model. Earth and Planetary Science Letters, and the Origin of Early Metazoan Life. Geological Society of Austra-
295, 451– 461. lia, Extended Abstracts, 91, 17 – 22.
Planavksy, N. J., Rouxel, O. J., Bekker, A., Lalonde, S. V., Konhau- Warren, S. G., Brandt, R. E., Grenfell, T. C. & McKay, C. P. 2002.
ser, K. O., Reinhard, C. T. & Lyons, T. W. 2010. The evolution of Snowball Earth: ice thickness on the tropical ocean. Journal of Geo-
the marine phosphate reservoir. Nature, 467, 1088– 1090. physical Research, 107(C10), 3167, doi: 10.1029/2001JC001123.
Pyle, L. J., Narbonne, G. M., James, N. P., Dalrymple, R. W. & Weil, A. B., Geismann, J. W. & Van der Voo, R. 2004. Paleomagnetism
Kaufman, A. J. 2004. Integrated Ediacaran chronostratigraphy, of the Neoproterozoic Chuar Group, Grand Canyon Supergroup,
Wernecke Mountains, northwestern Canada. Precambrian Research, Arizona: implications for Laurentia’s Neoproterozoic APWP and
132, 1– 27. Rodinia break-up. Precambrian Research, 129, 71 –92.
Raiswell, R. & Canfield, D. E. 1998. Sources of iron for pyrite Weil, A. B., Geismann, J. W. & Ashby, J. M. 2006. A new paleomagnetic
formation in marine sediments. American Journal of Science, 298, pole for the Uinta Mountain supergroup, Central Rocky Mountain
219– 245. States, USA. Precambrian Research, 147, 234– 259.
Ramsay, J. G. & Huber, M. I. 1987. The Techniques of Modern Structural Wingham, D. J., Siegert, M. J., Shepherd, A. & Muir, A. S. 2006.
Geology, Vol. 2. Academic Press, New York. Rapid discharge connects Antarctic subglacial lakes. Nature, 440,
Ross, G. M. 1991. Tectonic setting of the Windermere Supergroup 1033–1036.
revisited. Geology, 19, 1125– 1128. Yeo, G. M. 1981. The Late Proterozoic Rapitan glaciation in the northern
Ross, G. M., Villeneuve, M. E. 1997. U– Pb geochronology of stranger Cordillera. In: Campbell, F. H. A. (ed.) Proterozoic Basins of
stones in Neoproterozoic diamictites, Canadian Cordillera: impli- Canada. Geological Survey of Canada Paper, 81 –10, 25 –46.
cations for provenance and ages of deposition. In: Radiogenic Age Yeo, G. M. 1984. The Rapitan Group: relevance to the global association
and Isotopic Studies: Report 10. Geological Survey of Canada of Late Proterozoic glaciation and Fe-formation. PhD thesis, Univer-
Current Research, 1997-F, 141– 155. sity of Western Ontario, London (Ontario).
Ross, G. M., McMechan, M. E. & Hein, F. J. 1989. Proterozoic history: Yeo, G. M. 1986. Fe-formation in the late Proterozoic Rapitan Group,
the birth of the miogeocline. In: Rickets, B. D. (ed.) Western Canada Yukon and Northwest Territories. In: Morin, J. A. (ed.) Mineral
Sedimentary Basin: A Case History. Canadian Society of Petroleum Deposits of Northern Cordillera. Canadian Institute of Mining and
Geologists, Calgary, 79 – 104. Metallurgy, Special Volume, 37, 142– 153.
Ruelle, J. C. L. 1982. Depositional environments and genesis of strati- Young, G. M. 1976. Fe-formation and glaciogenic rocks of the Rapitan
form copper deposits of the Redstone Copper Belt, Mackenzie Moun- Group, Northwest Territories, Canada. Precambrian Research, 3,
tains, N.W.T. In: Hutchinson, R. W., Spence, C. D. & Franklin, 137– 158.
J. M. (eds) Precambrian Sulfide Deposits. Geological Association Young, G. M. 1982. The late Proterozoic Tindir Group, east-central
of Canada Special Paper, 25, 701–737. Alaska: evolution of a continental margin. Geological Society of
Shen, Y., Zhang, T. & Hoffman, P. F. 2008. On the co-evolution of America Bulletin, 93, 759–783.
Ediacaran oceans and animals. Proceedings of the National Young, G. M. 1988. Proterozoic plate tectonics, glaciation and
Academy of Sciences USA, 105, 7376–7381. Fe-formations. Sedimentary Geology, 58, 127–144.
Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal Young, G. M. 2002. Stratigraphic and tectonic settings of Proterozoic
of existing models and the plumeworld hypothesis. Terra Nova, 17, glaciogenic rocks and banded Fe-formations: relevance to the
299 – 310. snowball Earth debate. Journal of African Earth Sciences, 35,
Siegert, M. J., Carter, S., Tabacco, I., Popov, S. & Blankenship, D. 451– 466.
2005. A revised inventory of Antarctic sub-glacial lakes. Antarctic Ziegler, P. 1959. Frühpaläozoische Tillite im östlichen Yukon-
Science, 17, 453– 460. Territorium (Kanada). Eclogae Geologicae Helvetiae, 52, 735– 741.
Chapter 37

The record of Neoproterozoic glaciations in the Windermere Supergroup,


southern Canadian Cordillera

MARK D. SMITH1, EMMANUELLE ARNAUD2*, R.W.C. ARNOTT1 & GERALD M. ROSS3


1
Department of Earth Sciences, University of Ottawa, Ottawa, Ontario, K1N 6N5, Canada
2
School of Environmental Sciences, University of Guelph, Guelph, Ontario, N1G 2W1, Canada
3
Kupa’a Farm, P.O.Box 458, Kula, Hawai’i, 96790, USA
*Corresponding author (e-mail: earnaud@uoguelph.ca)

Abstract: The Neoproterozoic Windermere Supergroup (WSG) is exposed over an area of 35 000 km2 in the southern Canadian
Cordillera, and consists primarily of deep-marine meta-sedimentary rocks interpreted to have been deposited during rifting and sub-
sequent post-rift thermal relaxation. The main exposures of the WSG occur within thrust panels and structural culminations of the
eastern Cordilleran orogen. Within the thick stratigraphic succession (c. 9 km) are three units of glaciogenic affinity: Toby, Vreeland
and Old Fort Point (OFP) formations. The Toby Formation (Fm.) is composed of up to 2500 m of diamictite, interbedded with conglom-
erate, sandstone, mudstone, carbonate and mafic volcanic rocks. The Vreeland Formation ranges from 350 m to 2000 m in thickness and
consists of diamictite, interbedded with mudstone, sandstone and conglomerate. The OFP ranges from 60 to 450 m in thickness and con-
sists of a distinctive threefold stratigraphic package of basal siltstone grading upward into limestone –siltstone rhythmite, organic-rich
mudstone and an overlying heterolithic unit of diamictite, breccia, conglomerate, sandstone, siltstone to mudstone and limestone. A
locally prominent unconformity marks the base of the OFP upper member. Both the Toby and Vreeland formations represent remobilized
glacially derived marine sediments deposited by sediment-gravity flows. Deposition of the Toby Fm. was fault-controlled during an
active tectonic phase (rifting), whereas the Vreeland Fm. accumulated during the subsequent quiescent phase (post-rift) with limited
structural control. The OFP is interpreted to be a regionally extensive deep-marine post-glacial marker temporally associated with the
glaciogenic Vreeland Fm. Although direct geochronological ages for WSG units in southwestern Canada are generally absent, high-
precision radiometric ages of underlying and overlying igneous events constrain the relative maximum and minimum timing of depo-
sition from c. 740– 728 Ma to c. 569 Ma. At the base of the WSG succession, the Toby Fm. may be as young as c. 685 Ma based on
ages obtained from potential stratigraphic correlatives in the USA. There is no direct age constraint for the deposition of the Vreeland
Fm.; its minimum timing is based on its stratigraphic relationship with the post-glacial OFP. The middle member of the OFP was pre-
cisely dated at 607.8 + 4.7 Ma using the Re–Os method, placing it in the Ediacaran Period. Published geochemical and stable isotopic
data are similarly limited for all three units with only some d34Spy values available from one section of the OFP. Recent work has focused
on detailed sedimentological and stratigraphic studies of the Toby and OFP formations with future efforts being directed towards inte-
grated geochemical and isotopic research. Additional geochronological constraints are needed to refine palaeogeographical models and
strengthen regional correlations with other North American Cordilleran glaciogenic units.

The record of Neoproterozoic glaciogenic sedimentation in the including mapping of the Toby Fm. and related structures in the
southern Canadian Cordillera (Fig. 37.1) is preserved in three Paradise Mine (Atkinson 1975), Delphine Creek (Root 1987),
units within the Windermere Supergroup (WSG): a glacial interval Mount Forster (Bennett 1986), Creston/Salmo (Glover & Price
at the base of the succession is inferred from the diamictite-bearing 1976) and Crawford Bay/Columbia Point areas (Lis & Price
Toby Fm., with another event higher in the succession inferred 1976). Ongoing investigations are focusing on the relationship
from the diamictite-bearing Vreeland Fm. (glacial) and Old Fort between Toby Fm. facies variability and structural trends as well
Point Fm. (post-glacial) (Fig. 37.2) (Ross et al. 1989, 1995). as isotopic analyses of associated strata.

Toby Fm. Vreeland Fm.

The Toby Fm. (also known as the Toby Conglomerate) is best The Vreeland Fm. has received little study (due largely to inac-
exposed in the Purcell Mountains of southeastern British Colum- cessibility) despite its impressive exposures and thickness in the
bia. The variability of its lithofacies has precluded the identifi- Pine and Monkman Pass regions, central British Columbia (from
cation of a type section; the formation was named by Walker 548050 N to 578000 N) (Fig. 37.1). It was originally mapped by
(1926) after well-exposed outcrops located along Toby Creek, Slind & Perkins (1966) as an unnamed Precambrian conglomer-
west of Invermere, British Columbia (508300 N, 1168020 W) ate –schist. Later, it was briefly examined by Stelck et al. (1978),
(Fig. 37.1, A). This area has been extensively modified by Proter- but was not really described until regional mapping conducted
ozoic, Devonian and Mesozoic deformation and the structural by McMechan (1987) and McMechan & Thompson (1985,
context of these rocks is often complicated (e.g. Root 1987). 1995a, b). It was initially referred to as a diamictite within the
Previous studies of the Toby Fm. are relatively few, with Misinchinka Group or the Mount Vreeland –Paksumo Pass dia-
primary emphasis on regional mapping relative to Proterozoic- mictites (McMechan 1990), until Hein & McMechan (1994)
and Mesozoic-age structures (Walker 1926; Leech 1954; Aalto referred to it as the Vreeland Fm. McMechan (2000) provided
1971; Reesor 1973; Root 1987; Pope 1990; Warren 1997). The the first detailed stratigraphic description and palaeogeographical
glacial origin of the Toby Fm. was first proposed based on a interpretation in the Monkman Pass area. No type section has
more detailed study of facies types and variability, petrography, been defined for the Vreeland Fm., although the Mount Vreeland
as well as trends in clast characteristics in the Toby Fm. itself sections (548350 N, 1218320 W) in the Snake Indian thrust sheet
(Aalto 1971). Additional mapping of specific regions has followed, (Fig. 37.1,B) and the Paksumo Pass section (548200 N,

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 413– 423. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.37
414 M. D. SMITH ET AL.

Fig. 37.1. Generalized geological map


showing outcrop distribution of
Windermere Supergroup in western Canada
and specific outcrop localities (A– K)
discussed in the text (modified from Ross
et al. 1995). A comprehensive list of
outcrop locations for the Toby Fm. can be
found in Aalto (1971) and Reesor (1973).
Outcrop localities for the Vreeland Fm. are
from McMechan (2000) and for the OFP
from Ross & Murphy (1988) and Smith
(2009). Mtns, Mountains.

1208490 W) in the Wapiti Pass thrust sheet (Fig. 37.1,C) of the (Charlesworth et al. 1967). A more representative and better
Monkman Pass area are thought to be representative exposures exposed section crops out c. 15 km west of Jasper along the
(McMechan 2000). major highway (528520 2100 N, 1188170 3500 W). The regionally
extensive nature of the OFP has led to various names: (i) basal
Hector Fm. (Walcott 1910) or Mount Temple and Taylor Lake
Old Fort Point Fm. members (e.g. Gussow 1957) in the Lake Louise area, Alberta;
(ii) Old Fort Point Fm. (e.g. Weiner 1966; Charlesworth et al.
The Old Fort Point Fm. (OFP) is the preferred name to describe the 1967) in the Jasper area, Alberta; (iii) Middle Miette Marker
unique stratigraphic marker in the WSG that is exposed locally (e.g. McDonough 1989) in the Mount Robson area, British Colum-
over the entire outcrop region of southwestern Canada bia; (iv) Kaza Group marker (e.g. Ross & Murphy 1988) in the
(Fig. 37.1,D – K). Walcott (1910) was the first to describe the Cariboo Mountains, British Columbia; (v) Baird Brook Division
OFP from the Lake Louise area of Alberta using its distinctive (e.g. Kubli 1990) in the Dogtooth Range of the Purcell Mountains,
purple and green colours and lithology (siltstone, limestone, British Columbia; (vi) Comedy Creek unit (Grasby & Brown
breccia) to separate the Precambrian Corral Creek (lower con- 1993), Selkirk Mountains, British Columbia; (vii) upper and
glomerate to sandstone) and Hector (upper slate) formations. lower markers (e.g. Warren 1997) in the Purcell Mountains,
The name and type section (528520 1600 N, 1188030 4000 W) come British Columbia (Fig. 37.1). Although there were early attempts
from outcrops at a prominent landmark near Jasper, Alberta to correlate between certain locations (Charlesworth et al. 1967;
WINDERMERE SUPERGROUP 415

Fig. 37.2. Comparative stratigraphic


columns for the Windermere Supergroup
from the Monkman Pass area and the
Cariboo-Purcell-Rocky Mountains
(modified from Ross et al. 1995).
Nomenclature of rift and post-rift strata
(Kaza-Cariboo, Horsethief Creek and
Miette groups refer to similar strata,
depending on location within the region).
Relative stratigraphic position of the
glaciogenic Toby, Vreeland and Old Fort
Point formations and proposed correlation
between the basal Framstead and Old Fort
Point formations are shown. Ages shown
are discussed in the text and, in
stratigraphic order, are Malton Gneiss
(McDonough & Parrish 1991); rift-related
volcanic rocks from Idaho (Lund et al.
2003; Fanning & Link 2004); Old Fort
Point Fm. (Kendall et al. 2004); and Hamill
Group volcanic rocks (Colpron et al. 2002).
Bracketed ages indicate that the ages are
not from rocks of the southern Canadian
Cordillera, but rather interpreted
correlatives of the Irene Fm.

Aitken 1969), it was Ross & Murphy (1988) who first recognized are found in the southern Cariboo and western Purcell Mountains,
the basin-wide correlation of the OFP (and equivalents), its palaeo- as well as the Shuswap Complex (e.g. Simony et al. 1980;
geographical significance and importance in Neoproterozoic event Ross et al. 1995). Areas of low metamorphic grade and relatively
stratigraphy. Subsequent work by Ross et al. (1995) examined the uncomplicated tectonic deformation show exceptional preservation
S-isotopic evolution of the OFP and WSG strata in the Cariboo of primary sedimentary textures and structures. This allowed the
Mountains, British Columbia (Fig. 37.1). A regionally comprehen- development of a consistent internal stratigraphy of the WSG and
sive study of the OFP integrating sedimentology, stratigraphy, reconstruction of the Windermere basin in southwestern Canada
geochemistry and stable isotopes has recently been completed (Ross & Murphy 1988; Ross et al. 1989, 1995; Ross 2000).
(Smith 2009). The WSG is thought to represent deposition in two tectonic
phases (e.g. Stewart 1972; Ross 1991). The first phase was charac-
terized by rifting during the break-up of the supercontinent
Structural framework and basin setting Rodinia, whereas the second phase occurred during post-rift
thermal relaxation. At the base of the WSG, the Toby Fm. together
The WSG forms part of a long arcuate belt of semicontinuous out- with the Irene Fm. volcanic rocks are interpreted to have accumu-
crops in western North America that extends from the Yukon – lated during the rifting phase (Aalto 1971; Glover & Price 1976;
Alaska border region to northwestern Mexico (e.g. Ross et al. Root 1987). The variable thickness of the Toby Fm. is attributed
1989). In southwestern Canada, exposures crop out over an area to syn-sedimentary fault-controlled deposition (Aalto 1971; Root
of 35 000 km2 in a series of thrust sheets in the Main Ranges of 1987). Based on the nature of Proterozoic-age faults, Root
the western Fold-and-Thrust Belt and the Omineca Belt of the (1987) suggested that the basin underwent two episodes of exten-
southern Canadian Cordillera (Fig. 37.1). Major structures affect- sion; one trending approximately east –west, resulting in a series of
ing WSG strata in the region include the Southern Rocky Mountain NNE and NNW faults and another trending NW –SE, resulting in a
Trench (Fig. 37.1) and a broad gently north-plunging structure in series of NE– SW normal faults.
the Purcell Mountains (the Purcell Anticlinorium). Proterozoic- The Vreeland Fm. and OFP both accumulated during the post-
age structures that affected WSG deposition include regionally sig- rift phase with possible local structural control (McMechan
nificant transfer faults (e.g Mount Forster and Moyie) and associ- 2000; Smith 2009). During this later phase, the basin is thought
ated uplifted blocks (e.g. Windermere High and Montania; Reesor to be an elongate NW-flowing turbidite system (Ross et al. 1989,
1973; Lis & Price 1976; Root 1987; Warren 1997). 1995; Ross 1991), based on consistent palaeocurrent data and
All WSG strata were subjected to Mesozoic orogenic defor- bimodal pattern (.2.6 Ga and 1.9–1.75 Ga) of U –Pb detrital
mation that resulted in a wide range of structural complexity and zircon provenance (Ross & Bowring 1990; Ross & Parrish 1991).
metamorphic grade between different structural panels. Extensive Basement clasts from the Vreeland Fm. yielded ages between
tracts of sub-greenschist to greenschist grade exist in the western 1865 to 1842 Ma (Ross & Villeneuve 1997), suggesting derivation
Main Ranges of the Rocky Mountains (Lake Louise, Jasper, north- from a source area to the north and NE (McMechan 2000).
ern Cariboo Mountains and eastern Purcells Mountains), whereas Although there is broad agreement about these two tectonic
higher-grade metamorphic rocks (biotite to upper amphibolite) phases, the exact tectonic setting of the Windermere basin and
416 M. D. SMITH ET AL.

its evolution remains controversial. Some argue that the upper part to mudstone-dominated upper slope units and shelf platform car-
of the WSG in the southern Canadian Cordillera accumulated on a bonates (Ross et al. 1989). A significant regional unconformity
continental passive margin with a substantial Proto-Pacific Ocean marks the end of WSG deposition and the beginning of a latest
(Ross 1991; Ross et al. 1995; Dalrymple & Narbonne 1996). Neoproterozoic to Cambrian rift– drift succession with the
Others prefer an intracontinental rift and restricted ocean basin clastic and volcanic sediments of the upper Cariboo, Hamill and
setting for the southern WSG, and a late Neoproterozoic –early Gog groups.
Cambrian inception for the Proto-Pacific Ocean (Colpron et al.
2002). The controversy and uncertainty is in part due to the pres-
ence of two intervals of rift-related igneous rocks within the Glaciogenic deposits and associated strata
southern Canadian Cordillera (Bond & Kominz 1984; Ross 1991;
Colpron et al. 2002), diachronous dates for the timing of rifting Toby Fm.
along the whole of Laurentia’s Pacific margin (see discussion in
Lund et al. 2003), and tectonic deformation that currently precludes The Toby Fm. consists primarily of diamictite interbedded with
certain identification of a western Windermere basin margin. conglomerate, sandstone, dolomitized or recrystallized limestone
and mudstone (Fig. 37.3) (Aalto 1971). Its thickness is highly vari-
able, ranging from 0 to 2500 m. Lithofacies, in general, and sedi-
Stratigraphy mentary characteristics of diamictite, in particular (including clast
concentration, clast size, matrix and clast lithology, lateral conti-
In the southern Canadian Cordillera, the WSG is a thick (c. 9 km), nuity and bed thickness) are highly variable over short distances,
unconformity-bounded succession of predominantly coarse- a fact that has been stressed by all workers in this region. The
grained deep-marine siliciclastic rocks with subordinate carbonate following descriptions rely heavily upon the regional study of
and mafic volcanic rocks (Fig. 37.2). Older units associated with Aalto (1971), with additional observations based on more recent
the rift phase (e.g. Toby and Irene formations) are more limited fieldwork by E. Arnaud and K. Root.
in areal extent and tend to show greater lateral facies variations Diamictite is predominantly massive, matrix- or clast-supported
compared with units associated with the post-rift phase (e.g. with sub-angular to sub-rounded clasts up to 2 m, floating in a
Ross et al. 1989, 1995; Ross 1991; Warren 1997). At the base of muddy sandstone or sandy mudstone matrix. Clast lithology
the WSG, the Toby Fm. unconformably overlies shallow-water includes quartzite, volcanic greenstone, slate, dolomite and chert
deposits of the Mount Nelson Fm. (Neoproterozoic/Mesoprotero- derived from the underlying Mount Nelson Fm. as well as rare
zoic?, Root 1987; Ross & Villeneuve 2003) or deep-marine sedi- extrabasinal granite (Leech 1954; Reesor 1973; Loveridge et al.
mentary rocks of the Mesoproterozoic Purcell Supergroup (Aalto 1981). The diamictite matrix exhibits a similar lithological
1971; Root 1987). The Toby Fm. is associated with localized variability. Tectonic overprinting typically obscures clast fabric
rift-related mafic volcanic rocks of the Irene Fm. and a laterally in many outcrops. Where relatively unaltered, clasts show no pre-
discontinuous carbonate horizon at its top (e.g. Root 1987; ferred orientation. Diamictite units vary in thickness from several
Warren 1997). These basal units of the WSG are largely restricted metres to tens of metres. In outcrops with good lateral exposure,
to the Purcell Mountains region (Fig. 37.1). Potential high-grade basal contacts are relatively planar, and conformable or erosional.
metamorphic equivalents of the Toby (e.g. Simony et al. 1980; Conglomerate occurs in lenticular or planar units (centimetre to
Murphy et al. 1991) and Irene volcanic rocks (e.g. Simony et al. metre scale) interbedded with diamictite and sandstone. It is distin-
1980; Sevigny 1988) may exist in the Shuswap Complex and guished from the diamictite by its coarse sandstone matrix, and
northern Cariboo Mountains, although the stratigraphic control is consistently high clast concentration (65 –80%). It is generally
poor (Fig. 37.1). unorganized and massive with occasional subtle coarse-tail
The nomenclature of WSG succession that overlies the Toby grading. Sandstone varies from coarse- to fine-grained, moderately
Fm. in the southern Canadian Cordillera is rather complex: a to poorly sorted, and occurs in planar beds or lenticular units
direct result of correlation challenges across thrust-bound panels (centimetre to metre scale). It is most commonly massive or lami-
with varying levels of metamorphism, general absence of reliable nated, although some sandstone units exhibit ripple cross-
stratigraphic markers and poor geochronological constraints. The lamination, climbing ripples or normal grading. Dolomitized or
Horsethief Creek Group (Purcell Mountains), Kaza and lower/ recrystallized limestone (centimetre to metre scale thickness)
middle Cariboo groups (Cariboo Mountains), Miette Group with varying amounts of quartz grains is also observed within
(Rocky Mountains) and Misinchinka Group (Rocky Mountains) all and at the top of the Toby Fm. at various sites within the region.
refer to parts of the same c. 5 km deep-marine turbidite system These carbonate strata are planar –tabular and can be laterally
exposed over 35 000 km2 (e.g. Ross et al. 1989) (Figs 37.1 & persistent over hundreds of metres, although their overall distri-
37.2). Lithologically, the thick package is dominated by coarse- bution is patchy. Finally, mudstone is relatively common through-
grained arkosic sandstone and granule conglomerate interbedded out the Toby Fm. It is massive or laminated and occurs in relatively
with mudstone-dominated intervals or minor carbonate. planar tabular units (centimetre to metre scale thickness). Outsized
The Vreeland Fm. of the Misinchinka Group is a local excep- clasts are found within these mudstone units, and some are seen to
tion, being composed of up to 2000 m of diamictite interbedded depress or puncture underlying laminations (Aalto 1971).
with mudstone, sandstone and conglomerate (McMechan 2000). In terms of associated strata, the underlying Mount Nelson and
Mapping of the Vreeland Fm. has shown that the diamictite units Dutch Creek formations consist of thick packages of carbonate and
undergo a lateral facies change into typical deep-marine coarse- siliciclastic strata (kilometre scale; Walker 1926; Reesor 1973;
grained turbidites of the WSG in both a westward (McMechan Root 1987). Slate predominates in the Dutch Creek Fm.,
1987) and southward direction (McMechan & Thompson 1995a). whereas thick packages of white quartzite and dolomite inter-
Throughout much of the basin, the overall monotony of the bedded with intervals of argillite, siltstone and conglomerate
thick-bedded, immature turbidites of the Horsethief Creek, characterize the Mount Nelson Fm. In the Purcell Mountains, the
Miette and Kaza groups is interrupted by the fine-grained, region- Toby Fm. is overlain by the Horsethief Creek Group, which con-
ally widespread OFP (and correlative units) (Ross & Murphy sists of a thick package (,100–2000 þ m) of mudstone, sand-
1988; Ross et al. 1995). Nowhere is the OFP observed to be in stone, conglomerate, calcareous mudstone and dolostone
contact with diamictite of the Vreeland Fm.; rather, the OFP over- (Walker 1926; Root 1987; Kubli 1990; Warren 1997). Further to
lies WSG turbiditic strata interpreted as deep-marine lateral the south, the Toby Fm. is interbedded and overlain by volcanic
equivalents. The OFP is overlain by coarse-grained basin-floor tur- greenstone of the Irene Fm., and coarse- to fine-grained clastic
bidite deposits that form a kilometre-scale shoaling upwards trend and carbonate lithofacies of the Monk Fm. (Aalto 1971).
WINDERMERE SUPERGROUP 417

Vreeland Fm.

Description of the Vreeland Fm. is based entirely on regional


mapping studies in the Pine and Monkman Pass areas (McMechan
& Thompson 1985, 1995a, b; McMechan 1987) and geological
descriptions from the Snake Indian and Wapiti Pass thrust sheets
(Fig. 37.1) (McMechan 1990, 2000). The Vreeland Fm. comprises
a thick succession of diamictite interbedded with mudstone, sand-
stone and conglomerate (Fig. 37.4). Diamictite units are generally
massive (c. 98%) with a sandstone or siltstone to mudstone matrix

Fig. 37.3. Stratigraphic columns of the Toby Fm. from the Mount Brewer
(c. 508230 N, 1168140 W) and Paradise Mine localities (c. 508280 N, 1168180 W)
with lithofacies representative of other outcrops; however, their lateral
continuity, stratigraphic thickness and stratigraphic distribution vary widely
across the Purcell Mountains. Note the different scales used in each log. Facies Fig. 37.4. Stratigraphic columns of the Vreeland Fm. near Mount Vreeland in
codes used: (first letters) D, diamictite; G, conglomerate; S, sandstone; F, the Snake Indian thrust sheet, and Paksumo Pass in the Wapiti thrust sheet,
fine-grained facies; (second letters in diamictite and conglomerate) m, matrix modified from McMechan (2000). St, cross-trough bedded sandstone; Fld,
supported; c, clast-supported; (third letter in any lithofacies) m, massive; l, laminated fine-grained facies with lonestones. See Figure 37.3 caption for other
laminated; d, deformed; r, rippled; h, horizontally bedded; s, stratified. facies codes.
418 M. D. SMITH ET AL.

and form laterally extensive (up to 2 km) tabular sheets that range
from a few metres to .40 m thick. The basal contact of massive dia-
mictite units is sharp and locally erosive. Subtle normal grading is
observed locally in diamictite in addition to rare stratified diamictite
defined by clast layers, or sandstone/siltstone stringers and lenses.
Clasts are lithologically diverse and consist of both intra-basinal
(mudstone, sandstone, carbonate) and extra-basinal (felsic or mafic
plutonic, volcanic, quartzite) varieties. Differing provenance trends
are apparent in the two main areas based on clast lithology.
Interstratified mudstone units are up to 38 m thick with rare sand-
stone interbeds, and common lonestones. Sandstone units are
sharp-based and range from very fine- to very coarse-grained and
very thin- to thick-bedded. Sedimentary structures include scours,
cross-stratification, graded bedding and intraclast rip-ups. Con-
glomerate and sandstone units commonly form lenticular channel-
like deposits. Conglomerate is matrix- and clast-supported with a
sandstone matrix, which contrasts with the typical siltstone to mud-
stone matrix of the diamictite. Clast lithologies are similarly
diverse, rounded to angular, and range up to boulder size.
In the Snake Indian thrust sheet, the top of the Vreeland Fm. is
marked by a pyrite-rich zone up to 30 cm thick and locally the
diamictites are overlain by a thin (few metres), parallel laminated
limestone. The conformably overlying Framstead Fm. is
mudstone-dominated, but locally is composed of sandstone to con-
glomerate or limestone. Locally, large (up to 650 m long) carbon-
ate olistoliths occur on, or just above the basal contact with the
Vreeland Fm., and also near its upper contact. These olisoliths
occur as discrete, randomly oriented blocks that lack internal
deformation. Lithologies are predominantly tan-weathered
shallow-water carbonates and include laminated dolomite with
rare layers of teepee structures and bladed calcite, stromatolitic
dolomite, sandy to pebbly dolomite, dolomite conglomerate, and
quartzose fenestral and vuggy dolomite.

Old Fort Point Fm.

The following sedimentological and stratigraphic description


stems mostly from a recent detailed regional study (Smith 2009),
but also incorporates work of earlier authors (Fig. 37.1,D –K)
(Walcott 1910; Weiner 1966; Charlesworth et al. 1967; Aitken
1969; Murphy 1986; Pell & Simony 1987; Ross & Murphy
1988; McDonough 1989; Deschesne 1990; Kubli 1990; Grasby
& Brown 1993; Ross et al. 1995; Warren 1997; Ross & Ferguson
2003a, b). The OFP comprises three distinctive lithological
members, which despite variable metamorphic grade, form a
consistent stratigraphic relationship throughout the southern
Canadian Cordillera (Fig. 37.5).
The lower member is a purple, green, grey or red-brown fine-
grained unit that ranges from 50 to 125 m thick. The basal
portion is composed of siltstone to mudstone that grades upward
into rhythmic couplets of limestone –siltstone. Beds typically
range from ,1 cm to 10 cm in thickness and exhibit common trac- Fig. 37.5. Stratigraphic columns of the Old Fort Point Fm., simplified from
tional sedimentary structures including planar lamination, lenticu- Smith (2009). At the Geikie Siding section in Jasper National Park, the
regionally widespread lower two members are exposed and overlain sharply by
lar bedding, ripple cross-lamination, normal grading and minor
a thinly developed upper member. The Re– Os isochron age was obtained from
scours. Subordinate lithofacies include very fine- to fine-grained
the middle member at this section (Kendall et al. 2004). At the Upper
sandstone interstratified with uncommon limestone-clast breccia
Boomerang section on the Lake Louise Ski Resort, only a thickly developed
beds. Palaeocurrents measured from 3D current-ripples generally Upper Member is exposed, with complete erosion of the lower two members.
indicate flow toward the SW to NW. Str, stratified (ripple cross-laminated and cross-bedded). See Figure 37.3 for
The middle member ranges from 2 to 15 m in thickness and other facies codes.
consists of a dark grey organic-rich mudstone/pelite. The basal
contact with the underlying lower member is usually gradational
over a few metres. The unit is characterized by alternating siltstone In contrast to the lower two members, the upper member is litho-
and mudstone laminae with subtle normal grading. Isolated, thin- logically diverse and highly variable in thickness (,0.5–165 m).
bedded dark-grey massive or planar-laminated limestone and The basal contact with underlying strata is always sharp and,
very fine ripple-stratified sandstone occur locally. This member locally, it completely erodes the lower two members. Lithologies
is regionally extensive. Its fissile nature typically results in poor include diamictite, breccia, conglomerate, mudstone, siltstone,
or covered exposures, although notable exceptions occur (e.g. sandstone, quartz arenite, calcareous arenite, arenaceous limestone
Ross et al. 1995; Kendall et al. 2004). and limestone. Diamictite and breccia to conglomerate are
WINDERMERE SUPERGROUP 419

generally sharp, commonly erosively based, massive beds with of the Vreeland Fm. diamictite. The contact is commonly marked
mudstone/siltstone or well-sorted, coarse quartz-rich sandstone by a distinctive change in colour (e.g. grey to purple) and lithology
matrix. Bed thickness (,0.5 to .10 m), clast size (centimetre to (appearance of lower member siltstones). However, at some
metre) and clast lithologies are variable. Clasts include fragments locations the lower and middle members have been eroded and
of the OFP members and shallow-marine carbonates, some with the upper member forms a sharp unconformity over older WSG
rare bladed calcite crystals. Mudstone and siltstone exhibit strata. The top of the OFP is conformably overlain by younger
planar laminations, micro-scours, subtle normal grading, large WSG strata. In many cases the contact is identified by a sharp
single chlorite flakes and rare scours. This lithofacies is locally change in grain size and/or framework mineralogy (e.g. OFP
thick (c. 100 m) and monotonous with rare interbeds of diamictite, quartz arenite to WSG arkosic sandstone). The challenge arises
breccia, conglomerate or sandstone. Sandstone in the upper where OFP mudstone units are overlain directly by a younger mud-
member consists of a range of textures: fine- to very coarse- stone unit of the WSG. The contact is taken to coincide with the
grained, poor- to well-sorted and mineralogical maturity (imma- termination of organic-rich limestone beds or the appearance of
ture to supermature). The well-sorted, mature quartz arenites are common sandstone interbeds. Field observations can sometimes
interbedded with dark, organic-rich limestone beds and these be confirmed by a decrease in gamma-ray counts or correlative
units can exhibit a range of compositions between the two end geochemical trends (e.g. decrease in total organic carbon (TOC),
members (e.g. calcareous arenites or arenaceous limestones). Sedi- Mo) (Smith 2009).
mentary structures include massive beds, planar lamination and
cross-stratification. Rare palaeocurrent measurements from the
upper member are consistently towards the SW to NW. Soft- Chemostratigraphy
sediment deformation features such as load structures, convolute
or contorted bedding and ductile folding are common in both The only published geochemical or isotopic data from the three
fine- and coarse-grained facies of the upper member. units is the S-isotopic study of Ross et al. (1995) on pyrite in the
post-rift strata of the WSG, Cariboo Mountains, British Columbia
(Fig. 37.1,E). A total of 170 samples were analysed for d34Spy and
Boundary relations with overlying and underlying these showed a broad range of isotopic values (c. 50‰) (Ross et al.
non-glacial units 1995). As part of that study, a suite of 36 samples were collected
from a measured section that includes the OFP. A strong corre-
Toby Fm. lation between lithology (and inferred sedimentation rate) and
pyrite isotopic composition was reported (Ross et al. 1995). For
The Toby Fm. rests unconformably on various members of the example, the most negative d34Spy values ( –31.9‰ to –25.3‰)
Mount Nelson Fm. or the underlying Dutch Creek Fm. In corresponded to siltstone and mudstone of the OFP, whereas
outcrop, the unconformity is commonly subtle (,108), although the interval in the underlying Kaza Group with the highest
a distinctive angular unconformity with tilted underlying Mount sandstone –mudstone ratio was the most positive (around
Nelson strata is evident in some places. The upper contact of the þ11.9‰ to þ14.5‰) (Ross et al. 1995).
Toby Fm. with the Horsethief Creek Group is conformable and New isotopic data (d18Ocarb, d13Ccarb, d13Corg, d34Spy) from OFP
gradational throughout the Purcell Mountains and typically estab- sections and lithogeochemical (major, trace and rare-earth-
lished based on the loss of diamictite. To the west and south of the element analyses) data from a regional study of WSG mudstones
Purcell Mountains, the Toby Fm. is conformably overlain by have recently been obtained (Smith 2009). Preliminary analysis
pillow lavas of the Irene Fm. as well as conglomerate and diamic- shows lower member limestone units exhibiting negative
tite facies of the lower Monk Fm. (Aalto 1971). Considering there d13Ccarb isotopic values and the middle member characterized by
are localized lenses of conglomerate and diamictite in the Irene distinctive chemical (e.g. TOC, Mo, V/Cr) and isotopic (e.g.
Fm., and that the upper contact of the Toby Fm. is defined in the d13Corg) signatures.
Purcell Mountains with the loss of diamictite facies, the upper
contact of the Toby Fm. may actually occur several tens of
metres above the Irene volcanic rocks (Aalto 1971). Other characteristics

Economic deposits associated with the glaciogenic strata are


Vreeland Fm. limited. Local copper mineralization has been observed below,
or in the basal parts of the OFP (Ross 2000; Smith 2009).
In the western Snake Indian thrust sheet, the basal contact of the The only description of possible Precambrian fossils from gla-
Vreeland Fm. is not exposed (McMechan 2000). In the eastern ciogenic strata are some problematic discoid structures at the
Wapiti Pass thrust sheet, the basal diamictite appears to interfinger base of the OFP near Arnica Lake, west of Banff, Alberta
with the upper 150 m of the fine-grained facies of the underlying (Fig. 37.1). The structures are c. 3 mm in size, lentil-shaped with
Paksumo Fm. (McMechan & Thompson 1995b). The Vreeland either hypo- or epirelief, have concentric wrinkles and a five-point
Fm. is conformably overlain by the Framstead Fm., and in the star-like figure inside the concentric pattern (Hofmann 1971;
western exposures by a carbonate olistolith-bearing unit, whereas Gussow 1973; Smith 2009). The specimens are attributed to
in the eastern exposures it is overlain by a sandstone unit (McMe- Chuaria circularis (Hofmann 1971; Gussow 1973), but they lack
chan 2000). The carbonate olistolith-bearing unit correlates from any carbonaceous material that would strengthen a biogenic
west to east where it overlies the lowermost Framstead Fm. sand- origin and certainly require further detailed work. Ediacaran
stone. This suggests at the western Snake Indian thrust sheet fauna (including Namacalathus and Cloudina assemblages
locality that either local erosion of the sandstone unit occurred similar to those found in the Nama Group of Namibia) have other-
or it is actually a lateral equivalent of diamictite of the Vreeland wise been reported from the uppermost part of the WSG (Hofmann
Fm. (McMechan 1990). & Mountjoy 2001 and references therein).

Old Fort Point Fm. Palaeolatitude and palaeogeography

The OFP conformably and gradationally overlies deep-marine tur- There are no published palaeolatitute data for the WSG in south-
biditic strata of the WSG interpreted to be lateral facies equivalent western Canada. A near-equatorial palaeogeographical position
420 M. D. SMITH ET AL.

is based on palaeomagnetic data obtained from broadly correlative localized) of extrabasinal clasts, striated clasts, clast clusters
WSG strata in the Mackenzie Mountains, northern Canadian Cor- within diamictites and outsized clasts interpreted as far-travelled
dillera (Park 1997). and ice-rafted glacial debris (Aalto 1971; Warren 1997). Others
favoured the resedimentation of locally derived material formed
along a fault-bounded margin undergoing extension with limited
Geochronological constraints or no glacial influence (Walker 1926; Reesor 1973; Root 1987).
The predominance of intrabasinal clasts, the mounting evidence
The predominantly siliciclastic nature of the WSG in the southern for structural controls on facies type and variability (ongoing
Canadian Cordillera has led to a paucity of high-precision geochro- research), in addition to the localized and rare occurrence of
nological constraints. Currently, the local maximum depositional glacial indicators, suggest that tectonic activity imparted a
age of the WSG in the region is constrained by a U– Pb zircon primary control on deposition with localized glacial influence.
date of 736 þ 23/ –17 Ma [mean square weighted deviation Carbonate strata within and at the top of the Toby Fm. have
(MSWD) 2.36, four-point TIMS regression] from orthogneiss received limited attention to date. Preliminary results from
basement rocks of the Malton Gneiss Complex (McDonough & ongoing work suggest that primary carbonate was accumulating
Parrish 1991). This age corresponds well with U –Pb zircon on topographic highs while coarse-grained clastics accumulated
dates obtained from WSG basement rocks elsewhere in the in fault-bounded basins. Stable isotopic analyses of these
Canadian Cordillera, such as 740 + 36 Ma (Parrish & Scammell carbonates have so far yielded limited information, with current
1988) and 728 þ 8/– 7 Ma (Evenchick et al. 1984; see Ross research focused on assessing the effects of metamorphism on
et al. 1995 for a full review). A U –Pb zircon age of 569.6 + isotopic signatures.
5.3 Ma (three-point 207Pb/206Pb weighted average) from the
syn-rift felsic volcanic rocks of the Hamill Group that unconform-
ably overlie the WSG constrains the minimum timing of deposition Vreeland Fm.
(Colpron et al. 2002).
No direct ages have been obtained from the Toby Fm. or Irene The Vreeland Fm. is interpreted to have been deposited in a mid-
volcanic rocks in southern British Columbia. Mafic metavolcanic slope, glacially influenced marine setting (McMechan 2000). The
rocks of the Huckleberry Fm. in northeastern Washington, inter- diamictite units are interpreted to be resedimented sediment-
preted to be Irene Fm. correlatives were imprecisely dated with a gravity flow deposits (e.g. unconfined debris flows) of glaciogenic
preliminary age of 762 + 44 Ma (MSWD 0.06) from a three- debris or the result of rainout processes from fine-grained sus-
point Sm –Nd isochron of one whole-rock and two pyroxene sep- pended sediment plumes and ice-rafted debris (McMechan 2000).
arates (Devlin et al. 1988). Other possible treatments of the Sm – The argument for a glacial origin is based on large, sub-angular
Nd data produced even larger error estimates: 795 + 115 Ma extrabasinal basement clasts in the diamictite and the occurrence
(MSWD 6.71) from all the data excluding one pyroxene separ- of lonestones interpreted as dropstones in the mudstone units of
ate; 674 + 212 Ma (MSWD 1.0) from a seven-point whole-rock the Snake Indian thrust sheet (McMechan 2000). The absence of
isochron; 719 + 200 Ma (MSWD 1.0) from two whole-rock dropstones in the Wapiti Pass thrust sheet Vreeland sediments
and two pyroxene separates (Devlin et al. 1988). More recent suggests that there was little to no ice-rafted material at this palaeo-
work on possible correlative rift-related volcanic rock units in geographical location (McMechan 2000). The considerable strike
Idaho has obtained precise SHRIMP U –Pb zircon ages of length (c. 400 km) of diamictite suggests glacial erosion over a
685 + 7 Ma, 684 + 4 Ma (Lund et al. 2003) and 709 + 5 Ma large area, whereas its thickness (c. 2000 m) suggests active subsi-
(Fanning & Link 2004). A comparable U – Pb zircon age of dence accompanied by local faulting (McMechan 2000).
688.9 þ 9.5/ –6.2 Ma was obtained from syn-rift felsic volcani- In the Snake Indian thrust sheet, diamictite units are overlain
clastics of the Gataga Volcanics, northern British Columbia locally by a discontinuous thin-laminated grey limestone (cap car-
(Ferri et al. 1999). In contrast, high-precision U –Pb (ID-TIMS) bonate?) deposited during post-glacial sea-level rise and concomi-
dates were obtained from possible correlative strata in the tant shutdown of coarse clastic sediment flux to the basin
Central Ogilvie Mountains, NW Canada (716.47 + 0.24 Ma; (McMechan 1990, 2000). The large shallow-marine carbonate
Macdonald et al. 2010). olistoliths in the Framstead Fm. suggest continued uplift or fault-
No absolute age constraints are available for the Vreeland Fm. ing causing instability and downslope movement of parts of the
Its relative age is based on regional mapping and lithostratigraphic carbonate shelf edge (McMechan 2000), although simple high-
correlations within the WSG (McMechan 1987, 1990, 2000). stand shedding during post-glacial eustastic rise could also be a
Kendall et al. (2004) obtained two comparable five-point Re –Os potential delivery mechanism.
isochron ages from organic-rich black mudstone of the OFP
middle member in the Jasper area, Alberta (Fig. 37.1): an impre-
cise 634 + 57 Ma (MSWD 65) using the conventional inverse Old Fort Point Fm.
aqua regia digestion method, and a more precise 607.8 + 4.7 Ma
(MSWD 1.2) using the CrO3-H2SO4 digestion technique that is The OFP was deposited along a deep-marine basin-slope to basin-
thought to best reflect depositional age of the OFP. floor transect and overlies coarse-grained turbidites of the WSG.
The OFP is interpreted to be related to an Ediacaran-aged post-
glacial eustatic rise and a shutdown of the supply of coarse, imma-
Discussion ture siliciclastic sediment into the basin (Ross & Murphy 1988).
Fine-grained, mostly turbiditic strata of the lower and middle
Toby Fm. members are regionally uniform in thickness and lithofacies,
suggesting synchronous, basin-wide deep-marine deposition
The variable level of erosion below the sub-Toby unconformity during transgression and highstand (Ross & Murphy 1988; Ross
suggests a period of significant erosion and uplift preceded WSG et al. 1995; Smith 2009).
sedimentation. The depositional setting of the Toby Fm. itself Conversely, the variable lithology and thickness of the upper
has been controversial, specifically regarding the extent of member appears to be controlled by a combination of more local
glacial influence on deposition. Most lithologies record remobili- factors, including palaeogeographical location, topography along
zation of sediment by sediment-gravity flows. Some authors the slope, erosional mass wasting, structural activity and/or rela-
have favoured a glacial setting for these reworked deposits based tive sea-level changes (Smith 2009). Deposition of the upper
on the regional extent of the Toby Fm., the presence (albeit member is interpreted to have coincided with a fall of a relative
WINDERMERE SUPERGROUP 421

sea level caused, at least in part, by regional uplift (Smith 2009). In geochronological constraints (Lund et al. 2003; Fanning & Link
western North America, other Ediacaran-aged uplift or erosion 2004). The timing of the younger glaciogenic event is
features have been observed, suggesting some type of renewed constrained by the c. 608 Ma OFP (Kendall et al. 2004), which
extensional activity during the post-rift phase (e.g. Warren 1997; provides a relative minimum age for Vreeland diamictite depo-
Fedo & Cooper 2001; Clapham & Corsetti 2005). Where observa- sition based on current regional stratigraphic correlations.
ble, the locally sharp, basal contact is interpreted as a sequence
boundary (possibly overprinting the basal surface of forced M. D. S.’s PhD research on the OFP is a product of the Windermere Consortium,
regression) within the OFP (Smith 2009). Related erosional mass initiated by R. W. C. A. and G. M. R. and jointly funded by the Natural Sciences
wasting, submarine canyon incision and/or syn-sedimentary and Engineering Research Council of Canada and industry partners (Anadarko
growth faults controlled initial sediment transport fairways Petroleum, Canadian Natural Resources Ltd, Devon Canada Ltd, Encana Corp.,
(lowstand) that were subsequently filled during a renewed relative Husky Energy Corp., Nexen Inc. and Shell). M. D. S. has received additional
sea-level rise (transgression and highstand) (Smith 2009). The funding from an Ontario Graduate Scholarship (OGS), a Canadian Society of
quartz-rich lithologies of the upper member compared to the Petroleum Geologists Scholarship and two Grants-In-Aid from the American
Association of Petroleum Geologists. E. A.’s work on the Toby Fm. is funded
arkosic sandstone of other WSG turbidites reflect winnowing and
by the Natural Sciences and Engineering Research Council of Canada. Parks
maturation from residence on the shelf during the earlier transgres-
Canada and B. C. Parks are thanked for granting research and collections
sion and highstand (Ross 2000; Smith 2009).
permits to M. D. S. Critical reviews and detailed comments by Brian Kendall
and Paul Link improved the content and clarity of the manuscript. This represents
a contribution of the IUGS- and UNESCO-funded IGCP (International
Regional correlations Geoscience Programme) project #512.

The Toby, Vreeland and OFP formations have been correlated


with other units along the North American Cordillera based on References
lithostratigraphic similarities (see Lund et al. 2003 or Colpron
et al. 2002 for most recent reviews). The Toby Fm. has been cor- Aalto, K. R. 1971. Glacial marine sedimentation and stratigraphy of the
related with the Kingston Peak Fm. (Death Valley), the Edwards- Toby conglomerate (upper Proterozoic), southeastern British Colum-
burg and Pocatello formations (Idaho), the Shedroof Conglomerate bia, northwestern Idaho and northeastern Washington. Canadian
(NW Washington) and the Sayunei and Shezal formations (Mack- Journal of Earth Sciences, 8, 753– 787.
enzie Mountains, NW Canada) (Gabrielse & Campbell 1991; Link Aitken, J. D. 1969. Documentation of the sub-Cambrian unconformity,
et al. 1993; Ross et al. 1995). The Vreeland Fm. has been corre- Rocky Mountains main ranges, Alberta. Canadian Journal of Earth
lated to the similar diamictite-bearing Toobally (Pigage & Mac- Sciences, 6, 193–200.
Naughton 2004) and Icebrook (Aitken 1991; James et al. 2001) Aitken, J. D. 1991. Two late Proterozoic glaciations, Mackenzie Moun-
formations in the northern Canadian Cordillera. The Mount tains, northwestern Canada. Geology, 19, 445– 448.
Lloyd George diamictite in north-central British Columbia (Eisba- Atkinson, S. J. 1975. Surface geology of the Paradise Basin (82K/8W).
cher 1981a, b) is of uncertain age and may correlate with either the In: Geology in British Columbia. British Columbia Ministry of Mines
Toby or the Vreeland formations. and Petroleum Resources 1975: G7– G12.
The OFP overlies turbidites of the WSG that are interpreted to Bennett, S. 1986. Geology of the Mount Forster map area. British
be correlative with the Vreeland Fm. (McMechan 1990, 2000; Columbia Ministry of Energy, Mines and Petroleum Resources,
Preliminary Map 62.
Ross et al. 1995). The OFP lower member rhythmic limestone –
Bond, G. C. & Kominz, M. A. 1984. Construction of tectonic subsidence
siltstone couplets are interpreted as a deep-marine equivalent of
curves for the early Paleozoic miogeocline, southern Canadian Rocky
the thin limestone overlying the Vreeland Fm. (McMechan Mountains-implications for subsidence mechanisms, age of breakup,
2000) and of the cap-carbonate succession overlying the Ice and crustal thinning. Geological Society of America Bulletin, 95,
Brook Fm. (Aitken 1991; James et al. 2001). The organic-rich 155– 173.
OFP middle member potentially correlates with the pyrite-rich Charlesworth, H. A. K., Weiner, J. L. et al. 1967. Precambrian Geo-
horizon at the top of the Vreeland Fm. (McMechan 2000) and rep- logy of the Jasper Region. Research Council of Alberta Bulletin 23.
resents maximum flooding conditions and anoxic (?) bottom-water Clapham, M. E. & Corsetti, F. A. 2005. Deep valley incision in the term-
conditions during earliest post-glacial highstand. The correlative inal Neoproterozoic (Ediacaran) Johnnie Fm., eastern California,
link between the glaciogenic Ice Brook/Vreeland formations and USA: tectonically or glacially driven? Precambrian Research, 141,
the post-glacial OFP is based primarily on stratigraphic relation- 154– 164.
ships and on the consistent stratigraphic presence of associated Colpron, M., Logan, J. M. & Mortensen, J. K. 2002. U– Pb zircon age
carbonates and carbonate olistoliths possessing distinctive bladed constraint for late Neoproterozoic rifting and initiation of the lower
calcite crystals (Ross et al. 1995). Paleozoic passive margin of western Laurentia. Canadian Journal
The basal glaciogenic event in the southern Canadian Cordillera of Earth Sciences, 39, 133– 143.
is represented by deposition of the Toby Fm., which is thought to Dalrymple, R. W. & Narbonne, G. M. 1996. Continental slope sedi-
be younger than c. 740– 728 Ma (see Ross et al. 1995 for full mentation in the Sheepbed Fm. (Neoproterozoic, Windermere Super-
review). Various data treatments of the Sm –Nd ages (c. 795– group), Mackenzie Mountains, N.W.T. Canadian Journal of Earth
674 Ma) from correlative volcanic rocks (Devlin et al. 1988) are Sciences, 33, 848–862.
considered too imprecise to be useful. Current precise ages from Dechesne, R. G. 1990. Geology of the Ptarmigan Creek map area (east
half) and adjacent regions, Main Ranges, Rocky Mountains, British
potentially equivalent strata in Idaho and NW Canada provide
Columbia. Geological Survey of Canada Paper 90-1D, 81 – 89.
better absolute timing constraints on the glacial episodes because
Devlin, W. J., Brueckner, H. K. & Bond, G. C. 1988. New isotopic data
they were obtained from volcanic rocks intercalated with diamic- and a preliminary age for volcanics near the base of the Windermere
tite units, as opposed to unconformably underlying basement Supergroup, northeastern Washington, U.S.A. Canadian Journal of
rocks. However, further geochronological studies are needed Earth Sciences, 25, 1906–1911.
to determine the timing of glacial conditions in the southern Eisbacher, G. H. 1981a. Sedimentary tectonics and glacial record in
Canadian Cordillera relative to the timing of glacial conditions the Windermere Supergroup, Mackenzie Mountains, northwestern
in Idaho and NW Canada. The disparity in radiometric age con- Canada. Geological Survey of Canada Paper 80-27.
straints for rift-related magmatic rocks along the Cordilleran Eisbacher, G. H. 1981b. The Late Precambrian Mount Lloyd George dia-
margin suggests a diachronous, protracted history of crustal exten- mictites, northern British Columbia. In: Hambrey, M. J. & Harland,
sion and magmatism and thus underscores the need for caution W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge
when correlating Cordilleran glacial deposits without precise University Press, Cambridge, 728–730.
422 M. D. SMITH ET AL.

Evenchick, C. A., Parrish, R. R. & Gabrielse, H. 1984. Precambrian Supergroup, central Idaho; implications for rifting of western Lauren-
gneiss and Late Proterozoic sedimentation in north-central British tia and synchroneity of Sturtian glacial deposits. Geological Society
Columbia. Geology, 12, 233– 237. of America Bulletin, 115, 349– 372.
Fanning, C. M. & Link, P. K. 2004. U –Pb SHRIMP ages of Neoproter- Macdonald, F. A., Schmitz, M. D. et al. 2010. Calibrating the
ozoic (Sturtian) glaciogenic Pocatello Fm., southeastern Idaho. Cryogenian. Science, 327, 1241– 1243.
Geology, 32, 881–884. McDonough, M. R. 1989. The structural geology and strain history of
Fedo, C. M. & Cooper, J. D. 2001. Sedimentology and sequence stratigra- the northern Selwyn Range, Rocky Mountains, near Valemount,
phy of Neoproterozoic and Cambrian units across a craton-margin British Columbia. PhD thesis, University of Calgary.
hinge zone, southeastern California, and implications for the early McDonough, M. R. & Parrish, R. R. 1991. Proterozoic gneisses of the
evolution of the Cordilleran margin. Sedimentary Geology, 141– Malton Complex, near Valemount, British Columbia; U– Pb ages
142, 501– 522. and Nd isotopic signatures. Canadian Journal of Earth Sciences,
Ferri, F., Rees, C. J., Nelson, J. L. & Legun, A. S. 1999. Geology and 28, 1202– 1216.
mineral deposits of the northern Kechika Trough between Gataga McMechan, M. E. 1987. Stratigraphy and structure of the Mount Selwyn
River and the 60th parallel. British Columbia Ministry of Energy area, Rocky Mountains, northeastern British Columbia. Geological
and Mines Bulletin 107. Survey of Canada Paper 85-28.
Gabrielse, H. & Campbell, R. B. 1991. Upper Proterozoic assemblages. McMechan, M. E. 1990. Upper Proterozoic to Middle Cambrian history
In: Gabrielse, H. & Yorath, C. J. (eds) Geology of the Cordilleran of the Peace River Arch; evidence from the Rocky Mountains. Bulle-
Orogen. Geological Survey of Canada, Geology of Canada, 4, tin of Canadian Petroleum Geology, 38A, 36 – 44.
125– 150. McMechan, M. E. 2000. Vreeland diamictites; Neoproterozoic glacio-
Glover, J. K. & Price, R. A. 1976. Stratigraphy and structure of the genic slope deposits, Rocky Mountains, Northeast British Columbia.
Windermere Supergroup, southern Kootenay Arc, British Columbia. Bulletin of Canadian Petroleum Geology, 48, 246– 261.
Geological Survey of Canada Paper 76-1B, 21 –23. McMechan, M. E. & Thompson, R. I. 1985. Geology of southeast
Grasby, S. E. & Brown, R. L. 1993. New correlations of the Monkman Pass area (93I/SW), British Columbia. Geological
Hadrynian Windermere Supergroup in the northern Selkirk Moun- Survey of Canada Open File 1150.
tains, British Columbia. Geological Survey of Canada Paper 93-1A, McMechan, M. E. & Thompson, R. I. 1995a. Geology, Ovington Creek,
199– 206. west of sixth meridian, British Columbia. Geological Survey of
Gussow, W. C. 1957. Cambrian and Precambrian geology of southern Canada Map 1873A.
Alberta. Alberta Society of Petroleum Geologists Guidebook, McMechan, M. E. & Thompson, R. I. 1995b. Geology, Wapiti Pass, west
Seventh Annual Field Conference, 7 – 19. of sixth meridian, British Columbia. Geological Survey of Canada
Gussow, W. C. 1973. Chuaria sp. cf. C. circularis Walcott from the Map 1872A.
Precambrian Hector Fm., Banff National Park, Alberta, Canada. Murphy, D. C. 1986. Stratigraphy and structure of the east-central
Journal of Paleontology, 47, 1108– 1112. Cariboo Mountains, British Columbia, and implications for the geo-
Hein, F. J. & McMechan, M. E. 1994. Proterozoic – Lower Cambrian logical evolution of the southeastern Canadian Cordillera. PhD
strata of the Western Canada Sedimentary Basin. In: Mossop, G. D. thesis, Carleton University.
& Shetsen, I. (comps) Geological Atlas of the Western Canada Murphy, D. C., Walker, R. T. & Parrish, R. R. 1991. Age and geo-
Sedimentary Basin. Canadian Society of Petroleum Geologists, logical setting of Gold Creek Gneiss, crystalline basement of the
Calgary, 57 –68. Windermere Supergroup, Cariboo Mountains, British Columbia.
Hofmann, H. J. 1971. Precambrian fossils, pseudofossils and Problema- Canadian Journal of Earth Sciences, 28, 1217– 1231.
tica in Canada. Geological Survey of Canada Bulletin 189. Park, J. K. 1997. Paleomagnetic evidence for low-latitude glaciation
Hofmann, H. J. & Mountjoy, E. W. 2001. Namacalathus-Cloudina during deposition of the Neoproterozoic Rapitan Group, Mackenzie
assemblage in Neoproterozoic Miette Group (Byng Fm.), British Mountains, N.W.T., Canada. Canadian Journal of Earth Sciences,
Columbia; Canada’s oldest shelly fossils. Geology, 13, 819– 821. 34, 34 – 49.
James, N. P., Narbonne, G. M. & Kyser, T. K. 2001. Late Neoprotero- Parrish, R. R. & Scammell, R. J. 1988. The age of the Mount Copeland
zoic cap carbonates; Mackenzie Mountains, northwestern Canada; syenite gneiss and its metamorphic zircons, Monashee Complex,
precipitation and global glacial meltdown. Canadian Journal of southeastern British Columbia. Geological Survey of Canada Paper
Earth Sciences, 38, 1229– 1262. 88-2, 21 –28.
Kendall, B. S., Creaser, R. A., Ross, G. M. & Selby, D. 2004. Con- Pell, J. & Simony, P. S. 1987. New correlations of Hadrynian strata,
straints on the timing of Marinoan ‘Snowball Earth’ glaciation by south-central British Columbia. Canadian Journal of Earth Sciences,
187
Re– 187Os dating of a Neoproterozoic, post-glacial black shale in 24, 302– 313.
Western Canada. Earth and Planetary Science Letters, 222, Pigage, L. C. & MacNaughton, R. B. 2004. Reconnaissance geology of
729– 740. northern Toobally Lake (95D/8), southeast Yukon. In: Emond, D. S.
Kubli, T. E. 1990. Geology of the Dogtooth Range, northern Purcell & Lewis, L. L. (eds) Yukon Exploration and Geology 2003. Yukon
Mountains, British Columbia. PhD thesis, University of Calgary. Geological Survey, 199–219.
Leech, G. B. 1954. Canal Flats, British Columbia. Geological Survey of Pope, A. 1990. Geology and mineral deposits of the Toby-Horsethief
Canada Paper 54-7. Creek map area, northern Purcell Mountains, southeast British
Link, P. K., Christie-Blick, N. et al. 1993. Middle and Late Proterozoic Columbia (82K). Ministry of Energy, Mines and Petroleum
stratified rocks of the western United States Cordillera, Colorado Resources, Mineral Resources Division Open File Report 1990-26,
Plateau, and Basin and Range Province. In: Reed, J., Sims, P., 1 –53.
Houston, R. S., Rankin, D. W., Link, P. K., Van Schmus, W. R. Reesor, J. E. 1973. Geology of the Lardeau Map-area, east-half, British
& Bickford, M. E. (eds) Precambrian: Conterminous United Columbia. Geological Survey of Canada Memoir 369.
States. Geological Society of America Decade of North American Root, K. G. 1987. Geology of the Delphine Creek area, southeastern
Geology Series, C-3, 474–690. British Columbia: implications for the Proterozoic and Paleozoic
Lis, M. G. & Price, R. A. 1976. Large-scale block faulting during depo- development of the Cordilleran divergent margin. PhD thesis, Uni-
sition of the Windermere Supergroup (Hadrynian) in southeastern versity of Calgary.
British Columbia. Geological Survey of Canada Paper 76-1A, Ross, G. M. 1991. Tectonic setting of the Windermere Supergroup
135– 136. revisited. Geology, 19, 1125– 1128.
Loveridge, W. D., Leech, G. B., Stevens, R. D. & Sullivan, R. W. Ross, G. M. 2000. The Neoproterozoic Windermere Supergroup: an
1981. Zircon and monazite age of a granitic clast in Toby Conglom- on-land continental margin turbidite system. Geological Survey of
erate (Windermere Supergroup), Canal Flats, British Columbia. Canada Open File 3932.
Geological Survey of Canada Paper 81-1C, 131 – 134. Ross, G. M. & Murphy, D. C. 1988. Transgressive stratigraphy, anoxia
Lund, K., Aleinikoff, J. N., Evans, K. V. & Fanning, C. M. 2003. and regional correlations within the late Precambrian Windermere
SHRIMP U –Pb geochronology of Neoproterozoic Windermere grit of the southern Canadian Cordillera. Geology, 16, 139– 143.
WINDERMERE SUPERGROUP 423

Ross, G. M. & Bowring, S. A. 1990. Detrital zircon geochronology of the Simony, P. S., Ghent, E. D., Craw, D., Mitchell, W. & Robbins, D. B.
Windermere Supergroup and the tectonic assembly of the southern 1980. Structural and metamorphic evolution of the northeast flank of
Canadian Cordillera. Journal of Geology, 98, 879– 893. the Shuswap Complex, southern Canoe River area, British Columbia.
Ross, G. M. & Parrish, R. R. 1991. Detrital zircon geochronology In: Coney, P. J. (ed.) Cordilleran Core Complexes. Geological
of metasedimentary rocks in the southern Omineca Belt, Society of America Memoir 153.
Canadian Cordillera. Canadian Journal of Earth Sciences, 28, Slind, O. L. & Perkins, G. D. 1966. Lower Paleozoic and Proterozoic
1254– 1270. sediments of the Rocky Mountains between Jasper, Alberta and
Ross, G. M. & Villeneuve, M. E. 1997. U– Pb geochronology of stran- Pine River, British Columbia. Bulletin of Canadian Petroleum
ger stones in Neoproterozoic diamictites, Canadian Cordillera; Geology, 14, 442– 468.
implications for provenance and ages of deposition. Geological Smith, M. D. 2009. Stratigraphic and geochemical evolution of the Old
Survey of Canada Paper 1997-F, 141– 155. Fort Point Fm., southern Canadian Cordillera: the deep-marine
Ross, G. M. & Villeneuve, M. E. 2003. Provenance of the Mesoproter- perspective of Ediacaran post-glacial environmental change. PhD
ozoic (1.45 Ga) Belt basin (western North America): another piece in thesis, University of Ottawa.
the pre-Rodinia paleogeographic puzzle. Geological Society of Stelck, C. R., Burwash, R. A. & Stelck, D. R. 1978. The Vreeland
America Bulletin, 115, 1191– 1217. High; a Cordilleran expression of the Peace River Arch. Bulletin of
Ross, G. M. & Ferguson, C. R. 2003a. Geology and structure cross- Canadian Petroleum Geology, 26, 87 –104.
sections, Eddy, British Columbia. Geological Survey of Canada, Stewart, J. H. 1972. Initial deposits in the Cordilleran geosyncline;
‘A’ series map 1967A. evidence of a Late Precambrian (,850 m.y.) continental separation.
Ross, G. M. & Ferguson, C. R. 2003b. Geology and structure cross- Geological Society of America Bulletin, 83, 1345– 1360.
sections, Lanezi Lake, British Columbia. Geological Survey of Walcott, C. D. 1910. Pre-Cambrian rocks of the Bow River Valley, Alberta,
Canada, ‘A’ series map 2001A. Canada. Smithsonian Miscellaneous Collections, 53, 423–431.
Ross, G. M., McMechan, M. E. & Hein, F. J. 1989. Proterozoic History: Walker, J. F. 1926. Geology and Mineral Deposits of the Windermere
the Birth of the Miogeocline. In: Ricketts, B. D. (ed.) The Western Map-area, British Columbia. Geological Survey of Canada Memoir
Canadian Sedimentary Basin. Canadian Society of Petroleum Geo- 148.
logists, Calgary, 79 – 104. Warren, M. J. 1997. Crustal extension and subsequent crustal thickening
Ross, G. M., Bloch, J. D. & Krouse, H. R. 1995. Neoproterozoic strata along the Cordilleran rifted margin of ancestral North America, west-
of the southern Canadian Cordillera and the isotopic evolution of ern Purcell Mountains, southeastern British Columbia. PhD thesis,
seawater sulfate. Precambrian Research, 73, 71– 99. Queen’s University. Available at: http://www.collectionscanada.
Sevigny, J. H. 1988. Geochemistry of late Proterozoic amphibolites and ca/obj/s4/f2/dsk3/ftp04/nq22501.pdf.
ultramafic rocks, southeastern Canadian Cordillera. Canadian Weiner, J. L. 1966. The Old Fort Point Fm., Jasper, Alberta. PhD thesis,
Journal of Earth Sciences, 25, 1323– 1337. University of Alberta.
Chapter 38

Neoproterozoic strata of southeastern Idaho and Utah: record of Cryogenian


rifting and glaciation

PAUL KARL LINK1* & NICHOLAS CHRISTIE-BLICK2


1
Department of Geosciences, Idaho State University, Pocatello, ID 83209-8072, USA
2
Department of Earth and Environmental Sciences and Lamont-Doherty Earth Observatory of Columbia University,
Palisades, NY 10964-8000, USA
*Corresponding author (e-mail: linkpaul@isu.edu)

Abstract: Neoproterozoic strata in southeastern Idaho and Utah include the ,766 Ma Uinta Mountain Group and Big Cottonwood
Formation (Fm.) deposited in an east-trending rift basin and, to the west, the lower part of a westward-thickening rift to passive-
margin succession that initiated c. 720 Ma. The latter contains a lower diamictite and volcanic succession, with a complex stratigraphic
interval of Cryogenian marine glacial deposits (Pocatello and Mineral Fork formations and correlatives). This is overlain by a mostly
terrigenous succession of ,667 Ma strata assigned to the upper member of the Pocatello Fm. and Brigham Group in southeastern
Idaho, to the Kelley Canyon Fm. and Brigham Group in northern and western Utah, and to the McCoy Creek Group and Prospect Moun-
tain Quartzite in adjacent Nevada. Although the Brigham Group and correlative deposits contain no direct evidence for glaciation, widely
developed, though stratigraphically restricted, incised valleys, with erosional relief from a few metres to as much as 160 m, are inferred to
represent subsequent times of Cryogenian glacially lowered sea level. Overall interpretations of the stratigraphy and sedimentology of
these rocks have changed little in the past 10–15 years. The most important recent advances relate to U– Pb geochronology. In strata that
lie unconformably below demonstrable glacial deposits, the lower Uinta Mountain Group (formerly thought to be c. 900 Ma) contains
populations of detrital zircons as young as 766 + 5 Ma. Cryogenian magmatism north of the Snake River Plain in central Idaho is recog-
nized near House Mountain, east of Boise at c. 725 + 5 Ma, in the Pioneer Mountains Core Complex at about 695 Ma, and in central and
east-central Idaho at 685–650 Ma. Clasts interpreted to be from the rift-related Bannock Volcanic Member of the Pocatello Fm. are dated
at 717 + 4 Ma and 701 + 4 Ma. The overlying diamictite-bearing Scout Mountain Member contains a mafic lapilli tuff near the base
(686 + 4 Ma) and a reworked fallout tuff near the top (667 + 5 Ma). Strongly negative C-isotope data have been obtained from
some of the carbonate rocks, although the latter constitute only a small fraction of the succession. Palaeomagnetic data are available
only for the Uinta Mountain Group, and suggest an equatorial palaeolatitude.

Neoproterozoic glaciogenic rocks, locally in excess of 1 km thick, (Smith et al. 1994; Young 2002; Lorentz et al. 2004; Corsetti et al.
are exposed widely but discontinuously in southeastern Idaho, and 2007; Dehler et al. 2007).
northern and western Utah (Fig. 38.1). The deposits are assigned to
a plethora of local formal and informal stratigraphic units, for his-
torical reasons and because of their varied expression. Names used Structural framework
for commonly correlated diamictite-bearing units include Poca-
tello Fm. in Idaho, and Mineral Fork Fm., Sheeprock Group, for- The rocks are thought to have accumulated in rift-related basins
mation of Perry Canyon, Trout Creek sequence (units 3 and 5), associated with the development of a passive continental margin
and Horse Canyon Fm. in Utah (Fig. 38.2). Crittenden et al. in western North America between c. 665 Ma and c. 520 Ma,
(1971, 1983) and Link et al. (1993, 1994) provide regional and in part atop erosional topography with as much as 900 m of
reviews of diamictites and their interpretation. Important papers local relief (Stewart 1972; Stewart & Suczek 1977; Bond et al.
for specific locations include Blackwelder (1932), Crittenden 1983, 1985; Christie-Blick & Levy 1989; Levy & Christie-Blick
et al. (1952), Ojakangas & Matsch (1980), Blick (1981), Christie- 1991a; Ross 1991; Christie-Blick 1983a, 1997). They crop out
Blick (1982, 1983a, 1983b, 1985, 1997), Christie-Blick & Link today within and along the eastern flank of the late Jurassic to
(1988), Christie-Blick & Levy (1989) and Yonkee et al. (2000a) early Cenozoic Cordilleran thrust-and-fold belt, and across the
for northern and west-central Utah; Ludlum (1942), Trimble eastern edge of the late Cenozoic Basin and Range extensional
(1976), Link (1981, 1983, 1987; Link et al. 2005) for the Pocatello province (Fig. 38.1; Armstrong & Oriel 1965; Armstrong 1968;
area of Idaho; and Misch & Hazzard (1962), Bick (1966) and Levy & Christie-Blick 1989; Allmendinger 1992; Wernicke
Rodgers (1994) for westernmost Utah and eastern Nevada. Type 1992; DeCelles 2004; DeCelles & Coogan 2006). The structurally
localities are described by Misch & Hazzard (1962; Trout Creek lowest thrust sheets encompassing thick glacial and associated
sequence and McCoy Creek Group), Bick (1966; Horse Canyon deposits at the present level of exposure (generally on the
Fm.), Crittenden et al. (1971; Kelley Canyon Fm., Brigham eastern side of the thrust belt) belong to the Willard –Paris –
Group and Pocatello Fm.), Christie-Blick (1982, 1983a; Mineral Putnam system in northern Utah and southeastern Idaho, and to
Fork Fm. and Sheeprock Group), Crittenden et al. (1983; Pocatello the Tintic – Sheeprock –Canyon Range system of west-central
Fm.), Link (1983, 1987; Pocatello Fm.) and Link et al. (1985; Utah (Levy & Christie-Blick 1989; DeCelles 2004; Fig. 38.1).
Brigham Group). For example, the Pocatello Fm. in the Bannock Range of southeast-
The stratigraphy and sedimentology of these strata were worked ern Idaho and the formation of Perry Canyon near Ogden in the
out primarily in the 1970s and early 1980s. The correlations first northern Wasatch Range of Utah represent a 200 km strike-parallel
synthesized by Crittenden et al. (1971) were based on regional outcrop band within the Putnam-Paris thrust sheet. Although the
field mapping, and have largely stood the test of time as summar- rocks are generally foliated, with pervasive development of chlor-
ized in Link et al. (1993). More recent research has focused on U – ite and locally biotite in the greenschist facies, detrital zircon
Pb geochronology (Fanning & Link 2004, 2008) and geochemistry geochronology and C-isotope studies have been successfully

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 425– 436. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.38
426 P. K. LINK & N. CHRISTIE-BLICK

Fig. 38.1. Map showing areas of outcrop of Neoproterozoic rocks of Idaho and Utah, with key locations shown (after Link et al. 1994). Inset map is present geography.
Main map is palinspastic reconstruction after Levy & Christie-Blick (1989).

conducted (Fanning & Link 2004, 2008; Lorentz et al. 2004; complex, the entire Neoproterozoic succession is overturned
Corsetti et al. 2007). One or more generations of Neogene exten- beneath the Pole Canyon thrust in the southern Sheeprock Moun-
sional faults cut all of these ranges. In many cases, therefore, tains, over a lateral distance of c. 10 km (Christie-Blick 1983b). In
rocks transported eastward during Cretaceous thrusting have the northern Wasatch Range, the formation of Perry Canyon crops
been translated westward during Neogene extension (Levy & out in an east-vergent overturned fold above the Willard thrust
Christie-Blick 1989). fault. At Portneuf Narrows, SE of Pocatello, the type section of
Detailed field studies suggest that primary sedimentary features the Pocatello Fm. is exposed in an overturned fold, cut by a Creta-
and correlations are discernable through the deformation. In the ceous tear fault.
structurally highest hinterland thrust sheet of the Deep Creek Neoproterozoic rocks of the central Wasatch Range and Uinta
Range, close to the Utah – Nevada state line, garnet and staurolite Mountains are parautochthonous with respect to cratonic North
grade rocks are present in the Trout Creek sequence (Nelson America, having been displaced eastward no more than a few kilo-
1966; Rodgers 1994). East of there, but in the same thrust metres by mostly blind structures (Bruhn et al. 1986). In the Uinta
Mountain Group some strata are basically unmetamorphosed and
retain organic carbon (Dehler & Sprinkel 2005). The Big Cottonwood
(Ma)

Allochthon Parautochthon Fm. in the Wasatch Range is at greenschist facies. The least deformed
C Age

Deep Creek Central


Range SE Idaho N. Wasatch Sheeprock Wasatch Uinta Mtns and metamorphosed glacial deposits are found in the Mineral Fork
Prospect Camelback Geertsen Prospect Tintic Fm Tintic
Fm. of the central Wasatch parautochthon, except within the
Windermere Supergroup

Mountain Mountain Canyon Mountain


v ? aureole adjacent to Oligocene stocks (Christie-Blick 1983a).
Ediacaran

McCoy Browns Hole Fm 580 v


600 Creek Gp Mutual Fm
Inkom Fm ‘Mutual’ Fm
?
Neoproterozoic

Caddy Canyon Fm
fault? v 667 Stratigraphy
d Poca- d d d Mineral
d Trout d tello v Fm of v Sheep-
700 Creek v 686 Fm Perry rock Fork Fm
d sq d Canyon d Group
<717 Neoproterozoic and lower Cambrian, predominantly siliciclastic
fault? not exposed thrust ? ? rocks in southeastern Idaho and adjacent Utah are divisible into
Cryogenian

Big Uinta Mtn.


Cottonwood Gp three intervals. Pre-glacial deposits are best represented by
Fm < 766
800 locally conglomeratic sandstone and siltstone of the Uinta Moun-
Palaeoproterozoic to tain Group in the Uinta Mountains (Figs 38.2 & 38.3) and by
Archaean basement
Brigham Group 580 radiometric age (Ma) comparable quartzite and argillite of the Big Cottonwood Fm. in
and correlatives v volcanics the central Wasatch Range. Glacial and associated deposits are
Diamictite-bearing interval d glacial diamictite
represented by the Pocatello Fm. and correlatives (Figs 38.2 &
Uinta Mountain Group unconformity
and correlatives
38.4; Crittenden et al. 1971, 1983). Terminal Neoproterozoic to
lower Cambrian quartzite, minor siltstone and minor carbonate
Fig. 38.2. Utah and Idaho Neoproterozoic correlation chart showing are ‘post-glacial’ in terms of preserved facies. However, they are
stratigraphic names, ages, and locations of diamictites, volcanic rocks and relevant to the theme of this volume because they contain an indir-
carbonate strata. ect record of sea-level change (incised valleys) that may reflect the
NEOPROTEROZOIC IDAHO AND UTAH 427

Fig. 38.3. Stratigraphy of the Uinta


Mountain Group in the central Uinta
Mountains (Kings Peak quadrangle) and in
the eastern range near Brown’s Park (after
Dehler et al. 2007, and E. M. Kingsbury,
Idaho State University, pers. comm. 2008).
SB, sequence boundary.

waxing and waning of ice sheets elsewhere on the planet. The rocks folded and less easily studied than the Uinta Mountain Group in
are assigned in most places to the Brigham Group (Figs 38.2 & 38.5; the steep-sided Wasatch Range canyons in which it is exposed.
Crittenden et al. 1971; Christie-Blick 1982; Link et al. 1985). Cor- The most distinctive attribute of these apparently non-descript
relative strata in the Deep Creek Range are the McCoy Creek Group rocks is the presence of tidal rhythmites (Chan et al. 1994;
and overlying Prospect Mountain Quartzite (Fig. 38.2; Misch & Ehlers et al. 1997; Ehlers & Chan 1999). Direct correlation
Hazzard 1962). All of these Neoproterozoic and lower Cambrian between the Uinta Mountain Group and Big Cottonwood Fm. is
strata are broadly equivalent to the Windermere Supergroup of not possible because younger rocks intervene between available
Washington and western Canada and to the Kingston Peak Fm. exposures. However, the successions are thought to be of broadly
and overlying rocks in Death Valley (Link et al. 1993; Lund the same age on the basis of gross lithostratigraphy and provenance
et al. 2003, 2011; Hoffman & Halverson 2011; Mrofka & (Condie et al. 2001; Dehler et al. 2007), a stratigraphic position
Kennedy 2011; Petterson et al. 2011; Smith et al. 2011). below glacial deposits in the case of the latter, and the east –west
alignment of the sedimentary basins in which the successions are
thought to have accumulated (Christie-Blick 1997). Further, detri-
Uinta Mountain Group and Big Cottonwood Fm. tal zircons in both units show the same populations, with the
youngest population in the Uinta Mountain Group at ,766 Ma
The Cryogenian but pre-glacial Uinta Mountain Group consists of (Dehler et al. 2007).
4–7 km of pervasively cross-bedded arkosic and quartzose sand-
stone, with subordinate medial and upper intervals of mudrock
(Fig. 38.3; Hansen 1965; Sanderson & Wiley 1986; Stone 1993; Pocatello Fm. and correlative units
Link et al. 1993; Dehler & Sprinkel 2005; Dehler et al. 2010).
The Big Cottonwood Fm. is of similar thickness (5 km) and lithol- The Pocatello Fm. of southeastern Idaho consists of a lower inter-
ogy (interstratified quartzite and argillite), but is more strongly val of mafic volcanic flows, fragmental volcanic rocks and minor
428 P. K. LINK & N. CHRISTIE-BLICK

Fig. 38.4. Stratigraphy of diamictite and volcanic succession, southeastern Idaho and northern and western Utah (after Link et al. 1994). Available radiometric ages are
from Fanning & Link (2004, 2008).

intrusives (Bannock Volcanic Member); a heterolithic medial unit The formation of Perry Canyon of northern Utah consists of
(Scout Mountain Member) that includes two intervals of diamic- more than 2 km of diamictite, greywacke, argillite, sandstone
tite, cobble conglomerate, ferruginous sandstone and at least two and mafic volcanic rocks (Crittenden et al. 1983). These strata
zircon-bearing tuff beds; and an unnamed upper member consist- crop out widely though discontinuously in the Willard thrust
ing primarily of laminated shale or phyllite (Fig. 38.4; Ludlum sheet (Ogden area, Huntsville, Little Mountain and Fremont
1942; Crittenden et al. 1971, 1983; Trimble 1976; Link 1981, Island in Fig. 38.1). Neither the base nor the top is exposed at
1983, 1987; Link et al. 2005; Rodgers et al. 2006). A thin lami- most locations. North and east of Ogden, in the northern
nated carbonate, locally resedimented as multiple layers of breccia, Wasatch Range, however, the Perry Canyon locally overlies the
is present at the base of the upper member. The lower diamictite Palaeoproterozoic Facer Fm., and underlies a poorly developed
contains abundant intrabasinal volcanic clasts, whereas the upper laminated carbonate. The latter passes upward into several
contains clasts of quartzite (locally striated), granitic rocks and hundred metres of siltstone (Kelley Canyon Fm.; Crittenden
felsic volcanic rocks interpreted to have been derived from the sub- et al. 1971). Diamictite clasts consist primarily of granitic rocks
jacent Bannock Volcanic Member. and quartzite (Crittenden et al. 1983).
Lying unconformably on the Big Cottonwood Fm. in the vicin- More than 2 km of diamictite, greywacke, conglomerate, quart-
ity of Mineral Fork in the central Wasatch Range, the Mineral Fork zite and argillite is present in the Otts Canyon Fm. and overlying
Fm. consists of as much as 800 m of diamictite, siltstone, sand- Dutch Peak Fm. of the Sheeprock Group in the Sheeprock Moun-
stone and conglomerate (Fig. 38.4; Crittenden et al. 1952; Ojakan- tains in west-central Utah (Christie-Blick 1982, 1997; Crittenden
gas & Matsch 1980; Christie-Blick 1983a, 1997; Link et al. 1994). et al. 1983). The lower part of the Otts Canyon Fm., stratigraphi-
The diamictite, which consists primarily of rounded clasts of quart- cally below all exposed diamictite, consists of phyllitic argillite.
zite, carbonate, other sedimentary rocks and minor igneous rocks The uppermost part of the Otts Canyon Fm. is intruded by mafic
in a sandy matrix, contains striated stones, lonestones in laminated sills. The Dutch Peak Fm. is overlain directly by shale of the
host rock, and sheets and lenses of texturally mature sandstone. Kelley Canyon Fm. without an intervening carbonate. Diamictite
These strata are unconformably overlain in turn by quartzite of in the Otts Canyon Fm. consists predominantly of quartzite
Mutual Fm. and Tintic Quartzite (Ediacaran to lower Cambrian). clasts in a phyllitic matrix (Christie-Blick 1982). Substantially
Comparable deposits, also assigned to the Mineral Fork Fm., are thicker diamictite and greywacke of the Dutch Peak Fm. consist
present in the Charleston-Nebo thrust sheet of the southern primarily of clasts of granitic rocks, carbonate and quartzite in a
Wasatch Range (small outcrops not shown in Fig. 38.1), and in phyllitic to sandy matrix, with notable lateral variations in the rela-
the parautochthon of Antelope Island. In the case of the latter, the tive abundance of clast types.
Mineral Fork Fm. rests not on the Big Cottonwood Fm., but on Diamictite is present in units 3 and 5 of the lower part of the
gneisses of Archaean to Palaeoproterozoic age, and it is overlain Trout Creek sequence of Misch & Hazzard (1962; Horse Canyon
more or less conformably by a well-developed laminated carbonate Fm. of Bick 1966) in the Deep Creek Range on the Utah –
and argillite correlative with the upper member of the Pocatello Fm. Nevada state line (Rodgers 1994). Details of the stratigraphic
at Pocatello (Kelley Canyon Fm.; Christie-Blick 1983a; Crittenden succession are less well established than at other locations
et al. 1983; Bryant 1988; Yonkee et al. 2000a). Diamictite at this owing to structural complexity and higher metamorphic grade
location is coarse-grained, with clasts composed largely of gneiss. (Christie-Blick 1982).
NEOPROTEROZOIC IDAHO AND UTAH 429

A striking feature of the Brigham Group, in addition to its


overall lithostratigraphy, is the presence at many localities of an
unconformity (sequence boundary) with prominent conglomerate-
filled incised valleys at or near the top of the Caddy Canyon Quart-
zite from Pocatello south and west to the Sheeprock Mountains
(Christie-Blick et al. 1988; Christie-Blick & Levy 1989; Levy &
Christie-Blick 1991b; Levy et al. 1994; Christie-Blick 1997).
The valleys are typically as much as several tens of metres deep
(up to 160 m deep in the southern Sheeprock Mountains). A
second unconformity, at or near the base of the Mutual Fm., is
recognized on the basis of a regionally persistent abrupt upward
coarsening of facies, and the presence of incised valleys in the
southern Sheeprock Mountains and central Wasatch Range
(Christie-Blick 1997).

Glaciogenic deposits and associated strata

Glaciogenic deposits in southeastern Idaho and Utah consist of a


heterogeneous assemblage of sedimentary facies associations
(see Link et al. 1994 for the most recent review) including
massive diamictite, stratified diamictite and graded sandstone, dia-
mictite and laminated mature sandstone, and carbonate, shale and
sandstone. The massive diamictite association is characterized by
diamictite as much as hundreds of metres thick with little or no
stratification. Where present, bedding is indistinct, and defined
by subtle variations in clast abundance, inverse and normal
grading, and lenses and interbeds of stratified sandstone or argil-
lite. This association is best developed in the upper Scout Moun-
tain Member, lower part of the Mineral Fork Fm. and in the
formation of Perry Canyon (Fig. 38.4). The stratified diamictite
and graded sandstone association includes bedded diamictite,
Fig. 38.5. Pocatello area (Pocatello Fm., Blackrock Canyon Limestone and disorganized clast-supported conglomerate, massive to graded
Brigham Group) stratigraphy showing available C-isotope data (after Lorentz sandstone with parallel lamination and cross-lamination, and
et al. 2004). In the C-isotope curve, interval ‘a’ is the cap carbonate that overlies rhythmically bedded and laminated argillite. The association is
the diamictite of the Scout Mountain Member, interval ‘b’ is the Blackrock best expressed in the lower part of the Scout Mountain Member
Canyon Limestone, and interval ‘c’ is from thin dolomites within the marine and formation of Perry Canyon and in western locations (Shee-
Caddy Canyon Quartzite. Qz, quartzite. prock Mountains and Deep Creek Range). The diamictite and
laminated mature sandstone association is characterized by strati-
fied diamictite with contorted lenses of conglomerate and finer-
grained deposits, and parallel-laminated medium- to coarse-
Brigham Group grained sandstone. Much of the Mineral Fork Fm. and parts of
the Dutch Peak Fm. and Trout Creek sequence is composed of
In most areas of southeastern Idaho and northern Utah, many hun- this association. The carbonate, shale and sandstone association
dreds of metres of shale and argillite directly overlie glacial strata consists of generally laminated limestone and dolomite as much
(upper member of the Pocatello Fm., Kelley Canyon Fm., and as several metres thick, laminated, cross-laminated and graded silt-
Trout Creek sequence unit 6). These pass upwards into a quartzose stone, and fine- to coarse-grained sandstone. It is present in the
succession up to several kilometres thick of late Neoproterozoic upper Scout Mountain Member, and the upper formation of
(late Cryogenian and Ediacaran) to early Cambrian age. At most Perry Canyon.
locations in southeastern Idaho and northern Utah, these strata Striated clasts and outsized clasts in laminated facies are
are assigned to the Brigham Group (Crittenden et al. 1971; present widely in the massive and stratified diamictite facies
Christie-Blick 1982; Stewart 1982; Link et al. 1985, 1987, 1993; associations, though they are not common. The best examples
Levy & Christie-Blick 1991b). In westernmost Utah and eastern are found in the Mineral Fork Fm. of the central Wasatch
Nevada, the rocks are placed in the upper Trout Creek sequence, Range (Christie-Blick 1983a), particularly at the Mineral Fork
McCoy Creek Group and Prospect Mountain Quartzite (Misch & locality (striated clasts) and on the north side of Little Cotton-
Hazzard 1962; Fig. 38.2). wood Canyon (outsized clasts in laminated facies). Such isolated
The stratigraphy of the Brigham Group varies in detail, particu- stones are present locally in the Dutch Peak Fm., and in the for-
larly in the distribution of carbonate rocks (e.g. Blackrock Canyon mation of Perry Canyon on Fremont Island. Striated clasts have
Limestone in Idaho) and volcanic rocks (e.g. Browns Hole Fm. in been observed also in the Dutch Peak Fm. (Sheeprock Moun-
the northern Wasatch Range), and in the abundance of siltstone at tains) and in the upper diamictite of the Scout Mountain
the level of the Caddy Canyon Quartzite (Christie-Blick 1982). Member in Idaho.
The interval that encompasses the upper part of the Caddy
Canyon Quartzite (as much as 2 km of orthoquartzite), Inkom
Fm. (up to 150 m of olive drab to greyish red or liver-coloured silt- Boundary relations with overlying and underlying
stone) and Mutual Fm. (several hundred metres of greyish red non-glacial units
pebbly quartzite) contains incised valley-fill conglomerates, corre-
lated with confidence from range to range over two states The lower contacts of diamictite-bearing strata range from uncon-
(Fig. 38.2; Crittenden et al. 1971; Christie-Blick 1982). The suc- formable to concordant and perhaps conformable. Diamictite rests
cession at Pocatello is representative (Fig. 38.5). upon an unconformity at most exposures of the Mineral Fork Fm.
430 P. K. LINK & N. CHRISTIE-BLICK

(Fig. 38.4) and locally at the base of the formation of Perry diagenetic alteration. Lorentz et al. (2004) nevertheless took the
Canyon. A striated and grooved surface with roches moutonnées general consistency of C-isotope data and the absence of a diage-
is preserved beneath the Mineral Fork Fm. on the north side of netic trend in the isotopic cross-plot for section 3 to indicate that
Big Cottonwood Canyon (Christie-Blick 1983a, 1997). The age measured d13C values provide a reasonable approximation for
of the Big Cottonwood Fm., which underlies the Mineral Fork at the isotopic composition of seawater at the time of deposition.
that locality and all others except Antelope Island, is uncertain. C-isotope values cluster between þ1.0 and –1.5‰ in the Black-
By correlation with the ,766 Ma Uinta Mountain Group, the rock Canyon Limestone in Idaho (level ‘b’ in Fig. 38.5; Corsetti
Big Cottonwood is perhaps no more than a few tens of millions et al. 2007). Thin layers of dolomite in the middle of the Caddy
of years older than the glacial deposits (Fig. 38.3). We know that Canyon Quartzite in Idaho are characterized by d13C values
the Big Cottonwood was sufficiently lithified to maintain a palaeo- between þ3.9 and þ8.8‰ (three measurements, level ‘c’ in
gradient of ,408 at valley walls without deformation, to preserve Fig. 38.5; Smith et al. 1994; Corsetti et al. 2007).
glacial grooves, and to provide a source for well-rounded clasts C-isotope data have been obtained also from organic-rich shales
in the Mineral Fork Fm. in the Red Pine Shale in the upper part of the Cryogenian Uinta
Elsewhere (Antelope Island, and locally in the northern Wasatch Mountain Group, stratigraphically below the glacial deposits
Range, Yonkee et al. 2000b), diamictite overlies igneous and (Dehler et al. 2007). Whole-rock d13C values for organic matter
metasedimentary rocks of Archaean to Palaeoproterozoic age. At range from –16.9 to –30.8‰, and are comparable to values
most locations, the lower contacts of the Pocatello Fm., the for- obtained from other Cryogenian marine successions, including
mation of Perry Canyon and Sheeprock Group are not exposed. the Chuar Group of the Grand Canyon Supergroup (Dehler et al.
Diamictite-bearing intervals in those sections are concordant 2005a). Total organic carbon for the Red Pine Shale varies from
with underlying non-diamictic deposits (Fig. 38.4). The base of 0.07 to 5.9% (Dehler et al. 2007).
the Trout Creek sequence (Horse Canyon Fm.) in the Deep Mudstones in the upper part of the Mineral Fork Fm. are unu-
Creek Range is faulted. sually Fe-rich (c. 15% Fe2O3), and comparable with Fe-rich gla-
Upper contacts with non-glacial deposits are gradational to thick ciogenic deposits of the Rapitan Group in the northern Canadian
siltstone or shale in the case of the Scout Mountain Member Cordillera (Young 2002). The carbonate-corrected CIA for dia-
(Idaho), the Mineral Fork Fm. on Antelope Island, the formation mictite in the Mineral Fork Fm. ranges from 65 to 70, values
of Perry Canyon (northern Wasatch Mountains), the Dutch Peak that are significantly higher than those reported for the glaciogenic
Fm. (Sheeprock Mountains) and the Trout Creek sequence (Deep Palaeoproterozoic Gowganda Fm. of Ontario, Canada (CIA , 60).
Creek Range; Fig. 38.4). The upper contact is unconformable at The Gowganda data were taken to indicate reduced chemical
most other exposures of the Mineral Fork Fm., and not exposed weathering under frigid conditions. Data from the Mineral Fork
at Little Mountain or on Fremont Island. Fm. are consistent with the incorporation of a high proportion of
A laminated dolostone (in excess of 10 m thick) overlies diamic- weathered silicates (sedimentary rocks). About 94% of clasts
tite of the Mineral Fork Fm. on Antelope Island (Christie-Blick larger than 1 cm are sedimentary (mean of 22 counts excluding
1983a; Yonkee et al. 2000a). A correlative considerably thinner the Antelope Island locality; Christie-Blick 1983a). CIA values
carbonate is exposed in the northern Wasatch Range, where it over- for shales and argillites in the Uinta Mountain Group and Big
lies a variety of glacial and non-glacial facies (Crittenden et al. Cottonwood Fm. are higher than for the Mineral Fork Fm.
1983). A comparably thin laminated carbonate, which overlies the (mostly 75–85), reflecting significant chemical weathering of
upper diamictite of the Scout Mountain Member east of Pocatello their sources (Condie et al. 2001).
(Link 1983, 1987; Corsetti et al. 2007), is present both in situ and Provenance data are available for both pre-glacial and post-
as a sedimentary breccia. The laminated dolomite at Pocatello glacial strata. Neodymium (Nd) and Sr-isotope data from shales
passes up into an upward-fining succession containing several beds of the Uinta Mountain Group are consistent with a Laurentian pro-
of limestone and an epiclastic tuff (Fanning & Link 2004, 2008; venance, with both Archaean and mixed Proterozoic components
Lorentz et al. 2004). Laminated carbonates above diamictite are (Ball & Farmer 1998; Condie et al. 2001; Mueller et al. 2007).
not present in the Sheeprock Mountains or in the Deep Creek Range. A Nd-isotope study of the Trout Creek and McCoy Creek succes-
sions, above the glacial interval, suggest a Palaeoproterozoic
source within the Mojave and Yavapai provinces (Farmer & Ball
Chemostratigraphy 1997). Detrital zircons from Brigham Group sandstone are of
Grenvillian age (1250 – 1000 Ma), with smaller populations of
Cryogenian and Ediacaran strata in Idaho and Utah have been the Mesoproterozoic, Palaeoproterozoic and Archaean age. These
subject of several chemostratigraphic and other geochemical and are interpreted to represent a trans-continental provenance
provenance studies. C-isotope data have been obtained from (Stewart et al. 2001).
three relatively thin carbonate-bearing intervals in the post-glacial
part of the succession (Smith et al. 1994; Lorentz et al. 2004; Cor-
setti et al. 2007), and from organic matter from the upper part of the Other characteristics
pre-glacial Uinta Mountain Group (Dehler et al. 2007). Chemical
Index of Alteration (CIA) data are available for the glaciogenic Organic-walled spheres and aggregates c. 5– 20 mm in diameter
Mineral Fork Fm. (Young 2002), and for the Uinta Mountain [Bavlinella faveolata (Shepeleva) Vidal] have been described
Group and Big Cottonwood Fm. (Condie et al. 2001). Provenance from the Red Pine Shale and older formations of the Uinta Moun-
has been studied widely in both pre-glacial and post-glacial depos- tain Group, along with locally abundant leiosphaerid acritarchs
its (Farmer & Ball 1997; Ball & Farmer 1998; Condie et al. 2001; and filaments (Vidal & Ford 1985; Nagy & Porter 2005; Sprinkel
Mueller et al. 2007). & Waanders 2005; Dehler et al. 2007). Similar unicells, dyads and
C-isotope ratios range from –2.9 to –6.9‰ within carbonate aggregates, also referred to as Bavlinella, were recognized by
beds in the uppermost part of the Scout Mountain Member, includ- Knoll et al. (1981) in mudrocks of the Mineral Fork Fm. In
ing the thin laminated dolostone (level ‘a’ and ‘cap’ in Fig. 38.5; some cases, organic material appears to have been pyritized or to
Smith et al. 1994; Lorentz et al. 2004; Corsetti et al. 2007). The be composed of abiotic framboidal pyrite (Dehler et al. 2007;
data cluster around –4.5 to –5.5‰ in two of three sections (sec- N. J. Butterfield, pers. comm. 2009). However, the existence of
tions 2 and 3) sampled by Lorentz et al. (2004), without a well- at least some microfossils is supported by a restricted stratigraphic
defined stratigraphic trend. Data are more scattered in a third range, the presence of transitional morphologies, and in the case
section (their section 1). O-isotope values of –13.6 to – 22.5‰ of the Mineral Fork, the absence of pyrite (Knoll et al. 1981;
and a positive correlation with d13C in sections 1 and 2 suggest S. Porter, pers. comm. 2009). Trace fossils, including Skolithos,
NEOPROTEROZOIC IDAHO AND UTAH 431

are present widely in the upper part of the Brigham Group (Camel- Pb concordia age of 686 + 4 Ma (Fanning & Link 2008; a separate
back Mountain Quartzite and correlatives; Fig. 38.4), consistent older population of zircons (709 + 5 Ma) was reported by Fanning
with a Cambrian age. & Link 2004). Much or all of the Scout Mountain Member is thus
younger than 686 Ma. Above the upper Scout Mountain diamictite
and the overlying laminated carbonate, an epiclastic tuff contains
Palaeolatitude and palaeogeography zircons dated as 667 + 5 Ma (Fanning & Link 2004).
The more stratigraphically complete sections of northern and
Hematite-cemented sedimentary rocks in the Uinta Mountain west-central Utah remain to be dated. Attempts to separate zircons
Group yield a well-defined palaeomagnetic pole with a mean of from a clast of rhyolite obtained from conglomerate near the base
0.88N, 161.38E (a95 ¼ 4.68; n ¼ 9 sampling localities consisting of the Dutch Peak Fm. (Sheeprock Mountains) and from a suite of
of 79 sites; Weil et al. 2006). The characteristic remanent magne- intermediate-composition volcanic clasts collected from the for-
tization (ChRM) is east-directed (or antipode), of low positive or mation of Perry Canyon at Little Mountain were not successful.
negative inclination, and typically unblocked over a narrow Cryogenian magmatism is also recognized north of the Snake
range of high laboratory temperatures between 660 8C and River Plain in central Idaho. Three samples of felsic orthogneiss
680 8C. A second magnetization is north- to NE-directed with exposed in Wildhorse Creek of the Pioneer Mountains yield
moderate to steep inclination. The presence of dual polarities SHRIMP U –Pb zircon upper-concordia intercept ages of
suggests that the ChRM was acquired close to the time of depo- 692.3 + 5.2, 695.7 + 8.0 and 696.5 + 9.0 Ma (location in
sition, and over a span long enough to average out secular variation Fig. 38.1; K.M. Durk-Autenrieth, Idaho State University, pers.
of the geomagnetic field. The second magnetization is inferred to comm. 2007). The age is thus close to 695 Ma. A somewhat
be a recent or modern overprint. The data are consistent with a older U –Pb age (725 + 5 Ma) was obtained for the House Moun-
low palaeolatitude for Laurentia during deposition of Uinta Moun- tain orthogneiss near the southern edge of the Atlanta lobe of the
tain Group sediment, with a tight counterclockwise apparent polar Idaho batholith (location shown on Fig. 38.l; M. Schmitz, Boise
wander path in the south Pacific, and with deposition over an inter- State University, pers. comm. 2006). In central western Idaho,
val that was sufficiently short (tens of millions of years) for the 684 Ma volcanic rocks overlie diamictite near Edwardsburg
palaeolatitude not to have changed significantly during deposition. (Lund et al. 2003).
Sampling for this most recent study was more comprehensive than Lund et al. (2010) report ages of 665–650 Ma for Cryogenian
that undertaken by Bressler (1981), but inferences about pole alkalic plutonic rocks of the Beaverhead –Big Creek belt in
location and palaeolatitude are not statistically distinguishable. central Idaho. As these are intrusive rocks, within exposures of
Palaeomagnetic data are not available for the glacial interval or Proterozoic basement, their relations to Neoproterozoic glacio-
for the later Cryogenian and Ediacaran in the area of interest. Suit- genic successions are not clear. However, these new ages extend
able rocks are not present in the glaciogenic section, which in most the locations of Cryogenian magmatism into a NW-trending
places is sufficiently deformed and metamorphosed (greenschist) swath across much of Idaho.
to be problematic. Nor has it been possible to achieve a positive The only other direct age constraint on the Neoproterozoic of
fold test, though that was tried twice in the Brigham Group near Utah and Idaho is a 580 +7 Ma age (Fig. 38.2; 40Ar/39Ar on horn-
Huntsville, Utah (as reported in Link et al. 1994). Based on a blende from trachyte from the upper Browns Hole Fm.; Christie-
global assessment of the reliability of the best available palaeo- Blick & Levy 1989). This unit overlies strata that contain incised
magnetic data from Cryogenian strata and interpreted positions valleys interpreted to represent sea-level drawdown associated
of continents, Evans et al. (2000) concluded that glacial deposits with a younger Cryogenian glaciation (Christie-Blick & Levy 1989).
in southeastern Idaho and Utah accumulated at a palaeolatitude
of ,58, and that western North America remained at low palaeo-
latitude through the end of the Ediacaran (cf. Torsvik et al. 1996). Discussion

The Neoproterozoic and early Cambrian palaeoenvironmental and


Geochronological constraints palaeogeographic evolution of Idaho and Utah reflect a combi-
nation of climatic and tectonic controls. Mostly marine glacial
A population of detrital zircons from the formation of Outlaw deposits are dated in southeastern Idaho as c. 686 + 4 Ma to
Trail, in the lower part of the eastern Uinta Mountain Group, c. 667 + 5 Ma (Cryogenian). Thicker, more complete sections in
yields a concordia age of 766 + 5 Ma, suggesting the rocks must northern and west-central Utah may include older glacial deposits,
be younger (Dehler et al. 2007). This is substantially younger but age data are not yet available for those rocks. Direct evidence
than previous estimates of c. 900 Ma for the upper part of the for a younger Cryogenian glaciation is not preserved in this region.
Uinta Mountain Group (Link et al. 1993; Stone 1993), and it However, stratigraphically restricted incised valleys as much as
places a very conservative upper bound on the age of pre-glacial 160 m deep are inferred to be at least in part of glacial-eustatic
sedimentation in Utah. Based on similarities in microfossils to origin. Available evidence suggests that all of the Neoproterozoic
Chuar Group strata in the Grand Canyon that contain a 740 Ma deposits accumulated at low palaeolatitude.
tuff bed, the Uinta Mountain Group is estimated to span 766–
740 Ma, and to be Cryogenian in age. The Big Cottonwood Fm.
is estimated to span the same time interval. Depositional settings and climatic controls
The timing of Neoproterozoic glaciation in the Pocatello Fm. of
southeastern Idaho is bracketed by SHRIMP U – Pb dating of The Uinta Mountain Group accumulated in braided fluvial and
volcanic clasts within diamictite and two intervals of tuff in the shallow marine environments (Wallace & Crittenden 1969; Crit-
Scout Mountain Member. Felsic volcanic clasts in diamictite tenden & Wallace 1973; Sanderson 1984; Dehler et al. 2005b,
near Pocatello are dated as 717 + 4 Ma (Fanning & Link 2004) 2007, 2010). Beginning at the base, marine intervals have been
and 701 + 4 Ma (Fanning & Link 2008). These cobble- to documented in the Jesse Ewing Canyon Fm. and in the informal
boulder-sized clasts are interpreted as having been eroded from formations of Red Castle, Mount Agassiz, Dead Horse Pass, Moo-
uplifted exposures of the underlying Bannock Volcanic Member sehorn Lake and Red Pine Shale (Fig. 38.3; Dehler et al. 2007;
of the Pocatello Fm. (Link 1983). On Oxford Mountain near the E. M. Kingsbury, pers. comm. 2009). Tidal rhythmites preserved
Idaho –Utah state line (Fig. 38.1), an epiclastic plagioclase-phyric in the coeval Big Cottonwood Fm. provide some of the best evi-
mafic tuff breccia interbedded with diamictite-bearing rocks near dence for a marine connection in spite of an intracratonic setting
the base of the Scout Mountain Member yielded a SHRIMP U – that must have been many hundreds of kilometres from the
432 P. K. LINK & N. CHRISTIE-BLICK

nearest oceanic crust (Chan et al. 1994; Ehlers et al. 1997; Ehlers passive continental margin, remnants of which are preserved today
& Chan 1999; Dehler et al. 2005b). from eastern Alaska to eastern California (Stewart 1972; Stewart &
Glaciogenic and associated deposits in southeastern Idaho and Suczek 1977; Bond et al. 1983, 1985; Christie-Blick & Levy 1989;
Utah accumulated for the most part in a deep marine setting Ross 1991; Levy & Christie-Blick 1991a). Details of the tectonic
(water depths as great as hundreds of metres; Christie-Blick 1983a, history are unresolved, particularly with respect to the timing of
1997; Crittenden et al. 1983; Link 1983; Link et al. 1994). This is continental separation. Evidence for crustal extension is necessarily
indicated by the overall character of the facies, by their association indirect owing to the difficulty in documenting stratigraphic growth
with turbidite sandstones and subaqueous extrusive volcanic rocks, in available outcrop and to the tendency for originally faulted basin
by the presence of ice-rafted dropstones and till clots (albeit margins to be offset by younger structures. The pre-glacial Uinta
sporadically), and in the Sheeprock Mountains by a progradational Mountain Group and Big Cottonwood Fm. are thought to have
stratigraphic architecture. The Mineral Fork Fm. is thought to have been deposited in an east-trending ,766 Ma rift basin, based
accumulated close to the grounding line of a partially buoyant ice upon abrupt thinning of these deposits towards the north across a
sheet, on the basis of glacial grooves and probable roches mouton- fault-controlled boundary (Sears et al. 1982; Christie-Blick 1997;
nées at the basal contact, syndepositional deformation, and abun- Mueller et al. 2007). Stone’s (1993) alternative interpretation of
dant well-stratified sandstone interpreted as subaqueous outwash. the Uinta Mountain Group in terms of a seaway that encroached
Direct evidence for grounding is not present elsewhere, although on an area of low topography fails to account for a stratigraphic
quartzite in the Otts Canyon and Dutch Peak formations of the thickness as great as 7 km. Abrupt thickness changes in the Shee-
Sheeprock Mountains may have accumulated in a glacial–fluvial prock Group of the Sheeprock Mountains are attributed to a combi-
to braid-delta setting (Link et al. 1994). The appreciably greater nation of facies change, progradation into a deep depocentre, and
water depths implied by most of the deposits, in comparison with tectonic tilting towards the SE at a hinged basin margin (Christie-
‘syn-rift’ strata at lower and higher stratigraphic levels, is thought Blick 1982, 1997; Crittenden et al. 1983). Inferred stratigraphic
to indicate an isostatic response to the weight of the nearby ice growth encompasses the interval of glaciation, although it may
sheet as well as tectonically driven subsidence. have begun earlier. Magmatism, assumed to be rift-related, is wide-
Up to several hundreds of metres of siltstone that in most places spread at this same stratigraphic level, ranging in age from c. 720
directly overlie glaciogenic deposits, locally with multiple beds of to c. 685 Ma, and as young as 650 Ma in the case of intrusive
intervening carbonate, are attributed to glacial –eustatic sea-level rocks in central Idaho. The Bannock Volcanic Member of the Poca-
rise combined with thermally driven tectonic subsidence. C- tello Fm. includes mafic volcanic rocks with a rift-related trace
isotope data from the thin laminated carbonate beds that overlie element signature (Harper & Link 1986).
Scout Mountain Member diamictite are comparable with the Facies and thickness variation at the level of the upper member
strongly depleted d13C values of cap carbonates elsewhere of the Pocatello Fm. and Brigham Group (post-glacial) are less
(Corsetti et al. 2007). Near-zero to positive d13C signatures in stra- pronounced, consistent with the onset of thermally driven subsi-
tigraphically higher carbonate layers are comparable with data dence of a passive margin as early as c. 665 Ma. However, quan-
from the Johnnie Fm. of Death Valley (Corsetti & Kaufman titative analysis of tectonic subsidence in Cambro-Ordovician
2003) and are consistent with a late Cryogenian age. strata indicates that a second pulse of thermal subsidence began
The Brigham Group and correlatives are dominated by braided as late as c. 520 Ma, if account is taken of recent changes to the
fluvial, braid delta and shallow marine deposits (Link et al. 1987; Cambrian timescale (Bond et al. 1983; Christie-Blick & Levy
Christie-Blick & Levy 1989; Levy & Christie-Blick 1991b; Levy 1989; Levy & Christie-Blick 1991a; Link et al. 1994). Early Cam-
et al. 1994; Christie-Blick 1997). Incised valleys at several brian timing for this second pulse is supported in the central
scales and horizons, at or near the top of the Caddy Canyon Quart- Wasatch Range by angular discordance of up to 108 between the
zite, represent approximately the same stratigraphic level as the Big Cottonwood Fm. and overlying Cambrian quartzite, which
glaciogenic Ice Brook Fm. in the Mackenzie Mountains of north- Christie-Blick & Levy (1989) and Christie-Blick (1997) took to
western Canada (Aitken 1991; Levy & Christie-Blick 1991b; Levy indicate rift-related reactivation of the northern bounding fault of
et al. 1994; Ross et al. 1995; Christie-Blick 1997). Taken together, the Big Cottonwood basin. The mismatch between the magnitude
these observations are consistent with a glacial– eustatic origin and of early Palaeozoic post-rift thermal subsidence and the minimal
a late Cryogenian age (655 –635 Ma; Hoffman & Li 2009). evidence for crustal extension after c. 665 Ma is best explained
The origin of the sequence boundary at or near the base of the in terms of inhomogeneous extension of the lithosphere in latest
Mutual Fm. is less clear (Levy & Christie-Blick 1991b; Christie- Neoproterozoic time (Christie-Blick & Levy 1989). Stratigraphic
Blick 1997). Eustatic and tectonic explanations are both permitted evidence for multiple rifting events and for one or more times of
by the only available age constraint (an 40Ar/39Ar age of passive margin formation at or after c. 665 Ma is inconsistent
580 + 7 Ma from the overlying Browns Hole Fm. at Huntsville, with palaeomagnetically based interpretations of a rift-to-drift
Utah). A glacial –eustatic origin is consistent with age estimates transition prior to 750 Ma (Wingate & Giddings 2000; Torsvik
for the Gaskiers glaciation of Newfoundland. (ID-TIMS dates on 2003; Weil et al. 2006). The latter requires assumptions about
ash beds below, within and above the glaciogenic Gaskiers Fm. the identity of counterpart blocks, and about the positioning of
constrain its age to c. 584–582 Ma; Bowring et al. 2003.) A diffi- those blocks at the time of continental break-up.
culty with the glacial –eustatic hypothesis is that while sea-level
rise after an initial drawdown accounts for the thickness of the
Mutual (hundreds of metres) over a broad area, it does not Regional correlations
explain why fluvial sedimentation continued. Christie-Blick
(1997) suggested that this might have to do with an increase in Available stratigraphic and geochronological data support regional
sediment supply. A difficulty with a tectonic interpretation for correlation. The pre-glacial Uinta Mountain Group and Big Cot-
the unconformity is that regional uplift reduces available sedimen- tonwood Fm. are comparable in terms of age and tectonic setting
tary accommodation. So renewed subsidence is then required to to the Chuar Group of the Grand Canyon region (Karlstrom
account for the considerable thickness of overlying quartzite. et al. 2000). Marine strata within the three successions are approxi-
mately coeval (c. 766 –740 Ma), suggesting that they may have
accumulated within the same epicontinental ‘ChUMP seaway’,
Tectonic setting which Dehler et al. (2005b, 2007) regarded as encompassing at
least part of the Pahrump Group in the Death Valley area of
The rocks of southeastern Idaho and Utah are inferred to have accu- eastern California. Cryogenian glacial deposits of Idaho and
mulated in rift-related basins associated with the development of a Utah are inferred to correlate approximately with the Kingston
NEOPROTEROZOIC IDAHO AND UTAH 433

Peak Fm. (upper Pahrump Group; Stewart & Suczek 1977; Link and Crustal Evolution of the Western United States. Prentice-Hall,
et al. 1993; Corsetti et al. 2007; Mrofka & Kennedy 2011; Petter- Englewood Cliffs, New Jersey, 432– 445.
son et al. 2011) and the Edwardsburg Fm. in central Idaho (Lund Chan, M. A., Kvale, E. P., Archer, A. W. & Sonett, C. 1994. Oldest
et al. 2003, 2011), although a glacial origin for the diamictite- direct evidence of lunar– solar tidal forcing encoded in sedimentary
bearing Edwardsburg Fm. has not been demonstrated. The succes- rhythmites, Proterozoic Big Cottonwood Formation, Central Utah.
sions in southeastern Idaho and Utah have also traditionally been Geology, 22, 791– 794.
broadly correlated with the Windermere Supergroup in Canada Christie-Blick, N. 1982. Upper Proterozoic and Lower Cambrian
(Link et al. 1993; Lund et al. 2003), although age control is not rocks of the Sheeprock Mountains, Utah: regional correlation
and significance. Geological Society of America Bulletin, 93,
yet adequate to explore diachrony in the timing of either glaciation
735– 750.
or tectonism in western North America, as predicted by the Christie-Blick, N. 1983a. Glacial-marine and subglacial sedimentation,
‘zipper-rift’ hypothesis of Eyles & Januszczak (2004). Upper Proterozoic Mineral Fork Formation, Utah. In: Molnia, B. F.
(ed.) Glacial Marine Sedimentation. Plenum Press, New York,
Over the years, our research in the Neoproterozoic of the western United States 703– 776.
has been supported by the National Science Foundation (most recently NSF Christie-Blick, N. 1983b. Structural geology of the southern Sheeprock
08-19884 to Link), U.S. Geological Survey, and by the Donors of the Petroleum Mountains, Utah: regional significance. In: Miller, D. M., Todd, V.
Research Fund, administered by the American Chemical Society. We acknowl- R. & Howard, K. A. (eds) Tectonic and Stratigraphic Studies in the
edge mentors (J. C. Crowell and Max D. Crittenden, Jr) plus colleagues and Eastern Great Basin. Geological Society of America Memoir, 157,
students too numerous to mention individually, who have contributed to our emer- 101– 124.
ging understanding of these rocks. This represents a contribution of the IUGS- and Christie-Blick, N. 1985. Upper Proterozoic glacial-marine and subgla-
UNESCO-funded IGCP (International Geoscience Programme) project #512. cial deposits at Little Mountain, Utah. Brigham Young University
Geology Studies, 32, 9 –18.
Christie-Blick, N. 1997. Neoproterozoic sedimentation and tectonics in
west-central Utah. Brigham Young University Geology Studies, 42,
References 1– 30.
Christie-Blick, N. & Link, P. K. 1988. Glacial-marine sedimentation,
Aitken, J. D. 1991. The Ice Brook Formation and post-Rapitan, Late Mineral Fork Formation (Late Proterozoic) Utah. In: Holden,
Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. G. S. (ed.) Geological Society of America Annual Meeting Field
Geological Survey of Canada Bulletin, 404. Trip Guidebook 1988. Colorado School of Mines – Professional
Allmendinger, R. W. 1992. Fold and thrust tectonics of the western Contributions, 12, 259– 274.
United States exclusive of the accreted terranes. In: Burchfiel, Christie-Blick, N. & Levy, M. (eds) 1989. Late Proterozoic and Cam-
B. C., Lipman, P. W. & Zoback, M. L. (eds) The Cordilleran brian tectonics, sedimentation, and record of Metazoan radiation in
Orogen: Conterminous US, The Geology of North America. Geologi- the western United States. 28th International Geological Congress
cal Society of America, Boulder, Colorado, G-3, 583– 607. Field Trip Guidebook, American Geophysical Union, Washington
Armstrong, F. C. & Oriel, S. S. 1965. Tectonic development of Idaho- DC, T331.
Wyoming thrust belt. American Association of Petroleum Geologists Christie-Blick, N., Grotzinger, J. P. & von der Borch, C. C. 1988.
Bulletin, 49, 1847–1866. Sequence stratigraphy in Proterozoic successions. Geology, 16,
Armstrong, R. L. 1968. Sevier orogenic belt in Nevada and Utah. 100– 104.
Geological Society of America Bulletin, 79, 429–458. Condie, K. C., Lee, D. & Farmer, G. L. 2001. Tectonic setting and
Ball, T. T. & Farmer, G. L. 1998. Infilling history of a Neoproterozoic provenance of the Neoproterozoic Uinta Mountain and Big
intracratonic basin: Nd isotope provenance studies of the Uinta Cottonwood groups, northern Utah: constraints from geochemistry,
Mountain Group, western United States. Precambrian Research, Nd isotopes, and detrital modes. Sedimentary Geology, 141– 142,
87, 1 – 18. 443– 464.
Bick, K. F. 1966. Geology of the Deep Creek Mountains, Tooele and Juab Corsetti, F. A. & Kaufman, A. J. 2003. Stratigraphic investigations of
Counties, Utah. Utah Geological and Mineralogical Survey, Bulletin, carbon isotope anomalies and Neoproterozoic ice ages in Death
77, 120. Valley, California. Geological Society of America Bulletin, 115,
Blackwelder, E. 1932. An ancient glacial formation in Utah. Journal 916– 932.
of Geology, 40, 289–304. Corsetti, F. A., Link, P. K. & Lorentz, N. J. 2007. d13C chemostratigra-
Blick, N. 1981. Late Precambrian glaciation in Utah. In: Hambrey, M. J. phy of the Neoproterozoic succession near Pocatello, Idaho: impli-
& Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. cations for glacial chronology and regional correlations: In: Link,
Cambridge University Press, Cambridge, 740– 744. P. K. & Lewis, R. S. (eds) Proterozoic Geology of Western North
Bond, G. C., Kominz, M. A. & Devlin, W. J. 1983. Thermal subsidence America and Siberia: SEPM (Society for Sedimentary Geology)
and eustasy in the Lower Paleozoic miogeocline of western North Special Publication, 86, 193– 208.
America. Nature, 306, 775– 779. Crittenden, M. D., Jr. & Wallace, C. A. 1973. Possible equivalents of
Bond, G. C., Christie-Blick, N., Kominz, M. A. & Devlin, W. J. 1985. the Belt Supergroup in Utah. Idaho Bureau of Mines and Geology,
An Early Cambrian rift to post-rift transition in the Cordillera of Moscow ID, Belt Symposium, 1, 116–138.
western North America. Nature, 316, 742– 745. Crittenden, M. D., Sharp, B. J. & Calkins, F. C. 1952. Geology of
Bowring, S. A., Myrow, P., Landing, E., Ramezani, J. & Grotzinger, the Wasatch Mountains east of Salt Lake City, Parleys Canyon to
J. P. 2003. Geochronological constraints on terminal Neoproterozoic Traverse Range In: Marsell, R. E. (ed.) Geology of the Central
events and the rise of metazoans. European Geophysical Union Wasatch Mountains. Utah Geological and Mineralogical Survey
Annual Meeting, Nice, 2003. Geophysical Research Abstracts, 5, Guidebook to the Geology of Utah, 8, 1 –37.
13219. Crittenden, M. D., Jr., Schaeffer, F. E., Trimble, D. E. & Woodward,
Bressler, S. L. 1981. Preliminary paleomagnetic poles and correlation L. A. 1971. Nomenclature and correlation of some upper
of the Proterozoic Uinta Mountain Group, Utah and Colorado. Precambrian and basal Cambrian sequences in western Utah and
Earth & Planetary Science Letters, 55, 53 –64. southeastern Idaho. Geological Society of America Bulletin, 82,
Bruhn, R. L., Picard, M. D. & Isby, J. S. 1986. Tectonics and sedimen- 581– 602.
tology of Uinta Arch, western Uinta Mountains and Uinta Basin. In: Crittenden, M. D., Jr, Christie-Blick, N. & Link, P. K. 1983. Evi-
Peterson, J. A. (ed.) Paleotectonics, Sedimentation in the Rocky dence for two pulses of glaciation during the Late Proterozoic in
Mountain Region United States. American Association of Petroleum northern Utah. Geological Society of America Bulletin, 94, 437– 450.
Geologists Memoir, 41, 333–352. DeCelles, P. G. 2004. Jurassic to Eocene evolution of the Cordilleran
Bryant, B. 1988. Evolution and early Proterozoic history of the margin of thrust belt and foreland basin system, western U.S.A. American
the Archean continent in Utah. In: Ernst, W. G. (ed.) Metamorphism Journal of Science, 304, 105–168.
434 P. K. LINK & N. CHRISTIE-BLICK

DeCelles, P. G. & Coogan, J. C. 2006. Regional structure and the global carbon cycle, and ecosystem expansion by 740 Ma.
kinematic history of the Sevier fold-and-thrust belt, central Utah. Geology, 28, 619–622.
Geological Society of America Bulletin, 118, 841–864, Knoll, A. H., Blick, N. & Awramik, A. M. 1981. Stratigraphic and
doi: 10.1130/B25759.1. ecologic implications of late Precambrian microfossils from Utah.
Dehler, C. M. & Sprinkel, D. A. 2005. Revised stratigraphy and corre- American Journal of Science, 281, 247–263.
lation of the Neoproterozoic Uinta Mountain Group, northeastern Levy, M. & Christie Blick, N. 1989. Pre-Mesozoic palinspastic recon-
Utah. In: Dehler, C. M., Pederson, J. L., Sprinkel, D. A. & struction of the eastern Great Basin (western United States).
Kowallis, B. J. (eds) Uinta Mountain Geology. Utah Geological Science, 245, 1454– 1462.
Association Publication, 33, 17 –30. Levy, M. & Christie-Blick, N. 1991a. Tectonic subsidence of the early
Dehler, C. M., Elrick, M., Bloch, J., Crossey, L., Karlstrom, K. & Paleozoic passive continental margin in eastern California and
Des Marais, D. 2005a. High-resolution d13C stratigraphy of the southern Nevada. Geological Society of America Bulletin, 103,
Chuar Group (c. 770– 742 Ma), Grand Canyon: implications for 1590– 1606.
mid-Neoproterozoic climate change. Geological Society of America Levy, M. & Christie-Blick, N. 1991b. Late Proterozoic paleogeography
Bulletin, 117, 32– 45. of the eastern Great Basin. In: Cooper, J. D. & Stevens, C. H. (eds)
Dehler, C. M., Sprinkel, D. A. & Porter, S. M. 2005b. Neoproterozoic Paleozoic Paleogeography of the Western United States – II. Society
Uinta Mountain Group of northeastern Utah: Pre-Sturtian geographic, of Economic Paleontologists and Mineralogists, Los Angeles,
tectonic, and biologic evolution. In: Pederson, J. & Dehler, C. M. Pacific Section, 1, 371– 386.
(eds) Interior Western United States. Geological Society of America Levy, M., Christie-Blick, N. & Link, P. K. 1994. Neoproterozoic
Field Guide 6, 1 –25. incised valleys of the eastern Great Basin, Utah and Idaho: fluvial
Dehler, C. M., Porter, S. M., De Grey, L. D., Sprinkel, D. A. & response to changes in depositional base level. In: Dalrymple,
Brehm, A. 2007. The Neoproterozoic Uinta Mountain Group R. W., Boyd, R. & Zaitlin, B. A. (eds) Incised Valley Systems:
revisited: a synthesis of recent work on the Red Pine Shale and Origin and Sedimentary Sequences. SEPM (Society for Sedimentary
related undivided clastic strata, Northeastern Utah. In: Link, P. K. Geology) Special Publication, 51, 369– 382.
& Lewis, R. S. (eds) Proterozoic Geology of Western North Link, P. K. 1981. Upper Proterozoic diamictites in southeastern Idaho.
America and Siberia, SEPM Special Publication, 86, 151–166. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene
Dehler, C. M., Fanning, C. M., Link, P. K., Kingsbury, E. M. & Rybc- Glacial Record. Cambridge University Press, Cambridge, 736– 739.
zynski, D. 2010. Incipient Rodinia breakup, marine transgression, Link, P. K. 1983. Glacial and tectonically influenced sedimentation in the
and peri-Gondwanan sediment source in western Laurentia at Upper Proterozoic Pocatello Formation, southeastern Idaho. In:
,766– 742 Ma: new SHRIMP data from the Uinta Mountain Miller, D. M., Todd, V. R. & Howard, K. A. (eds) Tectonic and
Group and Big Cottonwood Formation, northern Utah. Geological Stratigraphic Studies in the Eastern Great Basin. Geological
Society of America Bulletin, 122, 1686– 1699, doi: 10.1130/B30094. Society of America Memoir, 157, 165– 181.
Ehlers, T. & Chan, M. 1999. Tidal sedimentology and estuarine depo- Link, P. K. 1987. The Late Proterozoic Pocatello Formation; A record of
sition of the Proterozoic Big Cottonwood Formation, Utah. Journal continental rifting and glacial marine sedimentation, Portneuf
of Sedimentary Research, 69, 1169–1180. Narrows, southeastern Idaho. In: Beus, S. S. (ed.) Centennial Field
Ehlers, T. A., Chan, M. A. & Link, P. K. 1997. Proterozoic tidal, glacial Guide Vol. 2. Rocky Mountain Section of the Geological Society of
and fluvial sedimentation in Big Cottonwood Canyon, Utah. Brigham America, Decade of North American Geology, Boulder, CO,
Young University Geology Studies, 42, 31 –58. 139– 142.
Evans, D. A. D., Li, Z. X., Kirschvink, J. L. & Wingate, M. T. D. 2000. Link, P. K., LeFebre, G. B., Pogue, K. R. & Burgel, W. D. 1985. Struc-
A high-quality mid-Neoproterozoic paleomagnetic pole from South tural geology between the Putnam Thrust and the Snake River Plain.
China, with implications for ice ages and the breakup configuration In: Kerns, G. & Kerns, R. (eds) Orogenic Patterns and Stratigraphy
of Rodinia. Precambrian Research, 100, 313–334. of North-Central Utah and Southeastern Idaho. Utah Geological
Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for Association Publication, 14, 67 –74.
Neoproterozoic glaciations during the breakup of Rodinia after Link, P. K., Jansen, S. T., Halimdihardja, P., Lande, A. C. & Zahn, P.
750 Ma. Earth-Science Reviews, 65, 1– 73. D. 1987. Stratigraphy of the Brigham Group (Late Proterozoic –
Fanning, C. M. & Link, P. K. 2004. U –Pb SHRIMP ages of Neoproter- Cambrian), Bannock, Portneuf, and Bear River Ranges, southeastern
ozoic (Sturtian) glaciogenic Pocatello Formation, southeastern Idaho. Idaho. In: Miller, W. R. (ed.) The Thrust Belt Revisited. Wyoming
Geology, 32, 881–884. Geological Association Guidebook, 38, 133–148.
Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian Link, P. K., Christie-Blick, N. et al. 1993. Middle and Late Proterozoic
glaciation; data from the Adelaide Geosyncline, South Australia stratified rocks of the western United States Cordillera, Colorado
and Pocatello Formation, Idaho, USA. In: Gallagher, S. J. & Plateau, and Basin and Range Province. In: Reed, J., Sims, P.,
Wallace, M. W. (eds) Neoproterozoic extreme climates and the Houston, R. S., Rankin, D. W., Link, P. K., Van Schmus, W. R.
origin of early metazoan life. Geological Society of Australia & Bickford, M. E. (eds) Precambrian: Conterminous United
Extended Abstracts, 91, 57 – 62. States. Geological Society of America Decade of North American
Farmer, G. L. & Ball, T. T. 1997. Sources of Middle Proterozoic to Early Geology Series, C-3, 474–690.
Cambrian siliciclastic sediments in the Great Basin: a Nd isotope Link, P. K., Miller, J. M. G. & Christie-Blick, N. 1994. Glacial-marine
study. Geological Society of America Bulletin, 109, 1193–1205. facies in a continental rift environment: Neoproterozoic rocks of the
Hansen, W. R. 1965. Geology of the Flaming Gorge Area Utah- western United States Cordillera. In: Deynoux, M., Miller,
Colorado-Wyoming. US Geological Survey Professional Paper, J. M. G., Domack, E. W., Eyles, N., Fairchild, I .J. & Young,
490, 196. G. M. (eds) Earth’s Glacial Record. International Geological
Harper, G. D. & Link, P. K. 1986. Geochemistry of Upper Proterozoic Correlation Project 260. Cambridge University Press, Cambridge,
rift-related volcanics, northern Utah and southeastern Idaho. 29 – 59.
Geology, 14, 864–867. Link, P. K., Corsetti, F. A. & Lorentz, N. J. 2005. Pocatello Formation
Hoffman, P. F. & Li, Z.-X. 2009. A palaeogeographic context for Neo- and overlying strata, southeastern Idaho: snowball Earth diamictites,
proterozoic glaciation. Palaeogeography, Palaeoclimatology, cap carbonates, and Neoproterozoic isotopic profiles. In: Pederson,
Palaeoecology, 277, 158–172. J. & Dehler, C. M. (eds) Interior Western United States. Geological
Hoffman, P. F. & Halverson, G. P. 2011. Neoproterozoic glacial record Society of America Field Guide, 6, 251– 259, doi: 10.1130/
in the Mackenzie Mountains, northern Canadian Cordillera. In: 2005.fld006(12).
Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Lorentz, N. J., Corsetti, F. A. & Link, P. K. 2004. Seafloor precipitates
Geological Record of Neoproterozoic Glaciations. Geological and C-isotope stratigraphy from the Neoproterozoic Scout Mountain
Society, London, Memoirs, 36, 397 –411. Member of the Pocatello Formation, southeast Idaho: implications
Karlstrom, K. E., Bowring, S. A. et al. 2000. Chuar Group of the for Neoproterozoic earth system behavior. Precambrian Research,
Grand Canyon: Record of breakup of Rodinia, associated change in 130, 57– 70.
NEOPROTEROZOIC IDAHO AND UTAH 435

Ludlum, J. C. 1942. Pre-Cambrian formations at Pocatello, Idaho. Smith, L. H., Kaufman, A. J., Knoll, A. H. & Link, P. K. 1994.
Journal of Geology, 50, 85 – 95. Chemostratigraphy of predominantly siliciclastic Neoproterozoic
Lund, K., Aleinikoff, J. N., Evans, K. V. & Fanning, C. M. 2003. successions: a case study of the Pocatello Formation and
SHRIMP U–Pb geochronology of Neoproterozoic Windermere Lower Brigham Group, Idaho, USA. Geological Magazine, 131,
Supergroup, central Idaho: implications for rifting of western Lauren- 301– 314.
tia and synchroneity of Sturtian glacial deposits. Geological Society Smith, M. D. S., Arnaud, E., Arnott, R. W. C. & Ross, G. M. 2011.
of America Bulletin, 115, 349–372. The record of Neoproterozoic glaciation in the Windermere Super-
Lund, K., Aleinikoff, J. N., Evans, K. V., duBray, E. A., Dewitt, E. H. group, southern Canadian Cordillera. In: Arnaud, E., Halverson,
& Unruh, D. M. 2010. SHRIMP U–Pb dating of recurrent Cryogen- G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
ian and Late Cambrian – Early Ordovician alkalic magmatism in terozoic Glaciations. Geological Society, London, Memoirs, 36,
central Idaho: implications for Rodinian rift tectonics. Geological 413– 423.
Society of American Bulletin, 122, 430– 453. Sprinkel, D. A. & Waanders, G. 2005. Stratigraphy, organic microfos-
Lund, K., Aleinikoff, J. N. & Evans, K. V. 2011. The Edwardsburg sils, and thermal maturation of the Neoproterozoic Uinta Mountain
Formation and related rocks, Windermere Supergroup, central Group in eastern Uinta Mountains, northeastern Utah. In: Dehler,
Idaho, U.S.A. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, C. M., Pederson, J. L., Sprinkel, D. A. & Kowallis, B. J. (eds)
G. (eds) The Geological Record of Neoproterozoic Glaciations. Geo- Uinta Mountain Geology. Utah Geological Association Publication,
logical Society, London, Memoirs, 36, 437–447. 33, 63– 73.
Misch, P. & Hazzard, J. C. 1962. Stratigraphy and metamorphism of Stewart, J. H. 1972. Initial deposits in the Cordilleran geosyncline:
late Precambrian rocks of central-east Nevada and adjacent Utah. evidence of a Late Precambrian (,850 m.y.) continental separation.
American Association of Petroleum Geologists Bulletin, 46, Geological Society of America Bulletin, 83, 1345– 1360.
310– 316. Stewart, J. H. 1982. Regional relations of Proterozoic Z and Lower Cam-
Mueller, P. A., Foster, D. A., Mogk, D. W., Wooden, J. L., Kamenov, brian rocks in the western United States. In: Cooper, J. D., Troxel,
G. D. & Vogl, J. J. 2007. Detrital mineral chronology of the Uinta B. W. & Wright, L. A. (eds) Geology of Selected Areas in the San
Mountain Group: implications for the Grenville flood in southwestern Bernardino Mountains, Western Mojave Desert, and Southern
Laurentia. Geology, 35, 431– 434. Great Basin. Guidebook for Fieldtrip No. 9, Cordilleran Section
Mrofka, D. & Kennedy, M. 2011. The Kingston Peak Formation in the Geological Society of America. Death Valley Publishing Company,
eastern Death Valley Region. In: Arnaud, E., Halverson, G. P. & Shoshone Ca.,171– 186.
Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Stewart, J. H. & Suczek, C. A. 1977. Cambrian and latest Precambrian
Glaciations. Geological Society, London, Memoirs, 36, 449– 458. paleogeography and tectonics in the western United States. In:
Nagy, R. M. & Porter, S. M. 2005. Paleontology of the Neoproterozoic Stewart, J. H., Stevens, C. H. & Fritsche, A. E. (eds) Paleozoic
Uinta Mountain Group. In: Dehler, C. M., Pederson, J. L., Sprin- Paleogeography of the Western United States. Society of Economic
kel, D. A. & Kowallis, B. J. (eds) Uinta Mountain Geology. Utah Paleontologists and Mineralogists, Pacific Section, Pacific Coast
Geological Association Publication, 33, 49 –62. Paleogeography Symposium, I, 1 – 18.
Nelson, R. B. 1966. Structural development of northernmost Snake Stewart, J. H., Gehrels, G. E., Barth, A. P., Link, P. K., Christie-
Range, Kern Mountains, and Deep Creek Range, Nevada and Utah. Blick, N. & Wrucke, C. T. 2001. Detrital zircon provenance of
American Association of Petroleum Geologists Bulletin, 50, Mesoproterozoic to Cambrian arenites in the western United States
921– 951. and northwestern Mexico. Geological Society of America Bulletin,
Ojakangas, R. W. & Matsch, C. L. 1980. Upper Precambrian (Eocam- 113, 1343–1356.
brian) Mineral Fork Tillite of Utah: a continental glacial and glacio- Stone, D. E. 1993. Tectonic evolution of the Uinta Mountains:
marine sequence. Geological Society of America Bulletin, 91, palinspastic restoration of a structural cross section along longitude
495– 501. 1098150 , Utah. Utah Geological Survey Miscellaneous Publication,
Petterson, R., Prave, A. R. & Wernicke, B. P. 2011 Glaciogenic and 93-8.
related strata of the Neoproterozoic Kingston Peak Formation in the Torsvik, T. H. 2003. The Rodinia jigsaw puzzle. Science, 300,
Panamint Range, Death Valley region, California. In: Arnaud, E., 1379–1381.
Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Torsvik, T. H., Smethurst, M. A. et al. 1996. Continental break-up and
Record of Neoproterozoic Glaciations. Geological Society, London, collision in the Neoproterozoic and Palaeozoic – a tale of Baltica and
Memoirs, 36, 459–466. Laurentia. Earth-Science Reviews, 40, 229–258.
Rodgers, D. W. 1994. Stratigraphy, correlation, and depositional environ- Trimble, D. E. 1976. Geology of the Michaud and Pocatello quadrangles,
ments of Upper Proterozoic and Lower Cambrian rocks of the Bannock and Power Counties, Idaho. US Geological Survey Bulletin,
southern Deep Creek Range, Utah. In: Kerns, G. J. & Kerns, R. 1400, 88.
L., Jr. (eds) Geology of Northwest Utah, Southern Idaho and Vidal, G. & Ford, T. D. 1985. Microbiotas from the Late Proterozoic
Northeast Nevada. Utah Geological Association Publication, 13, Chuar Group (northern Arizona) and Uinta Mountain Group (Utah)
79 – 91. and their chronostratigraphic implications. Precambrian Research,
Rodgers, D. W., Long, S. P., McQuarrie, N., Burgel, W. D. & 28, 349– 389.
Hersely, C. F. 2006. Geologic Map of the Inkom Quadrangle, Wallace, C. A. & Crittenden, M. D. 1969. The stratigraphy, deposi-
Bannock County, Idaho, Idaho Geological Survey Map, T-06-2, tional environment and correlation of the Precambrian Uinta Moun-
scale 1:24,000. tain Group, western Uinta Mountains, Utah. In: Lindsey, J. B. (ed.)
Ross, G. M. 1991. Tectonic setting of the Windermere Supergroup Geologic Guidebook of the Uinta Mountains. Intermountain Associ-
revisited. Geology, 19, 1125– 1128. ation of Geologists Field Conference, 16, 127–142.
Ross, G. M., Block, J. D. & Krouse, H. R. 1995. Neoproterozoic strata Weil, A. B., Geissman, J. W. & Ashby, J. M. 2006. A new palaeomag-
of the southern Canadian Cordillera and the isotopic evolution of netic pole for the Neoproterozoic Uinta Mountain Group, Central
seawater sulfate. Precambrian Research, 73, 71– 99. Rocky Mountain States, USA. Precambrian Research, 147,
Sanderson, I. D. 1984. The Mount Watson Formation, an interpreted 234– 259.
braided-fluvial deposit in the Uinta Mountain Group (upper Precam- Wernicke, B. 1992. Cenozoic extensional tectonics of the U.S. Cordillera.
brian), Utah. The Mountain Geologist, 21, 157– 164. In: Burchfiel, B. C., Lipman, P. W. & Zoback, M. L. (eds)
Sanderson, I. D. & Wiley, M. T. 1986. The Jesse Ewing Canyon The Cordilleran Orogen: Conterminous U.S. Geological Society of
Formation, an interpreted alluvial fan deposit in the basal Uinta America, Boulder, Colorado, The Geology of North America, G-3,
Mountain Group (Middle Proterozoic), Utah. The Mountain Geol- 553– 581.
ogist, 23, 77– 89. Wingate, M. T. D. & Giddings, J. W. 2000. Age and paleomagnetism
Sears, J. W., Graff, P. J. & Holden, G. S. 1982. Tectonic evolution of of the Mundine Well dyke swarm, Western Australia. Implications
lower Proterozoic rocks, Uinta Mountains Utah and Colorado. Geo- for an Australia-Laurentia connection at 755 Ma. Precambrian
logical Society of America Bulletin, 93, 990–997. Research, 100, 335–357.
436 P. K. LINK & N. CHRISTIE-BLICK

Yonkee, W. A., Willis, G. C. & Doelling, H. H. 2000a. Proterozoic and Antelope Island, Utah. In: King, J. K. & Willis, G. C. (eds) The
Cambrian sedimentary and low-grade metasedimentary rocks on Geology of Antelope Island. Utah Geological Survey Publication,
Antelope Island, Utah. In: King, J. K. & Willis, G. C. (eds) The 00-1, 5– 36.
Geology of Antelope Island. Utah Geological Survey Publication, Young, G. M. 2002. Geochemical investigation of a Neopro-
00-1, 37– 47. terozoic glacial unit; the Mineral Fork Formation in the
Yonkee, W. A., Willis, G. C. & Doelling, H. H. 2000b. Petrology and Wasatch Range, Utah. Geological Society of America Bulletin, 114,
geologic history of the Precambrian Farmington Canyon Complex, 387– 399.
Chapter 39

The Edwardsburg Formation and related rocks, Windermere Supergroup,


central Idaho, USA

KAREN LUND1 *, JOHN N. ALEINIKOFF2 & KARL V. EVANS1


1
US Geological Survey, MS 973, Federal Center Box 25046, Denver, CO 80225, USA
2
US Geological Survey, MS 964, Federal Center Box 25046, Denver, CO 80225, USA
*Corresponding author (e-mail: klund@usgs.gov)

Abstract: In central Idaho, Neoproterozoic stratified rocks are engulfed by the Late Cretaceous Idaho batholith and by Eocene volcanic
and plutonic rocks of the Challis event. Studied sections in the Gospel Peaks and Big Creek areas of west-central Idaho are in roof pen-
dants of the Idaho batholith. A drill core section studied from near Challis, east-central Idaho, lies beneath the Challis Volcanic Group and
is not exposed at the surface. Metamorphic and deformational overprinting, as well as widespread dismembering by the younger igneous
rocks, conceals many primary details. Despite this, these rocks provide important links for regional correlations and have produced
critical geochronological data for two Neoproterozoic glacial periods in the North American Cordillera.
At the base of the section, the more than 700-m-thick Edwardsburg Formation (Fm.) contains interlayered diamictite and volcanic
rocks. There are two diamictite-bearing members in the Edwardsburg Fm. that are closely related in time. Each of the diamictites is
associated with intermediate composition tuff or flow rocks and the diamictites are separated by mafic volcanic rocks. SHRIMP U–
Pb dating indicates that the lower diamictite is about 685 + 7 Ma, whereas the upper diamictite is 684 + 4 Ma. The diamictite units
are part of a cycle of rocks from coarse clastic, to fine clastic, to carbonate rocks that, by correlation to better preserved sections, are
thought to record an older Cryogenian glacial to interglacial period in the northern US Cordillera.
The more than 75-m-thick diamictite of Daugherty Gulch is dated at 664 + 6 Ma. This unit is preserved only in drill core and the
palaeoenvironmental interpretation and local stratigraphic relations are non-unique. Thus, the date for this diamictite may provide a
date for a newly recognized glaciogenic horizon or may be a minimum age for the diamictite in the Edwardsburg Fm.
The c. 1000-m-thick Moores Lake Fm. is an amphibolite facies diamictite in which glacial features have not been observed. However,
it is part of a sedimentary cycle from unsorted siliclastic deposits to mud and carbonate deposits. Using lithostratigraphy and available
geochronology, the Moores Lake Fm. is correlated with a younger succession of Cryogenian glaciogenic rocks in southeastern Idaho.
Traditional correlations of Neoproterozoic rocks in the Cordillera recognize two levels of Cryogenian diamictites. The Edwardsburg
and Moores Lake diamictites along the middle Cordillera fit well into the scenario of two glacial events. Because of the correlations, dates
that provide ages for the diamictites in central Idaho (and corroborated in southeastern Idaho, Link & Fanning 2008) could constrain the
age of correlated glaciogenic deposits elsewhere in the Cordillera. However, in the absence of dates for the glaciogenic diamictites in
Canadian and southern US Cordilleran sections, the correlations are considered possible but uncertain.

Two belts of Neoproterozoic continental margin facies rocks are stratigraphic trends of the Cryogenian –Ediacaran Margin in the
recognized in the western North American Cordillera, one along northern US Cordillera (Lund et al. 2010) and produce important
the Canadian Cordillera (Hoffman & Halverson 2011; Smith geochronological data that potentially constrain the ages of other
et al. 2011) and another along the southern US Cordillera in the Cryogenian glaciogenic units.
Great Basin (Link & Christie-Blick 2011; Mrofka & Kennedy The best-preserved exposures of Neoproterozoic rocks in this
2011; Petterson et al. 2011) (Fig. 39.1). Within those belts, two region are in roof pendants within the Idaho batholith, in the
horizons of glaciogenic strata are widely correlated (Crittenden Gospel Peaks and near the seasonal community of Big Creek,
et al. 1972; Gabrielse 1972; Stewart 1972). Using the nomencla- west-central Idaho (Fig. 39.2). There are other smaller or less com-
ture for the more completely preserved succession in the Canadian plete exposures in east-central Idaho.
Cordillera, the preferred stratigraphic terminology for upper Cryo- In west-central Idaho, the first mapping of the Gospel Peaks roof
genian through Ediacaran continental margin strata (including gla- pendant resulted in a description of lithological map units and in
ciogenic rocks) of the western North American Cordillera is the interpretation that they were Neoproterozoic to early Palaeo-
‘Windermere Supergroup’ (Ross 1991; Link et al. 1993). In the zoic in age, but did not provide an interpretation of the sedimentary
well-known compilations of the Windermere Supergroup, there environment (Lund 1984). Early study of the section near Big
was a major gap in the belts of the northern US Cordillera, from Creek resulted in the description of volcanic rocks, diamictite
northeastern Washington to southeastern Idaho. However, and associated carbonate and quartzite units as local lithological
mapping and direct dating of volcanic rocks in a NW-striking units within thick Mesoproterozoic formations (Leonard 1962).
belt of metamorphosed Neoproterozoic rocks that cross central Further studies of the Big Creek section included geological
Idaho confirm that this ‘gap’ region contains thick, widely distrib- mapping and a successful dating study (Fig. 39.3). These studies
uted, but previously unrecognized exposures of the Windermere resulted in a description of regionally distributed formations,
Supergroup, including diamictite-bearing units within Edwards- interpretation of depositional settings, documentation of Neopro-
burg and Moores Lake (Fig. 39.1). The metamorphic grade and terozoic ages and correlations to other Cryogenian glaciogenic
(or) restricted exposure of some of the central Idaho Neoprotero- units in the Cordillera (Lund et al. 2003; Lund 2004).
zoic rocks limit sedimentological, isotopic, palaeoclimatic and In east-central Idaho, coarse lithic tuff (Daugherty Gulch,
stratigraphic interpretations. However, these studies, combined Fig. 39.2) was described from drill core and suggested to be
with limited dating and lithostratigraphic evaluations in the Neoproterozoic in age (Jacob 1990; Oliver & Blackwell 2002).
Canadian and Great Basin Cordillera segments, resulted in This drill hole is in a region dominated by a series of
continental-scale correlation of units (Lund et al. 2003; Lund fragmentary exposures of structurally juxtaposed, poorly dated,
2008). Further, these rocks contribute to the reconstruction of Neoproterozoic(?) to Middle Ordovician(?) miogeoclinal

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 437– 447. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.39
438 K. LUND ET AL.

Fig. 39.1. Map and stratigraphic column


showing Windermere Supergroup and
related rocks in central Idaho, as well as
distribution of equivalent rocks along the
North American Cordillera. Grey pattern,
layered rocks of Windermere Supergroup
or equivalent (see Lund et al. 2003 for
sources); black dots, Neoproterozoic
syenite-diorite suites in central Idaho (665–
650 Ma); black star, diamictite of
Daugherty Gulch (664 Ma).

successions (Hobbs & Hays 1990; Hobbs et al. 1991). Geologi- volcanic-plutonic rocks (Fisher et al. 1992; Evans & Green
cal and geochronological study confirmed that this is a Cryogen- 2003; Lund 2004). Because of this, the Windermere Supergroup
ian diamictite, providing another link between similar strata in is now exposed in roof pendants and thrust– fault disrupted sec-
central Idaho and additional data for correlations with Cryogen- tions that strike NW across central Idaho within regional-scale
ian glaciogenic units in the Cordillera (Lund et al. 2010). thrust-bounded panels (Fig. 39.2). Metamorphism and ductile
deformation overprint and disguise much of the direct evidence
needed for primary sedimentological and stratigraphic interpret-
Structural framework ations of strata as well as for reconstructions of the tectonic
setting of the Neoproterozoic depositional basin.
In central Idaho, rocks of the Windermere Supergroup were Rocks of the Windermere Supergroup underwent regional
involved in intense Cretaceous thrust faulting and folding prior metamorphism at upper greenschist to amphibolite facies con-
to, and during emplacement of, the engulfing Cretaceous Idaho ditions. In most places, layering is compositional due to a combi-
batholith and prior to cover by, and intrusion of, Eocene Challis nation of metamorphism and ductile deformation. However,
THE EDWARDSBURG FORMATION 439

116° 115° 114°

suture

45°30’ Gospel
Peaks
River

Mo o
Ida

nta
h
Salmon

na
Fig. 39.3 Salmon
Big Creek
45°

McCall
113o

Daugherty
Gulch

Challis

0 50 km
Quaternary-Miocene sedimentary and volcanic rocks Ordovician-Cambrian syenite-diorite suite
Eocene -Cretaceous intrusive and volcanic rocks Cryogenian syenite-diorite suite
Jurassic-Permian volcanic and sedimentary rocks Neoproterozoic volcanic and metasedimentary rocks
Palaeozoic sedimentary rocks Mesoproterozoic metasedimentary rocks

Fig. 39.2. Generalized geological map showing location of Neoproterozoic rocks in roof pendants, central Idaho. Roof pendants discussed in the text and the location for
Daugherty Gulch drill hole are shown.

primary sedimentary structures are preserved locally such that stra- east-central Idaho (west and east of about longitude 1158, respect-
tigraphic younging can be determined. The preserved younging ively, Fig. 39.2). The diamictite in the Moores Lake Fm. in west-
indicators allow reconstruction of stratigraphy and mesoscopic central Idaho, is part of the younger diamictite-bearing package.
structures. In the Gospel Peaks roof pendant, Windermere Super- The following descriptions are from Lund et al. (2003, 2010)
group rocks are repeated in two thrust sheets, one as part of an and Lund (2004).
upright section and the other in a map-scale (kilometre-scale wave- The basal unit of the Windermere Supergroup in central Idaho is
length) syncline. In the Big Creek roof pendant, graded volcani- the upper greenschist to lower amphibolite facies Edwardsburg
clastic sandstone beds indicate that the exposures are steeply Fm. (Fig. 39.1). Four lithological members are described in the for-
overturned, top-to-the-SW (Fig. 39.3). At Daugherty Gulch, the mation from the most complete exposure near Big Creek
drill hole intersected upright, moderately dipping rocks that are (Fig. 39.3). The three lower members are also exposed in two
at the toe of a thrust sheet not exposed at the surface because of thrust sheets in the Gospel Peaks roof pendant. At the base of the
younger normal fault reactivation of thrust faults and cover by formation, the Wind River Meadows Member is composed of dia-
younger units. mictite and local intertongued rhyodacite lithic tuff. The succeed-
Across central Idaho, igneous rocks are interlayered with or geo- ing Golden Cup Member is mafic volcanic and volcaniclastic
graphically associated with diamictite-bearing units. As well as rocks. The Placer Creek Member is composed of two distinct
providing datable rocks, these units provide information on the types of diamictite. Intertongued with, and at the top of the
tectonic setting of the local Windermere basin. Bimodal volcanic Placer Creek Member, is the Hogback Rhyolite. This extrusive
rock interlayered with diamictite in the Edwardsburg Fm. and rhyolite is locally preserved in the Big Creek roof pendant
with the diamictite of Daugherty Gulch, as well as a series of (Fig. 39.3).
bimodal alkalic plutonic suites, are aligned along the same NW A younger, cleaved, lower-greenschist facies volcaniclastic dia-
trend (Lund et al. 2010). These aligned alkalic igneous rocks mictite is present in a drill hole near Daugherty Gulch, 10 km west
provide evidence for significant, long-lived, intracontinental of Challis, east-central Idaho (Fig. 39.2). In the lower part of the
extension related to the NW-trending sedimentary basin into core, the 1143-m-deep drill hole intersected 75 m of diamictite
which the central Idaho Windermere rocks were deposited. without reaching the base of the unit. The upper part of the dia-
mictite is an unsorted mixture of quartz-rich matrix and
sedimentary-rock clasts. Below that, most of the unit is unsorted
Stratigraphy rock with volcanic and sedimentary clasts in a matrix of mixed
sedimentary and tuffaceous lithic grains and sericitized clay to
The stratigraphy of the Windermere Supergroup in central Idaho silt. The deepest part of the core contains rock with mostly tuffa-
(Fig. 39.1) is based on composite sections from several roof pen- ceous matrix. Because the units cut by the drill hole do not crop
dants and isolated exposures. The older units include two possibly out, total unit thickness is not known.
glaciogenic diamictite-bearing members in the Edwardsburg Fm. Above the volcanic and diamictite rocks of the Edwardsburg
of west-central Idaho and the diamictite of Daugherty Gulch in Fm., the Moores Station Fm. is composed of thick carbonaceous
440 K. LUND ET AL.

Fig. 39.3. Geological map of Big Creek roof pendant, central Idaho, showing locations of dated samples from Edwardsburg Fm., modified from Lund et al. (2003).
00KL040, location of dated rhyodacite tuff of Wind River Meadows Member; 97KE074, location of dated Hogback Rhyolite Member. Members (M) of the Edwardsburg
Fm. shown in inset: Hogback Rhyolite, Placer Creek, Golden Cup, Wind River Meadows.
THE EDWARDSBURG FORMATION 441

phyllite with lesser interlayered marble lenses. A section of coarse-grained amphibolite containing relict pyroxene (now plagi-
similar, but lower metamorphic grade, phyllitic carbonate and car- oclase and chlorite) and plagioclase phenocrysts (Leonard 1962)
bonaceous siltite was intersected in the drill hole above the volca- and fine-grained, indistinctly layered amphibolite containing horn-
niclastic diamictite of Daugherty Gulch. Because the upper units of blende, biotite, plagioclase and quartz. The massive amphibolite
the Daugherty Gulch drill core were incompletely retained and contains zones of stretched, calcite-filled cavities suggesting
archived, a reported upper unit has not been examined. This amygdaloidal texture. Possible fragmental and pillow textures
upper unit is described as grey, sandy and phyllitic dolostone are found in the coarser-grained porphyritic amphibolites.
(Jacob 1990). Below this dolostone lie two units, (i) dark grey, Above the mafic volcanic rocks of the Golden Cup Member are
pyrite-bearing, carbonaceous phyllite, and calcareous and dolomi- two diamictite units of the Placer Creek Member that in total are
tic marble and (ii) white, muscovite-bearing, medium-grained, c. 100 m thick. The lower diamictite contains a heterogeneous
dolomitic marble (Jacob 1990). These rocks were interpreted to mixture of intrabasinal mafic and felsic volcanic clasts, indicating
be equivalent to three nearby Cambrian (?) units (Jacob 1990). that the deposits are probably the result of local reworking of vol-
However, the Cambrian correlations are problematic because the canic flows. The upper diamictite contains a mixture of extrabasinal
rocks in the core are more cleaved, are of a slightly higher meta- quartzite and pebble-sized clasts of calc-silicate composition
morphic grade, and have different preserved bedding character- (similar in composition to underlying older sedimentary rocks),
istics than nearby exposed Cambrian rocks. At the time the hole together with intrabasinal mafic and rhyodacitic volcanic pebbles.
was drilled, there were no published descriptions of Neoprotero- The pebbles are stretched parallel to cleavage and most are
zoic carbonate or phyllite rocks in central Idaho but, based on c. 2 cm long.
evaluation of the Neoproterozoic units, correlation between The Hogback Rhyolite is an c. 200-m-thick local unit that is located
rocks in the upper part of the Daugherty Gulch core and the at the same stratigraphic level as diamictite of the Placer Creek
Moores Station Fm. is preferred (Fig. 39.1; Lund et al. 2010). Member elsewhere. The Hogback Rhyolite has relict potassium feld-
An upper level of possible glaciogenic and post-glacial rocks is spar, plagioclase and dipyramidal quartz phenocrysts; groundmass
preserved in the Moores Lake and Missouri Ridge Formations crystals of aligned biotite and muscovite define the metamorphic foli-
(Fig. 39.1). The Moores Lake Fm. is best preserved in the Gospel ation. Possible flow and cooling textures, including relict fiamme and
Peaks roof pendant where it comprises coarse quartzite cobble dia- pumice fragments as well as flow breccia and water-worked layers,
mictite with a micaceous schist and metasandstone matrix. The are preserved. These features indicate that, although exposure of
diamictite of the Moores Lake Fm. is succeeded by the Missouri this rhyolite is restricted, it is an extrusive flow unit that is part of
Ridge Fm., a fine-grained siliciclastic and carbonate unit. the stratigraphic section rather than a younger plug or sill.

Volcaniclastic diamictite of Daugherty Gulch


Glaciogenic deposits and associated strata
Diamictite in the Daugherty Gulch drill hole is of two types charac-
Edwardsburg Fm.
terized by lower greenschist facies and prominent cleavage. The
top 2.4 m (8 ft) of this unit in the drill core is light greenish, chlori-
The Edwardsburg Fm. is estimated to be c. 700 m thick on the basis
tic, unsorted siltite to metasandstone matrix with grit to pebble-
of mapped exposures, but aggregate thickness of its members is as
sized clasts that are dominantly quartzite and metasandstone.
much as 1200 m thick.
Below this, the drill hole cut 72.5 m (181 ft) of cleaved, hetero-
At the base of the formation (Fig. 39.1), the Wind River
geneous-clast volcaniclastic diamictite to the bottom of the hole.
Meadows Member is about 200 m thick. The lower part of the
The matrix of this lower predominant diamictite comprises
Wind River Meadows Member is composed of minor fine-grained
unsorted material in the clay to sand size fractions. The matrix con-
quartzite and thin mafic volcanic layers. The upper part of the
tains 20–30% fractured quartz grains, some strained. Another,
member is composed of diamictite that is laterally gradational
about 25%, is lithic fragments, including quartz siltites, volcanic
with rhyodacite flows and volcaniclastic conglomerate.
chips and carbonate grains. The rest of the matrix is a mixture of
The diamictite in the Wind River Meadows Member is massive
fine-grained sericite, quartz, opaque grains and indistinguishable
to weakly layered metagreywacke with cobbles and pebbles of
sericitic and carbonate altered material that originated as a tuffac-
quartzite and calc-silicate compositions. The clasts are mostly
eous fraction. Towards the bottom of the core, the matrix becomes
,3 cm long, but can be as much as 20 cm long, sub-angular to
more tuffaceous, although there are few relict feldspar or mafic
rounded, and stretched parallel to the overprinting cleavage or
minerals. Clasts make up c. 50% of the rock and are as large as
foliation. As measured from map exposure, the diamictite is as
25 cm in diameter. These are dominantly rhyolite, lesser quartzite
much as 200 m thick. The unit is massive and there is no clear
and metasandstone (but these form the large clasts), minor carbon-
remnant compositional layering. Clasts in the diamictite are
aceous phyllite and dolostone, and a few plutonic clasts. The pro-
similar to the underlying Neoproterozoic(?) and Mesoproterozoic
portion of volcanic to sedimentary clasts increases downward in the
metasedimentary rocks.
core. Volcanic clasts are light yellow, pink, green and grey. Relict
The diamictite of the Wind River Meadows Member may inter-
volcanic textures, such as fiamme, are common in the clasts. Meta-
tongue with, or be regionally gradational into, rhyodacite flow
sandstone clasts are dominantly pink (although some are buff ), and
breccia, tuff, graded volcaniclastic sandstone and volcaniclastic
well rounded. Quartzite pebbles are similar to the known Mesopro-
conglomerate of heterogeneous composition. Where present, rhyo-
terozoic Gunsight and Swauger formations exposed in the vicinity
dacite flow and flow-breccia rocks locally dominate the base of the
of Daugherty Gulch. The less abundant phyllite and dolostone
Wind River Meadows Member. Volcaniclastic conglomerate con-
pebbles did not originate from the Mesoproterozoic formations of
tains a heterogeneous mixture of rhyodacite clasts and minor quart-
the region but, rather, are from other Neoproterozoic units. Volca-
zite clasts in water-worked tuffaceous matrix. The composition of
nic clasts are markedly elongated and aligned parallel to cleavage.
the clasts indicates that both the local rhyodacite and older quart-
They are commonly plastically deformed where in contact with
zite units were being eroded and redeposited. Graded volcaniclas-
rounded but not stretched quartzitic clasts.
tic sandstone beds associated with the volcaniclastic conglomerate
provide younging indicators, suggesting that the exposures NW of
Big Creek are steeply overturned with top to the SW (Fig. 39.3). Moores Station Fm.
Metamorphosed mafic volcanic rocks of the Golden Cup
Member form the middle of the Edwardsburg Fm. and are c. The Moores Station Fm. overlies the Edwardsburg Fm. across
400 m thick. The mafic volcanic rocks are massive medium- to west-central Idaho (Fig. 39.1). It is composed of two dominant
442 K. LUND ET AL.

rock types with a total thickness of c. 1000 m. The major rock type At the top of the Edwardsburg Fm., the coarse unsorted debris of
is dark grey to black (weathering to orange), fine-grained, graphi- the diamictite abruptly changes to the fine grained carbonaceous
tic, pyrite- and apatite-rich siliceous and calcareous phyllite. The phyllite and carbonate rocks of the Moores Station Fm. A
phyllite is interlayered with grey, calc-silicate bearing calcitic similar thick section of pyritic, carbonaceous phyllite and carbon-
and dolomitic marble lenses that, although probably both ate rocks above the Daugherty Gulch diamictite and is believed to
thickened and thinned by deformation, are as much as 30 m be correlative to the Moores Station Fm.
thick. Although fine-scale layering is readily observed, bedding Local preservation of the underlying fine-grained metasand-
character is not preserved because of the metamorphic mineral stone of the Goldman Cut Fm. suggests that the Moores Lake
growth and deformation. Fm. overlies an angular unconformity. The upper contact of the
In the Daugherty Gulch drill core of east-central Idaho, there is Moores Lake Fm. with the overlying Missouri Ridge Fm. is a
a 1082-m-thick section of lower greenschist facies, interlayered narrow gradational zone of calc-silicate-dominant metasiltites.
with carbonate and carbonaceous phyllite that is also considered The 450-m-thick Missouri Ridge Fm. is a ribbon-laminated calc-
part of the Moores Station Fm. (Lund et al. 2010). The lowest silicate gneiss and marble section. This succession from a diamic-
part of this section includes 59.7 m of white, muscovite-bearing, tite unit to a carbonate-silicate is similar to the cycle of Edwards-
medium-grained, dolomitic marble. Above this marble, 873.7 m burg Fm. diamictites to the Moores Station carbonate rocks except
of cleaved, dark grey, pyrite-bearing, carbonaceous phyllite is that the Missouri Ridge carbonate rocks are not as carbonaceous
gradationally interlayered on the centimetre scale with lighter and pyritic as the lower-cycle Moores Station carbonate rocks.
grey, calcareous and dolomitic phyllitic marble. Soft-sediment
bedding features are preserved, including load structures that
indicate the section is upright. The phyllite is folded and some Chemostratigraphy
folds are preserved in the core. Spaced cleavage is defined by
sericite foliation and is generally about 508 steeper than The metamorphic overprint and amount of regional magmatic and
bedding. An upper 148.3 m of sandy, phyllitic dolostone was hydrothermal activity probably make the central Idaho sections
not examined for this study (because this part of the core was inappropriate for chemostratigraphic analyses. However, C-
not archived). isotopic constraints available for the correlative Pocatello Fm.,
southeastern Idaho (Smith et al. 1994; Lorentz et al. 2004), bear
on interpretations for similar aged central Idaho units (see discus-
Moores Lake Fm. sion). A 667–686 Ma thin carbonate layer, lying directly above
diamictite in the upper Scout Mountain Member of the Pocatello
The Moores Lake Fm. is an c. 1000-m-thick unit that includes an Fm., yields negative d13C values and is interpreted as a cap carbon-
upper quartzite and a lower diamictite. Lenses of quartzite-cobble ate (Corsetti et al. 2007). Carbonate rocks associated with a
diamictite with cobbles up to 15 cm across occur near the base of 667 Ma tuff and a sandstone –shale succession at the top of the
the formation. The diamictite matrix is muscovite-feldspar-quartz Scout Mountain Member also have negative d13C values similar
metasandstone and biotite-muscovite-felspar-quartz schist. The to values from cap-carbonate rocks overlying Neoproterozoic
diamictite is massive and there is also no apparent compositional glacial deposits (Lorenz et al. 2004). However, this upper carbon-
layering. Upper parts of the Moores Lake Fm. are more than ate does not directly overlie rocks interpreted as glaciogenic and
95% recrystallized quartz; minor feldspar and mica make up the the negative d13C values have been interpreted as the possible
rest of the rock. The upper part of the unit is massive and result of overturn of an anoxic ocean rather than the result of post-
bedding is difficult to identify. glacial processes (Lorenz et al. 2004).

Boundary relations with overlying and underlying Other characteristics


non-glacial units
Limited whole-rock geochemical data are available for volcanic
The Edwardsburg Fm. is underlain by three formations of undeter- rocks interlayered with diamictite in the Edwardsburg Fm. (Lund
mined age within the late Mesoproterozoic to early Neoproterozoic et al. 2003, 2010). Although interpretation of this whole-rock
period (Fig. 39.1). These units are staurolite-garnet schist, felds- data is not ideal because of the metamorphic and hydrothermal
pathic quartzite, and calc-silicate gneiss to marble, from oldest overprints, trace-element data for alternative analyses in these
to youngest. The clasts in the diamictite of the Wind River rocks are incomplete (Lund et al. 2010). The metatuffs of the
Meadows Member of the Edwarsdsburg Fm. probably originated Edwardsburg Fm. are marginally alkalic to alkali-calcic and
from these underlying older metamorphic rocks. This suggests range in composition from dacite to trachydacite (as plotted on
that the subjacent older units were exposed to erosion prior to IUGS system, Le Bas & Strecheisen 1991). SiO2 concentrations
deposition of the Edwardsburg Fm. Local, low-angle discordance average about 65%. Two of the samples are Ti-rich (average
between the Edwardsburg Fm. and underlying rocks also suggest TiO2 of 1.3%), which is in agreement with the alkalic to alkali-
that the base of the Edwardsburg Fm. is an unconformity. At the calcic classification. Samples are mildly peraluminous, sodic to
top of the Edwardsburg Fm., the Placer Creek Fm. and Hogback very sodic (average 4.6%), have average to rich Fe concentrations
Rhyolite are variably preserved, suggesting an erosional top to (average 6.8%). The mafic volcanic rocks of the Edwardsburg
the formation or lateral facies variation. Fm. are basaltic (by IUGS system of Le Bas & Strecheisen
The base of the diamictite of Daugherty Gulch was not reached 1991; or picritic based on De la Roche et al. 1980), metaluminous,
by drilling so there is no information about the complete thickness average to potassic (average 0.35%), and have average to rich Mg
of the unit or its relations to underlying rocks. The heterogeneous concentrations (average 12.0%). High TiO2 concentrations
clasts of the diamictite, which includes volcanic, metasandstone, (average of 2.4%) support an alkalic parentage from a basaltic to
and carbonaceous and carbonate rocks, indicate that a varied rock- ultramafic source. The samples range from strongly to marginally
type source terrain of several ages was exposed and eroded to form undersaturated in silica. The Cryogenian mafic volcanic rocks in
this unit. The upper contact with the overlying white dolostone was the correlative Pocatello Fm. of southeastern Idaho (Harper &
described as fractured but conformable (Jacob 1990). However, Link 1986) and the broadly correlative Huckleberry Fm. of north-
without exposures that can be mapped in three dimensions, eastern Washington and southern British Columbia (Devlin et al.
interpretations of an erosional unconformity or a thrust fault for 1988) are chemically similar to these metabasalts of the Edwards-
the upper contact cannot be ruled out. burg Fm.
THE EDWARDSBURG FORMATION 443

Palaeolatitude and palaeogeography (a) 800

206Pb/238U age (Ma)


0.080
700
The metamorphism and multiple deformation events that have
1100
overprinted the central Idaho Windermere Supergroup rocks 0.076 600

have likely reset any palaeomagnetic signatures. As a result,

207Pb/206Pb
500
1000
direct evidence of palaeolatitude cannot be determined from 0.072 Age = 687 ± 10 Ma
(MSWD = 2.7)
central Idaho rocks. A palaeopole for Laurentia was developed
0.068 900
using data from 723 Ma sills thought to be correlative with
Rapitan Group diamictite in the Mackenzie Mountains, northern Concordia Age = 685 ± 7 Ma
800
0.064 prob. of concordance = 0.93
Canadian Cordillera (Park 1997). From these data, a low palaeola- n=11 (of 17)
titude was determined for Laurentia at 723 Ma, the presumed age 00KL040
0.060
of the Rapitan Group diamictites and of the traditionally correlated rhyodacite, Wind River
500
Cryogenian diamictite units and associated mafic volcanic rocks of Meadows Member
0.056
the Cordillera. However, the ages of, and possible correlations for, 4 6 8 10 12
the central Idaho diamictites (Fig. 39.4) put this interpretation in 238U/206Pb
question (Lund et al. 2003).
(b)

Pb/238U age (Ma)


0.080 700
Geochronological constraints
1100 660
0.076
U –Pb Sensitive High Resolution Ion Microprobe (SHRIMP)

207Pb/206Pb
620
zircon dates were obtained from volcanic members of the 1000

206
0.072 Age = 682 ± 6 Ma
Edwardsburg Fm. An additional zircon date was acquired from tuf- (MSWD = 0.97)
faceous matrix near the base of the diamictite at Daugherty Gulch. 0.068 900
A sample from the base of the rhyodacite flows in the Wind Concordia Age = 684 ± 4 Ma
800
River Meadows Member was dated by Lund et al. (2003) from a 0.064 prob. of concordance = 0.37
n=15 (of 15)
locality about 2 km north of Big Creek (location 00KL040,
97KE074
Fig. 39.3). Zircons from this rhyodacite contain low concentrations 0.060
Hogback Rhyolite 600
of U (,49 –606 ppm, mostly ,200 ppm). The weighted average 500
of the 206Pb– 238U ages, using 14 of 17 analyses, is 687 + 10 Ma 0.056
4 6 8 10 12
(Fig. 39.4a). The data are somewhat scattered, as indicated by 238U/206Pb
the mean square of weighted deviates (MSWD) of 2.7. One analy-
sis provides a significantly younger age (595 + 24 Ma), and two
analyses yield older ages (742 + 20 and 1181 + 28 Ma). A (c) 0.085

Pb/238U age (Ma)


720

Concordia Age calculated from 11 (of 17) analyses is 680


0.075
685 + 7 Ma (probability of concordance ¼ 0.93).
640
A sample from near the middle of the Hogback Rhyolite was
207Pb/206Pb

dated by Lund et al. (2003) from a locality about 0.5 km east of 0.065
600
Age = 664 ± 6 Ma
206
Big Creek (location 97KE074, Fig. 39.3). Zircons from the 750 (MSWD = 0.75)
Hogback Rhyolite contain relatively high U concentrations 550
0.055
(,327–1394 ppm) and have isotopic ratios that form a single 450
population on a concordia plot (Fig. 39.4b). The weighted
average of the 206Pb – 238U ages from 15 analyses is 682 + 6 Ma 0.045 DG1 Concordia Age = 667 ± 6 Ma
(MSWD ¼ 0.97). Alternatively, the Concordia Age calculated tuffaceous matrix prob. of concordance = 0.78
from 15 (of 15) analyses is 684 + 4 Ma (probability of n=15 (of 16)
daimictite of Daugherty Gulch
concordance ¼ 0.37). Thus, the two volcanic samples from the 0.035
7 9 11 13 15
Edwardsburg Fm. are about the same age (within uncertainty):
(i) Wind River Meadows Member, 685 + 7 Ma and (ii) 238U/206Pb
Hogback Rhyolite Member, 684 + 4 Ma (Lund et al. 2003).
Fig. 39.4. Terra-Wasserburg plots of SHRIMP U–Pb data for Cryogenian
Tuffaceous matrix from the 1140.3 m (3741 ft) depth in the
volcanic rocks in central Idaho. (a) Rhyodacite, Wind River Meadows Member,
Daugherty Gulch core (c. 55.2 m below the top of the diamictite
sample 00KL040. (b) Rhyolite, Placer Creek Member, Edwardsburg Fm.,
and 1.5 m above the bottom of the hole) was dated by the sample 97KE074. (c) Tuffaceous matrix, diamictite of Daugherty Gulch,
SHRIMP U –Pb method (Lund et al. 2010). Using weighted sample DG1. Locations for (a) and (b) in Figure 39.3. Location for (c) in
averages of selected individual 206Pb– 238U ages, SHRIMP U –Pb Figure 39.2.
isotopic data from 15 grains form a coherent grouping with an
age of 664 + 6 Ma (Fig. 39.4c). A Concordia Age calculated
from 15 (of 16) analyses is 667 + 6 Ma (probability of
Clast compositions in the diamictite units provide some evi-
concordance ¼ 0.78). This age is approximately the same as ages
dence of the provenance of the rocks. Many clasts are generally
for syenite-diorite plutonic suites that crop out across central
recognizable as locally derived from underlying sedimentary
Idaho, suggesting a regional magmatic event (Lund et al. 2010).
rocks. In the Placer Lake Member and the diamictite of Daugherty
Gulch, clasts of sedimentary rock are mixed with clasts of contem-
Discussion poraneous or slightly older volcanic rocks and closely associated
rip-up sedimentary clasts. These relations suggest that Neoproter-
Depositional setting ozoic and Mesoproterozoic rocks were locally exposed to erosion
and then included as clasts.
Because metamorphism and deformation overprinted original Direct indicators of glacial sedimentation, such as glacial facet-
sedimentary structures, palaeoenvironmental and palaeoclimatic ing of clasts, striated clasts or bedding compaction below drop-
interpretations for the central Idaho Windermere rocks are not stones, are not recognized in the diamictites of the Edwardsburg
straightforward. Fm., Daugherty Gulch drill core, or Moores Lake Fm. Possible
444 K. LUND ET AL.

origins for such poorly sorted deposits include local debris flows, (667 + 5 Ma, Fanning & Link 2008). The Scout Mountain tuff
fault-scarp debris or glaciogenic conditions (see discussion in is interpreted to provide a minimum age for underlying Scout
Miller 1985). Using the available data from Cryogenian rocks of Mountain glaciogenic rocks that are thought to be correlative
central Idaho, a glaciogenic origin is a permissible, yet uncertain with the Edwardsburg diamictites (Fanning & Link 2004, 2008).
interpretation. The palaeoenvironmental significance of these Because this Scout Mountain tuff is not associated with recognized
deposits is primarily based on the stratigraphic patterns of diamic- glaciogenic rocks (it is stratigraphically separated from the two
tite succeeded by thick, fine-grained carbonaceous siliciclastic and underlying glaciogenic diamictites), a directly overlying carbonate
carbonate units in the central Idaho sections. This stratigraphic unit with negative d13C values was interpreted in the context of
pattern is similar to stratigraphic successions described from oceanic processes rather than as a ‘cap’ carbonate related to a Neo-
better preserved, Cordilleran glaciogenic rocks to the north and proterozoic glaciation (Lorentz et al. 2004). The correlative ages
south (Figs 39.1 & 39.5). of the Daugherty Gulch diamictite and this Scout Mountain tuff
Textures (at microscopic to outcrop scales) in the volcanic rocks presents the possibility that the negative d13C values were
of the Edwardsburg Fm. provide evidence that these rocks are related to a 665 Ma glacial event or alternatively that the Daugh-
extrusive and interbedded with diamictites and metasandstones. erty Gulch diamictite is non-glacial. If the former is correct, the
In the case of the diamictite of Daugherty Gulch, where meta- diamictite of Daugherty Gulch presents a potentially different,
morphic grade is low, evidence of tuffaceous matrix is preserved. younger (c. 665 Ma), correlation point for undated glaciogenic
These data indicate that the diamictite was deposited synchro- diamictites that underlie thick carbonaceous shale and carbonate
nously with local volcanism and that dates on the volcanic rocks successions along the Cordillera (Fig. 39.5).
also provide ages for formation of associated diamictite units The thick carbonaceous shale and carbonate rocks of the Moores
(Fig. 39.4). Station Fm. which lie above the Edwardsburg Fm. and Daugherty
Gulch diamictites are important for reconstructing the Windermere
Supergroup stratigraphy of central Idaho because their special
Geochronological constraints layering and composition characteristics are not seen in older or
younger rocks in the region. For the same reasons, these rocks
The ages for the two Edwardsburg Fm. diamictites are 685+7 to serve as units that can be correlated along the Cordillera
684 + 4 Ma and, given the over lapping age constraints, may rep- (Fig. 39.5). Elsewhere, carbonate rocks at the base of these sections
resent two closely spaced events or just a single event. The age for are interpreted as cap carbonates and the section is interpreted
volcanism synchronous with diamictite at Daugherty Gulch is as the result of deepening water levels in post-glacial times (e.g.
664 + 4 Ma, but because of the lack of exposure, the exact relation Fanning & Link 2004).
between this diamictite and the older Edwardsburg Fm. diamictites Correlation between the Moores Lake Fm. in central Idaho
are undetermined. and the Caddy Canyon Quartzite in southeastern Idaho is based
on lithostratigraphic similarity and context (Fig. 39.5). The ages
of those deposits are only loosely constrained by dates between
Correlations along the North American Cordillera 685 and 665 Ma on the underlying volcanic rocks (Lund et al.
2003; Fanning & Link 2004, 2008) and 570 and 580 Ma on over-
Regional corrrelations based on lithostratigraphic similarity and lying volcanic rocks (Levy & Christie-Blick 1991; Link et al.
geochronological data are shown in Figure 39.5. Diamictites in 1993; Levy et al. 1994; Christie-Blick 1997). The Caddy
the upper part of the Scout Mountain Member, Pocatello Fm., Canyon Fm. is thought to represent late Cryogenian glaciogenic
southeastern Idaho, are bracketed between 701+4 (Bannock rocks on the basis of these geochronological data and the preserved
Volcanic Member clast in diamictite) and 667 + 5 Ma (tuff over- incised and refilled valleys that are interpreted as a product of
lying diamictite) in one section and determined to be younger than sea-level fall during a glacial event (Link et al. 1993; Levy et al.
686 + 4 Ma in another section (Fanning & Link 2004, 2008). 1994). Because of this interpretation, Moores Lake and Caddy
Thus, overlapping SHRIMP U –Pb ages of about 685 Ma now Canyon rocks can in turn be correlated with the late Cryogenian
provide chronostratigraphic correlations between Edwardsburg glacial interval in the northern Canadian Cordillera (Icebrook
Fm. diamictite units in central Idaho and upper Scout Mountain Fm., Fig. 39.5; Link et al. 1993; Lund et al. 2003; Fanning &
Member diamictite units in southeastern Idaho (Lund et al. Link 2008). However, without geochronological ties, these corre-
2003; Fanning & Link 2008). lations (especially for the Moores Lake Fm.) are possible, but
Diamictites and underlying mafic volcanic rocks in the Scout uncertain.
Mountain Member are correlated traditionally with other diamic-
tites associated with mafic volcanic rocks along the entire Cordil-
lera (Fig. 39.5). However, only the mafic volcanics in central Idaho Broader implications of Central Idaho geological record
are bracketed closely by dated units (685+7 to 684 + 4 Ma). The
age of the Bannock Volcanic Member in southeastern Idaho is The dates of c. 685 Ma for interlayered volcanic and diamictite
interpreted based on dates of 701+4 and 717 + 4 Ma from volca- rocks in the Edwardsburg Fm. provide the first direct dates for Cor-
nic clasts in an overlying diamictite, rather than based on direct dilleran diamictite-bearing units of Cryogenian age. Previously,
radiometric dating of the volcanic unit itself (Fanning & Link the older Cryogenian glaciogenic rocks of the Windermere Super-
2004, 2008). The generally accepted correlations between diamic- group were thought to be c. 760– 700 Ma on the basis of the few
tites from the Rapitan Group in the Mackenzie Mountains and available dates from underlying basement, intrusions that cut
other diamictite-bearing units along the Cordillera down to the older rocks, and inconclusively correlated rocks (see discussion
Great Basin (Figs 39.1 & 39.5) are permissive within the current in Lund et al. 2003). Although the Cordilleran Cryogenian glacio-
geochronological database. However, as the ages of the other Cor- genic deposits were thought to have formed during an older Cryo-
dilleran mafic volcanic and diamictite units are poorly constrained, genian (‘Sturtian’) glaciation, the recent dates from the
the diamictites in the Canadian Cordillera (Hoffman & Halverson Edwardsburg and Pocatello formations indicate that (much or all
2011; Smith et al. 2011) and SW USA (Mrofka & Kennedy 2011; of) the glaciation in Idaho occurred later than the commonly
Petterson et al. 2011) could be either older or younger than those in cited ages for a global, older Cryogenian glacial event. In the
the Edwardsburg Fm. absence of direct dating in many of the Cordilleran sections and
The diamictite of Daugherty Gulch (664 + 4 Ma, Figs 39.4 & the limited sedimentary evidence in central Idaho, it cannot be
39.5) is about the same age as a thin tuff in post-glacial sandstone- determined how widespread this event was. However, it is possible
shale rocks near the top of the Scout Mountain Member that the geochronological data for the Edwardsburg Fm. constrains
THE EDWARDSBURG FORMATION
Fig. 39.5. Correlation chart for Windermere Supergroup strata in central Idaho and the nearby Cordillera based on lithostratigraphic and geochronological data. Locations of strata are shown in Figure 39.1. Ages are
U– Pb zircon, except where otherwise noted. Diagram is constructed to tie rock units to geological time, but lack of dates throughout Cordillera results in much interpretation for duration of units and correlations. Cgl,
Conglomerate; Cyn, Canyon; Cr, Creek; Fm, Formation; Grp, Group; Lk, Lake; M, Member; Mdws, Meadows; Mtn, Mountain; Qtzt, Quartzite; Rhy, Rhyolite; Rv, River; Volc, Volcanic. 1Aitken (1991), Rainbird et al. (1996)
Ross et al. (1995), Kendall et al. (2004); 2 Devlin & Bond (1988); Hein & McMechan (1994); 3Devlin & Bond (1988), Devlin (1989), Hein & McMechan (1994), Colpron et al. (2002); 4Evans (1987), Miller & Whipple

445
(1989), Lindsey et al. (1990), Miller (1994); 5Lund et al. (2003); 6McCandless (1982), Evans & Green (2003), Lund et al. (2010); 7Crittenden et al. (1971, 1983), Link et al. (1993), Christie-Blick (1997); 8Crittenden et al. (1971,
1983), Link et al. (1993).
446 K. LUND ET AL.

not only the age of diamictite-bearing rocks in Idaho but also those Utah and southeastern Idaho. Geological Society of America Bulletin,
of other diamictite units along the Cordillera. 94, 437– 450.
Revising the age of the diamictite-bearing deposits in the north- Devlin, W. J. 1989. Stratigraphy and sedimentology of the Hamill Group
ern US Cordillera from 750– 700 Ma to c. 685– 665 Ma may affect in the northern Selkirk Mountains, British Columbia: evidence for
palaeolatitude reconstructions and global climate change models. latest Proterozoic – Early Cambrian extensional tectonism. Canadian
Park (1997) used 723 Ma sills thought to be correlative with the Journal of Earth Sciences, 26, 515–533.
glaciogenic Rapitan Group in the northern Canadian Cordillera Devlin, W. J. & Bond, G. C. 1988. The initiation of the early Paleozoic
to develop the palaeopole for Laurentia at 723 Ma. From these Cordilleran miogeocline: evidence from the uppermost Proterozoic –
Lower Cambrian Hamill Group of southeastern British Columbia.
data, he interpreted a low palaeolatitude for this older Cryogenian
Canadian Journal of Earth Sciences, 25, 1 –19.
glaciation (see discussion in Lund et al. 2003). However, direct age Evans, J. G. 1987. Geology of the Stensgar Mountain quadrangle, Stevens
constraints for the Cryogenian deposits in Idaho (Lund et al. 2003; County, Washington. US Geological Survey Bulletin, 1679, 23.
Fanning & Link 2004, 2008) indicate that the 723 Ma palaeopole Evans, K. V. & Green, G. N. 2003. Geologic Map of the Salmon National
for Laurentia would be too old for time of formation of the Forest and vicinity, east-central Idaho. US Geological Survey Geolo-
Idaho diamictites and thus would not provide an accurate palaeo- gic Investigations, I-2765.
latitude for that glaciation. Additionally, the palaeogeography for Fanning, C. M. & Link, P. K. 2004. U–Pb SHRIMP ages of Neopro-
the glaciogenic deposits needs to be reinterpreted due to the differ- terozoic (Sturtian) glaciogenic Pocatello Formation, southeastern
ence in supercontinent configurations at 750 Ma compared to Idaho. Geology, 32, 881– 884.
685 Ma. Ultimately, there is a broad range of possible ages for dia- Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian gla-
mictites in the Cordillera. Constraints are best for diamictites of ciation; data from the Adelaide geosyncline, South Australia and
685 to 665 Ma, but some less direct constraints may extend the Pocatello Formation, Idaho, USA. Selwyn Symposium 2008, Neo-
maximum age to c. 720 Ma. As such, models for Neoproterozoic proterozoic climates and origin of early life. Geological Society of
palaeogeography, isotopic composition of seawater curves, and Australia, Abstracts, 91, 57 – 62.
global glaciations (Snowball Earth events) that are partly based Fisher, F. S., McIntyre, D. H. & Johnson, K. M. 1992. Geologic Map of
on Cordilleran data must be used with care. the Challis 1oX2o quadrangle, Idaho. US Geological Survey Miscel-
laneous Investigation Series Map I-1819.
An early version of this manuscript was improved by comments from D. I. Gabrielse, H. 1972. Younger Precambrian of the Canadian Cordillera.
Bleiwas. Comments and suggestions by reviewers P. K. Link and M. Colpron American Journal of Science, 272, 521–536.
and suggestions and guidance from volume editor E. Arnaud are greatly appreci- Harper, G. D. & Link, P. K. 1986. Geochemistry of Upper Proterozoic
ated. This represents a contribution of the IUGS- and UNESCO-funded IGCP rift-related volcanics, northern Utah and southeastern Idaho.
(International Geoscience Programme) project #512. Geology, 14, 864–867.
Hein, F. J. & McMechan, M. E. 1994. Proterozoic– Lower Cambrian
strata of the western Canada sedimentary basin. In: Mossop, G. D.
& Shetsen, I. (eds) Geological Atlas of the Western Canada Sedi-
References mentary Basin. Canadian Society of Petroleum Geologists and
Alberta Research Council, 47 –57.
Aitken, J. D. 1991. The Ice Brook Formation and post-Rapitan, Late Pro- Hobbs, S. W. & Hays, W. H. 1990. Ordovician and older rocks of the
terozoic glaciation, Mackenzie Mountains, Northwest Territories. Bayhorse area, Custer County, Idaho. US Geological Survey Bulletin,
Geological Survey of Canada Bulletin 404, 43. 1891, 40.
Christie-Blick, N. 1997. Neoproterozoic sedimentation and tectonics in Hobbs, S. W., Hays, W. H. & McIntyre, D. H. 1991. Geologic Map of the
west-central Utah. In: Link, P. K. & Kowallis, B. J. (eds) Brigham Bayhorse area, central Custer County, Idaho. US Geological Survey
Young University Geology Studies. Salt Lake City, Utah, 1 –30. Miscellaneous Geologic Investigations Map I-1882.
Christie-Blick, N. & Levy, M. 1989. Stratigraphic and tectonic frame- Hoffman, P. F. & Halverson, G. P. 2011. Neoproterozoic glacial record
work of Upper Proterozoic and Cambrian rocks in the western in the Mackenzie Mountains, northern Canadian Cordillera. In:
United States. In: Christie-Blick, N. & Levy, M. (eds) Late Proter- Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
ozoic and Cambrian Tectonics, Sedimentation, and Record of logical Record of Neoproterozoic Glaciations. Geological Society,
Metazoan Radiation in the Western United States. American London, Memoirs, 36, 397–411.
Geophysical Union, Washington, DC, 28th International Geological Jacob, T. 1990. Late Proterozoic tuff near Challis, Idaho. In: Moye, F. J.
Congress, Field Trip Guidebook T331, 7 –21. (ed.) Geology and Ore Deposits of the Trans-Challis Fault System/
Colpron, M., Logan, J. M. & Mortensen, J. K. 2002. U –Pb zircon age Great Falls Tectonic Zone. Guidebook of the Fifteenth Annual
constraint for late Neoproterozoic rifting and initiation of the lower Tobacco Root Geological Society field Conference, 97 – 106.
Paleozoic passive margin of western Laurentia. Canadian Journal Kendall, B. S., Creaser, R. A., Ross, G. M. & Selby, D. 2004. Con-
of Earth Sciences, 39, 133– 143. straints on the timing of Marinoan ‘Snowball Earth’ glaciation by
187
Corsetti, F. A., Link, P. K. & Lorentz, N. J. 2007. d13C chemostratigra- Re – 187Os dating of a Neoproterozoic, post-glacial black shale in
phy of the Neoproterozoic succession near Pocatello, Idaho, USA: western Canada. Earth and Planetary Science Letters, 222, 729– 740.
implications for glacial chronology and regional corellations. In: Leonard, B. F. 1962. Old metavolcanic rocks of the Big Creek area,
Link, P. K. & Lewis, R. S. (eds) Proterozoic Geology of Western central Idaho, in Short papers in geology, hydrology and topography.
North America and Siberia, Belt Symposium IV, Society of Econ- US Geological Survey Professional Paper 450B, B11 –B15.
omic Paleontologists and Mineralogists, Special Publications, 86, Levy, M. & Christie-Blick, N. 1991. Late Proterozoic paleogeography
193– 205. of the eastern Great Basin. In: Cooper, J. D. & Stevens, C. H.
Crittenden, M. D. Jr, Schaefer, F. E., Trimble, D. E. & Woodward, (eds) Paleozoic Paleogeography of the Western United States – II.
L. A. 1971. Nomenclature and correlation of some Upper Precam- Pacific Section, Society of Economic Paleontologists and Mineralo-
brian and basal Cambrian sequences in western Utah and southeastern gists, Los Angeles, California, 371– 386.
Idaho. Geological Society of America Bulletin, 82, 581– 602. Levy, M., Christie-Blick, N. & Link, P. K. 1994. Neoproterozoic
Crittenden, M. D. Jr, Stewart, J. H. & Wallace, C. A. 1972. Regional incised valleys of the eastern Great Basin, Utah and Idaho: fluvial
correlation of Upper Precambrian strata in western North America. response to changes in depositional base level. In: Dalrymple,
24th International Geological Congress, 1972, 1, 334–341. R. W., Boyd, R. & Zaitlin, B. A. (eds) Incised-Valley Systems:
Crittenden, M. D. Jr & Wallace, C. A. 1973. Possible equivalents of Origin and Sedimentary Sequences. Society for Sedimentary
the Belt Supergroup in Utah. Belt Symposium 1973 Volume 1, Geology, Tulsa, OK, 369–382.
Idaho Bureau of Mines and Geology, Moscow, Idaho, 116– 138. Lindsey, K. A., Gaylord, D. R. & Groffman, L. R. 1990. Geology of the
Crittenden, M. D. Jr, Christie-Blick, N. & Link, P. K. 1983. Evidence Upper Proterozoic to Lower Cambrian Three Sisters Formation,
for two pulses of glaciation during the late Proterozoic in northern Gypsy Quartzite, and Addy Quartzite, Stevens and Pend Oreille
THE EDWARDSBURG FORMATION 447

counties, Washington, Washington Division of Geology and Earth Washington and Adjacent Areas. Washington Division of Geology
Resources Report of Investigations, 30, 37. and Earth Resources, 1– 21.
Link, P. K. & Christie-Blick, N. 2011. Neoproterozoic strata of South- Mrofka, D. & Kennedy, M. 2011. The Kingston Peak Formation in
eastern Idaho and Utah: record of Cryogenian rifting and glaciation. the eastern Death Valley Region. In: Arnaud, E., Halverson,
In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The G. P. & Shields-Zhou, G. (eds) The Geological Record of Neo-
Geological Record of Neoproterozoic Glaciations. Geological proterozoic Glaciations. Geological Society, London, Memoirs, 36,
Society, London, Memoirs, 36, 425–436. 449– 458.
Link, P. K., Christie-Blick, N. et al. 1993. Middle and Late Proterozoic Oliver, D. H. & Blackwell, D. D. 2002. Neoproterozoic diamictites in
stratified rocks of the western US Cordillera, Colorado Plateau, and central Idaho: evidence of felsic volcanism from embayed quartz
Basin and Range province. In: Reed, J. C. Jr, Bickford, M. E., grains. Geological Society of America Abstracts with Programs, 34,
Houston, R. S., Link, P. K., Rankin, D. W., Sims, P. K. & Van 89– 90.
Schmus, W. R. (eds) Precambrian: Conterminous US. Geological Park, J. K. 1997. Paleomagnetic evidence for low-latitude glaciation
Society of America, Boulder, Colorado, The Geology of North during deposition of the Neoproterozoic Rapitan Group, Mackenzie
America, 463– 595. Mountains, N.W.T., Canada. Canadian Journal of Earth Science,
Lund, K. 2004. Geology of the Payette National Forest and vicinity, 34, 34– 49.
west-central Idaho. US Geological Survey Professional Paper, Petterson, R., Prave, A. R. & Wernicke, B. P. 2011. Glaciogenic and
1666, 89. related strata of the Neoproterozoic Kingston Peak Formation in the
Lund, K. 2008. Geometry of the Neoproterozoic and Paleozoic rift Panamint Range, Death Valley region, California. In: Arnaud, E.,
Margin of western Laurentia: implications for mineral deposit Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
settings. Geosphere, 4, 429– 444. Record of Neoproterozoic Glaciations. Geological Society, London,
Lund, K., Aleinikoff, J. N., Evans, K. V. & Fanning, C. M. 2003. Memoirs, 36, 459–465.
SHRIMP U–Pb geochronology of Neoproterozoic Windermere Rainbird, R. H., Jefferson, C. W. & Young, G. M. 1996. The early
Supergroup, central Idaho: implications for regional synchroneity Neoproterozoic sedimentary succession B of northwestern Laurentia:
of Sturtian glaciation and associated rifting. Geological Society of correlations and paleogeographic significance. Geological Society
America Bulletin, 115, 349–372. of America Bulletin, 108, 454–470.
Lund, K., Aleinikoff, J. N., Evans, K. V., duBray, E. A., Dewitt, E. H. Ross, G. M., Bloch, J. D. & Krouse, H. R. 1995. Neoproterozoic strata
& Unruh, D. M. 2010. SHRIMP U–Pb dating of recurrent Cryogen- of the southern Canadian Cordillera and the isotopic evolution of
ian and Late Cambrian – Early Ordovician alkalic Magmatism in seawater sulfate. Precambrian Research, 73, 71 –99.
central Idaho: implications for Rodinian rift tectonics. Geological Smith, L. H., Kaufman, A. J., Knoll, A. H. & Link, P. K. 1994.
Society of America Bulletin, 122, 430– 453. Chemostratigrapy of predominantly siliciclastic Neoproterozoic
McCandless, D. O. 1982. A reevaluation of Cambrian through Middle successions: a case study of the Pocatello Formation and
Ordovician stratigraphy of the southern Lemhi Range. Master’s thesis, lower Brigham Group, Idaho, USA. Geological Magazine, 131,
The Pennsylvania State University, University Park, Pennsylvania, 157. 301– 314.
Miller, J. M. G. 1985. Glacial and syntectonic sedimentation: the upper Smith, M. D., Arnaud, E., Arnott, R. W. C. & Ross, G. M. 2011. The
Proterozoic Kingston Peak Formation, southern Panamint Range, record of Neoproterozoic glaciation in the Windermere Supergroup,
eastern California. Geological Society of America Bulletin, 96, southern Canadian Cordillera. In: Arnaud, E., Halverson, G. P.
1537– 1553. & Shields-Zhou, G. A. (eds) The Geological Record of Neopro-
Miller, F. K. 1994. The Windermere Group and Late Proterozoic tec- terozoic Glaciations. Geological Society, London, Memoirs, 36,
tonics in northeastern Washington and northern Idaho. Washington 413– 424.
Division of Geology and Earth Resources, 1 –19. Stewart, J. H. 1972. Initial deposits in the Cordilleran geosyncline;
Miller, F. K. & Whipple, J. W. 1989. The Deer Trail Group – is it part of evidence of a Late Precambrian (,850 m.y.) continental separation.
the Belt Supergroup. In: Joseph, N. L. (ed.) Geologic Guidebook for Geological Society of America Bulletin, 83, 1345– 1360.
Chapter 40

The Kingston Peak Formation in the eastern Death Valley region

DAVID MROFKA* & MARTIN KENNEDY


1
Department of Earth Sciences, University of California Riverside, Riverside, CA 92521, USA
*Corresponding author (e-mail: mrofkd01@ucr.edu)

Abstract: The late Neoproterozoic Kingston Peak Formation (Fm.) is a several-kilometre-thick sedimentary succession primarily influ-
enced by syndepositional tectonism and located in the region around Death Valley, California (Fig. 40.1). Its distribution is divisible into
an eastern facies assemblage, the subject of this paper, and a western facies assemblage covered in a separate chapter. The diamictite-
bearing Kingston Peak Fm. is bounded by the underlying shallow platform carbonates of the Beck Spring Fm. and overlain by the
Noonday Dolomite. There is an absence of direct palaeolatitude or radiometric age constraints and any correlation is based on broad
similarities with other coarse-grained strata (diamictite) located in a northward trending belt along the Cordilleran miogeocline. The
overlying Noonday Dolomite has been interpreted to be a late Cryogenian ‘cap carbonate’ and shares a set of unique facies associations
and isotopic and lithological characteristics with other late Neoproterozoic post-glacial carbonate intervals in Namibia, Canada, Australia
and Brazil. Research to date has focused on understanding local basin evolution, glacial sedimentology, correlation between the eastern
and western facies assemblages and initiation and development of the North American Cordillera. The intimate association of tectonic
and glacial facies with rapid local thickness and facies changes corresponding with syn-sedimentary faulting is the most distinctive strati-
graphic characteristic of the Kingston Peak Fm. The complex local stratigraphy complicates correlation both within the Death Valley
region as well as globally, and pending absolute age dates, does not fit easily with conventional Cryogenian Period glacial models
identifying two or more discrete ice ages.

The Kingston Peak Fm. (KPF) is the uppermost of the three for- Much of the interval of coarse-grained sediments comprising the
mations in the Pahrump Group (Hewett 1940) (Fig. 40.1), which KPF was first described in detail from the Panamint Range and
comprises the oldest sedimentary rocks preserved in the region. informally named in the Telescope group by Murphy (1930,
It crops out throughout the Death Valley region of southeastern 1932) as the Surprise Fm., Sourdough Limestone, Middle Park
California, where active extensional tectonism associated with Fm. and Wildrose Fm., from bottom to top respectively. These
the Basin and Range province along with an arid climate allow names have been retained but subsequently assigned to formal
clear lateral stratigraphic and facies relations to be observed. The member status (Johnson 1957; Carlisle et al. 1980; Labotka
mixed carbonate–siliciclastic sediments of the Crystal Spring et al. 1980) within Hewett’s (1940) Kingston Peak Fm.
Fm. and overlying shallow marine carbonate of the Beck Spring (Fig. 40.1). Noble (1934, 1941) described the eastern-KPF and
Dolomite comprise the basal and middle formations, respectively, likened it to the ‘Algonkian’ series in the Grand Canyon,
of the Pahrump Group. The KPF is overlain by the Noonday dolo- suggesting a correlation to Murphy’s (1932) Telescope group
mite (Fig. 40.1) (later the KPF) in the Panamint Range. Hazzard (1939) interpreted
Description of the KPF is complicated by its division into a the KPF to be glaciogenic based on the presence of striated clasts at
distinctive western facies assemblage (Petterson et al. 2011; the Gunsight Mine in the uppermost KPF.
western-KPF) along the western boundary of Death Valley in the From 1950 to the mid-1980s, relevant publications focused on
Panamint Range and a distinctive eastern facies assemblage (this the stratigraphy, sedimentology, palaeogeography and source
chapter; eastern-KPF) that itself is subdivided into northern and regions for the different facies of the KPF (Wright 1952, 1968;
southern facies (Fig. 40.1). The KPF also crops out in the Wright & Troxel 1966; Troxel 1967, 1982b; Wright et al. 1976,
Funeral Mountains further north (Miller 1983) but has undergone 1978, 1984). From the early 1980s onwards, publications primarily
amphibolite-grade metamorphism (Mattinson et al. 2007) and is addressed the glacial and tectonic features in the eastern and
not well-studied. Eastern and western facies assemblages of the western facies assemblages (Miller 1982, 1983, 1985, 1987;
KPF are dominated by similar coarse-grained siliciclastic rocks, Christie-Blick & Levy 1989; Link et al. 1993), providing interpret-
but lithostratigraphic correlation is complicated by lateral facies ations for the role of glaciation and extension in the deposition of
changes within each facies assemblage (Miller 1983), an overall the KPF as well as using the succession to help interpret the evol-
difference in the appearance of specific facies between both ution of the developing passive margin of the western Cordillera
regions and the presence in each region of distinctive carbonate (Levy & Christie-Blick 1989, 1991; Fedo & Cooper 2001).
intervals bounded by dissimilar facies. Prave (1999) proposed a chemo- and tectono-stratigraphic corre-
The eastern facies assemblage (also referred to as southeastern lation for the eastern and western facies assemblages of the KPF,
KPF in the literature) crops out in the southern Black Mountains suggesting the entire formation fit into a ‘Snowball Earth’ model
and in a number of hills and ranges SE of the Black Mountains, (Hoffman et al. 1998) with a stratigraphic record of two discrete
primarily in the Kingston Range (Fig. 40.1). More specifically, it ice ages. Most recently, Mrofka (2010) discussed the usefulness
crops out extensively in a readily accessible 30-km-long belt of carbonate intervals as timelines and the important relationship
along the northern and eastern flanks of the Kingston Range between syndepositional tectonism and preservation of the
(Hewett 1940) and is superbly exposed in a panel extending climate record in both the eastern- and western-KPF.
from the Silver Rule Mine (358480 1700 N, 1158560 3900 W, #6 in
Fig. 40.1) to Beck Canyon Divide (358480 1900 N, 1158550 3100 W,
#7 in Fig. 40.1). The northern Kingston Range is the location of
three characteristic KPF sections published in Hewett (1956; Structural framework
Fig. 40.1). It is the best location to examine the eastern-KPF
because of a relative lack of metamorphism, laterally persistent The KPF is concentrated in extensional basins, products of syn-
outcrops and visibility of rapid lateral facies changes. depositional tectonism (Mrofka 2010) likely associated with

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 449– 458. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.40
450 D. MROFKA & M. KENNEDY

Fig. 40.1. (a) Map showing regional


distribution of the western (white; see
Petterson et al. 2011 for more detail) and
eastern (northern, light-shaded; southern,
dark-shaded) facies assemblages of the
Kingston Peak Fm. Numbers represent
measured sections used for this study: 1,
Alexander Hills; 2, Gunsight Mine; 3, War
Eagle Mine; 4, Beck Canyon West; 5,
Crystal Spring; 6, Silver Rule Mine; 7, Beck
Canyon Divide; 8, Horsethief Spring; 9,
Kingston North; 10, Jupiter Mine; 11, Snow
White Mine; 12, Horsethief Mine; 13,
Mesquite North; 14, Mesquite South; 15,
Mesquite Small Block; 16, Winters Pass;
17, Southern Valjean Hills; 18, Sperry
Hills; 19, Saddle Peak Hills; 20, Saratoga
Hills; 21, Ibex Hills; 22, Eclipse Mine;
23, Virgin Spring Wash south; 24, Virgin
Spring Wash north; 25, Silurian Hills; 26,
southern Saddle Peak Hills; 27, Galena
Canyon; 28, Goler Wash; 29, Wood
Canyon; 30, Sourdough Canyon. Type
sections from Hewett (1956) are #7, #8 and
3 km south of #8. (b) Generalized
stratigraphy of the Pahrump Group showing
predominance of coarse-grained facies in
the Kingston Peak Fm. relative to the
bounding carbonate strata. (c) Schematic
stratigraphic logs comparing the different
facies assemblages of the Kingston Peak
Fm. across the Death Valley region.

rifting along the western margin of North America between 850 Cenozoic extension of the Basin and Range province (Davis
and 600 Ma (Stewart 1972; Christie-Blick & Levy 1989). During et al. 1993). Levy & Christie-Blick (1989) estimated c. 150%
extension and uplift along normal faults, resulting basins were extension to the original pre-Mesozoic basin after Mesozoic com-
filled with sediments derived from successively older stratigraphic pression and later Cenozoic extension. Estimates for Cenozoic
levels of the underlying Pahrump Group (Wright et al. 1976, 1992; extension range from c. 50–500% (Wright & Troxel 1967;
Burchfiel et al. 1992). Evidence for extension includes a prepon- Wernicke et al. 1988; Miller 1991), with higher estimates recently
derance of coarse-grained sediments including kilometre-scale called into question by the findings of Renik et al. (2008). At its
olistoliths (Miller 1985), syndepositional normal faults in the western boundary, in the southern Black Mountains, the eastern-
Kingston Range (Mrofka 2010), the southern Nopah Range KPF has undergone significant structural complication (Noble
(Wright et al. 1976, 1978) and the Panamint Range (Prave 1999) 1941; Troxel & Wright 1987; Miller 1991).
and tholeiitic (Hammond 1983) pillow lavas in the Panamint
Range (Miller 1985). Vertical offsets on the faults are .400 m
in the Kingston Range at Jupiter Hills (Fig. 40.1), as demonstrated Stratigraphy
by removal of the entire Beck Spring Dolomite on the footwall side
of faults, and possibly the removel of as much as 3 km of rock Underlying Pahrump Group
elsewhere (Burchfiel et al. 1992).
Sedimentation in the Death Valley region was relatively con- The lower and middle formations in the Pahrump Group
tinuous throughout the Palaeozoic (Abolins et al. 2000) but under- (Fig. 40.1b) record a transition from mixed carbonate –siliciclastic
went compressional shortening in the Permian (Snow 1992) and marine and fluvial facies of the Crystal Spring Fm. (Roberts 1974)
again during the Mesozoic (Levy & Christie-Blick 1989), with to the Beck Spring Dolomite, comprised of shallow-water carbon-
accompanying metamorphism in the Panamint Range ates in the north and mixed carbonate –siliciclastic fluvial –tidal
(western-KPF) (Labotka et al. 1980). In the Kingston Range, deposits south of the central Saddle Peak Hills (Marian &
north-northeastwards tilting strata of the eastern-KPF are generally Osborne 1992). In the Kingston Range, Marion & Osborne
offset along north –south trending normal faults and, in the (1992) divided the Beck Spring Dolomite into (i) a lower, lami-
southern Kingston Range, a detachment system occurs related to nated, cherty member, (ii) a laminated member with angular
KINGSTON PEAK FORMATION, EASTERN DEATH VALLEY 451

intraclasts and columnar stromatolites, (iii) a relatively thinner of the underlying KP1 Member or Virgin Spring limestone and
oolitic –pisolitic member and (iv) a partially silicified upper contains striated and faceted clasts. Clasts within the diamictite
member with abundant chert, shale lenses and stromatolites. are derived from the underlying Pahrump Group or basement
and the matrix is composed of coarse angular quartz sand and
illite (mica) or chlorite. Basal diamictite commonly contains
Kingston Peak Fm.: Eastern Facies Assemblage black limestone clasts and a carbonate-rich matrix likely derived
from the underlying Virgin Spring limestone. In the Saratoga
The coarse-grained siliciclastic eastern-KPF is separated into a Hills, southern Saddle Peak Hills and Alexander Hills, the diamic-
northern and southern facies based primarily on the distinct lithol- tite facies is interrupted by a 5– 20 m interval of finer-grained
ogy of clasts in diamictite from each area (Troxel 1967). The north- facies, variably including siltstone and sandstone with parallel
ern facies of the eastern-KPF is the focus of this contribution and is laminations, trough cross-bedding, steep bimodal cross-lamination
informally separated into the KP1 through KP4 members (Wright and normally graded pebble conglomerates with sandy tops.
1974) (Fig. 40.1a, c). The southern facies is limited in outcrop rela-
tive to the northern facies. Other prominent units within the KPF
include (i) a black laminated karstic limestone unit (Virgin KP3 Member, northern facies: sandstone, breccia and
Spring Limestone) that locally separates the KP1 and KP2 conglomerate
members, (ii) a discontinuous dolostone bed in the lower half of
the KP3 Member in the Kingston Range (Corsetti & Kaufman The KP3 Member is 15–2000 m thick, consists of interbedded silt-
2003) and (iii) lenses of diamictite that often overlie the KP4 stone and sandstone, diamictite with striated clasts, normally
Member (referred to informally as the Gunsight Member). graded conglomerate beds, kilometre-scale olistoliths and channel-
The basal KP1 and Virgin Spring Limestone of the northern filling sedimentary breccia. The lower KP3 Member is comprised
facies are separated from the glaciogenic units of KP2 through primarily of siltstone and sandstone interbedded with minor dia-
KP4 by a regional unconformity (Fig. 40.1c). Based on strati- mictite and conglomerate. Pebble- to cobble-sized outsized clasts
graphic and sedimentary characteristics, these two basal units commonly float in sandstone beds that grade laterally to conglom-
should constitute separate formations, but this has only come to erate or diamictite. Sedimentary structures include normally
light recently with the identification of the regional unconformity graded beds, convolute laminations and siltstone with intraclasts
(Mrofka 2010). The current stratigraphic scheme (Fig. 40.1c) and and flame structures. A 2– 3-m-thick oncolitic dolostone in the
member names are retained here until formal stratigraphic nomen- Kingston Range (358460 2700 N, 1158520 5900 W) is found near the
clature is published elsewhere. top of the finer-grained lower KP3 Member and marks a
coarsening-upwards transition to interbedded sandstone, normally
graded conglomerate and diamictite. As with the KP2 Member,
Noonday Dolomite diamictite intervals commonly contain black limestone clasts in
a black calcitic matrix. The middle to upper KP3 Member is
The Noonday Dolomite overlying the eastern-KPF is divided into a characterized by metre- to kilometre-scale mega-clasts and olisto-
lower, cream-coloured, laminated, microbial, dolomite member liths of the underlying Pahrump Group that form prominent ridges
and an upper, laminated, silty, dolomite member (Wright et al. in the Kingston Range around 358440 4300 N, 1158510 500 W and
1978). The laminations at the base several metres of the lower 358440 3500 N, 1158500 2200 W (Wright et al. 1976). Fe-rich (45%
member are parallel and horizontal, transitioning upwards to iron by weight) units at Sperry Wash (Abolins et al. 2000) are
arching laminations that define larger-scale microbial mounds found near Tertiary-age faulting and volcanic intrusions.
with synoptic relief of up to 200 m (Williams et al. 1974). Alterna-
tively, Summa (1993) suggested apparent mound ‘topography’
was due to an intra-formational erosion surface. The lower KP4 Member, northern facies: sedimentary breccia and
member contains distinctive vertical tubes possibly related to ver- conglomerate
tical transport of fluids (Cloud et al. 1974; Kennedy et al. 2001b) or
microbial processes (Corsetti & Grotzinger 2005) and centimetre- The KP4 Member gradationally overlies the KP3 Member and
to decimetre-scale pockets of sparry cement (Cloud et al. 1974; comprises 200– 1300 m of conglomerate, sedimentary breccia
Williams et al. 1974). In the east, the Noonday Dolomite tran- and monomictic mega-breccia (Figs 40.1 & 40.2). Conglomerate
sitions southwards to a siliciclastic-rich facies (Ibex Fm.), which and breccia beds are commonly normally graded and fill chan-
includes an arkosic siltstone member, a shaley-limestone nels 1–2 m deep and 10–50 m wide. Mega-breccia is massive,
member and a quartz-dolomite member (Williams et al. 1974). dominated by up to metre-scale angular blocks of Beck Spring
Dolomite and filling steeper and narrower channels than the
graded breccia and conglomerate beds. In the Kingston Range,
Glaciogenic deposits and associated strata sedimentary breccia composed entirely of Beck Spring Dolomite
was deposited adjacent to a syndepositional fault (358470 2200 N,
Glaciogenic strata occur in the KP2 –KP4 members of the 1158500 100 W, #10 in Fig. 40.1). Laterally, sedimentary breccia
Kingston Peak Fm. as well as in the southern facies assemblage and conglomerate is interbedded with normally graded to
(Fig. 40.1c). Deposition of the basal KP1 Member and overlying massive sandstone beds with sharp planar lower bed contacts. In
Virgin Spring limestone was not influenced by glacial processes footwall sections, the KPF is composed entirely of a diamictite
and both are therefore described subsequently within this section interval informally named the Gunsight Member (Troxel pers.
as associated strata. comm.). Near the Jupiter Mine (358470 2800 N, 1158500 100 W,
#10 in Fig. 40.1), the Gunsight Member includes channelized
sandstone with mudcracks and ripples overlain by diamictite
KP2 Member, northern facies: regional unconformity and with striated clasts.
diamictite deposition

The KP2 Member comprises a regionally extensive blanket of Glaciogenic deposits of the southern facies
massive- to diffusely-bedded cobble-boulder diamictite, which
varies from 10 to 250 m in thickness. It sharply overlies a region- Outcrops of the southern facies of the eastern-KPF (dark grey
ally extensive unconformity (Mrofka 2010), which defines the top shading in Fig. 40.1) are limited to a c. 8-km-wide belt of quartzite,
452 D. MROFKA & M. KENNEDY

Fig. 40.2. (a) Cross-section of four


measured sections showing footwall to
hanging wall transition and relationship
between northern and southern facies.
Numbers next to section names represent
approximate section thicknesses. Note:
different vertical scales are used. Northern
facies sections rest on the Beck Spring
Dolomite, except at the War Eagle mine,
where the Gunsight Member rests on the
Crystal Spring Fm. Uppermost unit shown
is the Noonday Dolomite. (b) Map showing
location of measured sections: 4, Beck
Canyon West; 7, Beck Canyon Divide;
12, Horsethief Mine; 25, Silurian Hills.
(c) Schematic diagram showing typical
extensional geometry of graben deposits
thought to apply to the eastern-KPF (after
Faerseth et al. 1997).

conglomerate and diamictite that crops out in the southern Saddle suggest the limestone in the Silurian Hills may alternatively be cor-
Peak Hills, the southern Salt Spring Hills and in the Silurian Hills. related to the Virgin Spring Limestone in the northern facies.
In the southern Salt Spring Hills, the southern facies is 1000 m Unlike northern facies diamictite, the diamictite in the southern
thick and dominantly composed of quartzite with c. 75 m of base- facies is dominated by clasts of granite and gneiss (Troxel
ment clast-bearing diamictite capping the lower third of the section 1967). Troxel (1982a) suggested this diamictite is interbedded
(Troxel 1967). In the Silurian Hills, the eastern-KPF is .2000 m with the northern facies in the southern Saddle Peak Hills.
thick (Kupfer 1960) and is floored by a dark parallel-laminated
karsted limestone. Kupfer (1960) correlated this limestone to the
Beck Spring Dolomite, but Prave (1999) correlated it to a discon- KP1 Member, northern facies: sandstone and siltstone
tinuous limestone below the Wildrose sub-Member in the western-
KPF. Prave’s (1999) correlation was based on similarities in d13C The non-glaciogenic KP1 Member is 1 to 180 m thick, composed
values and the appearance of a quartzite cobble conglomerate of centimetre-scale beds of parallel laminated sandstone and silt-
below both limestone intervals. The section above the limestone stone and underlies a regional erosional unconformity. Sedimen-
coarsens upwards and contains normally graded sandstone and tary structures include low-angle cross-lamination, beds with
conglomerate, diamictite and megaclasts (tens of metres scale). scoured bases, rare massive sandstone beds with mudstone chips
Similarities between the northern and southern facies include and a general coarsening and increase in carbonate cement upsec-
karsted and laminated limestone facies overlain by sandstone tion. In the southern Black Mountain (358540 4500 N, 1168380 5000 W,
beds, a coarsening-upwards trend above the karsted limestone #23 and 24 in Fig. 40.1), the KP1 Member varies in thickness by
intervals, and the presence of megaclasts. These similarities c. 30 m over a lateral distance of 100 m due to erosional truncation.
KINGSTON PEAK FORMATION, EASTERN DEATH VALLEY 453

Virgin Spring limestone, northern facies Field studies by the authors (see also Mrofka 2010) provide evi-
dence for uninterrupted deposition beginning at the base of the
The Virgin Spring limestone (Tucker 1986) sharply overlies the KP2 Member and continuing through the Noonday Dolomite,
KP1 Member, is erosionally truncated and karsted, dark, parallel- demonstrated by the following four sedimentary relationships.
laminae, and is preserved in only three localities. The limestone is First, KP4 Member sedimentary breccias in the Alexander Hills
17 m thick and best exposed in the Ibex Hills (358450 1800 N, are interbedded with the basal Noonday Dolomite (358450 5600 N,
1168260 1200 W, #21 in Fig. 40.1), but also crops out at Virgin 116860 5500 W). Second, in footwall sections the contact between
Spring Wash and in the Saratoga Hills where it is ,4 m thick the Noonday Dolomite and underlying strata is commonly inter-
and gradually truncated to the south (Fig. 40.1). Tucker (1986) rupted by a 1– 10 m layer of Gunsight Member diamictite; a
described the Virgin Spring limestone as comprising of similar diamictite interval appears conformable with the under-
centimetre- to decimetre-scale beds of parallel laminated lime- lying KP4 Member in hanging wall sections. Third, in the southern
stone interbedded with ,1-mm-thick sandstone laminae with Valjean Hills (358390 4000 N, 116870 2200 W, #17 in Fig. 40.1) and in
scoured bases and occasional normal grading. Petrographic and the Ibex Hills (DeYoung pers. comm.), Noonday Dolomite clasts
sedimentary features include ooids and convoluted or overturned are included in diamictite of the KP4 member or are in diamictite
beds attributed to mass sediment movement down the palaeoslope interbedded with KP4 Member sedimentary breccia. Alternatively,
(Tucker 1986). Corsetti & Kaufman (2005) interpreted this Ibex Hills interbedded
diamictite to post-date KPF deposition. Fourth, the base of the
Noonday Dolomite commonly contains clasts from the Beck
Boundary relations with overlying and underlying Spring Dolomite and Crystal Spring Fm. (i.e. 358450 4600 N,
non-glacial units 116860 5000 W). This relationship is consistently observed when
the Noonday Dolomite overlies the KP4 Member. Furthermore,
Contact with underlying units the Beck Spring Dolomite or Crystal Spring Fm. has not been
found in direct contact with the Noonday Dolomite. This likely
The contact underlying the glaciogenic interval of the KPF is indicates that during incipient Noonday Dolomite deposition,
complex and does not represent a single timeline, as it is located loose clasts from the surface of an unlithified KP4 Member were
above both an older and younger unconformity. The older uncon- reworked along with carbonate material from the flanks of
formity is preserved in hanging wall sections and separates the Noonday Dolomite mounds. Alternatively, the Pahrump Group
KP2 Member from a beveled contact with either the Virgin may have been exposed during Noonday Dolomite deposition,
Spring limestone or the underlying KP1 Member. The younger serving as a source of clasts in the basal Noonday Dolomite, and
unconformity occurs in footwall sections and variably separates contacts between the two are simply not exposed.
the KP4 Member from a beveled contact with either of the two
underlying formations of the Pahrump Group (Beck Springs
Dolomite or Crystal Springs Fm.) or the granitic basement. This Chemostratigraphy
younger unconformity developed due to erosion from uplift of
footwall blocks, which removed the older unconformity as well The most prominent carbonate units within the KPF succession are
as variable intervals of the Pahrump Group, down to the underlying the Virgin Spring limestone and the KP3 Member oncolitic dolo-
basement. stone bed in the northern facies, the Silurian Hills limestone in
In hanging wall sections where the contact between the non- the southern facies and the Sourdough Limestone in the
glacial units of the eastern-KPF and the underlying Beck Spring western-KPF (Fig. 40.1). Statistics for published isotopic data
Dolomite is preserved, it has been described as conformable, inter- for prominent carbonate units within and bounding the KPF are
fingering or unconformable (Christie-Blick & Levy 1989) and listed in Table 40.1 (Tucker 1983, 1986; Bergfeld et al. 1996;
shows no evidence of erosional truncation, although Kenny & Kennedy et al. 2001a; Corsetti & Kaufman 2003). Eastern-KPF
Knauth (2001) describe karstification of the upper Beck Spring carbonates have relatively enriched d13C and depleted d18O
Dolomite in some localities. In the Alexander Hills and Saratoga values, and data from published stratigraphic profiles (Tucker
Hills, the KP1 Member is described as transitional with the top
of the Beck Spring Dolomite over 10 m (Wright et al. 1992).
This relationship can be seen in the Alexander Hills (358460 200 N,
Table 40.1. Range of C- and O-isotopic values for carbonate associated with
116870 1000 W, #1 in Fig. 40.1) and in the southern Black Mounta-
the KPF
ins (358540 4500 N, 1168380 5000 W, #23 in Fig. 40.1) where there
is a sharp contact between the Beck Spring Dolomite and the
Unit Delta Min. Max. Average Median SD
KP1 Member sandstone, followed by interbedding between
centimetre-scale dolomite and sandstone beds over the next Beck Spring d13Ccarb – 4.4 5.8 2.7 3.0 1.8
several metres. Dolomite d18O –18.0 1.3 –6.0 – 5.5 3.8
n ¼ 259 d13Corg (n ¼ 45) –25.9 0.7 – 18.4 –18.6 4.3
KPF Oncolite d13Ccarb – 4.0 1.1 –1.6 – 2.0 1.5
Contact with the overlying Noonday Dolomite
Marker Bed d18O –11.2 – 2.1 –6.1 – 5.8 2.7
n ¼ 13 d13Corg
The contact between the eastern-KPF and the overlying Noonday Virgin Spring d13Ccarb 1.0 2.4 2.4 2.1 1.3
Dolomite is contentious and has been reported as regionally uncon- Limestone d18O –16.2 –12.5 – 15.4 –15.6 1.0
formable (Noble 1934; Wright et al. 1978), locally unconformable n ¼ 11 d13Corg
(Christie-Blick & Levy 1989) and locally conformable (Miller Sourdough d13Ccarb –7 3.9 –0.2 – 0.8 2.7
1987). An unconformable relationship has been suggested Limestone d18O –16.9 – 8.3 – 13.3 –13.6 2.1
because the Noonday Dolomite seems to cap successively older, n ¼ 45 d13Corg (n ¼ 9) –15.0 – 4.4 –7.9 – 6.7 3.7
seemingly tilted, strata (Cloud et al. 1974; Wright et al. 1976) Noonday d13Ccarb – 4.2 – 0.9 –2.8 – 2.7 1.1
between the Alexander Hills and the southern Nopah Range, ulti- Dolomite d18O –11.3 – 4.0 –6.9 – 6.5 3.1
mately straddling the contact between the Crystal Spring Fm. and n ¼ 22 d13Corg (n ¼ 9) –25.6 –17.6 – 21.3 –21.3 2.5
the basement at the War Eagle Mine. Alternatively, Prave (1999)
suggested the Gunsight Member infills erosional topography and Data from Tucker (1983, 1986); Bergfeld et al. (1996); Kennedy et al. (2001a)
is conformable with the overlying Noonday Dolomite. and Corsetti & Kaufman (2003).
454 D. MROFKA & M. KENNEDY

1986; Prave 1999; Corsetti & Kaufman 2003) for d13C and d18O Noonday Dolomite carbonate platforms, overlie a regional
identify no clear stratigraphic trends within the KPF succession. unconformity and record an abrupt change to siliciclastic sedi-
On the other hand, distinctive isotopic values and spatial continu- mentation resulting from tectonic uplift. The wedge-shaped
ity of the Beck Spring Dolomite and the Noonday Dolomite bracket packaging of the KPF strata next to tilted and erosionally trun-
the KPF and provide a clear stratigraphic and geochemical frame- cated segments of the underlying Pahrump Group and basement
work for the KPF. The Beck Spring Dolomite shows a several per is consistent with successions that accumulate during rotation of
mille enrichment in d13C above the base of the Formation (Corsetti hanging wall and footwall sections in extensional systems
& Kaufman 2003) and has enriched d13C and d18O values overall (Jackson & White 1988; Faerseth et al. 1990, p. 1291, fig. 7;
(Table 40.1). Noonday Dolomite d13C and d18O values for the over- Jackson et al. 2005). Coarse-grained facies, initiated with depo-
lying Noonday Dolomite show a moderate positive co-variation sition of the KP2 Member, are primarily the product of local tec-
(r 2 ¼ 0.3) and Kennedy et al. (1998, 2001a) pointed out the simi- tonic activity as indicated by (i) syndepositional normal faults
larity between the negative-to-positive d13C trend in the base of and erosional bevelling of footwall blocks, (ii) a systematic
the Noonday Dolomite and similar trends in other late Cryogenian and consistent pattern of coarsening upwards from sand and
cap carbonates. Hurtgen et al. (2004) published 34Ssulphate values angular cobble debrites to sedimentary breccia and kilometre-
for the Beck Spring Dolomite and the Noonday Dolomite in the scale olistoliths, (iii) a transition from marine debris-flow
range 11.0–27.4‰ and 15–35‰, respectively, and suggested the facies to terrestrial fanglomerates and (iv) a systematic pattern
values were evidence of ocean sulphate concentrations at 10% of of unroofing of the underlying Pahrump group as seen in the
modern values during the mid-Proterozoic with a transition to sequence of dominant clasts. A coarsening-upwards trend begin-
higher values in the late Proterozoic, possibly due to glaciation. ning in the upper KP3 Member and the transition to terrestrial
fanglomerates in the KP4 Member, especially near basin
margins, indicate shallowing as deposition outpaces subsidence
Other characteristics and creation of accommodation space. The occurrence of
striated clasts in the KP2 and KP3 members indicates a glacio-
Corsetti et al. (2003) documented complex microfossils preserved genic influence that is sporadic and limited to specific
in chert and carbonate from the oncolitic dolostone bed within the stratigraphic intervals.
KP3 Member, similar to microfossils identified in chert nodules The KP1 Member parallel-laminated sandstone was interpreted
from the Beck Spring Dolomite (Pierce & Cloud 1979; Horodyski by Tucker (1986) as having been deposited on a shelf by storm cur-
& Knauth 1994). Microfossil evidence was used to support the rents and the overlying laminated Virgin Spring limestone repre-
existence of active shallow-water microbial biota during a Cryo- senting a relatively deeper water carbonate environment with
genian glaciation, in which biological activity was thought to periodic input of sand and ooids by storms. The unconformity
have been non-existent (Hoffman et al. 1998). that truncates the Virgin Spring limestone might be a result of sea-
level fall from glaciation or tectonic uplift. Erosional truncation in
hanging wall sections indicates that while clasts from the under-
Palaeolatitude and palaeogeography lying Pahrump Group in the KP2 Member provide a clear signal
for initiation of tectonism, initial exposure and erosion across the
There is no published palaeolatitude data for the KPF itself and region may have been the result of ice growth and sea-level
palaeomagnetism test holes in the KP3 Member in the Alexander change prior to local tectonism, as indicated by the presence of
Hills yielded data that showed later remagnetization (Wright striated clasts.
2002, pers. comm.). Evans (2000, p. 365, fig. 3) estimated a c. 98 The KP2 Member was likely deposited in a glaciomarine
palaeolatitude for the KPF at 723 + 3 Ma based on correlation setting and underwent downslope reworking (Boulton &
with other Neoproterozoic strata along the North American Deynoux 1981) on the same broad shallow shelf on which the
Cordillera and proposed a nearly equivalent line of latitude with KP1 Member and Virgin Spring limestone were deposited.
the Toby Fm. (Christie & Fahrig 1983; Heaman et al. 1992). How- There is no evidence of the clinoform geometry or rapid lateral
ever, given Cryogenian Period glacial episodes occur between facies changes expected with ice-proximal debris aprons and
c. 750– 634 Ma (Condon et al. 2005; Kendall et al. 2006) and no evidence of laminations within the diamictite facies hosting
the lack of geochronological constraints on KPF deposition, the outsized clasts expected from rainout of ice-rafted debris. The
98 palaeolatitude is highly equivocal. KP3 Member records the greatest lateral thickness changes of
all the KPF members and represents the most active phase of tec-
tonism in the KPF, perhaps equivalent to a rift climax (Prosser
Geochronological constraints 1993). Diamictite intervals within the KP3 Member with striated
clasts are interbedded with coarse-grained deposits interpreted to
The KPF is poorly constrained by a date of 1.08 Ga from diabase be a result of gravity-driven debris flows of glaciogenic debris
within the middle member of the Crystal Spring Fm. (Heaman & down a rapidly tectonically steepening margin. The KP4
Grotzinger 1992), 500–1000 m below the base of the KPF, and Member is dominated by commonly monomictic channelized
the Precambrian–Cambrian Boundary in the lower member of the sedimentary breccia and likely represents terrestrial fanglomerate
Wood Canyon Fm. (Corsetti & Hagadorn 2000), 2000 m above (Hewett 1956). Lateral inter-fingering relations evident in the
the base of the Noonday Dolomite (Fig. 40.1). Deposition of Kingston Range between fanglomerate facies and turbiditic sand-
broadly similar (but not necessarily correlative) coarse-grained stone suggest the KP4 Member is a terrestrial equivalent of the
strata in Idaho is bounded by a lower limit of 717 + 4 Ma and more distal, marine KP3 Member, indicating fan-building pro-
701 + 4 Ma (Fanning & Link 2004, 2008; U–Pb SHRIMP on a gressed sub-aqueously and mixed with finer-grained distal
volcanic clast in diamictite) and upper limit of 685 + 7 Ma (Lund density deposits.
et al. 2003, U–Pb SHRIMP) and 667 + 5 Ma (Fanning & Link The Gunsight Member records a final pulse of glaciation and
2004; U–Pb SHRIMP based on tuff above overlying carbonate). subsequent deglacial flooding of KP4 Member terrestrial deposits.
The abrupt and conformable transition to the overlying, regionally
continuous, platformal Noonday Dolomite, as well as the thickness
Discussion of the dolomite’s microbial mounds (.200 m) suggest flooding
was a result of deglacial sea-level rise, and that continued trans-
Tectonic and glacial deposits of the eastern-KPF are associated gression and cessation of tectonism completely cut off any
with the underlying Beck Spring Dolomite and overlying source of siliciclastic sediments.
KINGSTON PEAK FORMATION, EASTERN DEATH VALLEY 455

Association with Neoproterozoic glaciation does not conform to a glacial – cap carbonate –interglacial–
glacial –cap carbonate cycle typical of other Neoproterozoic
The only direct evidence for glaciation in the KPF is striated clast- successions. Although associated with coarse-grained facies, the
bearing diamictite (Hazzard 1939; Miller 1985) found throughout palaeoclimatic significance of the carbonate intervals in the
the KP2 Member, interbedded in the lower half of the KP3 eastern KPF is unclear.
Member and commonly comprising the Gunsight Member.
Striated clasts are polished, comprised of siltstone, chert and quart-
zite, and rarely faceted. There is no systematic relationship Conclusion
between striated clast-bearing diamictite and bounding units; in
the KP2 Member diamictite is overlain sharply by sandstone of The Kingston Peak Fm. in the Death Valley region is one of the few
the basal KP3 Member, in the KP3 Member diamictite is inter- areas in the southern Cordillera where the association between
bedded with sandstone and sedimentary breccia and in the Gun- extension and resulting Neoproterozoic sedimentation is clear
sight Member it sharply overlies sedimentary breccia. Indirect (Burchfiel et al. 1992). This allows three competing hypotheses
evidence for a glacial influence (Crowell 1999) in the KPF was for the origin of Neoproterozoic glacial deposits to be tested: (1)
documented by Miller (1985) in both the western and eastern-KPF glacial sediments record a globally synchronous climate event
and includes evidence for rapid deposition, presence of diamictite (Snowball Earth); (2) glacial sediments record regional glaciation
and rapid lateral changes in facies and thickness. attributed to tectonism and terminated with the cessation of exten-
Sediments with striated clasts are often attributed to ice-rafted sion (i.e. zipper-rift model of Eyles & Januszczak (2004)); and (3)
debris in glaciomarine settings (Crowell 1999). At Sperry Wash glacial sediments comprise an incomplete record of either regional
(358420 1300 N, 1168140 3400 W, #18 in Fig. 40.1) outsized clasts glaciation or a long-term (50 –100 Ma) Cryogenian glacial era
within turbidite facies (Troxel 1982b) of the KP3 Member have (Allen & Etienne 2008) with sedimentary evidence of glaciation
been interpreted as dropstones (Abolins et al. 2000; Corsetti & only preserved during periods of tectonism and generation of
Kaufman 2003) or as lone clasts rolling down tectonically produced accommodation space. In hypothesis two and three, the preser-
slopes (Troxel 1982b). Several clasts appear to deform underlying vation of glaciogenic sediments is linked with regional tectonism
sediments and at least one pierces underlying sediments and may and therefore results in a diachronous record of glaciation.
have splash-marks (Corsetti pers. comm. 2008). However, outsized Each of the three hypotheses results in specific predictions for
clasts at Sperry Wash are grouped along common bedding the sedimentary record. In the first case, if the transgressive depos-
planes and laterally associated with diamictites and normally its of the Noonday Dolomite are related to global deglaciation and
graded conglomerate beds interpreted to be debris flows (Troxel not cessation of extension, syn-depositional faulting should con-
1982b). In other sections, similar outsized clast-bearing facies tinue through the Noonday Dolomite. This has not been observed.
are interbedded with conglomerate and host megaclasts (tens of In the second case, carbonate deposition is related to transgression
metres scale) and kilometre-scale olistoliths. The association during thermal subsidence when tectonism ends; rising sea level
with tectonically emplaced olistoliths, the bedding plane parallel confines glacial evidence up-dip or may be entirely lacking
orientation of many clasts and close association of outsize clasts because of slowdown in adiabatic cooling provided by uplift. In
to debrites beds with lonestones, suggests outsized clasts may the last case, extension and uplift provide accommodation space
alternatively represent the distal edges of debris flows or loose in shallow marine to terrestrial environments that rapidly preserve
clasts that tumbled down clast-laden slopes (Postma et al. 1988). glacial sediments and provide a record of glaciation and climate
The KP3 Member at Sperry Wash contains up to 45% Fe by change whose continuity and completeness is controlled by
weight (Abolins et al. 2000) and due to association between local tectonism.
Fe-rich sediments and Neoproterozoic glacial intervals (Young The intimate association between localized tectonically created
1976), has been interpreted to represent evidence for glaciation uplift, accommodation space, and conglomeratic facies bring into
(Stewart 1972; Abolins et al. 2000; Awramik et al. 2000). question the suggestion that KPF strata record a regional com-
However, there is no direct evidence for an oceanographic origin ponent of synchronous glacial sedimentation associated with
for the Fe. Alternatively, the Fe might be associated with nearby global Neoproterozoic ice ages (Prave 1999). The complete lack
Tertiary-aged volcanic intrusions. of geochronological data makes any global inference problematic.
Neoproterozoic glacial sediments are commonly intimately Evidence in the Kingston Peak Fm. provides more support for
associated with overlying carbonate facies (Hoffman et al. hypotheses two and three and consequently a strong argument
1998). These cap carbonates typically overlie an older and for diachronous glacial deposition and cap carbonate deposition.
younger glacial interval in many sections and host a variety of dis- These data support the argument that the Kingston Peak Fm.,
tinctive sedimentary and geochemical features (Kennedy et al. along with an increasing number of other Neoproterozoic deposits
1998; Hoffman et al. 2002). The Sourdough Limestone Member worldwide (Allen & Etienne 2008), record part of a continual and
in the western-KPF is characterized by graphite-rich parallel diachronous climate record (Mrofka 2010) spanning the Cryogen-
laminations and depleted d13C values (Table 40.1), characteristics ian Period.
shared among older Cryogenian cap carbonates (Kennedy et al.
1998), and has been interpreted to represent an older cap carbonate The National Science Foundation (EAR 0345207), NASA Exobiology
in the western-KPF (Prave 1999). In the eastern-KPF, the Virgin (NWG04GjJ42G), Geological Society of America Student Research Grant and
Spring and Silurian Hills limestone units are also characterized the American Association for Petroleum Geologists Grant-in-Aid provided
by graphite-rich parallel laminations but have enriched d13C support for D.M.’s field studies on the KPF. The authors appreciate the advice
values (Table 40.1). The Noonday Dolomite, interpreted to rep- and direction provided by B. Troxel and L. Wright. The authors also wish to
resent a younger cap carbonate (Prave 1999), overlies the KPF thank the following for their assistance in the field: K. Thompson, D. deYoung,
and shares a number of characteristics with other younger cap car- T. Bristow, Ganqing Jiang and C. Partin. This represents a contribution of
bonates, including its cream colour, tubestones, abundant marine the IUGS- and UNESCO-funded IGCP (International Geoscience Programme)
cements and depleted d13C values (Kennedy et al. 1998; Nogueira project #512.
et al. 2003).
Carbonate intervals in both eastern and western KPF intervals
are associated with coarse-grained facies but litho-stratigraphic References
correlation is complicated by dissimilar bounding lithofacies.
Deposition and a sporadic glacial influence are continuous from Abolins, M., Oskin, R., Prave, A. R., Summa, C. & Corsetti, F. A.
the KP2 Member through the Noonday Dolomite, so the interval 2000. Neoproterozoic glacial record in the Death Valley region,
456 D. MROFKA & M. KENNEDY

California and Nevada. In: Lageson, D. R., Peters, S. G. & Lahren, the Jurassic development of the northern North Sea. Journal of Struc-
M. M. (eds) GSA Field Guide 2: Great Basin and Sierra Nevada, tural Geology, 19, 1285–1302.
319– 335. Fanning, C. M. & Link, P. K. 2004. U –Pb SHRIMP ages of Neoproter-
Allen, P. A. & Etienne, J. L. 2008. Sedimentary challenge to Snowball ozoic (Sturtian) glaciogenic Pocatello Formation, southeastern Idaho.
Earth. Nature Geoscience, 1, 817–825. Geology, 32, 881–884.
Awramik, S. M., Corsetti, F. A. & Shapiro, R. 2000. Stromatolites and Fanning, C. M. & Link, P. K. 2008. Age constraints for the Sturtian
the pre-Phanerozoic to Cambrian history of the area south east of Glaciation; data from the Adelaide Geosyncline, South Australia
Death Valley. Bulletin of the San Bernardino County Museum, 47, and Pocatello Formation, Idaho, USA. In: Gallagher, S. J. &
65 –74. Wallace, M. W. (eds) Neoproterozoic extreme climates, the origin
Bergfeld, D., Nabelek, P. I. & Labotka, T. C. 1996. Carbon isotope of early metazoan life. Geological Society of Australia Extended
exchange during polymetamorphism in the Panamint Mountains, Abstracts, 91, 57 – 62.
California, USA. Journal of Metamorphic Geology, 14, 199– 212. Fedo, C. M. & Cooper, J. D. 2001. Sedimentology and sequence stratigra-
Boulton, G. S. & Deynoux, M. 1981. Sedimentation in glacial environ- phy of Neoproterozoic and Cambrian units across a craton-margin
ments and the identification of tills and tillites in ancient sedimentary hinge zone, southeastern California, and implications for the early
sequences. Precambrian Research, 15, 397– 422. evolution of the Cordilleran margin. Sedimentary Geology, 141,
Burchfiel, B. C., Cowan, D. S. & Davis, G. A. 1992. Tectonic overview 501– 522.
of the Cordilleran Orogen in the Western United States. In: Burch- Hammond, J. C. 1983. Late Precambrian diabase intrusions in the
fiel, B. C., Lipman, P. W. & Zoback, M. L. (eds) The Cordilleran southern Death Valley region California: their petrology, geochemis-
Orogen; Conterminous U.S.: The Geology of North America. try, and tectonic significance. PhD dissertation, University of
Geological Society of America, Boulder, G-3, 407– 480. Southern California.
Carlisle, D., Kettler, R. M. & Swanson, S. C. 1980. Geological study Hazzard, J. C. 1939. Possibility of pre-Cambric glaciation in southeast-
of uranium potential of the Kingston Peak Fm., Death Valley region, ern California. Pan-American Geologist, 71, 47– 48.
California. US Department of Energy Open File Report. Heaman, L. M. & Grotzinger, J. P. 1992. 1.08 Ga diabase sills in the
Christie, K. W. & Fahrig, W. F. 1983. Paleomagnetism of the Borden Pahrump Group, California; implications for development of the
dykes of Baffin Island and its bearing on the Grenville Loop. Cana- Cordilleran Miogeocline. Geology, 20, 637–640.
dian Journal of Earth Sciences, 20, 275–289. Heaman, L. M., Lecheminanta, A. N. & Rainbird, R. H. 1992. Nature
Christie-Blick, N. & Levy, M. 1989. Stratigraphic and tectonic frame- and timing of Franklin igneous events, Canada: implications for a
work of the upper Proterozoic and Cambrian rocks in the western Late Proterozoic mantle plume and the break-up of Laurentia.
United States. In: Christie-Blick, N., Levy, M., Mount, J. F., Earth and Planetary Science, 109, 117–131.
Signor, P. W. & Link, P. K. (eds) Late Proterozoic and Cambrian Hewett, D. F. 1940. New formation names to be used in the Kingston
Tectonics, Sedimentation, and Record of Metazoan Radiation in the Range, Ivanpah Quadrangle, California. Journal of the Washington
Western United States, Field Trip Guidebook T331, American Academy of Sciences, 30, 239–240.
Geophysical Union, Washington, DC, 7 –21. Hewett, D. F. 1956. Geology and mineral resources of the Ivanpah Quad-
Cloud, P., Wright, L. A., Williams, E. G., Diehl, P. E. & Walter, M. R. rangle, California and Nevada. United States Geological Survey
1974. Giant stromatolites and associated vertical tubes from the Professional Paper, 275, 23– 99.
upper Proterozoic Noonday Dolomite, Death Valley region, eastern Hoffman, P. F. & Schrag, D. P. 2002. The snowball earth hypothesis:
California. Geological Society of America Bulletin, 85, 1869– 1882. testing the limits of global change. Terra Nova, 14, 129–155.
Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005. Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P.
U –Pb ages from the Neoproterozoic Doushantuo Formation, China. 1998. A Neoproterozoic snowball earth. Science, 281, 1342–1346.
Science, 308, 95 –98. Horodyski, R. J. & Knauth, L. P. 1994. Life on land in the Precambrian.
Corsetti, F. A. & Hagadorn, J. W. 2000. Precambrian – Cambrian Science, 263, 494– 498.
transition: Death Valley, United States. Geology, 28, 299– 302. Hurtgen, M. T., Arthur, M. A. & Prave, A. R. 2004. The sulfur isotope
Corsetti, F. A. & Kaufman, A. J. 2003. Stratigraphic investigations of composition of carbonate-associated sulfate in Mesoproterozoic
carbon isotope anomalies and Neoproterozoic ice ages in Death to Neoproterozoic carbonates from Death Valley, California. In:
Valley, California. Geological Society of America Bulletin, 115, Amend, J. P., Edwards, K. J. & Lyons, T. W. (eds) Sulfur Biogeo-
916– 932. chemistry – Past and Present. Geological Society of America
Corsetti, F. A. & Grotzinger, J. P. 2005. Origin and significance of tube Special Paper, 379, 177–194.
structures in Neoproterozoic post-glacial cap carbonates: example Jackson, J. A. & White, N. J. 1988. Normal faulting in the upper conti-
from Noonday Dolomite, Death Valley, United States. Palaios, 20, nental crust: observations from regions of active extension. Journal
348– 362. of Structural Geology, 11, 15– 36.
Corsetti, F. A. & Kaufman, A. J. 2005. The relationship between the Jackson, C. A. L., Gawthorpe, R. L., Carr, I. D. & Sharp, I. R. 2005.
Neoproterozoic Noonday Dolomite and the Ibex Formation: New Normal faulting as a control on the stratigraphic development of
observations and their bearing on ‘snowball Earth’. Earth-Science shallow marine syn-rift sequences: the Nukhul and Lower Rudeis
Reviews, 73, 63– 78. Formations, Hammam Faraun fault block, Suez Rift, Egypt. Sedimen-
Corsetti, F. A., Awramik, S. M. & Pierce, D. 2003. A complex micro- tology, 52, 313– 338.
biota from snowball Earth times: microfossils from the Neoprotero- Johnson, B. K. 1957. Geology of a part of the Manly Peak Quadrangle,
zoic Kingston Peak Formation, Death Valley, USA. Proceedings of southern Panamint Range, California. University of California Publi-
the National Academy of Sciences of the United States of America, cations in Geological Sciences, 30, 353–423.
100, 4399– 4404. Kendall, B., Creaser, R. A. & Selby, D. 2006. Re– Os geochronology
Crowell, J. C. 1999. Pre-Mesozoic Ice Ages; Their Bearing on Under- of postglacial black shales in Australia; constraints on the timing of
standing the Climate System. Geological Society of America, ‘Sturtian’ glaciation. Geology, 34, 729–732.
Memoirs, 192, 1– 112. Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K. H. &
Davis, G. A., Fowler, T. K. et al. 1993. Pluton pinning of an active Arthur, M. A. 1998. Two or four Neoproterozoic glaciations.
Miocene detachment fault system, eastern Mojave Desert, California. Geology, 26, 1059–1063.
Geology, 21, 627–630. Kennedy, M. J., Christie-Blick, N. & Prave, A. R. 2001a. Carbon iso-
Evans, D. A. 2000. Stratigraphic, geochronological, and paleomagnetic topic composition of Neoproterozoic glacial carbonates as a test
constraints upon the Neoproterozoic climatic paradox. American of paleoceanographic models for snowball Earth phenomena.
Journal of Science, 300, 347– 433. Geology, 29, 1135–1138.
Eyles, N. 1993. Earth’s glacial record and its tectonic setting. Earth- Kennedy, M. J., Christie-Blick, N. & Sohl, L. E. 2001b. Are Protero-
Science Reviews, 35, 1– 248. zoic cap carbonates and isotopic excursions a record of gas hydrate
Faerseth, R. B., Knudsen, B. E., Liljedahl, T., Midboe, P. S. & destabilization following Earth’s coldest intervals? Geology, 29,
Soderstrom, B. 1997. Oblique rifting and sequential faulting in 443– 446.
KINGSTON PEAK FORMATION, EASTERN DEATH VALLEY 457

Kenny, R. & Knauth, L. P. 2001. Stable isotope variations in the Neo- Noble, L. F. 1941. Structural features of the Virgin Spring area, Death
proterozoic Beck Spring Dolomite and Mesoproterozoic Mescal Valley, California. Geological Society of America Bulletin, 52,
Limestone paleokarst: Implications for life on land in the Precam- 941– 999.
brian. Geological Society of America Bulletin, 113, 650–658. Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V. & Fair-
Kupfer, D. H. 1960. Thrust faulting and chaos structure, Silurian Hills, child, T. R. 2003. Soft-sediment deformation at the base of the
San Bernardino County, California. Geological Society of America Neoproterozoic Puga cap carbonate (southwestern Amazon craton,
Bulletin, 71, 181–214. Brazil): confirmation of rapid icehouse to greenhouse transition in
Labotka, T. C., Albee, A. L., Lanphere, M. A. & McDowell, S. D. snowball Earth. Geology, 31, 613– 616.
1980. Stratigraphy, structure, and metamorphism in the central Petterson, R., Prave, A. R. & Wernicke, B. P. 2011. Glaciogenic and
Panamint Mountains (Telescope Peak Quadrangle), Death Valley related strata of the Neoproterozoic Kingston Peak Formation in the
area, California. Geological Society of America Bulletin, 91 Part I, Panamint Range, Death Valley Region, California. In: Arnaud, E.,
125– 129. Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
Levy, M. & Christie-Blick, N. 1989. Pre-Mesozoic palinspastic recon- Record of Neoproterozoic Glaciations. Geological Society, London,
struction of the eastern Great Basin (Western United States). Science, Memoirs, 36, 459–465.
245, 1454–1462. Pierce, D. & Cloud, P. 1979. New microbial fossils from approximately
Levy, M. & Christie-Blick, N. 1991. Tectonic subsidence of the early 1.3 billion-year-old rocks of eastern California. Geomicrobiology
Paleozoic passive continental margin in eastern California and Journal, 1, 295– 309.
southern Nevada. Geological Society of America Bulletin, 103, Postma, G., Nemec, W. & Kleinspehn, K. L. 1988. Large floating clasts
1590– 1606. in turbidites: a mechanism for their emplacement. Sedimentary
Link, P. K., Christie-Blick, N. et al. 1993. Middle and late Proterozoic Geology, 58, 47 –61.
stratified rocks of the western U.S. Cordillera, Colorado Plateau, Prosser, S. 1993. Rift-related linked depositional systems and their seis-
and Basin and Range Province. In: Reed, J. C. Jr, Bickford, M. E. mic expression. Geological Society Special Publications, 71, 35– 66.
et al. (eds) Precambrian-Conterminous U.S.: The Geology of Prave, A. R. 1999. Two diamictites, two cap carbonates, two d13C excur-
North America, United States Geological Survey, Denver, CO, sions, two rifts; the Neoproterozoic Kingston Peak Formation, Death
463– 595. Valley, California. Geology, 27, 339–342.
Lund, K., Aleinikoff, J. N., Evans, K. V. & Fanning, C. M. 2003. Renik, B., Christie-Blick, N., Troxel, B., Wright, L. & Niemi, N.
SHRIMP U–Pb geochronology of Neoproterozoic Windermere 2008. Re-evaluation of the middle Miocene Eagle Mountain For-
Supergroup, central Idaho: implications for rifting of western Lauren- mation and its significance as a piercing point for the interpretation
tia and synchroneity of Sturtian glacial deposits. Geological Society of extreme extension across the Death Valley Region, California,
of America Bulletin, 115, 349–372. U.S.A. Journal of Sedimentary Research, 78, 199–219.
Mattinson, C. G., Colgan, J. P., Metcalf, J. R., Miller, E. L. & Roberts, M. T. 1974. Stratigraphy and depositional environments of the
Wooden, J. L. 2007. Late Cretaceous to Paleocene metamorphism Crystal Spring Formation, Southern Death Valley Region, California.
and magmatism in the Funeral Mountains metamorphic core Guidebook: Death Valley region, California and Nevada, Death
complex, Death Valley, California. In: Cloos, M., Carlson, Valley Publishing Company, Shoshone, 49 –57.
W. D., Gilbert, M. C., Liou, J. G. & Sorensen, S. S. Special Snow, J. K. 1992. Large-magnitude Permian shortening and continental-
Paper – Geological Society of America, 419, 205– 223. margin tectonics in the southern Cordillera. Geological Society of
Marian, M. L. & Osborne, R. H. 1992. Petrology, petrochemistry, and America Bulletin, 104, 80 –105.
stromatolites of the middle to late Proterozoic Beck Spring Dolomite, Stewart, J. H. 1972. Initial deposits in the Cordilleran geosyncline:
eastern Mojave Desert, California. Canadian Journal of Earth evidence of a late Precambrian (,850 m.y.) continental separation.
Sciences, 29, 2595–2609. Geological Society of America Bulletin, 83, 1345– 1360.
Miller, J. M. G. 1982. Kingston Peak Formation in the southern Panamint Summa, C. L. 1993. Sedimentologic, stratigraphic, and tectonic controls
Range; a glacial interpretation. In: Cooper, J. D., Troxel, B. W. & of a mixed carbonate-siliciclastic succession; Neoproterozoic
Wright, L. A. (eds) Geology of Selected Areas in the San Bernardino Johnnie Formation, Southeast California. PhD dissertation, Massa-
Mountains, Western Mojave Desert, and Southern Great Basin, chusetts Institute of Technology, Cambridge, MA.
California: Geological Society of America Cordilleran Section Field Troxel, B. W. 1967. Sedimentary rocks of late Precambrian and Cam-
Trip Guidebook and Volume. Death Valley Publishing Company, brian age in the southern Salt Spring Hills, southeastern Death
Shohone, California, 155–164. Valley, California. Short Contributions to California Geology:
Miller, J. M. G. 1983. Stratigraphy and sedimentology of the upper Special Report, California Division of Mines and Geology, 2, 33– 41.
Proterozoic Kingston Peak Formation, southern Panamint Range, Troxel, B. W. 1982a. Basin facies (Ibex Formation) of the Noonday
eastern California. PhD dissertation, University of California, Dolomite, southern Saddle Peak Hills, southern Death Valley,
Santa Barbara. California. In: Cooper, J. D., Troxel, B. W. & Wright, L. A.
Miller, J. M. G. 1985. Glacial and syntectonic sedimentation; the upper (eds) Geology of Selected Areas in the San Bernardino Mountains,
Proterozoic Kingston Peak Formation, southern Panamint Range, Western Mojave Desert, and Southern Great Basin, California: Geo-
eastern California. Geological Society of America Bulletin, 96, logical Society of America Cordilleran Section Field Trip Guidebook
1537– 1553. and Volume. Death Valley Publishing Company, Shoshone, 43– 48.
Miller, J. M. G. 1987. Paleotectonic and stratigraphic implications of the Troxel, B. W. 1982b. Description of the uppermost part of the Kingston
Kingston Peak –Noonday contact in the Panamint Range, eastern Peak Formation, Amargosa Rim Canyon, Death Valley region,
California. Journal of Geology, 95, 75 – 85. California. In: Cooper, J. D., Troxel, B. W. & Wright, L. A. (eds)
Miller, M. G. 1991. High-angle origin of the currently low-angle Geology of Selected Areas in the San Bernardino Mountains, Western
Badwater turtleback fault, Death Valley, California. Geology, 19, Mojave Desert, and Southern Great Basin, California: Geological
372– 375. Society of America Cordilleran Section Field Trip Guidebook and
Mrofka, D. D. 2010. Competing models for the timing of Cryogenian Gla- Volume. Death Valley Publishing Company, Shoshone, CA, 61– 70.
ciation: evidence from the Kingston Peak Formation, southeastern Troxel, B. W. & Wright, L. A. 1987. Tertiary extensional features,
California. PhD dissertation, University of California, Riverside. Death Valley region, eastern California. In: Hill, M. L. (ed.)
Murphy, F. M. 1930. Geology of the Panamint silver district, California. Centennial Field Guide-Cordilleran Section, Geological Society of
Economic Geology and the Bulletin of the Society of Economic Geol- America, Boulder, CO, 1, 121–132.
ogists, 25, 305– 325. Tucker, M. E. 1983. Diagenesis, geochemistry, and origin of a Precam-
Murphy, F. M. 1932. Geology of a part of the Panamint Range, report 27 brian dolomite: The Beck Spring Dolomite of eastern California.
of the state mineralogist. California Journal of Mines and Geology, Journal of Sedimentary Petrology, 53, 1097– 1119.
28, 329– 356. Tucker, M. E. 1986. Formerly aragonitic limestones associated with til-
Noble, L. F. 1934. Rock formations of Death Valley, California. Science, lites in the Late Proterozoic of Death Valley, California. Journal of
80, 173– 178. Sedimentary Petrology, 56, 818–830.
458 D. MROFKA & M. KENNEDY

Wernicke, B., Axen, G. J. & Snow, J. K. 1988. Basin and range exten- Wright, L. A., Troxel, B. W., Williams, E. G., Roberts, M. T. &
sional tectonics at the latitude of Las Vegas, Nevada. Geological Diehl, P. E. 1976. Precambrian sedimentary environments of the
Society of America Bulletin, 100, 1738– 1757. Death Valley region, eastern California. California Division of
Williams, E. G., Wright, L. A. & Troxel, B. W. 1974. The Noonday Mines and Geology Special Report, 106, 7 –15.
Dolomite and equivalent stratigraphic units, southern Death Valley Wright, L. A., Williams, E. G. & Cloud, P. 1978. Algal and cryptalgal
region, California. Guidebook: Death Valley Region, California structures and platform environments of the late pre-Phanerozoic
and Nevada. Death Valley Publishing Company, Shoshone, 73 –77. Noonday Dolomite, eastern California. Geological Society of
Wright, L. A. 1952. Geology of the Superior talc area, Death Valley, America Bulletin, 89, 321–333.
California. Special Report – California Division of Mines and Wright, L. A., Williams, E. G. & Troxel, B. W. 1984. Type section of
Geology, 20, 1– 22. the newly-named Proterozoic Ibex Formation, the basinal equivalent
Wright, L. A. 1968. Talc deposits of the southern Death Valley – King- of the Noonday Dolomite (Appendix II). Geology of the northern half
ston Range region, California. Special Report – California Division of the Confidence Hills 15-minute Quadrangle, Death Valley region,
of Mines and Geology, 95, 1 –79. eastern California, California Division of Mines and Geology, Map
Wright, L. A. 1974. Geology of the S.E. 1/4 Tecopa 15-minute quadran- Sheet 34, scale 1:24,000, 25 –31.
gle, San Bernardino and Inyo Counties, California. California Wright, L. A., Troxel, B. W. & Prave, A. R. 1992. Field traverse of
Division of Mines and Geology Map Sheet 20, 1:24,000. Proterozoic rock units, Alexander Hills and southern Nopah Range,
Wright, L. A. & Troxel, B. W. 1966. Strata of late Precambrian – Death Valley region, California. Geological Society of America
Cambrian age, Death Valley region, California –Nevada. Bulletin of Penrose Conference: Late Precambrian Plate Tectonics and the
the American Association of Petroleum Geologists, 50, 846– 857. Dawn of the Phanerozoic, Geological Society of America, 1 –11.
Wright, L. A. & Troxel, B. W. 1967. Limitations on right-lateral, strike– Young, G. M. 1976. Iron-formation and glaciogenic rocks of the Rapitan
slip displacement, Death Valley and Furnace Creek Fault Zones, Group, Northwest Territories, Canada. Precambrian Research, 3,
California. Geological Society of America Bulletin, 78, 933–950. 137– 158.
Chapter 41

Glaciogenic and related strata of the Neoproterozoic Kingston Peak Formation


in the Panamint Range, Death Valley region, California

RYAN PETTERSON1*, ANTHONY R. PRAVE2 & BRIAN P. WERNICKE1


1
Division of Geological and Planetary Sciences, California Institute of Technology, Mail Stop 100-23, Pasadena, CA, 91125, USA
2
Department of Earth Sciences, University of St Andrews, St Andrews, KY16 9AL, UK
*Corresponding author (e-mail: ryanpetterson@gmail.com)

Abstract: Glaciogenic deposits in the Death Valley region occur within the Neoproterozoic Kingston Peak Formation (Fm.). In the
Panamint Range, immediately west of Death Valley, the formation is as much as 1000 m thick and is continuously exposed for
nearly 100 km along the strike of the range. Although the strata are variably metamorphosed and locally exhibit pronounced ductile
strain, original sedimentary textures are well preserved in many places. Diamictite occurs in two distinct intervals, a lower one compris-
ing the Limekiln Spring and Surprise members, and an upper one, the Wildrose Sub-member of the South Park Member. Lonestones,
bullet-shaped and striated clasts, and rare dropstones within these members, along with the impressive lateral continuity of diamictic
units, support a glacial origin. Both diamictic intervals are succeeded by well-defined carbonates, the oldest is the Sourdough
Member of the Kingston Peak Fm. and the younger one is the Sentinel Peak Member of the overlying Noonday Dolomite. The strati-
graphic succession between the Sourdough Member and the Wildrose Sub-member (i.e. the Middle Park, Mountain Girl and Thorndike
sub-members of the South Park Member) is c. 300 m thick and includes lithologies recording deposition in braided fluvial to platform
carbonate settings. Lithostratigraphic and chemostratigraphic profiles of d13C for the Sourdough (– 3‰ to þ2‰, increasing upward) and
Sentinel Peak (– 3‰ +1‰) members suggest correlation with, respectively, the older Cryogenian (commonly referred to as ‘Sturtian’ in
previous literature) and younger Cryogenian (commonly referred to as ‘Marinoan’ in previous literature) cap-carbonate sequences recog-
nized worldwide. Potentially economic uranium deposits (secondary brannerite) occur in graphitic schist of the Limekiln Spring Member
and sub-economic uranium and thorium (hosted by detrital monazite) occur within quartz-pebble conglomerate in the South Park
Member. The Kingston Peak Fm. strata in the Panamint Range contain no fossils, radiometric age control or primary magnetizations.

Murphy (1932) was the first person to infer a glacial origin for a outcrop belt is the largest single exposure of Cryogenian strata
thick succession (1000 m) of Precambrian conglomeratic strata in the southwestern part of Laurentia, providing an unusual oppor-
exposed in the Telescope Peak area in the Panamint Range, a tunity to examine lateral transitions in facies and thickness in
major Basin and Range fault block on the western flank of Death detail.
Valley (Figs 41.1 & 41.2). Subsequently, Hewett (1940) documen-
ted equivalent rocks in the Kingston Range, about 50 km east of
Death Valley, and named them the Kingston Peak Fm., the young- Structural framework
est of three formations he assigned to the newly defined Pahrump
Group (the other two being the Crystal Spring Fm. and Beck The Panamint Range is a north –south-trending, east-tilted range
Spring Dolomite). This nomenclature has since been adopted block within the Basin and Range extensional province. The
and used throughout the Death Valley region. general structure of the range defines a NNW-trending anticline.
Kingston Peak Fm. strata are highly variable in thickness and The anticline core is 1.7 Ga gneiss, which is locally intruded by
sedimentary facies, owing to syn-depositional tectonism during a 1.4 Ga porphyritic quartz monzonite (Albee et al. 1981).
part (but not all) of their sedimentation history. Therefore, regional Various units of the Pahrump Group and younger strata overlie
correlation of units at the sub-formational level has proved chal- these basement rocks (collectively known as the World Beater
lenging (Miller et al. 1988; Prave 1999). Further, the recognition Complex; Labotka et al. 1980). Syn-depositional faulting and
and interpretation of the Sourdough Member (commonly refereed mafic magmatism have been identified within the Kingston Peak
to as Sourdough Limestone in the literature) of the Kingston Peak Fm. and are interpreted to be a manifestation of continental
Fm. and the Noonday Dolomite as cap-carbonate sequences related rifting (e.g. Stewart 1972; Wright et al. 1974; Hammond 1983),
to an older and younger Cryogenian glacial period, respectively, Contractile deformation, granitic magmatism and associated meta-
has remained a subject of debate (Prave 1999; Corsetti & morphism, and development of an east-directed thrust fault system
Kaufman 2003). This controversy, however, is becoming increas- in the Tucki Mountain area (Fig. 41.2; Wernicke et al. 1993)
ingly resolved by Petterson’s (2009) work, to be published in detail affected the Panamint Range in Jurassic and Cretaceous times
elsewhere, which shows that the Noonday Dolomite is the likely (Labotka et al. 1985). The final major deformation event was east-
equivalent to the younger Cryogenian cap sequence (commonly ward tilting of fault blocks along west-side-down normal faults
referred to as ‘Marinoan’), recognized worldwide and known to during Late Tertiary time, accompanied by intrusion of granitic
be c. 635 Ma in age (Hoffmann et al. 2004). plutons and associated dykes (Labotka et al. 1980; Hodges et al.
The strata east of Death Valley are described in detail in Mrofka 1989). Although the Neoproterozoic strata experienced green-
& Kennedy (2011). In that area, units are well exposed and have schist to amphibolite facies metamorphism and locally exhibit
experienced only minor metamorphism. However, lateral continu- pronounced ductile strain, original sedimentary textures are well
ity of exposures in eastern Death Valley is limited owing to strati- preserved in many places.
graphic variability, structural deformation, and the presence of The palaeogeographic setting of the Kingston Peak Fm. is that
wide areas of alluvium separating relatively small fault blocks. of an episodically active cratonic rift basin (Wright et al. 1974;
In contrast, Kingston Peak Fm. strata in the Panamint Range Labotka et al. 1980). The stratigraphic position of the Kingston
are well exposed for nearly 100 km along strike (Fig. 41.2). This Peak Fm. below more widespread deposits of the Cordilleran

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 459– 465. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.41
460 R. PETTERSON ET AL.

Fig. 41.1. Composite stratigraphic column


of Neoproterozoic units in the Death Valley
region (Stewart 1970; Wright et al. 1974;
Labotka et al. 1980; Heaman & Grotzinger
1992), showing the age constraints and key
markers discussed in the text. The left-hand
column shows a compilation of published
C-isotopic data discussed in text.

miogeocline has suggested to many workers that this phase of persistent and can be traced along most of the length of the Pana-
extensional deformation represents the early stages of continental mint Range.
break-up leading to the development of the early Palaeozoic
passive margin (Burchfiel & Davis 1972; Stewart 1972; Wright
et al. 1974; Levy & Christie-Blick 1991). The Limekiln Spring Member

The Limekiln Spring Member is the most stratigraphically


Stratigraphy complex member of the Kingston Peak Fm. in the Panamint
Range. Despite continuity of exposure of the Kingston Peak Fm.
In the Panamint Range, the Kingston Peak Fm. has been sub- as a whole, detailed mapping of this member is limited to a
divided into the Limekiln Spring, Surprise, Sourdough and South small area in the southern part of the range and a somewhat
Park members (Labotka et al. 1980; Miller 1985). The latter, orig- larger area in the vicinity of Telescope Peak (Fig. 41.2). Conse-
inally named by Murphy (1932), has been further sub-divided into, quently, the stratigraphic architecture of the Limekiln Spring
from oldest to youngest, the Middle Park, Mountain Girl, Thorn- Member is poorly understood. In southern exposures, which are
dike and Wildrose sub-members (Fig. 41.1). Diamictic strata not strongly deformed or metamorphosed, it consists of laterally
occur throughout the Limekiln Spring Member, the Surprise and vertically variable, interbedded diamictite, sandstone and
Member and the Wildrose Sub-member. With the exception of siltstone (Miller et al. 1988). In northern exposures, which are
the Limekiln Spring and Surprise members, which are lithologi- strongly deformed and metamorphosed, it is mostly immature
cally variable, the units of the Kingston Peak Fm. are laterally sandstone, pelitic schist, amphibolite, minor dolomitic marble
KINGSTON PEAK FORMATION, PANAMINT RANGE 461

Fig. 41.2. (a) Map showing distribution of


the Kingston Peak Fm. in the Death Valley
region. (b) Map showing distribution of the
four members of the Kingston Peak Fm. and
the Noonday Dolomite in the Panamint
Range.

and discontinuous lenses of metaconglomerate and breccia The Surprise Member


(Labotka 1978; Labotka et al. 1980; Miller 1985). Even though
structural thicknesses are highly variable, ranging from 0 to The Surprise Member is poorly bedded to massive diamictite in the
greater than 1000 m, at the scale of the Panamint Range, the Lime- southern Panamint Range and appears to pass northward into finer-
kiln Spring Member thins eastward beneath the overlying Surprise grained facies consisting of argillite and immature sandstone
Member. At least part of the thickness variability can be attributed (Labotka et al. 1980; Miller 1985). Locally, it contains a few
to the filling of pre-Limekiln-Spring palaeotopography formed on tens of metres of tholeiitic pillow basalt (Hammond 1983). The
the underlying crystalline basement (Labotka 1978; Labotka et al. Surprise Member ranges from 35 m in thickness near Goler
1980; Miller 1985). Wash in the southern Panamint Range (Miller 1985) to as much
462 R. PETTERSON ET AL.

as 1300 m west of Telescope Peak, although this may be partly the Noonday Dolomite
result of structural thickening (Labotka et al. 1980). In the central
and southern part of the Panamint Range, the Surprise Member is In contrast to the siliciclastic-dominated and diamictic-bearing
mostly massive diamictite dominated by quartzite and carbonate Kingston Peak Fm., the overlying Noonday Dolomite is predomi-
clasts supported by an argillaceous, sandy matrix; thin grey to nately carbonate. In the Panamint Range, the Noonday Dolomite
dark grey argillite beds are present locally, but these are minor. has been subdivided into three members, from oldest to youngest,
Clasts range from pebbles to boulders, with the quartzite clasts the Sentinel Peak, Radcliff, and Redlands (Murphy 1932; Johnson
generally rounded to sub-rounded whereas the carbonate clasts 1957; Labotka et al. 1980; Albee et al. 1981). Work by the authors
are sub-angular (Miller 1985). North of the latitude of Telescope has modified this nomenclature to include a fourth member, the
Peak, diamictite is uncommon and the dominant lithology is Mahogany Flats Member, which lies between the Radcliff
dark grey to black argillaceous siltstone and fine sandstone. In Member and strata equivalent to the type Redlands Member
addition, there is a significant amount of interbedded laminated (Petterson 2009); only the first three members will be discussed
limestone and argillaceous limestone (Labotka et al. 1980). here.
The transition between the diamictite-dominated facies in the The Sentinel Peak Member is an easily identifiable carbonate
southern and central parts of the Panamint Range and the argillite- marker unit exposed throughout the Panamint Range. It ranges
dominated, diamictite-poor facies in the northern part of the range from 0.5 m to as much as 70 m in thickness and, in its most com-
is structurally complex (Labotka 1978; Labotka et al. 1980). This plete development, can be sub-divided into three parts. Both the
leads to uncertainty as to whether the two facies are coeval. lower and upper parts consist of laminated dolostone and each is
typically 1– 5 m thick. The middle part makes up the remaining
thickness and is generally fine-grained dolostone with irregular
The Sourdough Member spar-filled vugs, locally containing mound-like features containing
tube structures (Hunt & Mabey 1966; Cloud et al. 1974; Corsetti &
The Sourdough Member is a 0.5 –45-m-thick, grey to dark grey Grotzinger 2005). In many places, the middle part is absent and the
laminated limestone. It commonly exhibits strong internal defor- upper and lower parts are amalgamated to form a single carbonate
mation, inhibiting confident identification of primary structures interval. In several localities where this occurs, a thin (as much as a
(Miller 1985). Laminations are defined by changes in calcite few metres thick) intraformational breccia occurs interbedded
grain size and abundance of muscovite and quartz, and, in areas within the laminated dolostone. We want to highlight that in
of higher metamorphic grade, finely disseminated graphite. Syn- earlier reports, in areas where the Wildrose Sub-member is
sedimentary deformation, isolated clasts, and lenticular beds of absent, the Thorndike Sub-member and the Sentinel Peak
clastic material have been reported (Miller 1985). Typically, car- Member were inadvertently mapped together, and their combined
bonate beds become interbedded with thin argillite beds upwards thickness was incorrectly assigned to the latter (e.g. the mapping of
from the base of the member. Albee et al. 1981 in the Telescope Peak area includes both units as
Sentinel Peak Member).
The Radcliff Member is highly variable in both thickness and
South Park Member (exclusive of the Wildrose Sub-member) lithology. It ranges from 100 to 250 m in thickness, and includes
limestone, dolostone, siltstone, sandstone and carbonate breccia.
Below the Wildrose Sub-member, the South Park Member All lithologies are thin-bedded, a characteristic feature of the Rad-
averages c. 300 m in thickness and contains three lithologically cliff Member. Its lower part is typically siliciclastic and consists of
distinct units that are mappable over the entire outcrop belt in laminated argillite and/or argillaceous arkose. These lithologies
the Panamint Range. They include 150–200 m of sandstone and pass gradationally upward to more abundant limestone with a dis-
pelite of the Middle Park Sub-member, 50–100 m of quartzite tinctive pale yellowish orange colour; metre-thick layers and
and conglomerate of the Mountain Girl Sub-member, and 50– lenses of intraformational breccia occur locally. The top of the
200 m of limestone and dolostone of the Thorndike Sub-member member includes quartz sandstone overlain by siltstone containing
(formerly referred to as the ‘un-named’ Limestone; Prave 1999; coarse carbonate breccia.
Corsetti & Kaufman 2003). Due to the details of stratigraphic The Mahogany Flats Member is primarily stromatolitic dolo-
architecture both within and above these units, thickness variations stone. It is about 200 m thick and exhibits a gradually increasing
and stratigraphic omissions of one or more units are common. For component of quartz sand in the upper half of the unit. Microbial
example, in the Wildrose Canyon area, the base of the Mountain structures are ubiquitous, and include (i) metre-scale mound struc-
Girl Sub-member defines an angular unconformity (c. 158) and tures; (ii) laterally linked heads; and (iii) branching columnar
variably sits on the Middle Park Sub-member, then on the Sour- stromatolites. Intermound fill includes both carbonate and quartz
dough and Surprise members (as discussed by Petterson 2009). sandstone.
In the Telescope Peak area, an erosion surface at the base of the
Wildrose Sub-member locally omits the entire lower portion of
the South Park Member (Miller 1987), although as discussed Glaciogenic deposits and associated strata
below, this surface does not appear to be an angular unconformity.
Diamictic units of the Surprise Member and the Wildrose Sub-
member are usually a few tens to a few hundred metres thick.
Wildrose Sub-member of the South Park Member One of the remarkable aspects of these units is their lateral persist-
ence, which permits them to be traceable along the strike of the
The Wildrose Sub-member is variably preserved at the top of the Panamint Range for at least 50–100 km. Clasts in both of the dia-
South Park Member and below the Noonday Dolomite. Where mictic intervals range from pebbles to cobbles in size, although
present, it is a massive, matrix-supported diamictite. Its thickness larger clasts up to 0.5 m are common and can be as much as 3 m
varies from 0 to 190 m (Miller 1985), but it is typically only a few in maximum dimension (Miller 1987). Clast shapes vary from
tens of metres thick. The diamictite in the Wildrose Sub-member is rounded to angular for all lithologies (although basement and
distinct from that in the Limekiln Spring-Surprise members in that quartzite clasts tend to be mostly the former whereas carbonate
it typically contains abundant gneiss clasts. However, there is a sig- clasts are typically the latter), and rare bullet-shaped quartzite
nificant internal variation in both clast and matrix compositions and/or basement clasts can be found. In general, the diamictic
such that, locally, Wildrose diamictite can be dominated by clasts units of the Surprise Member tend to be dominated by clasts
and matrix derived from the immediately underlying substrate. derived from the underlying Pahrump Group sedimentary rocks
KINGSTON PEAK FORMATION, PANAMINT RANGE 463

with the basement rocks being only minor. The opposite pro- although the Sentinel Peak Member locally sits on older units (Pet-
portion is generally the case for the clast composition in the Wild- terson 2009).
rose Sub-member, although clast composition can be nearly 100%
of the lithology of the immediately underlying substrate. The
matrix for the Surprise Member is a greenish grey argillite to argil- Chemostratigraphy
laceous sandstone. This is in contrast to the Wildrose Sub-member
in which the matrix varies from dark grey to black argillite and The Kingston Peak Fm. in the Panamint Range contains two main
arkosic fine sandstone where basement clasts predominate, to intervals of carbonate rocks, the Sourdough Member and Thorn-
carbonate-cemented medium to coarse greywacke sandstone dike Sub-member; minor occurrences of carbonate-bearing strata
where carbonate clasts (generally derived from the Thorndike Sub- also occur in parts of the Limekiln Spring, Surprise and South
member) predominate. The rarity of any type of layering is striking Park Members. In this discussion, we will also include the over-
in both the Surprise and Wildrose diamictic units; the diamictite is lying Sentinel Peak Member of the Noonday Dolomite (for discus-
largely massive, grading is poorly developed at best, and finer- sion of the chemostratigraphy of the underlying Beck Spring Fm.
grained, laterally discontinuous interbeds are rare. It is in the in the eastern Death Valley region, see Mrofka & Kennedy 2011).
latter intervals that rare lonestones and clasts that have pierced Although all these units have undergone variable metamorph-
laminae can be observed. ism, the C-isotopic data are consistent from section to section
and the trends are reproducible, independent of sedimentary
facies and the degree of metamorphism (Petterson 2009).
Boundary relations with overlying and underlying Further, the values obtained from the rocks in the Panamint
Range are nearly identical to those obtained by other workers in
non-glacial units the correlative, non-metamorphosed sections in eastern Death
Valley (Prave 1999; Corsetti & Kaufman 2003; Petterson 2009).
Boundary relations of the Limekiln Spring/Surprise Members Thus, the carbonate C-isotopic data can be confidently interpreted
as recording depositional values, and is therefore useful for both
In the Panamint Range, the Kingston Peak Fm. variably overlies intrabasinal and global correlations.
the Crystal Spring and Beck Spring formations and crystalline The Sourdough Member in the Redlands Canyon area of the
basement. In the southern part of the range, the Limekiln Spring southern Panamint Range yields d13C values between –2.6‰
Member sits directly on the crystalline basement (Miller 1985). and –1.1‰, although a single þ2.2‰ value is reported from the
Throughout much of the rest of the Panamint Range, the base of upper part of the unit (Prave 1999). Data from the Goler Wash
the Kingston Peak Fm. is not exposed and/or is modified by intru- and Pleasant Canyon areas provide d13C values ranging from
sion of Mesozoic and Cenozoic granitoids. In the Telescope Peak –2.4‰ to –1.6‰ and –3.1‰ to –2.9‰, respectively (Corsetti
area, the base of the Kingston Peak Fm. is obscured by metamorph- & Kaufman 2003).
ism and structural complexity, which hinders assessment of the Compared to the other carbonate units in the Kingston Peak
nature of the contact. The unit underlying the Kingston Peak Fm. Fm., the d13C values of the Thorndike Sub-member are strongly
in this area is a dolomitic marble correlated with the Beck positive. Values range from þ5.7‰ to þ6.3‰ in the Redlands
Spring Dolomite (Albee et al. 1981). The contact between them Canyon area (southern Panamint Range, Prave 1999) and
purportedly displays interfingering (Labotka et al. 1980, fig. 2a). þ4.7‰ to þ5.3‰ in the Tucki Mountain area (northern Panamint
However, in areas of lesser structural and metamorphic complex- Range, Corsetti & Kaufman 2003).
ity, the base of the Kingston Peak Fm. is observed to be uncon- d13C values for the Sentinel Peak Member in the Panamint
formable, including an angular discordance with underlying Range are between þ1.0‰ and – 3.0‰ (Petterson 2009). These
units of the Pahrump Group, and in other places, resting noncom- values match well those for the lower unit of the Noonday Dolo-
formably on the crystalline basement (Labotka et al. 1980). mite (i.e. the tube-bearing dolostone) in the eastern Death Valley
The contact between the Surprise and the Sourdough members is region, which has values near – 3.0‰ (Prave 1999; Corsetti &
sharp. However, in a number of places, the contact is reportedly Kaufman 2003), and recent work by Petterson (2009), to be pub-
marked by interbedded diamictite and limestone (Miller 1985); lished elsewhere, has shown that this unit is correlative with the
these occurrences, however, may be due to structural interleaving. Sentinel Peak Member.

Boundary relations of the Wildrose Sub-member Other characteristics

The base of the Wildrose Sub-member in the Panamint Range is A number of Neoproterozoic microfossil assemblages have been
sharp and erosive. Its substrate ranges from the Thorndike Sub- described from the Beck Spring and Kingston Peak formations
member to crystalline basement (Miller 1987, fig. 2). In the Tele- in the eastern Death Valley region (Pierce & Cloud 1978;
scope Peak area, the base of the Wildrose Sub-member cuts down Horodyski & Mankiewicz 1990; Awramik et al. 2000; Corsetti
section (from west to east) from the Thorndike Sub-member to the et al. 2003). However, no fossils have been reported from
Surprise Member (Miller 1987). In the southern part of the range, it Pahrump Group strata in the Panamint Range. This lack of
cuts down section from the Thorndike Sub-member to basement fossils is not surprising given both the degree of metamorphism
(Miller 1987; Prave 1999). Despite the diversity of substrates and the general scarcity of fossils elsewhere.
along this contact, there does not appear to be a significant Marginally economic levels of uranium (as secondary branner-
angular discordance between pre-Wildrose units and the lower ite, UTi2O6) occur in a graphitic schist unit of the Limekiln Spring
part of the Noonday Dolomite. Rather, as documented by Miller Member and sub-economic levels of uranium and thorium (as det-
(1987, figs 3 & 5), erosion of the pre-Wildrose units is rital monazite) are found in quartz pebble conglomerates in the
accompanied by an increase in thickness of the lower parts of lower part of the Mountain Girl Sub-member (Carlisle et al.
the Wildrose Sub-member, which reflects infilling of pre-Wildrose 1980). Iron ore has been mined from the Kingston Peak Fm. in
topographic lows, some with as much as 300 m of local relief. eastern Death Valley, but no such deposits occur in the Panamint
The top of the Kingston Peak Fm. in the Panamint Range is Range. Although many intrusion-related precious metal deposits
defined by the occurrence of the Sentinel Peak Member of the are present in the Panamint Range, no economically viable depos-
Noonday Dolomite. The contact is sharp, and the depositional its directly related to depositional processes or diagenesis
substrate is either the Wildrose or Thorndike sub-members, are known from the Kingston Peak Fm. in the Panamint Range.
464 R. PETTERSON ET AL.

Palaeolatitude and palaeogeography glacially influenced, either as glacimarine deposition, or in the case
of the Wildrose, perhaps as lodgement till (Miller 1985).
No palaeolatitude estimates have been determined from any of the A critical line of evidence that enables linking the Kingston
Kingston Peak Fm. rocks, either in the Panamint Range or in cor- Peak Fm. to Cryogenian strata elsewhere is the presence of two
relative strata from eastern Death Valley. Numerous attempts have carbonate units that overlie each of the diamictite-bearing units,
been made to isolate primary magnetizations, but have proved the Sourdough Member of the Kingston Peak Fm. and the Sentinel
unsuccessful, mostly displaying Mesozoic overprinting (J. Kirsch- Peak Member of the Noonday Dolomite. Although direct age
vink 2007, pers. comm.). control is lacking, the lithological and chemostratigraphical
characteristics of these two units permit correlations to older and
younger Crygenian cap carbonates known elsewhere (Prave
Geochronological constraints 1999; Petterson 2009). If correct, then the strongly positive d13C
values recorded in the Thorndike Sub-member suggest temporal
The geochronological constraints that exist for the Precambrian correlation with the Trezona anomaly (Halverson et al. 2005).
sedimentary succession in Death Valley are robust, but very The timing of the older Cryogenian glacial– cap carbonate
sparse. The maximum age of the section comes from diabase couplet (previously commonly referred to as ‘Sturtian’) remains
sills in the Crystal Spring Fm., which also occur as clasts in the unresolved (ages range from c. 750 to c. 643 Ma, and may
Limekiln Spring Member. Samples of the sills from two locations consist of multiple glaciations, Kendall et al. 2009) but the
in eastern Death Valley yield U –Pb baddeleyite ages of younger one (previously commonly referred to as ‘Marinoan’) is
1087 + 3 Ma and 1069 + 3 Ma (Heaman & Grotzinger 1992). now known to be c. 635 Ma in age (Hoffmann et al. 2004).
The Precambrian –Cambrian boundary occurs nearly 5 km above This global correlation suggests that the Kingston Peak Fm.
the Crystal Spring Fm., within the lower member of the Wood spans a significant interval of time without deposition of diamictite
Canyon Fm. (Corsetti & Hagadorn 2000). Chemostratigraphic or other glaciogenic rocks, as evident from the c. 300-m-thick
and lithological correlations to well-dated sections in Namibia South Park Member (exclusive of the Wildrose Sub-member).
(Hoffmann et al. 2004) may indicate an age of 635 Ma for the This permits defining at least two distinct intervals of glaciogenic
base of the Noonday Dolomite (Prave 1999; Petterson 2009). sedimentation, between which deposition of braided fluvial to
shallow marine carbonate units took place (Petterson 2009). In
the eastern Death Valley sections, the South Park strata are not
Discussion recognized, thereby raising the inevitable questions as to how
the sections in eastern Death Valley correlate to those in the
As noted previously, the palaeogeographic setting of the Kingston Panamint Range.
Peak Fm. is inferred to be one in which episodic extensional tec- Alternative global correlations have also been proposed for the
tonism influenced basinal development, replacing the relatively Kingston Peak Fm. In one interpretation, the Beck Spring Fm. is
stable carbonate platform setting of the underlying Beck Spring considered to be an older cap carbonate (Halverson et al. 2005),
Dolomite with the predominantly clastic deposition of the King- implying that the entire Kingston Peak is part of the younger Cryo-
ston Peak Fm. (Wright et al. 1974; Labotka et al. 1980). The King- genian glacial. In another, the Kingston Peak Fm.– Sentinel Peak
ston Peak Fm. in the Panamint Range does contain evidence for Member is inferred to be the older Cryogenian glacial–cap carbon-
deposition occurring along a surface that had pre-existing palaeo- ate couplet (Corsetti et al. 2007). All these ideas are currently
topography, which was infilled during deposition of the Limekiln being tested.
Spring and Surprise members. For example, in the Telescope Peak
area, a subaerially exposed basement high (the ‘World Beater We acknowledge support for this research from NSF grants EAR-01-07123 and
Island’ of Labotka et al. 1980) was postulated to have existed EAR-03-10413. This represents a contribution of the IUGS- and UNESCO-
during lower Kingston Peak time. However, the uniform character funded IGCP (International Geoscience Programme) project #512.
and widespread development of the overlying Sourdough Member
and Middle Park Sub-member implies a more uniform substrate References
and period of relative tectonic quiescence. Subsequently, a
renewed phase of tectonism during Mountain Girl Sub-member Albee, A. L., Labotka, T. C., Lanphere, M. A. & McDowell, S. D.
time resulted in tilting and development of an angular unconfor- 1981. Stratigraphy, structure, and metamorphism in the central Pana-
mity with associated locally developed coarse-grained facies (Pet- mint Mountains (Telescope Peak quadrangle), Death Valley area,
terson 2009). This gave way to a setting where widespread California: Summary. Geological Society of America Bulletin, 91,
development of carbonate strata occurred, as evident from the 125– 129.
Thorndike Sub-member. It was during these phases of relative tec- Awramik, S. M., Corsetti, F. A. & Shapiro, R. 2000. Stromatolites and
tonic quiescence that the Wildrose Sub-member was deposited and the pre-Phanerozoic to Cambrian history of the area south east of
overlain by the Noonday Dolomite. Death Valley. Bulletin of the San Bernardino County Museum, 47,
Evidence for a direct influence of glaciation in the Kingston 65 – 74.
Peak Fm. in the Panamint Range is sparse. A glaciogenic origin Burchfiel, B. C. & Davis, G. A. 1972. Structural framework and evol-
has been suggested for parts of the Limekiln Spring Member, the ution of the southern part of the Cordilleran orogen, western United
Surprise Member and the Wildrose Sub-member based on the pres- States. American Journal of Science, 272, 97 –118.
Carlisle, D., Kettler, R. M. & Swanson, S. C. 1980. Uranium- and
ence of laterally extensive and uniformly massive diamictic units,
thorium-bearing facies of the late Proterozoic Kingston Peak For-
the occurrence of rare lonestones and bullet-shaped and striated
mation, Death Valley region, California. In: Fife, D. L. & Brown,
clasts, and the even rarer presence of clasts that can be observed A. R. (eds) Geology and Mineral Wealth of the California Desert.
to pierce laminae. In the eastern Death Valley sections, striated South Coast Geological Society, Santa Ana, CA, 31 –51.
and faceted lonestones and dropstones are more abundant Cloud, P., Wright, L. A., Williams, E. G., Diehl, P. E. & Walter, M. R.
(Troxel 1982, p. 62; Miller 1985; Miller et al. 1988, fig. 6; 1974. Giant stromatolites and associated vertical tubes from the Upper
Mrofka & Kennedy 2011). However, in both regions, no direct Proterozoic Noonday Dolomite, Death Valley region, Eastern Califor-
influence of ice on deposition has yet been documented, such as nia. Geological Society of America Bulletin, 85, 1869–1882.
striated pavements at the base of diamictite units or glaciotectonic Corsetti, F. A. & Grotzinger, J. P. 2005. Origin and significance of tube
deformation features. Nevertheless, due to the difficulty of gener- structures in Neoproterozoic post-glacial cap carbonates: Example
ating sediment gravity flows of the dimensions shown by the dia- from Noonday Dolomite, Death Valley, United States. Palaois, 20,
mictic units by non-glacial mechanisms, deposition is presumably 348– 362.
KINGSTON PEAK FORMATION, PANAMINT RANGE 465

Corsetti, F. A. & Hagadorn, J. W. 2000. Precambrian – Cambrian tran- Levy, M. & Christie-Blick, N. 1991. Tectonic subsidence of the early
sition: Death Valley, United States. Geology, 28, 299– 302. Palaeozoic passive continental margin in eastern California and
Corsetti, F. A. & Kaufman, A. J. 2003. Stratigraphic investigations southern Nevada. Geological Society of America Bulletin, 103,
of carbon isotope anomalies and Neoproterozoic ice ages in Death 1590–1606.
Valley, California. Geological Society of America Bulletin, 115, Miller, J. M. G. 1985. Glacial and syntectonic sedimentation; the upper
916– 932. Proterozoic Kingston Peak Formation, southern Panamint Range,
Corsetti, F. A., Awramik, S. M. & Pierce, D. 2003. A complex micro- eastern California. Geological Society of America Bulletin, 96,
biota from snowball Earth times: microfossils from the Neoprotero- 1537–1553.
zoic Kingston Peak Formation, Death Valley, USA. Proceedings of Miller, J. M. G. 1987. Palaeotectonic and stratigraphic implications
the National Academy of Sciences of the United States of America, of the Kingston Peak-Noonday contact, Panamint Range, eastern
100, 4399–4404. California. Journal of Geology, 95, 75– 85.
Corsetti, F. A., Link, P. K. & Lorentz, N. J. 2007. d13C chemostratigra- Miller, J. M. G., Troxel, B. W. & Wright, L. A. 1988. Stratigraphy and
phy of the Neoproterozoic succession near Pocatello, Idaho, U.S.A.: palaeogeography of the Proterozoic Kingston Peak Formation, Death
implications for glacial chronology and regional correlations. In: Valley region, eastern California. In: Gregory, J. L. & Baldwin,
Link, P. K. & Lewis, R. S. (eds.) Proterozoic Geology of Western E. J. (eds) Geology of the Death Valley Region. South Coast Geologi-
North America and Siberia. Society for Sedimentary Geology cal Society, Santa Ana, CA, 118–142.
Special Publication, 86, 193–205. Mrofka, D. & Kennedy, M. 2011. The Kingston Peak Formation in
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice, the eastern Death Valley Region. In: Arnaud, E., Halverson,
A. H. N. 2005. Toward a Neoproterozoic composite carbon-isotope G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
record. Geological Society of America Bulletin, 117, 1181–1207. terozoic Glaciations. Geological Society, London, Memoirs, 36,
Hammond, J. L. G. 1983. Late Precambrian diabase intrusions in the 449– 458.
southern Death Valley region, California; their petrology, geochem- Murphy, F. M. 1932. Geology of a part of the Panamint Range,
istry, and tectonic significance. PhD thesis, University of Southern California, Report XXVII of the State Mineralogist. California
California. Department of Natural Resources, Division of Mines and Geology,
Heaman, L. M. & Grotzinger, J. P. 1992. 1.08 Ga Diabase sills in 28, 329– 356.
the Pahrump Group, California: implications for development of Petterson, R. 2009. I. Glaciogenic and related strata of the Neopro-
the Cordilleran miogeocline. Geology, 20, 637– 640. terozoic Kingston Peak Formation in the Panamint Range,
Hewett, D. F. 1940. New formation names to be used in the Kingston Death Valley region, California. II. The basal Ediacaran Noonday
Range, Ivanpah Quadrangle, California. Journal of the Washington Formation, and implications for Laurentian equivalents. III. Rifting
Academy of Sciences, 30, 239–240. of southwest Laurentia during the Sturtian-Marinoan intergla-
Hodges, K. V., McKenna, L. W. et al. 1989. Evolution of extensional cial: The Argenta orogeny. PhD thesis, California Institute of
basins and Basin and Range topography west of Death Valley, Technology.
California. Tectonics, 8, 453–467. Pierce, D. & Cloud, P. 1978. New microbial fossils from 1.3 billion-
Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L. year-old rocks of eastern California. Geomicrobiology Journal, 1,
2004. U– Pb zircon date from the Neoproterozoic Ghaub Formation, 295– 309.
Namibia: constraints on Marinoan glaciation. Geology, 32, 817– 820. Prave, A. R. 1999. Two diamictites, two cap carbonates, two d13C excur-
Horodyski, R. J. & Mankiewicz, C. 1990. Possible Late Proterozoic sions, two rifts: the Neoproterozoic Kingston Peak Formation, Death
skeletal algae from the Pahrump-Group, Kingston Range, southeast- Valley, California. Geology, 27, 339–342.
ern California. American Journal of Science, 290A, 149– 169. Stewart, J. H. 1970. Upper Precambrian and Lower Cambrian strata in
Hunt, C. B. & Mabey, D. R. 1966. General geology of Death Valley, the southern Great Basin, California and Nevada. US Geological
California; stratigraphy and structure. US Geological Survey Pro- Survey Professional Paper, 620, 206.
fessional Paper, 494-A, 162. Stewart, J. H. 1972. Initial deposits in Cordilleran geosyncline: evidence
Johnson, B. K. 1957. Geology of a part of the Manly Peak Quadrangle, of a Late Precambrian (less than 850 My) continental separation.
southern Panamint Range, California. University of California Publi- Geological Society of America Bulletin, 83, 1345– 1360.
cations in Geological Sciences, 30, 353– 423. Troxel, B. W. 1982. Description of the uppermost part of the
Kendall, B., Creaser, R. A., Calver, C. R., Raub, T. D. & Evans, Kingston Peak Formation, Amargosa Rim, Canyon, Death
D. A. D. 2009. Correlation of Sturtian diamictite successions in Valley region, California. In: Cooper, J. D., Troxel, B. W. &
southern Australia and northwestern Tasmania by Re– Os black Wright, L. A. (eds) Geology of Selected Areas in the San Bernardino
shale geochronology and the ambiguity of ‘sturtian’-type diamic- Mountains, Western Mojave Desert, and Southern Great Basin,
tite – cap carbonate pairs as chronostratigraphic marker horizons. California. The Death Valley Publishing Company, Shoshone, CA,
Precambrian Research, 172, 301– 310. 61– 70.
Labotka, T. C. 1978. Geology of the Telescope Peak Quadrangle, Cali- Wernicke, B., Snow, J. K., Hodges, K. V. & Walker, J. D. 1993.
fornia and late Mesozoic regional metamorphism, Death Valley Structural constraints on Neogene tectonism in the southern
area, California. PhD thesis, California Institute of Technology. Great Basin. In: Lahren, M. M., Trexler, J. H. Jr. & Spinosa, C.
Labotka, T. C., Albee, A. L., Lanphere, M. A. & Mcdowell, S. D. (eds) Crustal Evolution of the Great Basin and Sierra Nevada.
1980. Stratigraphy, structure, and metamorphism in the Central Geological Society of America Cordilleran/Rocky Mountain
Panamint Mountains (Telescope-Peak Quadrangle), Death-Valley Sections Meeting Fieldtrip Guidebook, University of Nevada,
Area, California – summary. Geological Society of America Bulletin, Reno, 453– 479.
91, 125– 129. Wright, L. A., Troxel, B. W., Williams, E. G., Roberts, M. T. &
Labotka, T. C., Warasila, R. L. & Spangler, R. R. 1985. Polymeta- Diehl, P. E. 1974. Precambrian sedimentary environments of the
morphism in the Panamint Mountains, California: a Ar39 –Ar40 Death Valley region, eastern California. Guidebook, Death Valley
study. Journal of Geophysical Research – Solid Earth and Planets, Region, California and Nevada. Death Valley Publishing Company,
90, 359– 371. Shoshone, CA, 27 –36.
Chapter 42

The deep-marine glaciogenic Gaskiers Formation, Newfoundland, Canada

SHANNON L. CARTO* & NICK EYLES


Department of Geology, University of Toronto, Toronto, Ontario, M1C 1A4 Canada
*Corresponding author (e-mail: shannon.carto@gmail.com)

Abstract: In eastern Canada, the Neoproterozoic Gaskiers Formation (Fm.) consists of a thick diamictite– turbidite succession (250–
300 m thick) that occurs within the deep-marine predominantly volcaniclastic turbidite units of the Conception Group (4– 5 km thick).
These rocks are well exposed on the coast of the Avalon Peninsula in eastern Newfoundland. The thick succession is considered by some
to represent the final Neoproterozoic glacial event, known as the Gaskiers Glaciation c. 582– 585 Ma. The Gaskiers Fm. appears to have
accumulated in a volcanically-influenced arc-related basin within the confines of the Proterozoic peri-Gondwanan Avalon terrane, before
rifting away from the Gondwanan margin in the early Palaeozoic, but it is not clear whether it was adjacent to the West African Craton or
the Amazon Craton. It has also been argued that Avalon had already rifted from Gondwana by the late Proterozoic.
Glacially striated clasts, dropstones, and chatter-marked garnets identify a glaciated source area, and clasts and matrix are of volcanic
origin; pyroclastic flows and volcanic bombs are also present. As a result, there is a growing consensus that the Gaskiers Fm. records local
glaciation of a high relief volcanic topography, owing its origin to the episodic downslope reworking of volcanic and glacial debris into a
deep, rapidly subsiding basin. The volcanic nature of these deposits implies that local volcanic activity was coeval with deposition. A
minimum age of the Gaskiers Fm. is constrained by a diverse assemblage of Ediacaran-type fossils in the upper Conception Group
dated at c. 565 + 3 Ma and in the lower St. John’s Group, which overlies the Conception Group. A maximum age for the Gaskiers is
provided by a U –Pb date of 606 þ3.7/ –2.9) Ma in the Harbour Main volcanic rocks underlying the Gaskiers Fm. A thin limestone
bed has also been identified at two localities in Conception Bay directly overlying the diamictite yielding strongly negative d13C values.

The Neoproterozoic Gaskiers Fm. is a 250 –300-m-thick succes- debris flows composed of glaciogenic debris transported down-
sion of glaciogenic diamictite with interbedded turbidite units, slope from a nearby grounded ice margin. A detailed lithofacies
which crop out on the coast of the southern Avalon Peninsula of analysis by Eyles & Eyles (1989) similarly concluded that the
eastern Newfoundland, Canada (Fig. 42.1). The Gaskiers Fm. Gaskiers Fm. was deposited as debris flows and turbidites in a
occurs towards the base of the 4– 5-km-thick succession of deep- deep-marine, volcanically influenced, mid-slope setting. This
marine volcaniclastic sedimentary rocks belonging to the Con- interpretation drew on the depositional context provided by turbi-
ception Group (Fig. 42.2a; Williams & King 1975). There are dite strata present conformably within, above and below the
excellent cliff exposures south of Cape St. Francis, along the Gaskiers Fm. indicating a deep-marine slope setting (Gardner &
southern margin of Conception Bay, in the south on the eastern Hiscott 1988). Because of the reworked deep-marine nature of the
shoreline of St. Mary’s Bay, and on the Colinet Islands. Outcrops Gaskiers Fm., it is difficult to infer much of the parent glacial
in the St. Mary’s Bay area in the vicinity of Double Road Point are environment onland other than to positively identify the presence
widely and informally regarded as the type section (Figs 42.1 & of wet-based glaciers on the peaks of an active volcanic arc.
42.2b). There, the diamictite comprises more than 80% of the The Gaskiers Fm. is well known for the associated diverse
total thickness of the exposed section, and is interbedded with tur- assemblage of Ediacara-type fossils present through the upper
bidite units containing dropstones as well as thin volcanic agglom- Drook and Mistaken Point formations dated at c. 565 + 3 Ma
erate units; volcanic bombs are also present (Fig. 42.2b; Eyles & (Benus 1988; Clapham & Narbonne 2002; Narbonne 2004) and
Eyles 1989; Myrow 1995). Within the Avalon terrane of New- in the lower St. John’s Group with U –Pb dates from 565 to
foundland, a number of other diamictite deposits have been con- 543 Ma (Myrow 1995). Moreover, recent analysis of the Fe
sidered to be potential time-equivalents to the Gaskiers Fm., content of the deep-marine sedimentary rocks of the Conception
such as within the Musgravetown Group, Rock Harbour Group Group and overlying St. John’s Group by Canfield et al. (2007)
and Connecting Point Group but no evidence has yet been ident- reveals that anoxic conditions persisted in the deep ocean before
ified that these are coeval or glacially influenced (Myrow 1995). and during the deposition of the Gaskiers Fm., subsequently
The Gaskiers Fm. is commonly considered to represent the last becoming oxic afterward. Canfield and colleagues argue that this
of the major Neoproterozoic glaciations known as the Gaskiers is suggestive of a causal link between the evolution of Edicaran
Glaciation c. 582 –585 Ma (Bowring et al. 2003; Knoll et al. biota and this oxygenation event.
2006). The Gaskiers diamictite units were labelled as ‘tilloids’
by McCartney (1967) in the sense of ‘non-glacial conglomeratic
mudstone resulting from extensive slides, slumps or mudflows’. Structural framework
Bruckner & Anderson (1971) identified striated and faceted
pebbles and Williams & King (1979) discovered outsized drop- The Avalon terrane of eastern Newfoundland, Canada, is com-
stones within the interbedded siltstone and sandstone laminations, posed of c. 15-km-thick relatively unmetamorphosed Proterozoic
demonstrating a glaciogenic source of debris. Other evidence of a sedimentary, volcanic and plutonic rocks (c. 750–570 Ma)
glacial origin rests on the presence of ‘chatter-marked’ garnets locally overlain by Cambrian – Ordovician (excluding the Carbon-
(Gravenor 1980) and ‘flat-iron’, bullet-shaped clasts produced by iferous sedimentary rocks of the Burin Peninsula) sedimentary
subglacial abrasion in the source area (Bruckner & Anderson rocks (King 1988). Murphy & Nance (1989) cited the close associ-
1971). Williams & King (1975) used the term ‘mixtite’, introduced ation of calc-alkaline and rift-related continental volcanic rocks as
by Schermerhorn (1966) for a ‘coarse, poorly sorted clastic rock evidence that the Avalon terrane developed in a volcanic-arc
without regard for composition or origin’ and interpreted the dia- setting along the active continental margin of West Gondwana
mictite units as largely the product of ice rafting. Gravenor (O’Brien et al. 1983; Nance et al.1991; McMenamin 1987;
(1980) and King (1988) argued that the Gaskiers deposits were Burrett et al. 1990; Forty & Cocks 2003). The abundance of

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 467– 473. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.42
468 S. L. CARTO & N. EYLES

(a)

Cape St. Francis

ay
nB
p tio
ce
n
Co
St. John’s (a)
Kitchuses

Avalon Harbour
Peninsula Main

Little
Colinet (b)
Island

Great
Colinet St. Mary’s
Island 46°55'

St. Mary’s Gaskiers


(c) Double Road Point
Bay Outcrop
D
0 km 20

M 40

St. Mary’s
47° 10' 90

(b)
M Mall Bay

Little
Colinet Gaskiers Point La Haye
Island
ne

Fig. 42.1. (a) General map of the Avalon


ticli
ne

Great
D Drook
ticli

Colinet
An

Island Peninsula in eastern Newfoundland, Canada


An

ce

V showing location of principal outcrops of


Pla
e Pt.

ing

the stratigraphic units of the Conception


Hay

ss

South
Cro

Group. (b) Map showing location of both


La

Point (c) Gaskiers


N
St. Mary’s
hidden and exposed outcrops of the
St. Mary’s
Bay Gaskiers Fm Gaskiers Fm. in the St. Mary’s Bay area,
Gaskiers Hidden
Holyrood Outcrop 90 Holyrood eastern Newfoundland, Canada. (c)
Pond Pond Enlargement of St. Mary’s Bay area
False 0 km 10 0 km 1 showing location of outcrops of Mall Bay,
Cape
St. Vincents Drook and Gaskiers formations on the east
46° 46' 46°52' side of St. Mary’s Bay (modified from Eyles
53°45' 53°20' 53°37' 53°34' & Eyles 1989).

bimodal Avalonian volcanic suites underlying the Conception inconsistencies observed in the Drook and Mistaken Point for-
Group, the high level of plutonic rocks and the abundance of ash mations exposed in the Bay Roberts and Spaniard’s Bay area, in
layers within the succession suggests the Gaskiers Fm. accumu- the north-central part of the Avalon Peninsula (palaeoflow direc-
lated in an arc-related basin (Myrow 1995). Possible basin types tion towards the east and SE) and exposures of these units at Mis-
include intra-arc (Dec et al. 1992), back-arc (Myrow et al. 1999) taken Point in southern Avalon (palaeoflow toward the NE). This
and fore-arc (Narbonne et al. 2001) basins. Based on petrological, topographical high is believed to be responsible for steering the
structural and geochronological data collected from the Con- turbidity currents in the NE direction along the axis of the basin
ception Group, it has been argued that the Gaskiers Fm. most throughout deposition of most of the Drook, Briscal and Mistaken
likely accumulated during the transition from a fore-arc (c. 630– Point formations in the southern Avalon Peninsula.
560 Ma) to strike –slip pull-apart basin (c. 560 –540 Ma) The rocks of the Avalon Peninsula were deformed during the
(Murphy et al. 1999; Nance et al. 2002; Wood et al. 2003). The late Proterozoic Avalon orogeny and subsequently deformed and
Fermeuse Fm. is thought to mark this change in tectonic regime metamorphosed during the mid-Palaeozoic Acadian orogeny
with the remainder of the St. John’s and Signal Hill groups (Hatcher & Goldberg 1991). The Acadian orogeny produced
accumulating in the strike –slip pull-apart basin. Support for this most of the folding, faulting and metamorphism recognized on
theory stems from the scarcity of volcanic ash in the Fermeuse the eastern Avalon Peninsula. The rocks of the Avalon Peninsula,
Fm. and higher stratigraphic units, indicating the cessation of including the diamictite units, have been metamorphosed under
volcanic activity expected if subduction no longer occurred. prehnite-pumpellyite facies, but sedimentary structures are very
Recent work by Ichaso et al. (2007) suggests that the present- well preserved (Papezik 1974). Outcrops of the Conception
day outcrops of the Harbour Main volcanic rocks may have been Group in the southern Avalon Peninsula on the east side of
a topographical high that segmented the Avalonian fore-arc St. Mary’s Bay have been folded into anticlines and synclines
region, creating a relatively confined sub-basin (termed the ‘west that have NE-trending axes (Eyles & Eyles 1989). The rocks of
Conception sub-basin’). This theory is based on palaeocurrent the Avalon Peninsula, including the Gaskiers Fm., are cleaved
THE GASKIERS FORMATION 469

Fig. 42.2. Schematic representation of


stratigraphy of the eastern Avalon Peninsula
and a detailed log through the Gaskiers Fm.
and associated Mall Bay and Drook
formations at Double Road Point, east side
of St. Mary’s Bay (see Fig. 40.1c for
location of detailed log; modified from
Eyles & Eyles 1989).

and cut by numerous faults, although the extent to which the the Holyrood plutonic complex of granite and gabbro (Anderson &
Gaskiers Fm. has been deformed varies throughout the region King 1981).
(King 1980). Deformation from the Acadian Orogeny is over- The Conception Group (4– 5 km thick) is divided into the Mall
printed by compression from the Alleghanian Orogeny (290 – Bay, Gaskiers, Drook, Briscal and Mistaken Point formations
250 Ma) in the Appalachian region (Osberg et al.1989; Rast 1989). (Fig. 42.2), but there is little lithofacies difference between these
siliceous and volcanogenic deep-water turbidite units, which
contain occasional tuff, rhyolite flows, agglomerates, mafic
Stratigraphy pillow lavas and mafic dykes (King 1988). The deep-marine silic-
eous turbidite units of the Mall Bay Fm. (.800 m) are overlain by
The Harbour Main Group (.1500 m thick) rests below the the diamictite-bearing Gaskiers Fm. (250 –300 m), which in turn
Conception Group and consists of mafic to acidic lavas, pyroclastic is overlain by the siliceous turbidite sequence of the Drook Fm.
rocks with interbedded continental sedimentary rocks, intruded by (1500 m) (King 1988). The Briscal Fm. (1200 m) is a grey to
470 S. L. CARTO & N. EYLES

green sandy thick-bedded unit that lies between the Drook and Drook Fm.
Mistaken Point formations in the southeastern Avalon Peninsula
(Ichaso et al. 2007). Pebbly mudstone diamictites similar to The Drook Fm. (1500 m) consists of a succession of medium- and
those of the Gaskiers Fm. occur in the overlying Drook and thick-bedded, fine-to coarse-grained, graded and massive sand-
Briscal formations, but these do not include striated clasts and stone units interbedded with graded argillite units, similar to
may be non-glaciogenic. The Mistaken Point Fm. (400 m) consists those of the Mall Bay Fm. (Eyles & Eyles 1989). Outcrops of
of a deep-marine siliclastic succession of graded beds composed the Drook Fm. are widely exposed in eastern St. Mary’s Bay, Gask-
of grey to pink sandstone units and green, red and purple shale iers, St. Vincent’s and Cape English, but the outcrop at Double
units (King 1988). The upper part of the formation is composed Road Point is poor (Eyles & Eyles 1989; King 1988). Soft-
of finely laminated light yellow to greenish grey tuffaceous sediment deformation structures and slump folds are common
layers (King 1988). Fossiliferous horizons are present throughout (Eyles & Eyles 1989). Beds vary in thickness from a few milli-
the formation. metres to 10 cm. Load-flame structures are moderately abundant,
The Conception Group is overlain conformably by the and ash beds are present but in small amounts (Ichaso et al. 2007).
8-km-thick molasses facies Hodgewater Group in the central
Avalon Peninsula and by the 8 –9-km-thick St. John’s and Signal
Hill groups to the east (Fig. 42.2). The St. John’s Group is domi- Associated volcanic rocks
nated by black shale, slate and silty sandstone. The Signal Hill
Group consists of grey sandstone and siltstone units at the base, Evidence of contemporaneous volcanic activity is present at
passing up into red sandstone and siltstone units, which are over- Double Road Point outcrop in the form of air-fall bombs, and
lain by red conglomerate and sandstone units (O’Brien et al. agglomerate and tuff layers. Beds of weathered volcanic agglom-
1983). These record shoaling of the basin prior to deposition of erates (,30 mm) occur within the thin-bedded, fine-grained sand-
extensive Cambrian –Ordovician shallow marine lithofacies stone units that are interbedded with the lower part of the
(Smith & Hiscott 1984; Myrow 1995). diamictite facies at this locality. Intercalations of tuff, as well as
a thin (,3 cm) bed of orange-weathering volcanic agglomerate
also occurs at Great Colinet Island (King 1980). A volcanic
Glaciogenic deposits and associated strata air-fall bomb, volcanic blocks and lapilli also occur within this
section (Anderson & King 1981; Eyles & Eyles 1989). Volcanic
Mall Bay Fm. ash beds, ranging from 0.2 to 30 cm in thickness, have been ident-
ified in low abundance in the Drook Fm. and in high abundance
The Mall Bay Fm. (.800 m thick) consists of mostly thin-bedded in the Mistaken Point Fm., exposed in the coastal sections in the
(10– 150 mm), fine-to-medium-grained, grey, brown and green Spaniard’s Bay –Bay Roberts area (Williams & King 1979;
graded sandstone units (Eyles & Eyles 1989). These units are Ichaso et al. 2007). Crystal tuffs are also present in the Trepassey
well exposed at Point La Haye and can be traced to the NE as a con- Fm., but are thinner and less common (Wood et al. 2003).
tinuous exposure up to the base of the overlying Gaskiers Fm. at
Double Road Point. Polygonal systems of clastic dykes, suggesting
rapid de-watering, occur on upper bedding surfaces of the Mall Associated carbonate
Bay Fm. (Eyles & Eyles 1989). Palaeocurrent directions are
southerly, and slump folds and other soft-sediment deformation Myrow & Kaufman (1999) reported a thin (,50 cm) limestone
structures are common (Eyles & Eyles 1989). bed and concretions overlying a thick diamictite succession
exposed discontinuously at Harbour Main and Kitchuses sites
in south Conception Bay. No limestone beds occur at the
The Gaskiers Fm. St. Mary’s Bay outcrops of the Gaskiers Fm. (or at other outcrops
on Little Colinet Island). The limestone at both localities have a
The Gaskiers Fm. is 250 –300 m thick and occurs conformably mottled nodular fabric with no preserved primary sedimentary or
between the Mall Bay and Drook formations (Fig. 42.2). biogenic structures, and are locally overprinted with tectonic
Detailed sedimentological analysis is provided by Eyles & fabrics (Myrow & Kaufman 1999). These outcrops were tenta-
Eyles (1989). The diamictite is mostly massive and unsorted, tively assigned to the Gaskiers Fm. on the basis of striated clasts,
with a distinct planar tabular geometry typical of debris flows. dropstones and evidence for deep-water conditions.
Some beds exhibit weak grading and are interbedded with thin
intraformational turbidite and tuff horizons (Anderson & King
1981). In some places, the diamictite shows subtle stratification, Boundary relations with overlying and underlying
suggesting the amalgamation of thinner beds. Beds of grey to non-glacial units
green thinly stratified and graded sandstone, siltstone and mud-
stone units separate diamictite beds, the latter of which are as Throughout the Avalon terrane, the Gaskiers Fm. is conformable
much as 40 m thick (Williams & King 1979; Eyles & Eyles with the underlying and overlying Mall Bay and Drook formations,
1989). These thinly stratified beds are continuous and can be respectively. Boundaries between the interbedded laminations and
traced for several tens of metres along strike, although some the underlying diamictite facies at the Double Road Point outcrop
beds are extensively folded and show evidence of slumping are generally sharp, planar contacts, but some units are extensively
(Eyles & Eyles 1989). Clasts are commonly 30–100 mm in folded (Eyles & Eyles 1989). Basal contacts of massive and
diameter, but rare larger clasts up to 0.8 m in diameter have crudely stratified diamictite facies are mostly sharp, planar or
also been observed (Myrow & Kaufman 1999). Clasts are com- gently undulating, and are locally erosive. The upper contacts of
posed of a wide variety of igneous lithologies such as basalt, diamictite units are sharp and planar (Eyles & Eyles 1989). The
granite, diorite and granophyre, as well as sedimentary litholo- Drook Fm. is conformably overlain by the Briscal Fm. at the
gies such as quartzite, siltstone and conglomerate fragments outcrop at Portugal Cove of the southern Avalon Peninsula
(Eyles & Eyles 1989). Intraclasts of mudstone, laminated silt- (King 1988). In turn, the Briscal Fm. in the southern Avalon
stone and carbonate are also present. Clast shape varies from Peninsula is conformably overlain by the Mistaken Point Fm.
well-rounded to highly angular; striated and faceted clasts are (Williams & King 1979). The Conception Group is overlain con-
common. The diamictite is also characterized by large-scale formably by the St. John’s Group and Signal Hill Group in the
slump structures. eastern Avalon Peninsula (e.g. Smith & Hiscott 1984).
THE GASKIERS FORMATION 471

Chemostratigraphy palaeolatitude of c. 348 þ88/– 78 (mean tilt-corrected inclination


of 538) at c. 580–570 Ma, similar to the palaeolatitude of the
The oxygenation state of the late Neoproterozoic global ocean was northern margin of the West African Craton (McNamara et al.
analysed by Canfield et al. (2007) by examining the Fe content 2001). These results were based on palaeomagnetic core samples
(reactive Fe:total Fe ratio, FeHR/FeT) of the deep-marine rocks collected from 21 sites, all of which yielded a positive conglomer-
in Newfoundland. Using Fe extraction techniques, Canfield et al. ate test, confirming a primary age of magnetization for the Marys-
discovered that both the upper Mall Bay Fm. and Gaskiers diamic- town Group as a whole (McNamara et al. 2001). Some samples
tite samples contained highly reactive FeHR/FeT ratios exceeding consisted of multiple specimens; magnetization directions from
0.38, indicative of anoxic deposition. Marine sediments deposited the samples within a site are averaged to produce the site mean
in oxygen-containing water columns are consistently below 0.38 of that site.
(Poulton & Raiswell 2002). The sedimentary rocks overlying the In contrast, Nance & Murphy (1996) have argued, based on Nd
Gaskiers diamictite yielded FeHR/FeT ratios lower than 0.38, isotopic data from crustally derived felsic igneous rocks of West
indicative of a long period of oxic marine conditions. Avalonia, that this terrane was close to the northern margin of
The isotopic composition of sulphide in pre-Gaskiers Fm. sedi- the Amazon Craton. Initial 1Nd values of West Avalonia (used to
ments was found to be greater than zero, indicating relatively small identify mantle source from an igneous rock) are strongly positive
fractionations from seawater sulphate caused by low rates of ( –0.4 to þ5.0), and the Tdm model age (age at which crust was
sulpate reduction under sulphate-limiting concentrations. Higher extracted from mantle) for West Avalonia is c. 0.8– 1.1 Ga.
fractionations were found in the Gaskiers Fm., and particularly Similar positive initial 1Nd values (þ0.2 to þ6.9) and similar
in the overlying Drook Fm., indicative of an increase in sulphate Tdm model ages of c. 0.9–1.2 Ga were identified in the Tocantins
concentration. To explain these results, Canfield et al. (2007) province of central Brazil (Pimental & Fuck 1992). As a result, a
suggest that increased oxygen enhanced the oxidative weathering palaeoposition adjacent to South America is favoured for West
of sulphide to sulphate on the continents, thus increasing the flux Avalonia during the Neoproterozoic. Samson et al. (2005) has
of sulphate to the ocean. also shown that the main age groups of detrital zirons from Avalo-
Isotopic analyses of a thin limestone bed overlying Gaskiers Fm. nia are predominantly Mesoproterozoic in age (c. 1.25– 1.15 Ga
deposits discovered at the Harbour Main and Kitchuses sites have and 1.65– 1.50 Ga), similar to the age pattern of detrital zircons
yielded negative values of d13C (– 7.5 to – 1.5‰), and has been found in the Amazonian Craton. Mesoproterozoic crust has not
interpreted as slope-deposited carbonate rocks that accumulated been identified in the West African Craton.
in response to sea-level rise (Myrow & Kaufman 1999). Other palaeogeographical reconstructions place Avalonia near
Laurentia at either a low latitude very near to the equator, or at a
high-latitude position in the southern hemisphere at c. 580 Ma
(Symons & Chaisson 1991; Trindade & Macouin 2007), but
Other characteristics
these reconstructions are incompatible with strongly different
Cambrian –Ordovician biotas in the two regions (e.g. Rehmer
A diverse assemblage, at least 30 species (Anderson & Conway
1981; Conway Morris & Rushton 1988).
Morris 1982) of Ediacaran-type fossils, exists within the upper
Conception Group (Drook and Mistaken Point formations) and
lower St. John’s Group (Trespassey and Fermeuse formations)
(Fig. 42.2a; Myrow 1995). The fossil assemblages are also Geochronological constraints
known from Charnwood Forest in central England (Jenkins
1992). As a result, a full taxonomic description has never been A maximum age for the Gaskiers Fm. is constrained by U – Pb
presented. At present, the following Ediacaran taxa have been zircon ages of 606.8 þ3.7/–2.9 Ma, 622.6 þ2.3/–2.0 Ma and
reported from the Mistaken Point Fm. (Bradgatia, Clapham 631 + 2 Ma from the underlying basement rocks of the underlying
et al. 2004; Charnia masoni, Anderson 1978; Ivesia, Clapham Harbour Main Group (Krogh et al. 1988). Zircons from ash beds in
et al. 2004; Thectardis avalonensis, Clapham et al. 2004). The the Ediacaran fossil-bearing Mistaken Point Fm. of the upper
Mistaken Point Fm. has yielded a date of 565 + 3 Ma for a vol- Conception Group, on the eastern Avalon Peninsula, have yielded
canic ash that covers the most fossilerous surface near the top a U –Pb date of 565 + 3 Ma (Benus 1988; Clapham et al. 2004;
of the formation (Benus 1988). Two species of the Ediacaran Narbonne 2004). The main age groups of detrital zirons from the
frond Charnia – C. masoni and C. wardi – represent the most basement rocks of Avalonia are predominantly Mesoproterozic
complex of the four species in the Drook Fm. (Narbonne & in age c. 1.25–1.15 Ga and 1.65 –1.50 Ga (Samson et al. 2005).
Gehling 2003).

Discussion
Palaeolatitude and palaeogeography
Traditionally, the Gaskiers Fm., which is conformable with the
The common view is that along with other exotic terranes, Avalo- overlying and underlying marine sedimentary sequences of the
nia originated along the margin of Gondwana, rifted from the Mall Bay and Drook formations, has been interpreted as the sedi-
Gondwanan margin in the early Palaeozoic and tectonically mentary product of floating ice in a marine shelf environment
accreted to eastern Laurentia in the mid-Palaeozoic (Nance & (King 1988). A glacial origin for the Gaskiers Fm. rests heavily
Murphy 1996). There is debate as to whether the Avalon terrane on the presence of dropstones, chatter-marked garnets, and striated
was adjacent to the West African or Amazonian margins of Gond- and faceted stones in the diamictite (Williams & King 1979; King
wana (Rast & Skehan 1983; Dalziel et al. 1994). In contrast, 1980). In addition to glaciogenic sediments, the Gaskiers Fm. also
Landing (1996) has argued that there is no lithological, biotic or contains pyroclastic and tuffaceous debris, as well as volcanic air
geological evidence that supports the contention that Avalon was fall bombs and interbeds of agglomerate flows indicating volcanic
contiguous with the margins of West Gondwana in the late Proter- eruptions coeval with glaciation (Eyles & Eyles 1989; Ichaso et al.
ozoic –Cambrian. Instead, Landing (1996) argued that Avalon was 2007). The Mall Bay and Drook formations have also been inter-
an insular continent by the Early Palaeozoic, having already rifted preted as base-of-slope or basin plain turbidite deposits (Gardiner
from Gondwana by the late Proterozoic. Palaeomagnetic investi- & Hiscott 1988); the same turbidite facies are interbedded with
gation of the Neoproterozoic Marystown Group (Burin Peninsula) the diamictite of the Gaskiers Fm. As a result, a clear picture
of southeastern Newfoundland suggests western Avalonia had a emerges of a deep basin receiving large volumes of volcaniclastic
472 S. L. CARTO & N. EYLES

sedimentary rocks that during glaciation was subjected to episodic In: Landing, E., Narbonne, G. M. & Myrow, P. (eds) Trace
downslope resedimentation of muddy glaciogenic sediment. It is Fossils, Small Shelly Fossils and the Precambrian– Cambrian
highly likely that repeated episodes of submarine mass wasting Boundary. New York State Museum and Geological Survey Bulletin
and downslope flow of glacially sourced debris may have been 463, 8– 9.
induced by slope failure, triggered by volcanic eruptions and Bowring, S. A., Myrow, P. M., Landing, E. & Ramezani, J. 2003.
melting of ice cover, producing mudflows akin to modern-day Geochronological constraints on terminal Neoproterozoic events
lahars (Fisher 1984). and the rise of metazoans. NASA Astrobiology Institute General
The Conception Group is one of several Neoproterozoic succes- Meeting Abstracts, 113– 114.
Bruckner, M. D. & Anderson, M. M. 1971. Late Precambrian glacial
sions deposited between 620 and c. 570 Ma in arc-related basins
deposits in southeastern Newfoundland, a preliminary note.
marginal to Gondwana, reflecting oblique subduction below the Geology Association of Canada, Proceedings, 24, 95– 102.
Gondwanan margin (Murphy et al. 2004). They are all character- Burrett, J., Long, J. A. & Stait, B. 1990. Early– Middle Palaeozoic bio-
ized by great thicknesses of volcanogenic turbidite, diamictite geography of Asian terranes derived from Gondwana. In: McKer-
and volcanic units. Debate has long surrounded the glacial status row, W. S. & Scotese, C. R. (eds) Palaeozoic Palaeogeography
of some age-equivalent diamictite units within several basins and Biogeography. Geological Society, London, Memoirs, 12,
such as in Armorica (Eyles 1990) and eastern North America 163– 174.
(Squantum Tillite; Carto & Eyles 2011). The case for a glaciated Canfield, D. E., Poulton, S. W. & Narbonne, G. M. 2007. Late
hinterland can be convincingly made for the Gaskiers Fm. with Neoproterozoic deep ocean oxygenation and the rise of animal life.
its numerous dropstones and striated clasts. The glacial origin of Science, 315, 92 – 94.
the age-equivalent deposits are arguable indicating that the Carto, S. L. & Eyles, N. 2011. The Squantum Member of the Boston
extent of glaciation around the outer volcanic Gondwanan Basin, Massachusetts, USA. In: Arnaud, E., Halverson, G. P. &
margin (during the so-called Gaskiers Glaciation) was probably Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
limited in extent, topographically constrained to volcanic peaks Glaciations. Geological Society, London, Memoirs, 36, 475–480.
and unrelated to any global cooling event. The Quaternary and Clapham, M. E. & Narbonne, G. M. 2002. Ediacaran epifaunal tiering.
modern Earth affords numerous examples of glaciated volcanic Geology, 30, 627–630.
peaks (Ehlers & Gibbard 2004). These ice masses do not extend Clapham, M. E., Narbonne, G. M., Gehling, J. G., Greentree, C. &
to sea level, but poorly sorted glaciogenic debris may be swept Anderson, M. M. 2004. Thectardis avalonensis: a new Ediacaran
into marine basins by lahars to be preserved as deep-marine fossil from the Mistaken Point biota, Newfoundland. Journal of
debris flows intercalated with thick turbidite successions. The pres- Paleontology, 78, 1031– 1036.
Conway Morris, S. & Rushton, A. W. A. 1988. Precambrian to Trema-
ence of non-glaciogenic debrites (lahars?) intercalated with turbi-
doc biotas in the Caledonides In: Harris, A. L. & Fettes, D. J. (eds)
dite units of the Mall Bay and Drook formations, above and below The Caledonian– Appalachian orogen. Geological Society, London,
the Gaskiers Fm. respectively, indicates continuity of marine Special Publications, 38, 93– 109.
depositional processes through time (Eyles & Eyles 1989). Sub- Dalziel, I. W. D., Dalla Salda, L. H. & Gahagan, L. M. 1994.
marine debris flow and turbidity current activity continued Paleozoic Laurentia – Gondwana interaction and the origin of the
throughout, but only within the Gaskiers Fm. was poorly sorted Appalachian – Andean mountain system. Geological Society
glacial debris with striated clasts supplied to the marine margin. American Bulletin, 106, 243–252.
The Ediacaran biota found in the Drook and Mistaken Point for- Dec, T., O’Brian, S. J. & Knight, I. 1992. Late Precambrian volcani-
mations and lower St. John’s Group (Trespassey and Fermeuse for- clastic deposits of the Avalonian Eastport Basin (Newfoundland
mations) (Myrow 1995) are thought to represent the oldest known Appalachians): petrofacies, detrital clinopyroxene geochemistry
complex, soft-bodied organisms, including the oldest definitive and plate tectonic implications. Precambrian Research, 59,
animals on Earth (Narbonne 1998). The Edicaran fossils are 243– 262.
thought to mark a critical stage in the evolution of life on Earth, Ehlers, J. & Gibbard, P. L. 2004. Quaternary Glaciations: Extent and
and the presence of these fossils above the Gaskiers Fm. has led Chronology, Part II: North America. Elsevier Science, Amsterdam,
to the theory that Neoproterozoic glaciations influenced eukaryote 450.
evolution and/or acted as an impetus for the evolution of animals Eyles, N. 1990. Marine debris flows: Late Precambrian ‘tillites’ of the
(Hoffman et al. 1998; Runnegar 2000). Avalonian – Cadomian orogenic belt. Palaeogeography, Palaeocli-
matology, Palaeoecology, 79, 73 –98.
N. Eyles thanks the Natural Sciences and Engineering Research Council of Eyles, N. & Eyles, C. H. 1989. Glacially-influenced deep-marine
Canada for long-term financial support and a postgraduate scholarship to sedimentation of the Late Precambrian Gaskiers Formation, New-
S. Carto, R. Slatt and B. Rogerson (formerly of Memorial University of New- foundland, Canada. Sedimentology, 36, 601–620.
foundland). R. Hiscott and C. Eyles are thanked for their assistance and discus- Fisher, R. V. 1984. Submarine volcaniclastic rocks. In: Kokelaar, B. P.
sions in the field. This represents a contribution of the IUGS- and UNESCO- & Howells, M. F. (eds) Marginal Basin Geology: Volcanic and
funded IGCP (International Geoscience Programme) project #512. Associated Sedimentary and Tectonic Processes in Modern and
Ancient Marginal Basins. Geological Society of London, Special
Publications, 16, 5 –27.
References Forty, R. A. & Cocks, L. R. M. 2003. Palaeontological evidence bearing
on global Ordovician– Silurian continental reconstructions. Earth
Anderson, M. M. 1978. Ediacara fauna. In: Lapedes, D. N. (eds) Science Reviews, 67, 247–307.
McGraw-Hill Yearbook of Science and Technology. McGraw-Hill, Gardiner, S. & Hiscott, R. N. 1988. Deep-water facies and depositional
New York, 146– 149. setting of the lower Conception Group (Hadrynian), southern Avalon
Anderson, M. M. & King, A. F. 1981. Precambrian tillites of the Con- Peninsula, Newfoundland. Canadian Journal of Earth Sciences, 25,
ception Group on the Avalon Peninsula, southeastern Newfoundland. 1579– 1594.
In: Hambrey, M. J. & Harland, W. B. (eds) Pre-Pleistocene Glacial Gravenor, C. P. 1980. Heavy minerals and sedimentological studies on
Record. Cambridge University Press, Cambridge, 760–763. the glaciogenic late Precambrian Gaskiers Formation of Newfound-
Anderson, M. M. & Conway Morris, S. 1982. A review, with descrip- land. Canadian Journal of Earth Sciences, 17, 1331–1341.
tions of four unusual forms, of the soft-bodied fauna of the Hatcher, R. D. & Goldberg, S. A. 1991. The Blue Ridge province.
Conception and St. John’s Groups (Late Precambrian), Avalon In: Horton, J. W. & Zullo, V. M. (eds) Geology of the
Peninsula, Newfoundland. Proceedings of the Third North American Carolinas. Carolina Geological Society, University of Tennessee
Paleontology Convention, 1, 1 –8. Press, 11– 35.
Benus, A. P. 1988. Sedimentological context of a deep-water Ediacaran Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P.
fauna: Mistaken Point, Avalon Zone, Eastern Newfoundland. 1998. A Neoproterozoic Snowball Earth. Science, 281, 1342– 1346.
THE GASKIERS FORMATION 473

Ichaso, A. A., Dalrymple, R. W. & Narbonne, G. M. 2007. Paleoenvir- Nance, R. D., Murphy, J. B., Strachan, R. A., Lemos, R. S. & Taylor,
onmental and basin analysis of the late Neoproterozoic (Edicaran) G. K. 1991. Late Proterozoic tectonostratigraphic evolution of the
upper Conception and St. John’s groups, west Conception Bay, Avalonian and Cadomian terranes. Precambrian Research, 53,
Newfoundland. Canadian Journal of Earth Sciences, 44, 25– 41. 41– 78.
Jenkins, R. J. F. 1992. Functional and ecological aspects of Ediacarian Nance, R. D. & Murphy, J. B. 1996. Basement isotopic signatures and
assemblages. In: Lipps, J. H. & Signor, P. (eds) Origin and Early Neoproterozoic paleogeography of Avalonian– Cadomian and
Evolution of the Metazoa. Plenum, New York, 131–176. related terranes in the circum-North Atlantic. In: Nance, R. D. &
King, A. F. 1980. The birth of the Caledonides: Late Precambrian rocks Thompson, M. D. (eds) Avalonian and Related Peri-Gondwanan
of the Avalon Peninsula, Newfoundland, and their correlatives in Terranes of the Circum-North Atlantic. Geological Society of
the Appalachian Orogen. In: Wones, D. R. (ed.) Proceedings of the America Special Papers, 304, 333–346.
Caledonides in the USA Memoirs of the Department of Geological Nance, R. D., Murphy, J. B. & Keppie, J. D. 2002. Cordilleran model for
Science, Virginia Polytechnic Institute, Blacksburg, Virginia, 2, 3– 8. the evolution of Avalonia. Tectonophysics, 352, 11 –32.
King, A. F. 1988. Late Precambrian sedimentation and related orogenesis O’Brien, S. J., Wardle, R. J. & King, A. F. 1983. The Avalon zone: a
of the Avalon Peninsula, Eastern Avalon Zone. Fieldtrip Guidebook Pan-African terrane in the Applachian orogen of Canada. Geological
A4, Geological Association of Canada, 84. Journal, 18, 195–222.
Knoll, A. H., Walter, M. R., Narbonne, G. M. & Christie-Blink, N. Osberg, P. H., Tull, J. F., Robinson, P. H. R & Butler, J. R. 1989. The
2006. The Ediacaran Period: a new addition to the geologic time Acadian orogen. In: Hatcher, R. D., Jr, Thomas, W. A. & Viele,
scale, Lethaia, 39, 13 – 30. G. W. (eds) The Appalachian– Ouachita Orogen in the United
Krogh, T. E., Strong, D. F., O’Brien, S. J. & Papezik, V. S. 1988. States: The Geology of North America, Geological Society of
Precise U– Pb zircon dates from the Avalon terrane in Newfoundland. America, F-2, 179–232.
Canadian Journal of Earth Sciences, 25, 442– 453. Papezik, V. S. 1974. Prehnite-pumpellyite facies metamorphism of the
Landing, E. 1996. Avalon – insular continent by the latest Precambrian. Late Precambrian rocks of the Avalon Peninsula, Newfoundland.
In: Lance, R. D. & Thompson, M. (eds) Avalonian and Related Canadian Minerology, 12, 463–468.
Peri-Gondwanan Terranes of the Circum-North Atlantic. Geological Pimentel, M. M. & Fuck, R. A. 1992. Neoproterozoic crustal accretion
Society of America, Special Paper, 304, 27 – 64. in central Brazil. Geology, 20, 375– 379.
McCartney, W. D. 1967. Whitbourne Map Area, Newfoundland. Poulton, S. W. & Raiswell, R. 2002. The low-temperature geochemical
Geological Survey of Canada, Memoir, 341, 1– 133. cycle of iron: from continental fluxes to marine sediment deposition,
McMenamin, M. A. S. 1987. On the emergence of animals. Scientific American Journal of Science, 302, 774–805.
American, 256, 84 – 92. Rast, N. 1989. The evolution of the Appalachian chain. In: Bally, A. W.
McNamara, A. K., Mac Niocaill, C., Van der Pluijm, B. A. & Van der & Palmer, A. R. (eds) The Geology of North America – An Over-
Voo, R. 2001. West African proximity of Avalon in the latest Precam- view. Geological Society of America, 4, 323–347.
brian. Geological Society of America Bulletin, 113, 1161– 1170. Rast, N. & Skehan, J. W. 1983. The evolution of the Avalon plate.
Murphy, J. B. & Nance, R. D. 1989. Model for the evolution of the Tectonophysics, 100, 257– 286.
Avalonian – Cadomian belt. Geology, 17, 735– 738. Rehmer, J. 1981. The Squantum tilloid Member of the Roxbury Conglom-
Murphy, J. B., Keppie, J. D., Dostal, J. & Nance, R. D. 1999. Neopro- erate of Boston, Massachusetts. In: Hambrey, M. J. & Harland,
terozoic –early Paleozoic evolution of Avalonia. In: Ramos, V. A. & W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge
Keppie, J. D. (eds) Laurentia –Gondwana Connections Before University Press, Cambridge, 756–759.
Pangea. Geological Society of America, Special Papers, 336, Runnegar, B. N. 2000. Loophole for Snowball Earth. Nature, 405,
253– 266. 403– 404.
Murphy, J. B., Pisarevsky, S. A., Nance, R. D. & Keppie, J. D. 2004. Samson, S. D., D’lemos, R. S., Miller, B. V. & Hamilton, M. A. 2005.
Neoproterozoic – Early Paleozoic evolution of peri-Gondwanan ter- Neoproterozic palaeogeography of the Cadomia and Avalon terranes:
ranes: implications for Laurentia – Gondwana connections. Inter- constraints from detrital zircons U– Pb ages. Journal of the Geologi-
national Journal of Earth Sciences, 93, 659–682. cal Society, 162, 65– 71.
Myrow, P. M. 1995. Neoproterozoic rocks of the Newfoundland Avalon Schermerhorn, L. J. G. 1966. Terminology of mixed coarse-fine sedi-
zone. Precambrian Research, 73, 123– 136. ments. Journal of Sedimentary Petrology, 36, 831–835.
Myrow, P. M. & Kaufman, A. J. 1999. A newly discovered cap carbonate Smith, S. A. & Hiscott, R. N. 1984. Latest Precambrian to Early
above Varanger-age glacial deposits in Newfoundland, Canada. Cambrian basin evolution, Fortune Bay, Newfoundland: fault-
Journal of Sedimentary Research, 69, 784–793. bounded basin to platform. Canadian Journal of Earth Sciences,
Myrow, P. M., Taylor, J. F., Miller, J. F., Ethington, R. L., Ripper- 21, 1379–1392.
dan, R. L. & Branchle, C. M. 1999. Stratigraphy, sedimentology, Symons, D. T. A. & Chaisson, A. D. 1991. Palaeomagnetism of the
and paleontology of the Cambrian – Ordovician of Colorado and Callander complex and the Cambrian apparent polar wander path
adjacent areas. Geological Society of America Field Guide, 1, for North America. Canadian Journal of Earth Sciences, 28,
157– 176. 355– 363.
Narbonne, G. M. 1998. The Ediacaran biota: a terminal Neoproterozoic Trindade, R. I. F. & Macouin, M. 2007. Palaeolatitude of glacial depos-
experiment in the evolution of life. Geological Survey of America its and palaeogeography of Neoproterozoic ice ages. Comptes Rendus
Today, 8, 1– 6. Geoscience, 339, 200–211.
Narbonne, G. M. 2004. Modular construction of early Ediacaran Williams, H. & King, A. F. 1975. Southern Avalon, Newfoundland,
complex life forms. Science, 305, 1141– 1144. Trepassey map area. Geological Survey of Canada Paper, 75, 11– 15.
Narbonne, G. M. & Gehling, J. G. 2003. Life after Snowball: the oldest Williams, H. & King, A. F. 1979. Trepassey Map Area, Newfoundland.
complex Ediacaran fossils. Geology, 31, 27– 30. Geological Survey of Canada, Memoir, 389, 1 –24.
Narbonne, G. M., Dalrymple, R. W., Gehling, J. G., Wood, D. A., Wood, D. A., Dalrymple, R. W., Narbonne, G. M., Gehling, J. G. &
Clapham, M. E. & Sala, R. A. 2001. Neoproterozoic fossils and Clapman, M. E. 2003. Paleoenvironmental analysis of the Late
environments of the Avalon Peninsula, Newfoundland. Field Trips Neoproterozoic Mistaken Point and Trepassey formations, south-
B5, Geological Association of Canada– Mineralogical Association eastern Newfoundland. Canadian Journal of Earth Sciences, 40,
of Canada Joint annual Meeting, St. John’s, NL. 1375–1391.
Chapter 43

The Squantum Member of the Boston Basin, Massachusetts, USA

SHANNON L. CARTO* & NICK EYLES


Department of Geology, University of Toronto, Toronto, Ontario, M1C 1A4, Canada
*Corresponding author (e-mail: shannon.carto@gmail.com)

Abstract: The Neoproterozoic diamictite-bearing Squantum Member is located in the Boston Basin in eastern Massachusetts, USA. The
Boston Basin forms part of the Avalonia island arc terrane (c. 650 Ma), and appears to have originated as a rift-type basin in an exten-
sional setting along the northern margin of Gondwana, although its exact position is debated. Inferred palaeoenvironmental reconstruc-
tions of the Boston Basin have alternated between a fluvial basin where ice played a major role in transporting much of the coarse material
and an evolving marine basin dominated by non-glacial subaqueous mass flow, submarine fans and turbidity-current deposition.
The age of the Squantum is bracketed between c. 595 and 570 Ma, and is correlated by some to the glaciogenic diamictite succession of
the Gaskiers Formation (eastern Newfoundland) as part of the putative global Gaskiers Glaciation c. 582–585 Ma. However, the Squan-
tum Member consists of diamictite, graded sandstone and siltstone units, and fine-grained laminated argillite/mudstone units typical of
debris flow and turbidite facies that accumulate in a submarine setting. A glacial influence is not readily identified and revolves around
early interpretations of the diamictite as being ‘till-like’, the presence of laminated horizons that resemble glaciolacustrine ‘varvites’ and
the disputed recognition of ice-rafted dropstones. There are no associated carbonates and, consequently, no geochemical data are
available in connection with the Squantum Member.

The famous Squantum diamictite occurs within the sedimentary poorly sorted and laminated facies (e.g. diamictite, mixtite, rhyth-
rocks of the Boston Bay Group, preserved within the Neoprotero- mite, symmictite; see Schermerhorn 1966) that avoided specific
zoic Boston Basin in eastern Massachusetts, USA (Figs 43.1 & genetic references. Some authors suggest the Squantum Member
43.2). It is formally known as the Squantum Member of the was deposited by gravity-debris flows that reworked glacial sedi-
Roxbury Conglomerate, but has also been referred to as the Squan- ment into deep water (Stuart et al. 1975).
tum ‘Tillite’, Squantum Tillite, Squantum Tilloid, the Roxbury/ Conventionally, the Boston Bay Group is divided into a lower
Brighton complex and Massive diamictite association (Billings Roxbury Conglomerate assemblage containing the diamictite-
1976; Smith & Socci 1990; Thompson 1993). The Squantum dominated Squantum Member and an upper Cambridge Argillite
Member has been very influential in the evolution of ideas con- Formation (Fig. 43.2; Billings et al. 1939; Skehan 1964; Rehmer
cerning pre-Pleistocene glacial deposits (Dodge 1875; Sayles & Roy 1976). The type area and best exposures of the Squantum
1914; Coleman 1926), continental drift and the reconstruction of Member are in the low coastal cliffs at Squantum Head in
Pangea (Wegner 1912), the recognition of submarine mass flow Quincy, Massachusetts, in the southern part of the Basin
processes (Crowell 1957; Dott 1961) and the terminology used (718000 W, 428180 N; Rehmer 1981). Outcrops are of limited
for poorly sorted rocks (Pettijohn 1957). extent and many locations cited in earlier literature have been
Dodge (1875) was first to suggest a glacial origin for the Squan- lost to urbanization. Outcrops at Squantum Head expose two dia-
tum Member. Shaler (1869) suggested an exclusively marine mictite facies interbedded with thinly laminated diamictite and
origin in view of the fine laminations, current structures and mudstone units.
graded bedding observed in the overlying Cambridge Formation
(Fm.), and parts of the Roxbury assemblage. He later suggested
a glacial origin for the conglomerate units within the Roxbury
assemblage (Shaler et al. 1899). Sayles & LaForge (1910) Structural framework
regarded the diamictite and laminated facies of the Squantum as
a Carboniferous –Permian glacial and varved glaciolacustrine The Squantum Member is one of several diamictite successions
deposit. Wegner (1912) considered the Squantum to be non-glacial preserved within the Late Proterozoic Avalonia– Cadomian oro-
on the basis that poorly sorted facies are not exclusively of glacial genic belt. This belt consists of a series of exotic terranes that occu-
origin and also because the deposit did not fit with any simple lati- pied positions adjacent to the margin of Gondwana during the
tudinally constrained ice sheet on a reconstructed Pangaea for Neoproterozoic (the so-called peri-Gondwana terranes) (e.g.
which he received much undue criticism. Its age was later deter- Murphy & Nance 1989). The remnants of these terranes can be
mined to be Neoproterozoic (for a review of historical ideas see widely traced in North America, Europe and Africa (McNamara
Eyles 2004). et al. 2001). The Squantum Member and its associated sedimentary
Some interpretations of the Boston Bay Group have character- and volcanic rocks, which occur within the 750 km2 fault-bounded
ized all or part of the succession as a subaerial alluvial fan or Boston Basin (Rehmer 1981), are part of the Western Avalonia
braided river system close to an ice margin (Rehmer & Roy terrane. The Boston Basin appears to be a rift-related basin, repre-
1976; Kaye 1984). Others attribute the Squantum Member to sub- sentative of an intra-arc (Nance 1990; Socci & Smith 1990) or a
aqueous debris flow processes entirely unrelated to ice (Pettijohn back-arc (Cardoza et al. 1990) basin that developed in an exten-
1957; Crowell 1957; Caldwell 1964; Dott 1961; Lindsay et al. sional or transtensional setting following the closing of the
1970; Billings 1976). Crowell (1957) proposed that the Squantum Cadomian Ocean (Avalonian orogeny) (Rast & Skehan 1983). It
diamictite had been generated by submarine slumping and mixing subsequently rifted from the Gondwanan margin in the early
of gravel and mud downslope. His work, and especially that of Dott Palaeozoic and drifted northward, to be later accreted to eastern
(1961), highlighted the limitations of ‘tillite’ as a descriptive term Laurentia c. 440 Ma (Murphy & Nance 1989; Murphy et al.
for such deposits because of its implication of direct glacial depo- 2004). The arc basin interpretation stems from the calc-alkaline
sition. He recommended the use of purely descriptive terms for character and trace element geochemistry of the granitoids

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 475– 480. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.43
476 S. L. CARTO & N. EYLES

45’N
Boston Basin

.
N.H
a ss.
Boston
M

Squantum Head
N Boston Basin,
Nantasket Massachusetts
50 km

R.I.
ss.
Ma
n.
Con

36’N Fig. 43.1. General schematic map of the


45’W 30’W location of the Boston Basin in
Massachusetts, USA, indicating the
location of principal outcrops mentioned in
the text (after Socci & Smith 1987).

(Dedham Granite) and felsic volcanic rocks (Mattapan-Lynn Vol- in rocks outside the Boston Basin to the NW (e.g. Clinton-
canic Complex) (Hermes et al. 1981) that underlie the strata of the Newbury and Bloody Bluff faults) (Skehan & Murray 1980).
Boston Bay Group.
The Boston Bay Group was affected by subsequent Alleghenian
deformation in the form of broad east –NE striking thrust faults and Stratigraphy
NE-trending folds (Billings 1979; Skehan & Murray 1980). The
metamorphic grade of the Boston Bay Group is subgreenschist The Boston Bay Group (c. 5 km thick; Bailey 1987) has tradition-
facies, and a slaty, well-developed, spaced cleavage, oriented ally been divided into two distinctive units: a lower Roxbury
approximately perpendicular to bedding, is present (Rehmer Conglomerate assemblage and the upper Cambridge Argillite For-
1981). Deformation induced by the Acadian orogeny is recorded mation (Rehmer & Roy 1976). The Roxbury is subdivided into
three sub-members: the basal Brookline Member (clast-supported
conglomerate with minor sand), the medial Dorchester Member
(mostly sandstone with minor conglomerate) and the upper Squan-
tum Member (diamictite-dominated) (Emerson 1917; LaForge
Cambridge Argillite 1932). There is agreement that the basin fill cannot simply be sub-
(2300 - 5500m) divided in layer-cake fashion on the basis of correlative facies
types and the complex interfingering of the various clastic litholo-
gies within the Roxbury assemblage (Fig. 43.2; Dott 1961; Billings
1976, 1979; Socci & Smith 1990).
The granitic basement of the Boston Basin, represented by the
c. 610 Ma Dedham Granite (Hepburn et al. 1993) and the
c. 599 Ma Westwood Granite (Thompson et al. 1996) is thought
Squantum Member to have been intruded into the older crust (Middlesex Fells Volca-
Dorchester Member nic Complex). The basement rocks are also represented by the
(100 - 500m) bimodal Mattapan-Lynn Volcanic Complex (c. 596 Ma), which
Roxbury unconformably underlies and pre-dates the Boston Bay Group in
Conglomerate most areas (Hon & Hepburn 1986). This complex is composed
of more than 300 m of rhyolitic, andesitic, basaltic, local kerato-
phyre flows, coarse granitic and syenitic dykes, volcanic breccia,
Brookline Member
scoria, and volcanic mudflows and tuffs (La Forge 1932; Thomp-
(150 - 1300m)
son 1993), and records the transition from a compressional to
extensional tectonic regime (Skehan & Murray 1980; Hon &
Fig. 43.2. Original ‘layer-cake’ stratigraphic framework of the Hepburn 1986). These basement units are overlain by the
lithostratigraphic units of the Boston Bay Group (modified from Socci & Smith calc-alkaline Brighton Volcanic rocks/Melaphyre (c. 580–
1987). Each stratigraphic unit was defined and recognized on the basis of the 650 Ma), which consist of the altered basalt and andesite flows,
percentage of conglomerate, diamictite, argillite, sandstone, siltstone and pyroclastic rocks, breccia, tuff, and intrusive rocks that sporadi-
matrix present. Recent sedimentological studies indicate that the sedimentary cally interfinger the Brookline and Dorchester members through-
units of the Boston Bay Group are complexly intercalated, with repetition of out the basin (Dott 1961; Billing 1976; Zartman & Naylor
facies within different members suggesting that the sedimentation of these rocks 1984). The Brighton Volcanic rocks record a later phase of
was diachronous and genetically related. volcanic activity.
THE SQUANTUM MEMBER OF THE BOSTON BASIN 477

Glaciogenic deposits and associated strata been reported by early workers but have not been confirmed by
later observers. Moreover, several previously identified dropstones
Brookline Member have been re-interpreted as having been emplaced by lateral
sediment-gravity or current processes (Dott 1961; Bailey &
The Brookline Member (150 –1300 m thick), at the base of the Bland 2001). The sand- and gravel-sized detritus is composed of
Roxbury Conglomerate, has been described as a ‘puddingstone’ volcanic, granitic and metasedimentary lithic fragments and is
consisting of a massive clast-supported pebble and cobble con- essentially the same composition as that of the other Roxbury
glomerate with interbedded argillite, sandstone and melaphyre units (Dott 1961). The presence of lapilli tuff beds has been ident-
units (Billing 1976; Reymer & Roy 1976). The matrix is composed ified in thin sections of strata genetically related to, and in close
of grey feldspathic sandstone, supporting well-rounded pebbles association with, the Squantum diamictite units at Squantum
and cobbles of quartzite granite, felsite, and quartz monzonite Head, showing that further analysis of the volcanic character of
with clast size ranging from 1 to 15 cm (Billing 1976). Smith & this deposit is warranted.
Socci (1990) describe this unit as a diamictite/sandstone/argillite
association, in reference to complex interbedding of laminated and
graded argillite and sandstone units and massive, matrix- and clast- Cambridge Fm.
supported diamictite units.
Overlying the Squantum Member, the Cambridge Fm. (known
formally as the Cambridge Argillite or the Cambridge Slate) con-
Dorchester Member sists of up to 5 km of laminated, dark to olive grey, graded siltstone
and sandstone beds (Billings 1976). Graded beds, starved ripples,
In the Boston Basin, purplish, greenish and grey siltstone and sand- scour marks, load casts and micro-faults are numerous. Soft-
stone units, as well as shale units above the Brookline Member, are sediment deformation structures, such as mega slump folds
traditionally assigned to the Dorchester Member (180– 500 m many metres in amplitude, and pinch and swell bedding, are also
thick) (Billings 1976). Smith & Socci (1990) also describe this common (Bailey 1987). Petrographically, quartz, feldspar,
unit as commonly containing full or partial Bouma sequences melaphyre fragments, chlorite and epidote dominate these facies,
and dominated by medium-to fine-grained argillite, containing indicating a volcanic/igneous source area. Discrete ash
lesser amounts of sandstone and conglomerate relative to the beds measuring a few centimetres to tens of centimetres in
Brookline Member. Evidence of penecontemporaneous fragmen- thickness have also been documented in this unit (LaForge 1932;
tation and contortion due to downslope slumping are common in Thompson & Bowring 2000).
these units. The same facies occur interbedded within the Squan-
tum diamictite units (Bailey 1987).
Boundary relations with overlying and underlying
Squantum Member non-glacial units

The Squantum diamictite has been described as a polymictic, het- The Boston Bay Group strata are fault-bounded to the north and
erogeneous and poorly sorted admixture of rare boulders up to west, and to the south rest unconformably on the bimodal Mattapan
1.2 m in diameter with pebbles, cobbles and sand in a silty-clay Volcanic Complex, which, in turn, overlies the Dedham Granite
matrix (Dott 1961; Socci & Smith 1990). It has also been described (Rehmer 1981). The upper and lower contacts of the Squantum
as a matrix-supported conglomerate (Billings 1976; Rehmer & Member are gradational in outcrop (LaForge 1932). Significantly,
Roy 1976) and as a pebbly mudstone (Crowell 1957). At the many authors note intricate facies variations and complex
type area at Squantum Head, these facies are interbedded with interfingering between the clastic lithofacies of the Roxbury Con-
laminated mudstone units (bed thickness ranging from 2 to glomerate and volcanic lithologies within the Roxbury assem-
10 cm) and lamina-thick diamictite horizons (bed thickness blage. Sayles (1914), Dott (1961), Rahm (1962) and Socci &
ranging from a few millimetres to 1 cm; Dott 1961; Socci & Smith (1987) have noted that Squantum-type rocks (less well-
Smith 1990), some of which contain a few outsized clasts that sorted facies of the Brookline and Dorchester lithologies as
have depressed the underlying laminae, having the appearance of defined by Dott 1961) appear to inter-finger with the Roxbury
dropped pebbles (Lindsay et al. 1970). and overlying Cambridge Argillite, and become more common
Most outcrops of the diamictite show chaotic bedding in the towards the NE (Dott 1961). The Brighton Volcanic rocks are
form of contorted and folded patches of sand and local clusters also interbedded throughout the Brookline and Dorchester
of clasts, as well as coherent slump blocks of mudstone (Dott Members in the southern portion of the basin (Zartman & Naylor
1961; Lindsay et al. 1970). These chaotically mixed facies are 1984; Thompson 1993).
readily distinguished from more homogeneous diamictite facies.
Diamictite units reportedly range in thickness from 18 to 215 m
(Rehmer 1981). These units are mostly massive and lenticular in Chemostratigraphy
form; some are crude to moderately well sorted and some
exhibit normal grading. There are also thinner diamictite and intra- Recently, Passchier & Erukanure (2010) determined the chemical
clastic rich pebbly horizons exposed in other parts of the Boston index of alteration (CIA) to assess if any evidence of climatic
Basin area (Bailey, pers. comm. 2008). changes is present in the Squantum Member. The geochemical
Clasts consist of multicoloured, locally derived felsic and mafic work targeted the fine-grained facies, including the matrix of the
volcanic rocks (Mattapan and/or Brighton type), granodiorite diamictite units at the type locality. The researchers found CIA
(Dedham-type), quartzite and intrabasinal clasts of massive, to be relatively high for the Squantum Member (CIA of 61 –75).
graded and laminated sandstone and siltstone (Rehmer 1981). This value is higher than those of Pleistocene glacial diamictite,
Much of the matrix has been derived from the latter facies. which has a typical CIA of 50– 55 (Nesbitt & Young 1982; Passch-
Clasts range from sub-rounded to angular clasts, 5– 60 cm in diam- ier & Krissek 2008). The CIA values from the Squantum diamictite
eter, to well-rounded clasts 3 –8 cm in diameter (Billings 1976). were interpreted to indicate that, at the time of deposition, the con-
Striated clasts (Sayles & LaForge 1910; Sayles 1914), chatter- tinental areas could not have been entirely ice covered and that
marked quartz grains (Rehmer & Hepburn 1974) and dropstones significant weathering occurred before the muds were supplied
(Cameron & Jeanne 1976; Wolfe 1976; Cameron 1979) have to the basin (Passchier & Erukanure 2010).
478 S. L. CARTO & N. EYLES

Other characteristics for the two youngest grains) to 618 Ma (date for the oldest grain
from the remaining nine grains; an older grain with a
207
Ediacaran (Vendian as per Lenk et al. 1982) microfossils (Bavli- Pb/206Pb date of 818 Ma was not included) (Thompson &
nella cf. faveolata) occur in the Cambridge Argillite Formation Bowring 2000; Thompson et al. 2007). In this study, similar
(Lenk et al.1982). These fossils are preserved as petrifactions in ages were yielded from a welded tuff clast, crystal-poor tuff
pyrite interspersed within organic laminae (Lenk et al. 1982). sample, and a sample of granophyre removed from the Squantum
The microfossils were not fragmented and do not show any evi- diamictite, c. 595 + 2 Ma, 600 + 1 Ma and 610 + 2 Ma, respect-
dence of having been reworked from sediment. Raised ring struc- ively (Thompson & Bowring 2000). The minimum age limit of
tures (5 –35 mm) have been described from outcrops of Boston the Squantum Member is thought to be c. 570 Ma (207Pb/206Pb
Bay Group sediments at Hewitts Cove and Slate Island (Bailey date) or younger based on the youngest detrital zircon component
& Bland 2001). Although similar in form and Neoproterozoic in from an ash bed in the overlying Cambridge Argillite, which
age, these are thought to be distinct from Ediacaran Aspidella ter- yielded a 207Pb/206Pb date of c. 570 Ma (Thompson & Bowring
ranovica described in Newfoundland and elsewhere (Bailey & 2000). Based on these results, the Squantum deposition is
Bland 2001). More recently, Passchier & Erukanure (2010) have bracketed between c. 595 and 570 Ma.
reported two dislocated stromatolite hemispheroids in outcrop in
the laminae exposed at the type section.
Discussion

There is an agreement that the Squantum Member is the product of


Palaeolatitude and palaeogeography
the downslope slumping and intermixing of fine and coarse facies in
a marine environment (Crowell 1957; Pettijohn 1957; Dott 1961;
Palaeogeographical reconstructions of Avalonia based on palaeo-
Caldwell 1964; Lindsay et al. 1970; Billings 1976; Smith & Socci
magnetic pole data are inconclusive, as they place Avalonia
1990). The close similarity of the Roxbury conglomerate and Squan-
either near Laurentia at low latitudes or at a high palaeolatitude
tum clast compositions has been cited by Dott (1961) as evidence
(Wu et al. 1986; McNamara et al. 2001; Trindade & Macouin
that Squantum Member diamictite units were derived by mixing of
2007). Most palaeomagnetic reconstructions have long placed
Roxbury gravels with mud as they slumped downslope and were
Avalonia along the northern margin of Gondwana (Van der Voo
redeposited (see also Crowell 1957), known to be a common submar-
1993; Torsvik et al. 1996), but whether it was adjacent to the
ine process (e.g. Nemec et al. 1984). These diamictites and the
West African craton or Amazonian craton remains uncertain.
associated thick turbidite succession of the Cambridge Argillite
Recent palaeomagnetic results from the c. 580 to 570 Ma volcanic
are readily placed into a submarine fan or slope environment. The
and sedimentary rocks of the Marystown Group of the Burin
volcanic and coarsely clastic character of the Boston Bay Group
Peninsula, Newfoundland (western Avalonia), has revealed that
points to deposition associated with volcanic activity, whereas the
Avalonia had a palaeolatitude of 348 þ88/–78 (mean tilt-corrected
extent of any glacial influence is arguable and was supported by
inclination of 538), coinciding with that of the northern margin of
indirect evidence at best. Past identification of a glaciogenic
West Africa (McNamara et al. 2001). This palaeolatitude was
source rested entirely on the identification of ‘dropstones’ (Wolfe
based on palaeomagnetic core samples collected from 21 sites,
1976; Cameron & Jeanne 1976; Cameron 1979) and striated
all of which yielded a positive conglomerate test, confirming a
pebbles (Sayles & LaForge 1910; Sayles 1914), which have not
primary age of magnetization for the Marytown Group as a
been substantiated by later observers (Dott 1961; Bailey 1987). At
whole. Demagnetization results revealed that two or more com-
present, no carbonates have been identified in this succession.
ponents of magnetization were present in almost all samples, yield-
Stratigraphic relationships among the units of the Boston Bay
ing curved trajectories in orthogonal plots. Some samples consist
Group have, as stated above, traditionally been described in terms
of multiple specimens; magnetization directions from the samples
of a simple, vertical layer-cake succession (e.g. Mattapan-Lynn
within a site are averaged to produce the site mean for that site.
Complex/Dedham-Brookline-Dorchester-Squantum-Cambridge).
A similar palaeoposition of Avalonia is supported by Thompson
This scheme was made on the basis of correlating like-for-like
et al. (2007), who suggested Avalonia occupies a midlatitude pos-
facies and in ignorance of any understanding of the origin of the
ition of 38 + 88 between c. 600 Ma and 575 Ma. This palaeolati-
facies and the depositional system in which these facies occur.
tude is based on samples collected at six locations (142 cores)
The Boston Basin is one of several Neoproterozoic successions
from Lynn and Mattapan volcanic complexes and three sites (33
deposited in arc-related basins that were found within peri-
cores) from the Squantum Member. Results from the Squantum
Gondwana terranes (e.g. Armorica). They are characterized by volu-
Member were corrected for bedding tilt and plunge. The samples
minous accumulations of volcanogenic turbidites, debrites and vol-
from each site passed reversal and fold tests, and are considered
canic rocks reflecting oblique subduction below the Gondwanan
to record primary Neoproterozoic magnetization at c. 595 Ma.
margin sometime between 620 and c. 590 Ma (Murphy et al.
Thompson et al. (2007) noted that two permissible palaeomagnetic
2004). The degree of any glacial influence on debrites elsewhere
positions exist for Avalonia. If placed at a palaeolatitude of 388S
within Armorica has long been debated, much like the controversy
it would border West Africa, whereas if it were placed 388N it
surrounding the Squantum Member deposits. Given its resemblance
would border Amazonia.
to till, past practice has simply been to label as glacial any deposit
Conversely, Samson et al. (2005) have shown that the main age
that is poorly sorted (see Eyles 1990, 1993, for review). It is signifi-
groups of detrital zircons from Avalonia cluster within the Meso-
cant that Crowell’s (1999) authoritative review of Earth’s Neoproter-
proterozic, similar to the age pattern of detrital zircons also found
ozoic glacial record does not include the Squantum Member.
in the Amazonian craton, and therefore place Avalonia at 388N
There are nonetheless close sedimentological similarities with
along side the Amazonian craton at c. 580 Ma.
the glacially and volcanically influenced debris flow/turbidite
succession of the Gaskiers Fm. in Newfoundland, Canada (Carto
& Eyles 2011), but the case for a glacial setting for the Boston
Geochronological constraints Basin strata is tenuous, unlike for the Gaskiers Fm., where
undoubted glacially striated and faceted clasts, as well as drop-
U –Pb zircon data derived from a sandstone bed from the Squan- stones, are found. The extent of glaciation around the outer Gond-
tum Member on the western end of Squantum Head (12 detrital wanan margin (the so-called Gaskiers Glaciation c. 580 Ma) was
zircon samples) yielded a weighted mean 207Pb/206Pb date probably limited to topographically controlled small glacier com-
ranging from c. 593 + 3 Ma (weighted mean 207Pb/206Pb date plexes on higher standing, arc-related volcanoes.
THE SQUANTUM MEMBER OF THE BOSTON BASIN 479

This represents a contribution of the IUGS- and UNESCO-funded IGCP (Inter- Hermes, O. D., Barosh, P. J. & Smith, P. V. 1981. Contact relationships
national Geoscience Programme) project #512. of the late Paleozoic Narragansett Pier Granite and country rock,
In: Boothroyd, J. C. & Hermes, O. D. (eds) Guidebook for Field
Studies in Rhode Island and Adjacent Areas: New England Intercol-
legiate Geological Conference, 73rd Annual Meeting, University of
References Rhode Island, Kingston, Rhode Island, 47– 67.
Hepburn, J. C., Hon, R., Dunning, G. R., Bailey, R. H. & Galli, K.
Bailey, R. H. 1987. Stratigraphy of the Boston Bay Group, Boston Area, 1993. The Avalon and Nashoba terranes (eastern margin of the Appa-
Massachusetts. In: Roy, D. C. (ed.) Northeastern Section of the Geo- lachian orogen in southeastern New England). In: Cheney, J. T. &
logical Society of America: Boulder, Colorado. Geological Society of Hepburn, J. C. (eds) Field Trip Guidebook for the Northeastern
America, Centennial Field Guide, 5, 209–212. United States. Boston GSA, Volume 2. Geology Department, Univer-
Bailey, R. H. & Bland, B. H. 2001. Recent developments in the study of sity of Massachusetts, Contribution 67, X.1– X.31.
the Boston Bay Group. In: West, D. P. Jr. & Bailey, R. H. (eds) Hon, R. & Hepburn, J. C. 1986. Igneous geochemistry and its impli-
Guidebook for Geological Field Trips in New England, 2001 cations for terrane analysis of the Avalonian event in southeastern
Annual Meeting of the Geological Society of America, Boston, New England, U.S.A. Atlantic Geoscience Society Abstract, 1986
Massachusetts, U-1 –U-23. colloquium. Maritime Sediments and Atlantic Geology, 22, 331.
Billings, M. P. 1976. Geology of the Boston Basin. In: Skehan, J. W. & Kaye, C. A. 1984. Boston Basin restudied. In: Hanson, L. S. (ed.)
Murray, D. P. (eds) Studies in New England Geology. Geological Geology of the Coastal Lowlands Boston, Massachusetts to Kenne-
Society of America, Memoir, 46, 5– 30. bunk, Maine. New England Intercollegiate Geological Conference
Billings, M. P. 1979. Bedrock geology of the Boston Basin. In: 76th Annual Meeting, 124–140.
Cameron, B. (ed.) Carboniferous Basins in Southeastern New LaForge, L. 1932. Geology of the Boston area, Massachusetts. US
England. Field guidebook for Trip No. 5, 9th International Congress Geological Survey Bulletin, 839, 105.
of Carboniferous Stratigraphy and Geology, 46 – 64. Lenk, C., Strother, P. K., Kaye, C. A. & Barghoorn, E. S. 1982. Pre-
Billings, M. P., Loomis, F. B. & Stewart, G. W. 1939. Carboniferous cambrian age of the Boston Basin: new evidence from microfossils.
topography in the vicinity of Boston, Massachusetts. Geological Science, 216, 619– 620.
Society of American Bulletin, 50, 1867– 1884. Lindsay, J. F., Summerson, C. H. & Barrett, P. J. 1970. A long-axis
Caldwell, D. W. 1964. The Squantum Formation: Paleozoic tillite or clast fabric comparison of the Squantum ‘Tillite’ and the Gowganda
tilliod? In: Skehan, S. J. (ed.) Guidebook to Field Trips in the Formation, Ontario. Journal of Sedimentary Petrology, 40, 475– 479.
Boston Area and Vicinity. New England Intercollegiate Geological McNamara, A. K., Niocaill, C. M., Van der Pluijm, B. A. & Van der
Conference, 56th Annual Meeting, 53 –60. Voo, R. 2001. West African proximity of the Avalon terrane in the
Cameron, B. 1979. General geology and the Carboniferous Basins of latest Precambrian. Geological Society of America Bulletin, 113,
eastern Massachusetts and Rhode Island. In: Cameron, B. (ed.) 1161–1170.
Carboniferous Basins of Southeastern New England, Field Guide- Murphy, J. B. & Nance, R. D. 1989. Model for the evolution of the
book for Trip No. 5, 9th International Congress of Carboniferous Avalonian– Cadomian Belt. Geology, 17, 735– 738.
Stratigraphy and Geology, 1– 6. Murphy, J. B., Pisarevsky, S. A., Nance, R. D. & Keppie, J. D. 2004.
Cameron, B. & Jeanne, R. A. 1976. New evidence for glaciation during Neoproterozoic– Early Paleozoic evolution of peri-Gondwanan
deposition of the Boston Bay. In: Cameron, B. (ed.) Geology of terranes: implications for Laurentia– Gondwana connections.
Southern New England, 68th Annual Meeting New England Inter- International Journal of Earth Sciences, 93, 659–682.
collegiate Geological Conference, Science Press, Princeton, NJ, Nance, R. D. 1990. Late Precambrian – Early Paleozoic arc –platform
117– 134. transitions in the Northern Appalachians; review and implications,
Cardoza, K. D., Hepburn, J. C. & Hon, R. 1990. Geochemical con- In: Socci, A. A., Skehan, J. W. & Smith, G. W. (eds) Geology
straints on the paleotectonic settings of two late Proterozoic mafic of the Composite Avalon Terrane of Southern New England:
volcanic suites, Boston-Avalon Zone, Eastern Massachusetts. In: Boulder, Colarado. Geological Society of America Special Paper,
Socci, A. D., Skehan, J. W. & Smith, G. W. (eds) Geology of the 245, 1– 11.
Composite Avalon Terrane of Southern New England. Geological Nemec, W., Steel, R. J., Porebski, S. J. & Spinnangr, A. 1984. Domba
Society of America Special Paper, 245, 113–131. Conglomerate, Devonian, Norway: process and lateral variability in
Carto, S. L. & Eyles, N. 2011. The deep marine glaciogenic Gaskiers a mass flow-dominated lacustrine fan-delta. In: Koster, E. H. &
Formation: Newfoundland, Canada. In: Arnaud, E., Halverson, Steel, R. J. (eds) Sedimentology of Gravels and Conglomerates.
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro- Canadian Society of Petroleum Geologists, Calgary, 295– 330.
terozoic Glaciations. Geological Society, London, Memoirs, 36, Nesbitt, H. W. & Young, G. M. 1982. Early Proterozoic climates and
467– 473. plate motions inferred from major element chemistry of lutites.
Coleman, A. P. 1926. Ice Ages: Recent and Ancient. Macmillan, London. Nature, 299, 715– 717.
Crowell, J. C. 1957. Origin of pebbly mudstones. Geological Society of Passchier, S. & Krissek, L. A. 2008. Oligocene –Miocene Antarctic
America Bulletin, 68, 993–1010. continental weathering record and paleoclimatic implications, Cape
Crowell, J. C. 1999. Pre-Mesozoic ice ages: their bearing on under- Roberts drilling Project, Ross Sea, Antarctica. Palaeogeography,
standing the climate system. Geological Society of America Palaeoclimatology, Palaeoecology, 260, 30– 40.
Memoir, 192, 106. Passchier, S. & Erukanure, E. 2010. Palaeoenvironments and weather-
Dodge, W. W. 1875. Notes on the geology of eastern Massachusetts. ing regime of the Neoproterozoic Squantum ‘Tillite’, Boston Basin:
Boston Society of Natural History, 17, 388–419. no evidence of a snowball Earth. Sedimentology, 57, 1526–1544.
Dott, R. H. 1961. Squantum ‘Tillite’, Massachusetts, evidence of glacia- Pettijohn, F. J. 1957. Sedimentary Rocks. Harper & Brothers, New York.
tion or subaqueous mass movement? Geological Society of America Rahm, D. A. 1962. Geology of the main drainage tunnel, Boston, Massa-
Bulletin, 72, 1289–1306. chusetts. Journal of the Boston Society of Civil Engineering, 49,
Emerson, B. K. 1917. Geology of Massachusetts and Rhode Island. 319– 368.
United States. Geological Survey Bulletin, 597, 56– 57. Rast, N. & Skehan, J. W. 1983. The evolution of the Avalonian Plate.
Eyles, N. 1990. Late Precambrian ‘tillites’ of the Avalonian –Cadomian Tectonophysics, 100, 257– 286.
belt; marine debris flows in an active tectonic setting. Palaeogeo- Rehmer, J. A. 1981. The Squantum tilloid member of the Roxbury
graphy, Palaeoclimatology, Palaeoecology, 79, 73 –98. Conglomerate of Boston, Massachusetts. In: Hambrey, M. J. &
Eyles, N. 1993. Earth’s glacial record and its tectonic setting. Earth Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial Record.
Science Reviews, 35, 1 –248. Cambridge University Press, London, 756– 759.
Eyles, N. 2004. Frozen in time: concepts of ‘global glaciation’ from 1837 Rehmer, J. A. & Hepburn, J. C. 1974. Quartz sand surface textural
(die Eiszeit) to 1998 (the Snowball Earth). Geoscience Canada, 31, evidence for a glacial origin of the Squantum ‘Tillite’, Boston
157– 166. Basin, Massachusetts. Geology, 2, 413– 415.
480 S. L. CARTO & N. EYLES

Rehmer, J. A. & Roy, D. C. 1976. The Boston Bay Group: the boulder the Squantum ‘Tillite’, Boston Basin, Massachusetts: comments
bed problem. In: Cameron, B. (ed.) Geology of Southeastern New and reply. Geology, 3, 153– 155.
England. New England Intercollegiate Geology Conference Thompson, M. D. 1993. Late Proterozoic stratigraphy and structure in the
Science Press, Princeton, New Jersey, 71 –91. Avalonian magmatic arc southwest of Boston, Massachusetts.
Samson, S. D., D’lemos, R. S., Miller, B. V. & Hamilton, M. A. 2005. American Journal of Science, 293, 725–743.
Neoproterozic palaeogeography of the Cadomia and Avalon terranes: Thompson, M. D. & Bowring, S. A. 2000. Age of the Squantum ‘tillite’,
constraints from detrital zircons U –Pb ages. Journal of the Geologi- Boston Basin, Massachusetts: U– Pb zircon constraints on terminal
cal Society, 162, 65 –71. Neoproterozoic glaciation. American Journal of Science, 300,
Sayles, R. W. 1914. The Squantum Tillite. Bulletin Harvard Museum 630– 655.
Comparative Zoology, 66, 141–175. Thompson, M. D., Hermes, O. D., Bowring, S. A., Isachsen, C. E.,
Sayles, R. W. & LaForge, L. 1910. The glacial origin of the Roxbury Besancon, J. R. & Kelly, K. L. 1996. Tectonostratigraphic impli-
Conglomerate. Science, 32, 723–724. cations of Late Proterozoic U–Pb zircon ages in the Avalon Zone
Schermerhorn, L. J. G. 1966. Terminology of mixed coarse-fine of southeastern New England. In: Nance, R. D. & Thompson, M.
sediments. Journal of Sedimentary Petrology, 36, 831– 836. D. (eds) Avalonian and Related Peri-Gondwanan Terranes of the
Shaler, N. S. 1869. Note on the geological section at Chestnut Hill Reser- Circum-North Atlantic. Geological Society of America Special
voir, Massachusetts. Boston Society of Natural History Proceedings, Papers, 304, 179– 191.
13, 172– 177. Thompson, M. D., Grunow, A. M. & Ramezani, J. 2007. Late
Shaler, N. S., Woodworth, J. B. & Foerste, A. F. 1899. Geology of Neoproterozoic paleogeography of the Southeastern New England
the Narragansett Basin. United States Geological Survey, Mono- Avalon Zone: insights from U– Pb geochronology and paleomagnet-
graph, 33. ism. Geological Society of America Bulletin, 119, 681– 696.
Skehan, S. J. 1964. Guidebook to Field Trips in the Boston Area and Torsvik, T. H., Smethurst, M. A. & Meert, J. G. 1996. Continental
Vicinity. New England Intercollegiate Geological Conference breakup and collision in the Neoproterozoic and Palaeozoic: a tale
(NEIGC), 56th Annual Meeting, Boston, Massachusetts. of Baltica and Laurentia. Earth Science Reviews, 40, 229– 258.
Skehan, S. J. & Murphy, D. P. 1980. Geological profile across southeast- Trindade, R. I. F. & Macouin, M. 2007. Paleolatitude of glacial deposits
ern New England. Tectonophysics, 69, 285– 319. and paleogeography of Neoproterozoic ice ages. Geoscience, 339,
Smith, G. W. & Socci, A. D. 1990. Late Precambrian sedimentary 200– 211.
geology of the Boston Basin. In: Socci, A. D., Skehan, J. W. & Van der Voo, R. 1993. Paleomagnetism of the Atlantic, Tethys and
Smith, G. W. (eds) Geology of the Composite Avalon Terrane of Iapetus Oceans. Cambridge University Press, Cambridge.
Southern New England. Geological Society of America, Special Wegner, A. 1912. The origin of the continents. Geologische Rundschau,
Papers, 245, 75 – 84. 3, 276– 292.
Socci, A. D. & Smith, G. W. 1987. Recent sedimentological interpret- Wolfe, W. 1976. Geology of Squaw Head, Squantum, Massachusetts In:
ations in the Avalon terrane of the Boston Basin, Massachusetts. Cameron, B. (ed.) Geology of Southeastern New England: Prince-
Maritime Sediments and Atlantic Geology, 23, 75– 84. ton, New Jersey, New England Intercollegiate Geologic Conference,
Socci, A. D. & Smith, G. W. 1990. Stratigraphic implications of 68th Annual Meeting, Guidebook, 107– 116.
facies within the Boston Basin. In: Socci, A. D., Skehan, J. W. & Wu, F., Van der Voo, R. & Johnson, R. J. E. 1986. Eocambrian
Smith, G. W. (eds) Geology of the Composite Avalon Terrane of palaeomagnetism of the Boston basin: evidence for a displaced
Southern New England. Geological Society of America Special terrane. Geophysical Research Letters, 13, 1450.
Paper, 245, 55 –74. Zartman, R. E. & Naylor, R. S. 1984. Structural implications of some
Stuart, J., Baker, H. W., Dott, H. R., Rehmer, J. A. & Hepburn, J. C. radiometric ages of igneous rocks in southeastern New England.
1975. Quartz sand surface textural evidence for a glacial origin of Geological Society of America Bulletin, 95, 522–539.
Chapter 44

The Chiquerı́o Formation, southern Peru

DAVID CHEW1* & CHRISTOPHER KIRKLAND2


1
Department of Geology, Trinity College Dublin, Dublin 2, Ireland
2
Laboratory for Isotope Geology, Swedish Museum of Natural History, S-104 05 Stockholm, Sweden
*Corresponding author (e-mail: chewd@tcd.ie)

Abstract: The Chiquerı́o Formation (Fm.) is a thick glaciogenic succession deposited unconformably on gneisses of the Arequipa massif
in southern Peru. It has undergone greenschist facies metamorphism during Early Palaeozoic orogenesis. The Chiquerı́o Fm. consists of
nearly 400 m of diamictite, sandstone, mudstone and carbonate, with a thin (11 m) cap dolostone at the top of the formation. It is overlain
by the San Juan Fm., a 2-km-thick carbonate succession. The thick glacially influenced succession was deposited in deep marine conditions
and consists mainly of massive diamictites (representing either ice-rafted debris or submarine debris flows) interbedded with turbiditic sand-
stones. Where internal lamination is present (e.g. bedding in the turbiditic packages), abundant dropstones can be recognized. There is no
evidence of shallow marine reworking of the succession. No absolute age constraints on the depositional timing of the Chiquerı́o Fm. exist,
because no volcanic tuffs have yet been identified. U–Pb dating of detrital zircons (U–Th–Pb SIMS) from the Chiquerı́o Fm. and the over-
lying San Juan Fm. suggest it is autochthonous with respect to Amazonia, as the detrital zircon age spectra suggest derivation from the
Amazonian craton. Detrital grains as young as c. 700 Ma have been documented in the post-glacial San Juan Fm. The sparse (chemo)strati-
graphic data available for the Chiquerı́o Fm. exhibit patterns similar to those observed generally in Neoproterozoic post-glacial carbonate
sequences. Palaeogeographic models for the deposition of the Chiquerı́o Fm. are critically dependent on the timing of the docking of the
basement of the Arequipa massif with the South American craton (Amazonia). Presently there are no palaeomagnetic constraints. More
research on the chronological and palaeogeographical constraints of this succession is required.

Supplementary material: Data are available at http://www.geolsoc.org.uk/SUP18479.

The Chiquerı́o Fm. crops out locally on the western coast of stratigraphic position of the Chiquerı́o Fm. above the Precambrian
southern Peru (Fig. 44.1a). It is best exposed at its type locality basement gneiss and the Late Ordovician age for the cross-cutting
5 km SE of the town of San Juan (Fig. 44.1b, 75880 W, 158240 S, San Nicolas batholith (Wilson 1975). Caldas (1978, 1979) con-
UTM 18L 482500 8301000). It rests unconformably on basement sidered the Marcona Fm. to be Late Precambrian –Early Palaeo-
gneiss (Fig. 44.1b), termed the Arequipa massif (Cobbing & zoic in age.
Pitcher 1972; Ramos 2008), and is cut by Early Palaeozoic intru- Subsequent workers (Shackleton et al. 1979; Cobbing 1981) res-
sions (Loewy et al. 2004). The Chiquerı́o Fm. is overlain uncon- urrected the terminology of the Marcona Mining Company (1968),
formably by Jurassic sedimentary rocks, although the oldest with the Marcona Fm. representing all the low-grade metasedi-
cover rocks at its type locality (Fig. 44.1b) are Neogene in age. mentary rock overlying the basement gneiss including the basal
There are very few data from other sections of the Chiquerı́o glaciogenic strata. Shackleton et al. (1979) regarded the Marcona
Fm. Caldas (1978) documents the presence of a well-exposed Fm. to be entirely Early Palaeozoic based on structural consider-
section of the Chiquerı́o Fm. in the Quebadra Jahuay (Fig. 44.1a, ations. He traced deformation events from the Marcona Fm. into
748510 W, 158280 S), while sporadic outcrops are encountered over- the underlying basement gneiss and established that the crystalline
lying the Arequipa massif basement in the vicinity of Marcona basement rocks had experienced an older deformation history. The
Mine (Fig. 44.1a, 75870 W, 158120 S). glaciogenic rocks were described by Cobbing (1981) in the IGCP
The first detailed study of these rocks was undertaken by the 38 volume on the Earth’s Pre-Pleistocene Glacial Record
Marcona Mining Company (1968). They defined the Marcona (Hambrey & Harland 1981). Cobbing (1981) considered them to
Fm. to include all the low-grade metasedimentary rock overlying be Early Palaeozoic in age. Subsequently, Injoque & Romero
the basement gneisses, and they considered the Marcona Fm. to (1986) described possible Precambrian stromatolites in the San
be Carboniferous in age. The basal member of the Marcona Fm. Juan Fm.
was termed the Justa Member, and described as conglomerate Loewy et al. (2003, 2004) undertook whole-rock Pb and U – Pb
with pebbles of gneissic basement. Wilson (1975) obtained four zircon geochronological analyses on the Chiquerı́o Fm. and the
K –Ar ages for the San Nicolas batholith, which cross-cuts the low- underlying gneisses of the Arequipa massif. They used, with modi-
grade metasedimentary rocks in the region. The ages obtained (K – fications, the stratigraphic terminology of Caldas (1978, 1979).
Ar hornblende ages of 442 + 10 Ma and 438 + 9 Ma and K –Ar The Chiquerı́o Fm. was defined as the basal, glaciogenic portion
biotite ages of 428 + 12 Ma and 421 + 11 Ma) demonstrated of the sequence and was considered to be Neoproterozoic in age.
that the Marcona Fm. must be pre-Late Ordovician in age. Loewy et al. (2003, 2004) were unable to find any significant
Caldas (1978), in the course of a regional mapping programme, differences between the San Juan and Marcona formations of
reinterpreted the stratigraphy of the Marcona Mining Company Caldas (1978, 1979), and considered the San Juan Fm. to include
(1968). The basal Justa Member of the Marcona Fm. was termed all the low-grade metasedimentary rock overlying the Chiquerı́o
the Chiquerı́o Fm., and the glaciogenic nature of this part of the Fm. at this locality. This stratigraphic nomenclature was adopted
sequence was recognized for the first time (Caldas 1978, 1979). by Chew et al. (2007a) and is used in this chapter. Chew et al.
The overlying dolomitic rocks were also assigned to a new for- (2007a) presented sedimentary observations, chemostratigraphic
mation, the San Juan Fm. The Marcona Fm. was redefined to rep- data and U –Pb detrital zircon analyses from the Chiquerı́o and
resent the phyllitic, siliciclastic unit lying above the San Juan Fm. San Juan formations. They considered the Chiquerı́o and San
Caldas (1978, 1979) inferred a Late Precambrian age for the Chi- Juan formations Late Neoproterozoic in age, and to be autochtho-
querı́o Fm. and the overlying dolomitic San Juan Fm. based on the nous with respect to the Amazonian craton. These considerations

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 481– 486. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.44
482 D. CHEW & C. KIRKLAND

(a) (b)
76° 75° 480 484 488
Early Palaeozoic plutons
San Juan and Chiquerío N
Formations
45
Arequipa massif
50
basement gneisses

SAN
Marcona 15° Punta
0 20 km 5 JUAN
Mine San Juan
8300
Ecuador 30

85
San Juan
Fig. 44.1b 54 48
Punta El Cenicero 50
Brazil
.3
Fig. 44
Quebadra Fault
Peru Jahuay Neogene
Fig. 44.1. (a) Location map and geological
San Juan Fm. 80 70
N
Chiquerío Fm. map of the western coast of southern Peru.
Punta
Arequipa massif . 44.2 (b) Geological map of the region around
16° Chiquerío Fig
8296
basement gneisses San Juan, modified from Caldas (1978,
0 1km
1979).

also constrained the docking history of the underlying Arequipa and 440 Ma (Loewy et al. 2004). The regional D3 and D4 events
massif to be Late Neoproterozoic or older in age, most likely jux- are considered to be Early Palaeozoic in age (Loewy et al.
taposing during the 1.3–1.0 Ga Grenville –Sunsas orogeny (Chew 2004), and are probably coeval with Famatinian (Early Ordovi-
et al. 2007a), as first postulated by Loewy et al. (2004). cian) metamorphism and subduction-related magmatism on the
western Gondwanan margin (Chew et al. 2007b).
Structural framework
Stratigraphy
Very limited work has been published on the nature of the basin in
which the sediments of the Chiquerı́o and San Juan formations The Chiquerı́o Fm. rests unconformably on gneisses of the Are-
accumulated. The present-day outcrop distribution of the Chi- quipa massif and consists of nearly 400 m of diamictite, sandstone,
querı́o and San Juan formations is restricted to a thin zone, mudstone and carbonate, with a thin (11 m) finely laminated dolos-
20 km wide, along a strike length of 200 km on the western tone and dolomicrite unit at the top of the formation (Fig. 44.2,
margin of the Arequipa massif in southern Peru (Fig. 44.1a). Chew et al. 2007a). It is overlain by the San Juan Fm., a
However, the initial geometry and tectonic setting of this Late 2-km-thick carbonate succession (Fig. 44.3). The San Juan Fm.
Neoproterozoic –?Early Palaeozoic basin is uncertain. Juvenile consists of several hundred metres of massive beige dolomite,
extensional magmatism (dacite dykes) has been dated at 635 + with subordinate thinly bedded limestone, black shale, graded
5 Ma in the basement of the Antofalla terrane in Northern Chile pebbly dolostone and phyllite (Fig. 44.3, Chew et al. 2007a). No
(Loewy et al. 2004). Late Neoproterozoic extension-related unconformities have been observed in either formation. At many
volcanism related to Rodinia break-up has been identified in the localities, the San Juan Fm. rests directly on the basement gneiss
Puncoviscana fold belt of northwestern Argentina (Omarini et al. (Caldas 1978).
1999), although the Puncoviscana Basin has a complex history
of extension, compression and magmatism (Ramos 2008) that
may not be directly comparable to the tectonic evolution of the Glaciogenic deposits and associated strata
basin where the sediments of the Chiquerı́o and San Juan for-
mations accumulated. Considering a Late Neoproterozoic age for The following description of the glaciogenic deposits of the Chi-
the Chiquerı́o and San Juan formations (e.g. Caldas 1978, 1979; querı́o and San Juan formations is derived chiefly from Chew
Chew et al. 2007a; Loewy et al. 2003, 2004), an extensional et al. (2007a), based on the well-exposed coastal section SE of
basin setting is most likely. the town of San Juan (Fig. 44.1b). The basal siliciclastic section
Shackleton et al. (1979) proposed a structural evolution of the of the Chiquerı́o Fm. is 348 m thick (Fig. 44.2), and consists pri-
Chiquerı́o and San Juan formations. Two deformation events are marily of massive diamictite with poorly developed internal strati-
recognized within these rocks. Two earlier deformation events fication. The matrix of the diamictite is a dark meta-siltstone,
(D1 and D2) are restricted to the underlying Arequipa massif base- whereas the majority of the clasts are granitic gneiss that superfi-
ment, so the deformation events affecting the Chiquerı́o and San cially resemble the underlying Arequipa massif basement
Juan formations are attributed to the regional D3 and D4 defor- (Loewy et al. 2004; Chew et al. 2007a). The clasts display no evi-
mation events (Shackleton et al. 1979). The earlier event (D3) pro- dence of faceting or striation. Clast types include weakly foliated,
duced a bedding-parallel schistosity (S3) formed by fine-grained K-feldspar rich granites, foliated grey-pink gneiss (sometimes
muscovite and biotite. Where suitable strain markers are present megacrystic), fine-medium grained sandstone blocks and clasts
(such as in the conglomeratic portions of the Chiquerı́o Fm.), a of amphibolite. Some of the gneissic clasts are greater than
strong lineation is aligned along this foliation surface, plunging 50 cm across. There is only very occasional evidence of stratifica-
moderately to the south. The S3 schistosity is crenulated and tion in the massive diamictite portion of the sequence. This strati-
folded by a second deformation (D4), which controls the large- fication is present in the form of thin, discontinuous mudstone
scale distribution of the units. Peak metamorphic conditions layers and occasional lenses, up to 3 m thick, of boulder conglom-
were attained during the D3 event, and the muscovite and biotite erate. These conglomeratic lenses are poorly sorted, and have a
assemblages present are indicative of the greenschist facies high concentration of clasts, approaching being clast supported.
(Shackleton et al. 1979). Undeformed fine-grained granite dykes Between 76 and 152 m above the basement contact there is a
cut the F3 and F4 fold axial planes, and one such granite dyke sequence of stratified diamictite interbedded with thin siltstone
has yielded a U –Pb zircon lower intercept age of between 468 and graded sandstone beds. The stratification in the diamictite is
THE CHIQUERÍO FORMATION 483

Fig. 44.2. (a) Stratigraphic section of the Chiquerı́o Fm. and the basal portion
of the San Juan Fm. The line of section is illustrated in Figure 44.1b. (b) Log of
the carbonate portion of the upper Chiquerı́o Fm. This section (359–410 m) is
indicated by a box in (a).

defined by either thin siltstone beds or mudstone lenses. This Fig. 44.3. Stratigraphic section and C-isotopic trends through the upper part of
internal stratification is much more pronounced and more continu- the Chiquerı́o Fm. and the San Juan Fm. Stratigraphic heights are in metres
ous than that seen in the underlying massive diamictite. The stra- above the Chiquerı́o Fm.– Arequipa massif basement contact. The line of the
tified diamictite contains abundant outsized clasts of granitic section is illustrated in Figure 44.1b.
gneiss that deflect underlying lamina. A large angular megaclast
of bedded sandstone, greater than 6 m across (Fig. 44.2) is also
present within diamictite in this part of the sequence. It appears (transitional over 1 m) to carbonate-dominated sedimentation
to pierce the crude lamination in the underlying diamictite. occurring at 348 m (Fig. 44.2). The dominant lithology is a calcar-
The upper part of the Chiquerı́o Fm. and the overlying San Juan eous diamictite with white dolostone and limestone clasts, and
Fm. are predominantly carbonate, with a relatively abrupt switch only minor amounts of granitic gneiss. The carbonate clasts are
484 D. CHEW & C. KIRKLAND

strongly flattened and stretched along the main bedding-parallel þ2.5‰. So far, no unequivocal glaciogenic strata nor significant
tectonic fabric (S3). Overlying the carbonate diamictite are 11 m sequence boundary associated with this younger, strongly nega-
of finely laminated (0.2 –5 cm), fine-grained pink dolostone and tive ( –5‰ to – 8‰) d13C excursion have been identified (Chew
dark dolomicrite in the upper Chiquerı́o Fm. (Chew et al. 2007a). et al. 2007a).
This dolomite unit shows no internal structure apart from prominent
lamination; it has no evidence of outsized clasts. The San Juan Fm.
overlies this fine-grained laminated dolostone –dolomicrite unit. Other characteristics (e.g. economic deposits, biomarkers)
The basal portions of the formation consist of several hundred
metres of predominantly massive beige dolomite (Fig. 44.3, The Marcona deposit (20 km north of San Juan, Fig. 44.1a) and the
Chew et al. 2007a). This is overlain by a lithologically varied associated Pampa de Pongo deposit (35 km east of San Juan) are
unit of black shale, massive dolomite, and thinly bedded graded the largest Fe accumulations, with associated copper and gold,
pebbly dolostone (950 –1093 m; Fig. 44.3). The overlying unit is along the western South America margin. The deposit substan-
170 m thick, and consists of thinly bedded limestone and dark tially post-dates the deposition of the glaciogenic strata, and is con-
micrite (Fig. 44.3). Above this unit, there is a thick (nearly 1 km) sidered to have formed during a phase of Mesozoic arc magmatism
sequence of massive dolomite that is only briefly interrupted by (Hawkes et al. 2002). Approximate resources include more than
the deposition of a thin package of graded pebbly dolostone and 1400 Mt of iron ore at Marcona and 1000 Mt of magnetite miner-
mudstone (1395 – 1487 m; Fig. 44.3). alization at Pampa de Pongo (Hawkes et al. 2002). The two depos-
its form part of a cluster of similar occurrences that together define
the ‘Marcona Fe –Cu District’. The larger Fe bodies are located
Boundary conditions with overlying and underlying within the Chiquerı́o, San Juan and Marcona formations, and
non-glacial units also by basaltic andesite, andesite and volcaniclastic rock of the
Middle to Upper Jurassic Rio Grande Fm. (Hawkes et al. 2002).
The basal contact of the Chiquerı́o Fm. is an unconformity with There are no biostratigraphic data available for the Chiquerı́o
gneissic rock of the Arequipa massif basement, best seen at low Fm. ‘Stromatolite-like’ structures have been recorded in the over-
tide, along the coast, 5 km SE of San Juan. At the regional scale, lying San Juan Fm. (Injoque & Romero 1986) and are corre-
it is unknown whether the unconformity surface is planar or has lated by the authors with late Neoproterozoic– Early Cambrian
topography. At many other localities (e.g. Punta San Juan, 2 km stromatolites.
west of the town of San Juan, Fig. 44.1b), the contact between
the Chiquerı́o Fm. and the basement gneiss is a fault. In addition, Palaeolatitude and palaeogeography
the entire Chiquerı́o Fm. is frequently excised, with the overlying
San Juan Fm. resting directly on the basement gneiss (Caldas Palaeolatitudinal and palaeogeographic constraints for the Chi-
1978). querı́o Fm. are sparse. There have been no palaeomagnetic
The upper contact of the Chiquerı́o Fm. has been placed at either studies on the Chiquerı́o Fm. Given that the Chiquerı́o and San
the base (e.g. Caldas 1978) or at the top of the finely laminated, Juan formations have experienced greenschist-facies metamorph-
11-m-thick unit of pink dolostones and dark dolomicrites (e.g. ism (Shackleton et al. 1979) and were subsequently intruded by
Cobbing 1981; Chew et al. 2007a). The Chiquerı́o and San Juan Early Palaeozoic plutons (Loewy et al. 2004), remagnetization
formations are overlain unconformably by unmetamorphosed by Early Palaeozoic-age metamorphism is likely.
Mesozoic and Cenozoic sediments (Caldas 1978). The detrital zircon data of Chew et al. (2007a) from the Chi-
querı́o and San Juan formations are consistent with derivation
from the Proto-Andean margin (Chew et al. 2007b). This would
Chemostratigraphy imply that both the glaciogenic strata and its underlying gneissic
basement were proximal to the South American craton (Amazonia)
Chew et al. (2007a) presented C- and O-isotope data for the during Late Neoproterozoic times (Chew et al. 2007a). This juxta-
upper part of the Chiquerı́o Fm. and the overlying San Juan position indicates that the Arequipa massif basement must have
Fm. (Fig. 44.3). These are the only chemostratigraphic data pre- accreted earlier, probably during the 1.3– 1.0 Ga Grenville –
sently available for the Chiquerı́o and San Juan formations. Sunsas Orogeny (Chew et al. 2007a) as first postulated by
Diagenetic overprinting of the original seawater isotopic signa- Loewy et al. (2004). Ramos (2008) and Loewy et al. (2004)
tures is difficult to assess. Detailed textural evidence to evaluate provide a detailed synthesis on the tectonic evolution and
diagenesis within the carbonate rocks is lacking as the rocks docking history of the Arequipa massif and the Antofalla terrane.
have undergone some recrystallization during greenschist-facies An autochthonous origin for the Chiquerı́o Fm., with respect to
metamorphism, and there is no alternative complete section cratonic South America, places crude palaeolatitudinal constraints
with which to compare lateral variations in the stable on these rocks. There is presently only one palaeomagnetic pole
isotope profile. The beginning of carbonate-dominated sedimen- for the Amazon craton in the Late Neoproterozoic (Tohver et al.
tation in the Chiquerı́o Fm. occurs at 348 m (Fig. 44.2), where 2006), derived from the palaeomagnetic study of the Neoprotero-
the dominant lithology is a calcareous diamictite with white zoic Puga cap carbonate (Trindade et al. 2003). The dolomite and
dolostone and limestone clasts. Both the clasts within the dia- limestone of the Puga Fm. from the SE Amazon craton preserve a
mictite and the interbedded limestone beds yield d13C values dual-polarity component that is interpreted as a primary magneti-
between 0‰ and þ2‰ (VPDB) (Fig. 44.3). Finely laminated zation. This implies a low palaeolatitude of 22 þ 6/ –58 for the
pink dolostone and dark dolomicrite overlie the carbonate dia- Amazonian block just after deposition of the Puga diamictites.
mictite. This dolostone unit yields consistent negative d13C Although direct ages for the Puga Fm. are not yet available,
values of –2‰ (Fig. 44.3, Chew et al. 2007a). 87
Sr/86Sr ratios and d13C results presented by de Alvarenga et al.
The overlying San Juan Fm. exhibits a recovery in d13C values (2004) suggest correlation with the c. 635 Ma (Hoffmann et al.
to between þ1‰ and þ2‰ (Fig. 44.3). The basal portions of the 2004; Condon et al. 2005) post-glacial units of the Congo craton.
formation consist of several hundred metres of predominantly
massive beige dolomite. Between 1075 m and 1250 m, a thinly
bedded limestone and dark micrite unit exhibits strongly negative Geochronological constraints
d13C values from –5‰ to –8‰ (five data points, Fig. 44.3, Chew
et al. 2007a). Above this unit, there is a return to deposition of To date, no tuffs have been recorded from the Chiquerı́o Fm.
massive dolomite and the d13C values range between þ1‰ and Existing age constraints include a minimum age of 468– 440 Ma
THE CHIQUERÍO FORMATION 485

c Proterozoic
subsidiary peak at c. 1000 Ma (Chew et al. 2007a, Fig. 44.4).
SJ-16 is a sample of diamictite matrix from the Chiquerı́o Fm. It
SJ-57, n=55/66 Pebbly limestone,
is also characterized by a restricted age distribution from 950 to
5 6 1300 Ma, with a prominent peak at c. 1200 Ma and a subsidiary
697 ± 11 Ma San Juan Formation peak at c. 1000 Ma (Chew et al. 2007a, Fig. 44.4). The detrital
4 5 zircon data from these Chiquerı́o Fm. samples yield very
minimal detritus, which could potentially be derived from the
4 underlying basement (the Palaeoproterozoic Arequipa massif,
3 1790 –2020 Ma; Loewy et al. 2004). Sample SJ-57 (55 grains) is
3 from a coarse pebbly limestone bed from the San Juan Fm.,
2 2 1412 m above the Chiquerı́o Fm. – Arequipa massif basement
contact and 178 m above the second negative C-isotope excursion
1 1 (Fig. 44.4). The majority of grains from this sample also lie in the
950– 1300 Ma range, with peaks at c. 1000 Ma and c. 1200 Ma.
There are also minor peaks within the c. 1600–2000 Ma and
9 SJ-16, n=62/71 Siltstone (tillite matrix) 14 c. 700 –830 Ma intervals (Chew et al. 2007a, Fig. 44.4). A
8 Chiquerío Formation c. 700 Ma grain provides a maximum age constraint for the depo-
932 ± 28 Ma 12 sition of this portion of the San Juan Fm.
7
6 10
5 8 Discussion
4
6 The depositional environment of the Chiquerı́o Fm. is most likely
3
2 4 deep marine, based on the lithofacies and the lack of high-energy
sedimentary structures. The stratified portion of the siliciclastic
1 2 section (between 76 and 152 m above the basement contact,
Fig. 44.2) contains abundant dropstones of granitic gneiss,
9
Turbiditic sandstone 10
8 SJ-11, n=35/46
which disrupt the lamination in graded turbiditic sandstone beds,
9 thus suggesting a glacially influenced marine environment
Chiquerío Formation 8
7 955 ± 18 Ma (Caldas 1978; Cobbing 1981; Loewy et al. 2004; Chew et al.
7 2007a). The depositional environment of the massive diamictite
6
6 portions straddling this stratified interval (Fig. 44.2) may represent
Probability x 10-3

5 ice-rafted debris and suspension settling of fine grained sediment,


5
4 or alternatively may have been produced by submarine debris
4
Frequency

3 flows. There is no sedimentary evidence in the siliciclastic


3 portion of the sequence (e.g. wave ripples, cross-bedding) of
2 2 shallow-water conditions (Chew et al. 2007a). The San Juan Fm.
1 1 consists predominantly of massive beige dolomite with little
0 0 internal structure. The possible Precambrian stromatolites
described by Injoque & Romero (1986) would suggest a shallow
400
500
600
700
800
900
1000
1100
1200
1300
1400
1500
1600
1700
1800
1900
2000

marine or intertidal environment for portions of the San Juan Fm.


The laminated dolostone facies at the top of the Chiquerı́o Fm.
and its associated large negative d13C excursion are characteristic
Age (Ma) of cap dolostone associated with Late Neoproterozoic glacials
Fig. 44.4. Zircon probability density distribution diagrams from the Chiquerı́o (Kennedy et al. 1998; Hoffman & Schrag 2002; Halverson et al.
Fm. (SJ-11, SJ-16) and the San Juan Fm. (SJ-57) (Chew et al. 2007a). Light 2005; Shields 2005), although negative anomalies in the late Neo-
grey curves represent all ages from each sample, and the dark curves represent proterozoic are not exclusively linked to ice ages (Le Guerroué
ages that are .90% concordant. The youngest detrital zircon age in each et al. 2006). Chew et al. (2007a) considered the Chiquerı́o Fm.
sample is shown within a black box. and the pronounced negative C-isotope excursion in the San Juan
Fm. to represent two distinct glacial events correlated to a ‘Stur-
tian–Marinoan’ couplet elsewhere in the world. Although no
based on a loosely defined U –Pb TIMS zircon lower intercept unequivocal glaciogenic strata nor a significant sequence boundary
from the cross-cutting, post-tectonic San Juan granite of the San indicative of a glacial event have been identified with the second
Nicolas batholith (Loewy et al. 2004), and maximum ages of negative C-isotope excursion, it may correlate with the negative
932 + 28 Ma and 955 + 18 Ma (the youngest detrital U –Pb Trezona anomaly, which immediately preceded the Marinoan gla-
SIMS zircon ages from the study of Chew et al. 2007a). ciation (Halverson et al. 2005). Alternatively, if Chiquerı́o Fm.
Loewy et al. (2004) also dated three clasts from the Chiquerı́o and the C-isotope excursion represent a Marinoan-age glacial
Fm. SE of San Juan by U –Pb TIMS zircon. Two clasts of pink, event and the Shuram/Wonoka isotopic anomaly (Halverson et al.
weakly foliated, K-feldspar rich megacrystic granite yielded ages 2005; Le Guerroué et al. 2006), then a depositional age of c.
of 1168 þ 9/–6 Ma, 1162 + 6 Ma, while a third clast of similar 635 Ma for the Chiquerio Fm. (Chew et al. 2007a) is inferred. In
composition but with a gneissic foliation yielded a poorly con- either case, a Late Neoproterozoic (,700 Ma) age is supported by
strained upper intercept of c. 1165 Ma. Chew et al. (2007a) under- the youngest detrital zircon population in the San Juan Fm. The
took U –Th –Pb SIMS analyses of detrital zircons from three Proto-Andean margin of Amazonia is characterized by abundant
samples from the Chiquerı́o and San Juan formations. Combined zircon detritus between 1300–900 Ma and 650–550 Ma (Chew
age (probability-density distribution) plots and histograms for 2007b). The absence of Late Neoproterozoic (700 Ma and
the three samples from that study are illustrated in Figure 44.4. younger) zircon in the Chiquerı́o Fm. may simply reflect that
Sample SJ-11 is from a thin graded turbiditic sandstone bed these glaciogenic strata were too old to accumulate such detritus,
from the Chiquerı́o Fm. It yields a restricted age distribution of or alternatively had a restricted sediment source, mainly derived
950– 1300 Ma, with a prominent peak at c. 1200 Ma and a from local basement.
486 D. CHEW & C. KIRKLAND

Further global correlation of the Chiquerı́o Fm. and the pro- Belt. In: Porter, T. M. (ed.) Hydrothermal Iron Oxide Copper –Gold
nounced negative C-isotope excursion in the San Juan Fm. are & Related Deposits: A Global Perspective. PGC Publishing, Ade-
hampered by the lack of consensus on the temporal range of the laide, Australia, 2, 115–130.
‘Sturtian’ glacial episode (see Hoffman & Li 2009 for a review), Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
the lack of absolute age constraints for the Chiquerı́o and San testing the limits of global change. Terra Nova, 14, 129–155.
Juan formations, and the relatively low-resolution sampling Hoffman, P. F. & Li, Z.-X. 2009. A palaeogeographic context for Neopro-
employed for the stable isotope study. Further areas of research terozoic glaciation. Palaeogeography, Palaeoclimatology, Palaeo-
that might prove beneficial in the future include higher-resolution ecology, 277, 158– 172.
Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
sampling for stable isotope analysis in the San Juan Fm. and a
2004. U– Pb zircon date from the Neoproterozoic Ghaub
comprehensive search for tuffs in the glaciogenic strata. Formation, Namibia: constraints on Marinoan glaciation. Geology,
32, 817– 820.
This study was funded by the Swiss National Science Foundation under a grant
Injoque, J. & Romero, L. 1986. Estromatolitos (?) en la formación San
held by U. Schaltegger. We are extremely grateful to Urs Schaltegger at the Uni-
Juan, San Juan de Marcona. Evidencia de estructuras fósiles precám-
versity of Geneva, the Geological Survey of Peru (INGEMMET) and C. Moreno
bricas en el Perú. De re metallic.De la minerı́a y los metales: Revista
of S. Marcos University in Lima for scientific advice and logistical support during
del Instituto Geológico Metalúrgico, 11, 4 –5.
the field seasons in Peru. The careful and insightful reviews of V. A. Ramos, E. Le
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K. H. &
Guerroué and editor E. Arnaud are gratefully acknowledged. This work represents
Arthur, M. A. 1998. Two or four Neoproterozoic glaciations?
a contribution of the IUGS- and UNESCO-funded IGCP (International
Geology, 26, 1059–1063.
Geoscience Programme) project #512.
Le Guerroué, E., Allen, P. A., Cozzi, A., Etienne, J. L. & Fanning,
C. M. 2006. 50 Myr recovery from the largest negative d13C excur-
References sion in the Ediacaran ocean. Terra Nova, 18, 147– 153.
Loewy, S. L., Connelly, J. N., Dalziel, I. W. D. & Gower, C. F. 2003.
Caldas, J. 1978. Geologı́a de los Cuadrángulos de San Juan, Acarı́ y Eastern Laurentia in Rodinia: constraints from whole-rock Pb and
Yauca, Hojas: 31-m, 31-n, 32-n. Instituto de Geologı́a y Minerı́a, U/Pb geochronology. Tectonophysics, 375, 169– 197.
Lima. Loewy, S. L., Connelly, J. N. & Dalziel, I. W. D. 2004. An orphaned
Caldas, J. 1979. Evidencias de una glaciación Precambriana en la costa basement block: The Arequipa –Antofalla basement of the central
sur del Perú. Segundo Congreso Geológico Chileno J, Arica, 29 – 37. Andean margin of South America. Geological Society of America
Chew, D. M., Kirkland, C. L., Schaltegger, U. & Goodhue, R. Bulletin, 116, 171–187.
2007a. Neoproterozoic glaciation in the Proto-Andes: tectonic impli- MARCONA MINING COMPANY. 1968. Geologic Map of the Marcona
cations and global correlation. Geology, 35, 1095– 1099. iron deposits.
Chew, D. M., Schaltegger, U., Košler, J., Whitehouse, M. J., Omarini, R. H., Sureda, R. J., Götze, H. J., Seilacher, A. & Pflüger,
Gutjahr, M., Spikings, R. A. & Miškovic, A. 2007b. U– Pb geo- F. 1999. Puncoviscana folded belt in northwestern Argentina: testi-
chronologic evidence for the evolution of the Gondwanan margin mony of Late Proterozoic Rodinia fragmentation and pre-Gondwana
of the north-central Andes. Geological Society of America Bulletin, collisional episodes. International Journal of Earth Sciences, 88,
119, 697– 711. 76 – 97.
Cobbing, E. J. 1981. Tillites at the base of the possible Early Palaeozoic Ramos, V. A. 2008. The basement of the Central Andes: the Arequipa and
Marcona Formation, southwest coastal Peru. In: Hambrey, M. J. & related terranes. Annual Review of Earth and Planetary Sciences, 36,
Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. 289– 324.
Cambridge University Press, Cambridge, 899– 901. Shackleton, R. M., Ries, A. C., Coward, M. P. & Cobbold, P. R. 1979.
Cobbing, E. J. & Pitcher, W. S. 1972. Plate tectonics and the Peruvian Structure, metamorphism and geochronology of the Arequipa
Andes. Nature, 246, 51– 53. Massif of coastal Peru. Journal of the Geological Society, 136,
Condon, D., Zhu, M. Y., Bowring, S., Wang, W., Yang, A. H. & Jin, 195– 214.
Y. G. 2005. U–Pb ages from the Meoproterozoic Doushantuo For- Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal
mation, China. Science, 308, 95 –98. of existing models and the plumeworld hypothesis. Terra Nova, 17,
de Alvarenga, C. J. S., Santos, R. V. & Dantas, E. L. 2004. C –O– Sr 299– 310.
isotopic stratigraphy of cap carbonates overlying Marinoan-age Tohver, E., D’Agrella-Filho, M. S. & Trindade, R. I. F. 2006. Paleo-
glacial diamictites in the Paraguay Belt, Brazil. Precambrian magnetic record of Africa and South America for the 1200–500 Ma
Research, 131, 1 –21. interval, and evaluation of Rodinia and Gondwana assemblies.
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Precambrian Research, 147, 193– 222.
Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon- Trindade, R. I. F., Font, E., D’Agrella-Filho, M. S., Nogueira,
isotope record. Geological Society of America Bulletin, 117, A. C. R. & Riccomini, C. 2003. Low-latitude and multiple geomag-
1181– 1207. netic reversals in the Neoproterozoic Puga cap carbonate, Amazon
Hambrey, M. J. & Harland, W. B. 1981. Earth’s Pre-Pleistocene craton. Terra Nova, 15, 441–446.
Glacial Record. Cambridge University Press, London & New York. Wilson, P. A. 1975. Potassium– argon age studies in Peru with particular
Hawkes, N., Clark, A. H. & Moody, T. C. 2002. Marcona and Pampa de reference to the chronology of emplacement of the coastal batholith.
Pongo: Giant Mesozoic Fe-(Cu, Au) deposits in the Peruvian Coastal PhD thesis, University of Liverpool.
Chapter 45

Glacially influenced sedimentation of the Puga Formation, Cuiabá Group


and Jacadigo Group, and associated carbonates of the Araras
and Corumbá groups, Paraguay Belt, Brazil

CARLOS J. S. ALVARENGA1 *, PAULO C. BOGGIANI2, MARLY BABINSKI2, MARCEL A. DARDENNE1,


MILENE F. FIGUEIREDO2, ELTON L. DANTAS1, ALEXANDRE UHLEIN3, ROBERTO V. SANTOS1,
ALCIDES N. SIAL4 & ROLAND TROMPETTE5
1
Instituto de Geociências, Universidade de Brası́lia, Campus Darcy Ribeiro, CEP 70910-900, Brası́lia, Brazil
2
Instituto de Geociências, Universidade de São Paulo, Rua do Lago, 562, CEP 05508-080, São Paulo, Brazil
3
Instituto de Geociências, Universidade Federal de Minas Gerais, Av. Antônio Carlos, 6627, Campus da Pampulha,
CEP 31270-901, Belo Horizonte, Brazil
4
Departmento de Geologia, Universidade Federal de Pernambuco, CEP 50732-970, Recife, PE, Brazil
5
CEREGE, University d’Aix-Marseille III, CNRS-FU017, BP 13545, Aix-en-Provence, Cedex 04, France
*Corresponding author (e-mail: alva1@unb.br)

Abstract: Discontinuous exposures of diamictite (.1000 km) termed the Puga Formation (Fm.) and interpreted as being related to a late
Cryogenian glacial event are known in the Paraguay Belt (Brazilian– Pan-African Orogeny) and part of the Amazon craton and Rio Apa
Block. These diamictite units and a mixed assemblage of sandstone, conglomerate and claystone were first named the Jangada Group.
Recent interpretations have shown that the diamictite of the Puga Formation passes laterally into metasediments of the Cuiabá Group,
interpreted as glacially influenced turbidites. The correlative Jacadigo Group in the southern Paraguay Belt includes a thick succession of
banded iron formation (BIF) bearing large boulders that have been interpreted as recording glacially influenced sedimentation. The dia-
mictite of the Puga Fm. is overlain by two different carbonate-bearing successions, the Corumbá Group (Cadieus, Cerradinho, Bocaina,
Tamengo and Guaicurus formations) in the south and the Araras Group (Mirassol d’Oeste, Guia, Nobres formations) in the north. Evi-
dence of glaciation in the Puga Fm. consists of striated and faceted pebbles and blocks, and dropstones in the turbidites. Sedimentary and
geochemical data from the associated carbonate reinforce the interpretation of a glacial origin. C, O and Sr isotope data from the northern
Paraguay Belt are consistent with the proposed late Cryogenian age for the Puga Fm. sedimentation.

Discontinuous outcrops of the Puga Fm. extend for more than by the economic manganesiferous bed and banded iron formation
1000 km along the Paraguay Belt and over the Amazonian (BIF) of the Santa Cruz Fm. (type section, 198 110 24.3600 S, 578
Craton and Rio Apa Block (Fig. 45.1). The glacial diamictite 360 31.4500 W). Fe formations bearing large granite boulders have
succession was first described from the isolated Puga Hill (198 been interpreted as recording a glacial influence (Barbosa 1949;
370 20.0300 S, 578 310 40.0100 W), SE of Corumbá (Fig. 45.1; Maciel Walde et al. 1981; Hoppe et al. 1987; Urban et al. 1992; Klein
1959), but its best exposures are located in the northern Paraguay & Ladeira 2004; Walde & Hagemann 2007). The Jacadigo
Belt, where the diamictite of the Puga Fm. passes laterally into tur- Group, in contrast to the huge exposure area of the Puga Fm.
bidites with dropstones of the Cuiabá Group (Fig. 45.2) (Alvarenga and Cuiabá Group in the northern Paraguay Belt, is restricted to
& Trompette 1992). A glaciogenic origin was first proposed by a few hills on the border of Bolivia and Brazil near Corumbá
Maciel (1959), but Alvarenga & Trompette (1992) were the first (Fig. 45.1).
to suggest a glaciomarine setting; they interpreted the fine-grained There are two Fe formation units in the southern Paraguay Belt.
sediments of the Cuiabá Group as glacially influenced turbidites. The most significant is the economic Fe formation and Mn ore
The diamictite exposures in the northern Paraguay Belt were deposits of the Jacadigo Group in Urucum Hill (Almeida 1945l
first described by Almeida (1964a, b) as part of Acorizal, Dorr II 1945); the second is an Fe formation in diamictite of the
Engenho, Bauxi and Marzagão formations within the Jangada Puga Formation (Boggiani et al. 2006; Piacentini et al. 2007).
Group. The type section was described between Jangada and Carbonate rocks of the Araras Group and mixed clastic-
Bauxi (Fig. 45.1). Quartzite and sandstone units below the diamic- carbonate rocks of the Corumbá Group overlie diamictite of the
tite (Puga Fm.) and above the slate (Cuiabá Group) were described Puga Fm. and in some areas are interpreted as cap carbonate
as the Bauxi Fm. (Almeida 1964a; Vieira 1965), whereas diamic- (Nogueira et al. 2003; Boggiani et al. 2003; Alvarenga et al.
tite and associated conglomerate, quartzite and slate were named 2004, 2008). Historically, ‘Corumbá and Arara Limestones’
the Jangada Group (Almeida 1964a, b; Rocha-Campos & Hasuı́ were the first names given to these carbonate rocks in the Paraguay
1981). The name ‘Jangada Group’ is not used in the current Belt (Evans 1894). Most studies about carbonate rocks that overlie
nomenclature. the diamictite of the Puga Fm. focus on isotope geochemistry,
In the southern Paraguay Belt, glacial deposits include the Jaca- chemostratigraphy (Boggiani et al. 2003; Alvarenga et al. 2004,
digo Group and the Puga Fm. These two units occur in isolated 2008; Font et al. 2006; Nogueira et al. 2007; Riccomni et al.
hills making it difficult to confirm the stratigraphic relationships 2007) and palaeomagnetic data (Trindade et al. 2003).
between them. The Jacadigo Group (Almeida 1945; Dorr II Outcrops of the Araras Group extend for more than 600 km
1945; Almeida 1946) is exposed in some hills south of the town along the Northern Paraguay Belt. This group is subdivided into
of Corumbá extending into Bolivia. The Urucum Fm. (lower three carbonate formations (Fig. 45.1, Table 45.1): (i) the Mirassol
unit) is characterized by conglomerate and arkose and is overlain d’Oeste Fm., described as a cap dolomite (Nogueira et al. 2003;

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 487– 497. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.45
488 C. J. S. ALVARENGA ET AL.

Fig. 45.1. Geological map of the Paraguay


Belt (modified from Alvarenga &
Trompette 1993; Boggiani 1998).

Alvarenga et al. 2004, 2008), (ii) the Guia Fm. (Almeida 1964a; (148 440 13.700 S, 578 490 52.4500 W) quarries and in a drill core
Hennies 1966), and (iii) the Nobres Fm. (Hennies 1966; Luz & from the Bauxi region (158 090 06.1400 S, 568 410 00.7400 W). The
Abreu 1978). Good exposures of the Mirassol d’Oeste Fm. over- Guia Fm. can be found in the Terconi and Tangará quarries
lying the Puga Fm. in the Northern Paraguay Belt can be found as well as at the Cimento Tocantins Mine (148 380 12.0200 S,
at the Terconi (158 400 41.8700 S, 588 040 18.2200 W) and Tangará 568 150 57.5000 W) in the Nobres region and the Nossa Senhora
PUGA FORMATION, PARAGUAY BELT 489

Platformal glaciomarine depositional system Glacial diamictite-turbidite

Slope depositional system Outer slope depositional system


Araras Group

Puga Formation
Bauxi Formation
Amazonian Craton
(3.0 - 1.0 Ga.)
Cuiabá Group

Diamictite Conglomerate and Fine-grained sediments


graded pebbly sandstone with occasional clasts

Fig. 45.2. Schematic cross-section showing the glacial facies relationships for the late Cryogenian glacial event (c. 635 Ma) along the southeastern edge of the
Amazonian Craton (modified from Alvarenga & Trompette 1992). Diamictites of the Puga Fm. are in the platformal glaciomarine system, while the diamictites of the
Cuiabá Group are in the slope and outer slope depositional system.

da Guia Quarry (158 210 15.0700 S, 568 110 24.4700 W). The Nobres Structural framework
Fm. is exposed in the Provı́ncia Serrana.
The Corumbá Group, in the southern Paraguay Belt, is subdi- Outcrops of the Neoproterozoic rocks extend along the southeast-
vided into five formations (Tables 45.2 & 45.3). The basal ern border of the Amazonian Craton as well as along the eastern
Cadieus and Cerradinho formations occur only in the Serra da Bod- margin of the Rio Apa Block, and contain a thick sequence of gla-
oquena (Fig. 45.1, Table 45.2; Boggiani 1998). Above this unit lies ciomarine, turbidite, carbonate and siliciclastic sedimentary rocks
the grey dolomite of the Bocaina Fm. (Almeida 1945), which is formed on an extensional continental margin. In the northern Para-
exposed in the Serra da Bodoquena and Corumbá regions and guay belt, the Puga Fm. is mostly subhorizontal and c. 100 m thick
which outlines the geographical extent of exposures of the on the craton, whereas in the central part of the belt, in the deepest
Corumbá Group in the southern Paraguay Belt. The pink carbonate part of the basin, the glacially influenced sediments (Cuiabá
of the Bocaina Fm. overlies the Puga Fm. at Puga Hill (Table 45.3). Group) are more than 3000 m thick (Alvarenga & Saes 1992;
The uppermost two units of the Corumbá Group are the Tamengo Alvarenga & Trompette 1992, 1993; Dantas et al. 2009). In the
and Guaicurus formations (Almeida 1965) exposed in the southern Paraguay Belt, the NE – SW-trending extensional struc-
Corumbá region and in the eastern part of Serra da Bodoquena. tures, with sub-vertical faults of a half-graben system, were filled
In the northern Paraguay belt, two Neoproterozoic diamictite- by conglomerate, arkose sandstone and jaspilite of the Jacadigo
bearing glacial intervals are described (Table 45.1): the units Group. The present landscape in the south is due to a later
described in this chapter and attributed to a late Cryogenian glacia- tectonic event and erosion during the Neogene subsidence of the
tion (c. 630 Ma), as well as the Serra Azul Fm. attributed to the Pantanal Basin, with blocks defined by vertical faults (Trompette
Ediacaran glaciation (580 Ma; Alvarenga et al. 2007; Figueiredo et al. 1998).
et al. 2008, 2011). Most rocks in the region have experienced The overlying Corumbá and Araras groups were deposited in a
low-grade metamorphism. In order to facilitate subsequent rock passive margin environment. The total thickness of the sedimen-
descriptions, the prefix ‘meta’ has been omitted. tary pile increases from c. 200 m in the western cratonic area

Table 45.1. Neoproterozoic lithostratigraphy nomenclature in the northern Paraguay Belt

Period Group Formation Lithology Ages


Diamantino Siltstone and arkose sandstone

Alto Paraguay Sepotuba Siltstone Rb–Sr: 569 ± 20 Ma


Raizama Sandstone
Ediacaran
Serra Azul Diamictite and mudstone

Nobres Dolostone and sandy dolostone


Araras Guia Limestone and mudstone
Mirassol d'Oeste Pink dolostone Pb/Pb: 633 ± 25 Ma

Puga Formation Diamictite


Cryogenian Diamictite,
conglomerate,
Cuiabá Group sandstone and
fine-grained rocks
490 C. J. S. ALVARENGA ET AL.

Table 45.2. Neoproterozoic lithostratigraphy nomenclature in the Serra da Bodoquena region, southern Paraguay Belt

Period Group Formation Lithology U–Pb Ages

Guaicurus Shale
Tamengo Limestone, mudstone and shale 543 ± 2 Ma

Corumbá Bocaina Dolostone


Ediacaran
Cerradinho Limestone, shale, sandstone
Cadieus Conglomerate, arkose and shale

Puga Diamictite
Cryogenian

(inner-shelf domain in Mirassol d’Oeste region) to more than Stratigraphy


1000 m eastwards (middle-outer shelf domain) along the Provı́ncia
Serrana (Alvarenga & Saes 1992; Alvarenga & Trompette 1993; The Puga Fm. and the Cuiabá and Jacadigo Groups include diamic-
Trompette 1994; Alvarenga et al. 2008, 2009). tite and fine-grained rocks with outsized clasts. In the north Para-
The Neoproterozoic rocks along the Paraguay Belt were folded guay Belt, the Puga Fm. and Cuiabá Group are overlain by the
and metamorphosed to very low-grade metamorphism during the carbonate rocks of the Mirassol d’Oeste Fm., Araras Group
Brasiliano/Pan-African Orogeny, in the Ediacaran – Cambrian (Table 45.1; Nogueira et al. 2003; Alvarenga et al. 2008, 2009).
Period (Alvarenga 1990; Trompette 1994; Pimentel et al. 1996; In the south Paraguay Belt glaciogenic deposits are named the
Trindade et al. 2003; Tohver et al. 2006). This event was followed Puga Fm. and the Jacadigo Group; these are in turn overlain by
by post-orogenic sub-alkaline granitic magmatism at c. 500 Ma mixed clastic and carbonate rocks of the Corumbá Group
(Almeida & Mantovani 1975). (Tables 45.2 & 45.3).
Deformation and metamorphism increase from imperceptible
effects on flat-lying rocks overlying the stable Amazonian craton
border to low-grade metamorphism within the Paraguay Belt, Puga Fm. and Cuiabá Group
determined by folding and crystallization of the illite parallel to
cleavage planes (Alvarenga et al. 1990; Alvarenga & Trompette The diamictite lying concordantly below the dolomite of the
1993). Despite low metamorphism, diamictite, quartzite, slate Bocaina Fm. originally described in the southern Paraguay Belt
and phyllite of the Cuiabá Group and Puga Fm., as well as the at Puga Hill (Maciel 1959) has been found in other areas of the
limestone, marl, dolostone and shale of the Araras and Corumbá Paraguay Belt, and is also named Puga Fm.
groups, have well-preserved sedimentary structures in most In the northern Paraguay Belt, the Puga Fm. consists of
outcrops. diamictite, sandstone and claystone units that are part of a

Table 45.3. Neoproterozoic lithostratigraphy nomenclature in the Corumbá and Puga hill region, southern Paraguay Belt

Period Group Formation Lithology U–Pb Ages

Guaicurus Shale

Corumbá Tamengo Limestone, mudstone and shale 543 ± 2 Ma


Ediacaran

Bocaina Dolostone

Puga* Diamictite

Cryogenian Santa Cruz BIF, Mn beds and arkose


Jacadigo
Urucum
Conglomerate and arkose

*Puga Fm. is thought to be coeval with the sedimentation of the Jacadigo Group, although isolated outcrops make it difficult to
confirm their stratigraphic relationship.
PUGA FORMATION, PARAGUAY BELT 491

marine succession (c. 100 m thick) close to the western border of


the Neoproterozoic basin. To the east, in the deeper parts of the
basin, layers of diamictite intercalated with conglomerate, sand-
stone and fine-grained rocks of the Cuiabá Group (c. 3000 m

Santa Cruz Formation


thick) were deposited as sediment gravity flows (Table 45.1;
Alvarenga & Trompette 1992).
In the Bauxi region (Fig. 45.1), the Puga Fm. contains diamictite
associated with conglomerate, sandstone and siltstone in its lower Mn
part followed by a massive diamictite. At this site, the massive dia-
mictite, interbedded with conglomerate and sandstone, passes to
the east into a thick fine-grained facies with a few isolated clasts
of the Cuiabá Group.

Jacadigo Group
The exposed section of the Puga Fm., at its type section on Puga Mn
Hill (Fig. 45.1), consists of a 95-m-thick massive diamictite with a
sandy matrix, which in the lower 50 m, contains a sandstone layer Mn
intercalated with a clay matrix massive diamictite (Maciel 1959).
A pink limestone of the Bocaina Fm. rests on the diamictite
(Table 45.3; Maciel 1959) and was interpreted as a cap carbonate
typical of other Neoproterozoic successions (Boggiani & Coimbra

Urucum Formation
1996; Boggiani et al. 2003).
Diamictite of the Puga Fm. at the Serra da Bodoquena (Fig. 45.1,
Table 45.2) are found along the axis of folded rocks, overlain by 200
carbonate rocks of the Corumbá Group (Almeida 1965). To the
east of the Serra da Bodoquena a massive foliated diamictite
with a ferruginous matrix is intercalated with centimetre-thick
hematite and magnetite layers, and sub-millimetre layers of 100
quartz and chert (Piacentini et al. 2007).

Jacadigo Group 0 (m)

Between Corumbá and Puga Hill (Fig. 45.1), the Neoproterozoic


exposures of the Jacadigo Group occur in isolated hills (up to Basement
900 m high), surrounded by lowlands of the wetland Pantanal (granite)
plain (Neogene Basin). The Jacadigo Group (Fig. 45.3) differs
from the Puga Fm. in the presence of significant ferruginous
and manganesiferous chemical deposits and the absence of
diamictite. The lower unit of the Jacadigo Group (Urucum
Fm.) comprises conglomerate and arkose sandstone, with
discontinuous conglomerate interbeds, c. 200 m thick. The Santa
Cruz Fm. is c. 400 m thick, and forms the upper part of the Jaca- 1 2 3 4 5
digo Group. The lower part of this formation starts with a continu-
ous layer of red-violet ferruginous-manganese arkose with a Fig. 45.3. Lithostratigraphy of the Jacadigo Group (modified from Almeida
jaspilite/hematite cement along Urucum Hill (Almeida 1945; 1945). 1, Arkose sandstone; 2, conglomerate and arkose sandstone; 3, siltstones;
Walde & Hagemann 2007); this unit was named the Banda 4, BIF with five Mn beds; 5, ferruginous arkose sandstone, sometimes with
Alta Fm. by Dorr II (1945). Towards the top of the Santa boulders and clasts.
Cruz Fm., there are bedded hematite-rich rocks and jasper
(BIF) with arkose and manganese ore intercalations bearing
large granite boulders. In Bolivia, the lower part of the Boqui
Group has been interpreted as correlative to the Jacadigo Group
(Litherland et al. 1986). formation reaches c. 250 m in thickness in the Nobres region. In
the Terconi and Tangará quarries (inner shelf carbonate basin),
this unit overlies the dolostone of the Mirassol d’Oeste Fm. and
Araras Group, northern Paraguay Belt presents at its base fan-like crystals interpreted as aragonite pseu-
domorphs (Alvarenga et al. 2008). Interbedded graded grainstone
The thickness of the Araras Group varies from 100 to 150 m in the and cross-bedded grainstone occur in the middle part of this
western area, overlying the cratonic region (inner shelf), to more formation. Locally, in the fore slope basin, an interbedded
than 1300 m eastwards and in the middle-outer shelf domain. 12-m-thick layer of dolostone (grainstone and breccia) was
Deeper-water rocks are chiefly marl limestone, rhythmite and silt- described at the Nossa Senhora da Guia Quarry (Alvarenga et al.
stone that mark the slope deposits of the basin (Alvarenga et al. 2004, 2008).
2008). This unit was subdivided into three formations The uppermost Nobres Fm. consists of a uniform, light grey
(Table 45.1). The lower Mirassol d’Oeste Fm. consists mostly of dolostone (c. 1100 m) that forms the main karstic relief
dolostone. The 18- to 32-m-thick type section at the Terconi of the region. This formation is exposed in the Provı́ncia
Quarry is characterized by laminated dolostone associated with Serrana but is absent in the eastern part of the Paraguay Belt
microbialites, giant wave ripples, tube-like structures and (Fig. 45.1). The base of this unit is characterized by 2– 4 m of
fan-like crystals (Nogueira et al. 2003; Nogueira & Riccomini dolomite brecciated layers intercalated with fine-grained and
2006; Alvarenga et al. 2008). laminated dolomite. Most of the sequence consists of a thick
The overlying Guia Fm. consists of dark grey laminated micrite succession of grainstone and packstone dolostone (Alvarenga
limestone with thin interbedded shale or marl laminae. This et al. 2000, 2004).
492 C. J. S. ALVARENGA ET AL.

The Corumbá Group, southern Paraguay belt The mudstone comprises an alternation of silt-shale, siltstone
and very fine-grained sandstone, with clasts cutting or disturbing
The Corumbá Group is exposed both in unfolded and folded the bedding (Alvarenga & Trompette 1992). Some thin beds and
domains. In the unfolded domains, this unit overlies the Rio Apa lenses, up to 0.3 m thick, made up of clast-rich diamictite, have
Craton, located on the western margin of the Paraguay erosive basal contacts. Lenses of fine-grained sandstone com-
River, whereas in the folded domains, it crops out in the eastern monly show load structures. Some massive or laminated sandstone
part of the Serra da Bodoquena (Fig. 45.1). The Corumbá intercalations, up to 15 m thick, include isolated clasts.
Group is the uppermost Neoproterozoic succession exposed in
the southern Paraguay Belt and was subdivided into five for-
mations: Cadieus, Cerradinho, Bocaina, Tamengo and Guaicurus Cuiabá Group
(Tables 45.2 & 45.3).
The Cadieus and Cerradinho formations occur only in the Serra The Cuiabá Group includes a thick sequence of fine-grained rocks,
da Bodoquena (Fig. 45.1) and comprise a mixed assemblage of sandstone, conglomerate and diamictite (Fig. 45.2; Alvarenga &
clastic and carbonate rocks that unconformably overlies the Rio Trompette 1992). Diamictite in this group is massive, with an
Apa Craton (Boggiani 1998). The Cadieus Fm. consists of con- abundant clay-silt matrix. Most clasts are millimetre to centimetre
glomerate and arkosic sandstone that was deposited in an alluvial scale, although a few reach 1 m in diameter. Some clasts of carbon-
fan atop the igneous-metamorphic basement. The proximal facies ate rocks are found in clast-poor diamictite.
of this fan grades upwards and laterally into the arkosic sandstone, Sandstone and coarse-grained rocks are widely distributed in
shale and grainstone that comprises the more distal alluvial facies this group. Conglomerate with beds 0.3–15 m thick is commonly
of the Cerradinho Fm. (Boggiani et al. 1993; Boggiani 1998). Cross- interbedded with sandstone and diamictite. The clast size ranges
bedding and hummocky cross-stratification found within grainstone from less than 1 cm to c. 10 cm. The matrix represents less than
facies in the upper portion of the Cerradinho Fm. suggest a rise in sea 10%, and consists of a clay-silt-sand mixture (Alvarenga &
level at the end of the rift phase (Boggiani 1998). Trompette 1992). Pebbly sandstone exhibits normal grading, start-
The Bocaina Fm. is present in the Serra da Bodoquena, Corumbá ing with a pebbly conglomerate and passing upward to a fine-
region and in Puga Hill, where it may reach thicknesses up to grained conglomerate followed by sandstone. The individual
300 m (Maciel 1959). This unit is placed directly above the lime- graded bed is normally 0.5–3.0 m thick, with a sharp erosive
stone of the Cerradinho Fm. at the Serra da Bodoquena and the basal contact. Clasts include quartz, feldspar, quartzite, mudstone,
diamictite of the Puga Fm. at Puga Hill. limestone and some granite. Flame, load casts and ball and pillow
The Tamengo Fm. is up to 100 m thick and consists of dark, structures are observed at the contact between two graded
organic-rich limestone and shale rhythmically interbedded with sequences, with an injection of sandy material into the overlying
uncommon limestone grainstone. In the Corumbá area, the pebbly sandstone layer. Fine-grained sandstone layers grade
organic-rich limestone and shale of this formation contain a rich upwards into mudstone. Occasionally, isolated clasts can be
Ediacaran microfossil assemblage that will be described below. found cutting the underlying layers of laminated claystone.
The dominance of shale in the Bodoquena region suggests In the Puga Fm., east of the Serra da Bodoquena, a BIF occurs
deeper waters when compared to the Corumbá region (Boggiani as a bed, c. 2 m thick, confined within a massive diamictite with
1998), thus explaining the absence of Cloudina lucianoi in the a ferruginous matrix. The BIF is formed by centimetre-thick
Serra da Bodoquena limestone-shale. layers of hematite and magnetite, alternating with millimetre-thick
The Guaicurus Fm. was first described by Almeida (1965) as a layers of quartz and chert (Piacentini et al. 2007).
thick shale succession on the top of the Corumbá Group that
extends over the eastern part of the Serra da Bodoquena and Jacadigo Group
along the Miranda River valley. It consists of fine-grained silici-
clastic rocks, mainly shale. Diamictite and pelite with clasts that A BIF at the top of the Jacadigo Group (Santa Cruz or Banda Alta
occur as lenses in the pelite of the Guaicurus Formation at formations) consists of a thick succession (up to 450 m thick) of
Laginha Mine, Corumbá, Mato Grosso do Sul, have been inter- interbedded (i) ferruginous arkose sandstone, (ii) hematite-rich
preted as sediment gravity-flow deposits in a slope setting (Bog- jaspilite (BIF), (iii) arkose and feldspathic sandstone, sometimes
giani et al. 2004). There are no equivalents of both the Tamengo containing isolated granite blocks and boulders, (iv) siltstone and
and Guaicurus formations in the northern part of the Paraguay ferruginous mudstone, (v) mudstone with banded jaspilite, and
belt (Fig. 45.1; Tables 45.1, 45.2 & 45.3). (vi) four manganese ore beds between 0.3 and 4 m thick
(Fig. 45.3). Beds vary from centimetres to a few metres in thick-
ness. Arkose and feldspathic sandstone ranges from massive to
Glaciogenic deposits and associated strata bedded, and is rarely graded. Some arkose beds exhibit isolated
granite-gneiss boulders, some of them bigger than 1.5 m in diam-
Puga Fm. eter. Deformation around the boulders is generally more pro-
nounced above them (Trompette et al. 1998).
In the northern Paraguay Belt, the Puga Fm. consists of a succes-
sion of diamictite units intercalated with sandstone, conglomerate
and mudstone. The matrix of the diamictite varies from sandy to Mirassol d’Oeste Fm.
clayey. Diamictite is massive or stratified and contains clasts
from a few centimetres up to 1 m in diameter. Clasts mainly of The Mirassol d’Oeste Fm. (20 –32 m thick) is the basal pink
basement rocks (granite, gneiss, quartzite, quartz, schists, etc.), dolostone of the Araras Group, resting on the diamictite of the
some of them deflecting the underlying laminae, faceted and Puga Fm. (Nogueira et al. 2003). The first 8 m of the formation
striated, and abundant detrital mica, are found in outcrop, close consists of laminated pinkish dolostone that grades upward
to the Palaeoproterozoic basement rocks. Massive diamictite through a diffuse and transitional contact to a grey laminated
shows great variations in the relative proportions of matrix dolostone. The basal dolostone has stratiform and wave microbia-
(muddy sand) and clasts. Occasionally, a crude stratification can lite with dispersed tube-like structures (Nogueira et al. 2003; Font
be observed. Some stratified diamictite consists of few- et al. 2006; Elie et al. 2007). The upper part of the dolostone has
centimetres-thick, massive diamictite beds associated with mud- an enigmatic wave bed form that was interpreted as a giant wave
stone with graded bedding and outsized clasts clearly deflecting ripple by Allen & Hoffman (2005) but described as a tepee-like
the underlying laminae. structure by Nogueira et al. (2003). Fan-like crystals interpreted
PUGA FORMATION, PARAGUAY BELT 493

as aragonite pseudomorphs are commonly found in the upper part -5 0 +5 +10


of the dolostone and in the overlying limestone of the Guia For-
mation (Nogueira et al. 2003; Alvarenga et al. 2008). Carbonate
has also been described from a drill core in the Bauxi area 1500
(Fig. 45.1) in the middle-outer shelf domain where the Araras
Group is c. 1300 m thick. In this area, 18-m-thick white laminated
dolostone overlies the diamictite of the Puga Formation in sharp
contact (Alvarenga et al. 2008). Other common sedimentary struc-
tures found in this dolostone include stromatolite, and breccia.

Basal member of the Bocaina Fm.

Maciel (1959) was the first to describe the carbonate of the Bocaina
Fm. that overlies the diamictite of the Puga Fm. at Puga Hill. The A
succession starts with a succession of 4-m-thick massive pink 1000
limestone overlaid by c. 10 m of a purple limestone/dolostone, R
mudstone and sandstone that grade upwards to grey dolostone A Nobres
and light grey dolostone. This upper dolostone is the common R FM.
rock of the Bocaina Fm., which also presents grainstone, pack- A
stone, breccia and stromatolites with poorly developed stratifica- S
tion (Almeida 1945; Boggiani et al. 1993).

Boundary relations with overlying and underlying G


non-glacial units R
The basal diamictite of the Puga Fm. lies unconformably above the 500 O
Palaeoproterozoic metamorphic rocks, in the western part of the U
basin, around Mirassol d’Oeste (Fig. 45.1). This contact is P 87
Sr/ 86
Sr
usually poorly exposed, but a road cut section exhibits 12 m of
0.7077
massive diamictite above the unconformity, followed upwards
by mudstone with isolated cobbles and pebbles, including some
0.7078
intercalations of massive diamictite, up to 0.80 m thick (Alvarenga Mirassol 0.7076
1990; Alvarenga & Trompette 1992). Guia d' Oeste Fm. 0.7075
Good exposures of the upper contact of the Puga Fm. with the Fm
Araras Group can be found at the Terconi and Guia quarries and 0.7075
0.7078
in a drill core in the Bauxi region (Alvarenga & Trompette 0.7075
1992; Nogueira et al. 2003; Alvarenga et al. 2004, 2008). The 0
best exposure of this contact (Terconi Quarry) exhibits dolostone (m) Puga Fm.
of the Mirassol d’Oeste Fm. overlying diamictite along a conform-
able sharp contact. This contact exhibits soft-sediment defor- -10 -5 0 +5 +10
mation (Nogueira et al. 2003). In the southern Paraguay Belt, the
13
pink limestone from the Bocaina Fm. overlies the diamictite at d C (PDB)
Puga Hill (Maciel 1959; Boggiani et al. 2003).
Fig. 45.4. Stratigraphic section and variations of d13Cpdb and 87Sr/86Sr for the
Araras Group (data from Nogueira et al. 2003, 2007; Alvarenga et al. 2004,
Chemostratigraphy 2008; Font et al. 2006; Figueiredo 2006).

The chemostratigraphic data discussed here take into account all


available C, O and Sr isotope data from the Araras and Corumbá range from –8.2 to –1.3‰, whereas 87Sr/86Sr ratios are high
carbonate successions (Gaucher et al. 2003; Boggiani et al. and variable with low Sr content (44 – 96 ppm) and high Mn –Sr
2003; Nogueira et al. 2003, 2007; Pinho et al. 2003; Alvarenga ratios (.18).
et al. 2004, 2008; Figueiredo 2006; Font et al. 2006; Misi et al. Dark grey laminated micritic limestone interbedded with the
2007; Riccomini et al. 2007). The discussion will focus on C iso- thin shale or marl layers of the Guia Fm. lie above the Mirassol
topes, because they are more resistant to post-depositional altera- d’Oeste Fm. This grey limestone presents increasing d13C values
tion. The Sr-isotope ratios were also considered for those across the first 20 m from the base, with a narrow range of C-
samples in which the Sr concentration was higher than 400 ppm. isotopic values (between –1.6 and þ0.1‰; Fig. 45.4). The
87
Sr/86Sr ratios for limestone samples with high Sr content
(750 –4351 ppm) and low Mn –Sr (,0.2) range between 0.70763
Northern Paraguay Belt and 0.70780 (Fig. 45.4).
The dolostone of the overlying Nobres Fm. presents a wide
The d13C profiles of the Mirassol d’Oeste Fm. were constructed range of C-isotopic values (Alvarenga et al. 2004; Nogueira
from three sections in the northern Paraguay Belt: Terconi and et al. 2007). In the Bauxi-Nobres area, for instance (Fig. 45.1),
Tangará quarries, both located on the western cratonic domain, the dolostone at the base of the sequence (900 m thick) is charac-
and the João Santos borehole, which is located in the fold belt terized by homogeneous and positive d13C values (þ1.9 to
near Bauxi (Figs 45.1 & 45.4). While the d13C values of the first þ2.7‰), while the upper dolostone, which consists of thin
20– 30 m in the cratonic region range between – 10.5‰ and layers of sandy dolostone, exhibit increasing d13C values up to
–3.0‰, the d13C values along the profile located in the folded þ9.6‰ (Fig. 45.4; Alvarenga et al. 2004; Figueiredo 2006). On
domain range from –4.8‰ to –1.7‰. Corresponding d18O ratios the other hand, in the Cáceres area, the lowermost 300 m of the
494 C. J. S. ALVARENGA ET AL.

dolostone placed on top of the Guia Fm. have d13C values ranging intercalations (Walde et al. 1981; Hoppe et al. 1987; Trompette
from – 2.2 to þ0.3‰ (Fig. 45.4; Nogueira et al. 2007). et al. 1998; Walde & Hagemann 2007). The manganese deposits
consist of four individual layers ranging from 0.3 to 4.0 m in thick-
ness. Beds of Mn ore, up to 5 m thick, are mined in the under-
Southern Paraguay Belt
ground Urucum Mine. Isolated granite clasts up to 1.5 m in
diameter have been identified in the arkose intercalations
Boggiani et al. (2003) presented d13C values close to –5‰ for the
(Trompette et al. 1998; Walde & Hagemann 2007).
first 12 m of laminated carbonate rocks overlying the Puga Fm. in
The BIFs in the eastern part of the Serra da Bodoquena are thin
the Puga Hill (Fig. 45.5). In the Jacadigo Group, thin limestone
beds (2 –3 m thick) in a massive diamictite with a ferruginous
intercalations in the BIF show d13C ratios between –5.2 and
(magnetite) matrix, suggesting the relationship of these sediments
–7.0‰ (Klein & Ladeira 2004).
with Fe precipitation (Piacentini et al. 2007). The magnetite matrix
The Tamengo Fm. occurs in the Corumbá area and along the
has an Fe content ranging from 15 to 72%, with an average of 27%
eastern part of the Serra da Bodoquena. The isotopic data of lime-
(Piacentini 2008).
stone from this unit were obtained at the Laginha, Saladeiro and
Corcal quarries, located in the Corumbá area (Figs 45.1 & 45.5;
Boggiani 1998; Boggiani et al. 2003; Misi et al. 2007). Limestone Palaeolatitude and palaeogeography
of the lower part of the Tamengo Formation records a negative
d13C excursion (–3.3 to –2.5‰), which is followed by limestone Palaeomagnetic data are only available for the Neoproterozoic
with positive d13C values (up to þ5.8‰). In the Serra da Bodo- Mirassol d’Oeste Fm. and are restricted to 19 samples from the
quena, located about 200 km SE of Corumbá, limestone correlated Terconi Quarry, in the northern part of the Paraguay belt (Trindade
to the Tamengo Fm. shows homogeneous d13C values around et al. 2003). These palaeomagnetic data indicated five polarity
þ3‰ (Boggiani 1998). The 87Sr/86Sr ratios of the Tamengo car- reversals within the first 20 m of the dolomite, suggesting a
bonate rocks from both areas are clustered between 0.7084 and primary magnetization and a low palaeolatitudes (22 þ 6/– 58)
0.7085 (Boggiani 1998; Babinski et al. 2008). for these rocks.

Other characteristics Geochronological constraints

Mn and Fe ore bodies with 36 billion metric tonnes of hematite, A U– Pb zircon age of 543 + 2 Ma (Ediacaran) has been obtained
and estimated Mn ore reserves of 608 million metric tonnes from an ash bed intercalated with the Cloudina bearing limestone
(Urban et al. 1992) are known around south Corumbá. of the Tamengo Fm. near Corumbá city (Babinski et al. 2008), thus
This mineralization occurs throughout the Santa Cruz providing a precise Upper Ediacaran age for this unit. In the north-
Fm., as banded, hematite-rich, jaspilite with arkose sandstone ern Paraguay belt (Terconi Quarry), a Pb/Pb isochron age of
633 + 25 Ma was obtained for the dolostone of the Mirassol
d’Oeste Fm. and the limestone of the Guia Fm. (Babinski et al.
-5 0 +5 +10 unpublished). This age is interpreted as the time of deposition
for these carbonate rocks. There are otherwise no radiometric
data available to directly constrain the age of the Puga Fm.
87Sr/ 86Sr
60 Other radiometric ages include the Nd isotope signature
obtained from clasts in diamictite, quartzite and phyllite in the
0.7086 Cuiabá Group and Puga Fm. by Dantas et al. (2009). The Tdm
40 model ages display an irregular distribution. Diamictite present
Tamengo Fm

variable Tdm values (1.4 –2.15 Ga), but homogeneous 1nd (t)
around – 8. The sandstone show Tdm model ages (c. 2.0 –2.1 Ga)
older than the ones determined for fine-grained rocks (Tdm
30 values of 1.7 and 1.8 Ga). This difference of Nd isotopes may be
0.7085 related to variations in the source region or may be the result of
mixing of material from different sources. Nevertheless, all of
20
these data indicate sources within the Amazonian Craton, in agree-
ment with the palaeogeographic interpretation, which points to the
northwestern basement as the main source of detritus (Alvarenga
0 & Trompette 1992; Dantas et al. 2009).
The limestone and shale of the Tamengo and Guaicurus for-
20 mations near the city of Corumbá contain a record of Ediacaran
fossils (Hahn et al. 1982; Walde et al. 1982; Zaine & Fairchild
1985, 1987; Gaucher et al. 2003). The microfossil assemblage is
dominated by scyphozoan Corumbella werneri (Hahn et al.
Bocaina Fm

0 1982; Walde et al. 1982), Babvinella faveolata, Vandalosphaeri-


dium sp., Cloudina lucianoi (Zaine & Fairchild 1985, 1987), Sol-
dadophycus bossii, Titanotheca and the vendotaenid Eoholynia
10 corumbensis sp. (Gaucher et al. 2003). The presence of these
fossils suggests a deposition during the 570–545 Ma period, con-
0 sistent with the U –Pb age obtained in the Tamengo Fm.
Puga Fm. (m)
-10 -5 0 +5 +10
Discussion
13
d C (PDB)
Northern Paraguay Belt depositional settings
Fig. 45.5. Composite of d13Cpdb and 87Sr/86Sr record for the Corumbá Group
from Corumbá area, Puga Hill and Serra da Bodoquena (data from Boggiani The glacially influenced succession of the Puga Fm. and Cuiabá
1998; Boggiani et al. 2003; Misi et al. 2007 and unpublished data). Group in the northern Paraguay Belt displays three main
PUGA FORMATION, PARAGUAY BELT 495

depositional settings: platform, slope and outer slope (Fig. 45.2). formations record proximal and distal alluvial deposition. The pink
The platformal deposits are reworked by sedimentary gravity limestone of the Bocaina Fm. has been described as a ‘cap’ carbon-
flows. The deposits on the inner shelf show dominant massive dia- ate (Boggiani et al. 2003; see discussion below). The Tamengo
mictite, alternating with sandstone and fine-grained sedimentary and Guaicurus formations comprise a trangressive– highstand
rocks that contain striated and faceted clasts (Alvarenga & Tromp- sequence, which drowned the carbonate shelf of the Tamengo
ette 1992). On the outer shelf, there is an association of massive Fm. (Boggiani & Coimbra 1996). Diamictite and pelite with clasts
diamictite, stratified diamictite and fine-grained sedimentary that occur as lenses in the pelites of the Guaicurus Fm. at Laginha
rocks with some clasts disrupting the underlying beds. The outer Mine, Corumbá, Mato Grosso do Sul, have been interpreted as sedi-
shelf succession has been interpreted as a succession of resedimen- ment gravity flow deposits in a slope setting (Boggiani et al. 2004).
ted glacial deposits with alternations of debris-flows and shallow
water turbidite events (Alvarenga & Trompette 1992).
Glaciomarine deposits reworked by sediment gravity flows and Correlations and geotectonic evolution of the Paraguay Belt
related to submarine fans characterize the slope depositional
system. Progressive sorting in the deeper portions of the fans are Despite the widespread areal distribution of diamictite underlying
demonstrated by the transition from diamictite to massive con- both carbonate successions (Araras and Corumbá groups), there is
glomerate and sandstone with local inverse and/or normal no consensus about the lithostratigraphic correlation between the
grading and normally graded, fine-grained turbidites. Sandstone northern and southern parts of the Paraguay Belt. In addition, the
and siltstone intercalations are consistent with inter-channel evolution of these two segments could have been diachronic.
deposits. Deposition on the outer slope system was dominated by Some differences and similarities observed between the rocks
fine-grained deposits related to low-density turbidity currents, in from the north and south will be mentioned here in order to
which a direct glacial influence is only indicated by isolated provide a picture of the current knowledge and controversies
clasts (dropstones). surrounding the geotectonic evolution of the Paraguay Belt.
This type of basin-filling suggests that the sedimentary source Glacial episodes have been proposed for both sectors of the
was located on the Amazonian Craton (Alvarenga & Trompette Paraguay Belt, but some differences are observed. In the northern
1992) and this was confirmed by Nd isotopic analysis (Dantas Paraguay Belt, two glacial events can be found: an older one rep-
et al. 2009). Shelf reworking during the glacial event helped resented by the Puga Fm. and the Cuiabá Group rocks (Alvarenga
develop submarine channels and turbidite deposits on the slope & Trompette 1992), and a younger one represented by the Serra
and outer slope (Alvarenga & Trompette 1992). Diamictite at the Azul Fm. and thought to be related to the Ediacaran glaciation
top of this succession is interpreted as the final stage of glaciation, (Alvarenga et al. 2007, 2009; Figueiredo et al. 2008, 2011). In con-
during which all deposits were transported into the basin by sedi- trast, the glacial origin of the Puga Fm. in the southern Paraguay
ment gravity flows and iceberg melting. Belt is debatable and no younger diamictite-bearing unit has
The Araras Group records carbonate deposition and is divided been found there.
into three lithostratigraphic units. The basal Mirassol d’Oeste Whereas carbonate lithofacies and sedimentary structures
Fm. was deposited directly on top of glacial diamictites of the typical of other Neoproterozoic cap-carbonate rocks have been
Puga Fm., and is interpreted as a transgression over previously gla- described in the North (Nogueira et al. 2003), none have been
ciated landscapes. The laminated limestone and shale of the middle found in the southern belt.
Guia Fm. are related to a deep-water system. The upper carbonate Although there is some evidence of diagenetic alteration of Sr
unit, Nobres Fm., consists of more than 1100 m of shallow-shelf values in the lower Mirassol d’Oeste Fm., the carbonate units
dolostone (breccia, grainstone and packstone) indicating high- found overlying diamictite of the Puga Fm. in both regions
energy environments (Alvarenga et al. 2004). exhibit d13C and 87Sr/86Sr values interpreted as representing
original seawater geochemistry; these data are consistent with a
north – south correlation for the Paraguay Belt and are typical of
Southern Paraguay Belt depositional settings other Neoproterozoic cap-carbonate rocks.
In contrast, the upper part of the Corumbá and Araras Groups
In the southern Paraguay Belt (Fig. 45.1), the rocks interpreted as differ in their lithostratigraphy. The limestone and shale from the
deposited during the late Cryogenian glacial event are the Jacadigo Tamengo and Guaicurus formations, both with microfossils
Group and the Puga Fm. The Jacadigo Group around Corumbá (Gaucher et al. 2003), represent a transgressive phase at the top
is interpreted as deposited in a Neoproterozoic tectonic graben of Corumbá Group. This is contrasted with the 1300 m of
system, coeval with the sedimentation of the Puga Formation carbonate-dominated deposition and no microfossils in the top of
(Trompette et al. 1998). The Santa Cruz Formation shows alterna- the Araras Group, in the northern part of the belt. A depositional
tions of autochthonous chemical sediments and siliciclastic sedi- age of 543 + 2 Ma (zircon U –Pb SHRIMP dating) was obtained
ments (fine-grained and arkose sandstone), which have been for the Tamengo Fm. (Babinski et al. 2008). If we consider that
interpreted as glacially influenced based on the presence of large the Bocaina Fm. was deposited after the late Cryogenian glacia-
granite boulders thought to be dropstones (Barbosa 1949; Walde tion, which is assumed to be 630 Ma, an unconformity would
et al. 1981; Urban et al. 1992; Klein & Ladeira 2004). Trompette have to exist between the Bocaina and Tamengo formations,
et al. (1998), however, remarked that these granite pebbles and because the deposition of the Bocaina Fm. (200 m thick, at
boulders generally do not cross-cut the underlying bedding, and most) could not have taken such a long time (c. 100 Ma). In con-
the deformation around the boulders is associated with compaction trast, the 1300-m-thick carbonate succession between the two
from overlying strata, being more pronounced above the boulders glacial units in the northern Paraguay Belt is thought to have
rather than beneath them. These data were used to deny a glacial lasted 50 Ma, assuming the two glacial units represent a late Cryo-
origin of these deposits, suggesting reworked gravitational fluxes genian (c. 630 Ma) and an Ediacaran age glaciation (580 Ma),
(Trompette et al. 1998). The diamictite of the Puga Fm. in the respectively (Alvarenga et al. 2007; Figueiredo et al. 2008).
southern Paraguay Belt does not show evidence of striated clasts Further evidence that undermines a north –south correlation for
or dropstones, but their widespread areal distribution underlying at least the upper part of the succession comes from the age of the
a carbonate succession (Araras and Corumbá groups) is strong evi- Alto Paraguay Group, which overlies the Araras Group. Although
dence for their stratigraphic correlations to other glacial deposits it is poorly constrained, a Rb –Sr whole-rock isochron age of c.
of the Puga Fm. in the Paraguay Belt. 569 Ma, was determined from the shale of the Sepotuba Formation
The Corumbá Group records mixed clastic and carbonate sedi- (Table 45.1; Cordani et al. 1985), suggesting that the units of the
mentation in a passive margin setting. The Cadieus and Cerradinho uppermost formations of the northern part of the belt are older
496 C. J. S. ALVARENGA ET AL.

than the ones from the southern part (e.g. the Tamengo Fm., Alvarenga, C. J. S. & Trompette, R. 1992. Glacially influenced
Table 45.2). sedimentation in the Later Proterozoic of Paraguay belt (Mato Grosso,
In addition, the geotectonic evolution of the two regions appears Brazil). Palaeogeography, Palaeoclimatology, Palaeoecology, 92,
to differ during deposition of the upper part of the succession. 85–105.
40
Ar– 39Ar ages ranging from 541 and 531 Ma were determined Alvarenga, C. J. S. & Trompette, R. 1993. Brasiliano tectonic of the
from the biotite of metavolcanics (Araés gold deposit in the Paraguay Belt: the structural development of the Cuiabá Region.
eastern end of the belt) and interpreted as the cooling ages follow- Revista Brasilieira de Geociências, 23, 18 –30.
ing regional metamorphism in the northern Paraguay Belt Alvarenga, C. J. S. de, Cathelineau, M. & Dubessy, J. 1990. Chron-
ology and reorientation of N2 – CH4, CO2 – H2O, and H2O-rich
(Geraldes et al. 2008). In contrast, in the southern part of the
fluid-inclusions trails in intra-metamorphic quartz veins from the
belt, the deposition of carbonate rocks and shale from the Cuiabá gold district, Brazil. Mineralogical Magazine, 54, 245–255.
Tamengo (c. 543 Ma) and Guaicurus formations was taking Alvarenga, C. J. S. de, Moura, C. A. V., Gorayeb, P. S. S. & Abreu,
place in a passive margin environment. These drastic differences F. A. M. 2000. Paraguay and Araguaia Belts. In: Cordani, U. G.,
in the two segments of the belt are still not understood, and a Milani, E. J., Thomaz Filho, A. & Campos, D. A. (eds) Tectonic
more detailed and profound study is needed in order to better Evolution of South America, Rio de Janeiro, 31st International
constrain the evolution of the Paraguay belt. Geological Congress, 183– 193.
Radiometric constraints on the Puga Fm. are limited to Edia- Alvarenga, C. J. S., Santos, R. V. & Dantas, E. L. 2004. C-O-Sr iso-
caran ages (543 Ma) obtained from the Tamengo Fm. in the topic stratigraphy of cap carbonates overlying Marinoan-age glacial
southern Paraguay Belt, although the significance of this data diamictites in the Paraguay Belt, Brazil. Precambrian Research,
depends on stronger evidence for north –south stratigraphic corre- 131, 1– 21.
lations. In the end, the Puga Fm. and the correlated Cuiabá Group Alvarenga, C. J. S. de, Figueiredo, M. F., Babinski, M. & Pinho,
in the northern Paraguay Belt are interpreted as related to the late F. E. C. 2007. Glacial diamictites of Serra Azul Formation (Edia-
Cryogenian glaciation because of the overlying dolomite lithofa- caran, Paraguay Belt): evidence of the Gaskiers glacial event in
cies, palaeomagnetic data and negative d13C values of the Mirassol Brazil. Journal of South American Earth Science, 23, 236–241.
d’Oeste Fm. (Nogueira et al. 2003; Alvarenga et al. 2004, 2008; Alvarenga, C. J. S. de, Dardenne, M. A. et al. 2008. Isotope stratigra-
Allen & Hoffman 2005). Limestone and mud-limestone deposited phy of Neoproterozoic cap carbonate in the Araras Group, Brazil.
above the Mirassol d’Oeste Fm. have 87Sr/86Sr ratios between Gondwana Research, 13, 469–479.
0.70763 and 0.70780, consistent with a seawater composition Alvarenga, C. J. S. de, Boggiani, P. C., Babinski, M., Dardenne,
from c. 630 Ma (Alvarenga et al. 2004, 2008). The age and M. A., Figueiredo, M. F., Santos, R. V. & Dantas, E. L. 2009.
The Amazonian Paleocontinent. In: Gaucher, C., Sial, A. N., Hal-
glacial origin of the Puga Fm. in the southern Paraguay Belt
verson, G. P. & Frimmel, H. E. (eds) Neoproterozoic–Cambrian
awaits further study to resolve existing controversies. Tectonics, Global Change and Evolution: A Focus on Southwestern
Gondwana. Developments in Precambrian Geology, 16, Elsevier,
This chapter is a contribution to the International Geological Programme (IGCP)
15 – 28.
Project 512 ‘Neoproterozoic Ice Ages’ and Project 478 ‘Neoproterozoic– Early
Babinski, M., Boggiani, P. C., Fanning, M., Simon, C. M. & Sial, A. N.
Palaeozoic Events in SW-Gondwana’. This research was supported by the Con-
2008. U– Pb SHRIMP geochronology and isotope chemostratigraphy
selho Nacional de Desenvolvimento Cientı́fico e Tecnológico (CNPq) and Funda-
(C, O, Sr) of the Tamengo Formation, southern Paraguay belt, Brazil.
ção de Amparo à Pesquisa do Estado de São Paulo (FAPESP). We thank
In: South American Symposium on Isotope Geology, 6, San Carlos de
Mineração Corumbaense Reunidas S.A. and Urucum Mineração S.A. for their
Bariloche, Argentina. Proceedings, CD-ROM.
support during fieldwork. A. C. Rocha-Campos, N. M. Chumakov, E. Tohver,
Barbosa, O. 1949. Contribuição à geologia da região Brasil-Bolı́via.
C. Riccomini, G. Shields and E. Arnaud are thanked for their careful and
Mineração e Metalurgia, 13, 271– 278.
helpful reviews. This represents a contribution of the IUGS- and UNESCO-
Boggiani, P. C. 1998. Análise estratigráfica da bacia Corumbá
funded IGCP (International Geoscience Programme) project #512.
(Neoproterozóico)-Mato Grosso do Sul. PhD thesis. Universidade
de São Paulo, Instituto de Geociências, Brazil.
Boggiani, P. C. & Coimbra, A. M. 1996. The Corumbá Group
References (Central South America) in the context of Late Neoproterozoic
global changes. Anais da Academia Brasileira de Ciências, 68,
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves in the 595– 596.
aftermath of a Neoproterozoic glaciation. Nature, 433, 123– 127. Boggiani, P. C., Fairchild, T. R. & Coimbra, A. M. 1993. O Grupo
Almeida, F. F. M. de 1945. Geologia do sudoeste matogrossense. Boletim Corumbá (Neoproterozóico–Cambriano) na região central da Serra
da Divisão de Geologia e Mineralogia, DNPM, 116, 1 –118. da Bodoquena (Faixa Paraguai) Mato Grosso do Sul. Revista Brasi-
Almeida, F. F. M. de 1946. Origem dos minérios de ferro e manganês de lieira de Geociências, 23, 301– 305.
Urucum (Corumbá Estado de Mato Grosso). Boletim Divisão de Boggiani, P. C., Ferreira, V. P. et al. 2003. The cap carbonate of the
Geologia e Mineralogia, DNPM, 119, 1– 57. Puga hill (central South America) in the context of the Post-Varanger
Almeida, F. F. M. de 1964a. Geologia do centro-oeste matogrossense. glaciation. In: South American Symposium on Isotope Geology, 4,
Boletim da Divisão de Geologia e Mineralogia, DNPM, 215, 1 – 137. Salvador, Brazil, Short Papers, 324–327.
Almeida, F. F. M. de 1964b. Glaciação Eocambriana em Mato Mato Boggiani, P. C., Fairchild, T. R. & Riccomini, C. 2004. New level of
Grosso. Notas Preliminares e Estudos, DNPM, 117, 10. diamictites in the Corumbá Group (Ediacaran), Paraguay Belt,
Almeida, F. F. M. de 1965. Geologia da Serra da Bodoquena (Mato South America. In: 1st Symposium on Neoproterozoic-Early Paleo-
Grosso), Brasil. Boletim da Divisão de Geologia e Mineralogia, zoic Events in SW-Gondwana. Extended Abstract, IGCP, Project
DNPM, 219, 1 –96. 478, Second Meeting, Brazil, 10 –13.
Almeida, F. F. M. de & Mantovani, M. S. M. 1975. Geologia e geocro- Cordani, U. G., Filho, T. A., Brito Neves, B. B. & Kawashita, K.
nologia do granito São Vicente, Mato Grosso. Anais da Academia 1985. On the applicability of the Rb– Sr method to argillaceous
Brasileira de Ciências, 47, 451–458. sedimentary rocks: some examples from Precambrian sequences of
Alvarenga, C. J. S. de 1990. Phénomènes sédimentaires, structuraux et Brazil. Giornale di Geologia, 47, 253–280.
circulation de fluides développés à la transiction chaı̂ne-craton. Dantas, E. L., Alvarenga, C. J. S., Santos, R. V. & Pimentel, M. M. 2009.
Exemple de la chaı̂ne Paraguai d 0 âge proterozoı̈que supérieur, Using Nd isotopes to understand the provenance of sedimentary rocks
Mato Grosso, Brésil. PhD thesis, Université d0 Aix-Marseille III, from a continental margin to a foreland basin in the Neoproterozoic
France. Paraguay Belt, Central Brazil. Precambrian Research, 170, 1–12.
Alvarenga, C. J. S. de & Saes, G. S. 1992. Stratigraphy and sedimentol- Dorr II, J. V. N. 1945. Manganese and iron deposits of Morro do Urucum,
ogy of the middle and Late Proterozoic in the southeast of the Mato Grosso, Brazil. United States Geological Survey Bulletin,
Amazonian Craton. Revista Brasileira de Geociências, 22, 493– 499. 946A, 1 – 47.
PUGA FORMATION, PARAGUAY BELT 497

Elie, M., Nogueira, A. C. R., Nédélec, A., Trindade, R. I. F. & Kenig, isotope fluctuations and paleoceanographic changes in the late Neo-
F. 2007. A red algal bloom in the aftermath of the Marinoan Snowball proterozoic Araras carbonate platform, southern Amazon craton,
Earth. Terra Nova, 19, 303– 308. Brazil. Chemical Geology, 80, 168– 190.
Evans, J. W. 1894. The geology of Matto Grosso (particulary the region Piacentini, T. 2008. A Formação Ferrı́fera da Formação
drained by Upper Paraguay). Quaternary Journal of the Geological Puga: Avaliação Regional dos Recursos da Serra da Bodoquena,
Society, London, 50, 85 –104. MS. MS thesis, Universidade de São Paulo, Instituto de Geociências,
Figueiredo, M. F. 2006. Quimioestratigrafia das rochas ediacaranas no Brasil.
extremo norte da Faixa Paraguai, Mato Grosso. MSc thesis, Instituto Piacentini, T., Boggiani, P. C., Yamamoto, J. K., Freitas, B. T. &
de Geociências, Universidade de São Paulo, Brazil. Campanha, G. A. C. 2007. Formação Ferrı́fera associada à sedi-
Figueiredo, M. F., Babinski, M., Alvarenga, C. J. S. & Pinho, F. E. C. mentação glaciogênica da Formação Puga (Marinoano) na Serra da
2008. Nova unidade litoestratigráfica registra glaciação ediacarana Bodoquena, MS. Revista Brasileira de Geociências, 37, 530–541.
em Mato Grosso: Formação Serra Azul. Geologia USP: Série Cientı́- Pimentel, M. M., Fuck, R. A. & Alvarenga, C. J. S. 1996. Post-
fica, 8, 65 – 75. Brasiliano (Pan-African) high-K granitic magmatism in central
Figueiredo, M., Babinski, M. & Alvarenga, C. J. S. 2011. The Serra Brazil: the role of late Precambrian –early Paleozoic extension.
Azul Formation, Paraguay Belt, Brazil. In: Arnaud, E., Halverson, Precambrian Research, 80, 217– 238.
G. P. & Shields-Zhou, G. (eds) The Geological Record of Neo- Pinho, F. E. C., Sial, A. N. & Figueiredo, M. F. 2003. Contribution
proterozoic Glaciations. Geological Society, London, Memoirs, 36, to the Neoproterozoic C and O isotopic record: carbonate rocks
499– 502. from the Paraguay Belt, Mato Grosso, Brazil. In: IV South American
Font, E., Nédélec, A., Trindade, R. I. F., Macouin, M. & Charrière, Symposium on Isotope Geology, Salvador, Brasil, Short Papers, 1,
A. 2006. Chemostratigraphy of the Neoproterozoic Mirassol d’Oeste 386– 389.
cap dolostone (Mato Grosso, Brazil): an alternative model for Riccomini, C., Nogueira, A. C. R. & Sial, A. N. 2007. Carbon and
Marinoan cap dolostone formation. Earth and Planetary Science oxygen isotope geochemistry of Ediacaran outer platform carbonates,
Letters, 250, 89 – 103. Paraguay Belt, central Brazil. Anais da Academia Brasileira de
Gaucher, C., Boggiani, P. C., Sprechmann, P., Sial, A. N. & Fair- Ciências, 79, 519–527.
child, T. 2003. Integrated correlation of the Vendian to Cambrian Rocha-Campos, A. C. & Hasuı́, Y. 1981. Late Precambrian Jangada
Arroyo del Soldado and Corumbá Groups (Uruguay and Brazil): Group and Puga Formation of central western Brazil. In: Hambrey,
palaeogeography, palaeoclimatic and palaeobiologic implications. M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial
Precambrian Research, 120, 241– 278. Record. Cambridge University Press, Cambridge, 916– 919.
Geraldes, M. C., Tassinari, C. C. G. et al. 2008. Isotopic evidence for Tohver, E., D’Agrella-Filho, M. S. & Trindade, R. I. F. 2006. Paleo-
the Late Brasiliano (500– 550 Ma) ore-forming mineralization of magnetic record of Africa and South America for the 1200–500 Ma
the Araés Gold Deposit, Brazil. International Geology Review, 50, interval, and evaluation of Rodina and Gondwana assemblies.
177– 190. Precambrian Research, 147, 193– 222.
Hahn, G., Hahn, R., Leonardos, O. H., Pflug, H. D. & Walde, D. H. G. Trindade, R. I. F., Font, E., D’Agrella-Filho, M. S., Nogueira,
1982. Körperlich erhaltene Scyphozoen – Reste aus dem A. C. R. & Riccomini, C. 2003. Low-latitude and multiple geomag-
Jungprakambrium Brasiliens. Geologica et Palaeontologica, 16, netic reversals in the Neoproterozoic Puga cap carbonate of Amazo-
1 – 18. nia. Terra Nova, 15, 441– 446.
Hennies, W. T. 1966. Geologia do centro-oeste matogrossense. PhD Trompette, R. 1994. Geology of Western Gondwana (2000 –500 Ma).
thesis, Escola Politécnica, Universidade de São Paulo, Brazil. Pan-African Brasiliano Aggregation of South America and Africa.
Hoppe, A., Schobbenhaus, C. & Walde, D. H. G. 1987. Precambrian Balkema, Rotterdam.
iron formation in Brazil. In: Appel, P. W. U. & LaBerge, G. L. Tompette, R., Alvarenga, C. J. S. de & Walde, D. 1998. Geological
(eds) Precambrian Iron Formation. Theophrastus Publications, evolution of the Neoproterozoic Corumbá graben system
Athenas, 347– 390. (Brazil). Depositional context of the stratified Fe and Mn ores of
Klein, C. & Ladeira, E. A. 2004. Geochemistry and mineralogy of the Jacadigo Group. Journal of South American Earth Sciences, 11,
Neoproterozoic banded iron-formations and some selected, siliceous 587– 597.
manganese formations from the Urucum District, Mato Grosso do Urban, H., Stribrny, B. & Lippolt, H. J. 1992. Iron and manganese
Sul, Brazil. Economic Geology, 99, 1233– 1244. deposits of the Urucum District, Mato Grosso do Sul, Brazil.
Litherland, M., Annels, R. N. et al. 1986. The Geology and Mineral Economic Geology, 87, 1375–1392.
Resources of the Bolivian Precambrian Shield. British Geological Vieira, A. J. 1965. Geologia do centro-oeste de Mato Grosso. Petrobrás-
Survey, Overseas Memoir, 9. DEBSP, Relatório Técnico Interno no. 303, 58.
Luz, J. S. & Abreu Filho, W. 1978. Aspectos geológico-econômicos da Walde, D. H. G. & Hagemann, S. G. 2007. The Neoproterozoic
Formação Araras e do Grupo Alto Paraguai – MT. In: An. 308 Con- Urucum/Mutúm Fe and Mn deposits in W-Brazil/SE-Bolivia:
gresso Brasileiro de Geologia, sociedade Brasileira de Geologia, assessment of ore deposits models. Zeitschriftdeutschen Gesellschaft
Recife, 4, 1816–1826. für Geowissenschaften, 158, 45 –55.
Maciel, P. 1959. Tilito Cambriano no Estado de Mato Grosso. Boletim da Walde, D. H. G., Gierth, E. & Leonardos, O. H. 1981. Stratigraphy and
Sociedade Brasileira de Geologia, 8, 31 –39. mineralogy of the manganese ores of Urucum, Mato Grosso, Brazil.
Misi, A., Kaufman, A. J. et al. 2007. Chemostratigraphic correlation of Geologische Rundschau, 70, 1077– 1085.
Neoproterozoic successions in South America. Chemical Geology, Walde, D. H. G., Leonardos, O. H., Hahn, G. & Pflug, H. D. 1982.
237, 143– 167. The first Precambrian megafossils from South America,
Nogueira, A. C. R & Riccomini, C. 2006. O Grupo Araras (Neoprotero- Corumbella werneri. Anais da Academia Brasileira de Ciências,
zoico) na parte norte da Faixa Paraguai e sul do Craton Amazônico, 54, 461.
Brazil. Revista Brasileira de Geociências, 36, 576–587. Zaine, M. F. & Fairchild, T. R. 1985. Comparison of Aulophycus Lucia-
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V. & noi Beurlen and Sommer from Ladário (MS) and genus Cloudina
Fairchild, T. R. 2003. Soft-sediment deformation at the base of germs, Ediacaran of Namibia. Anais da Academia Brasileira de Ciên-
Neoproterozoic Puga cap carbonate (southwestern Amazon craton, cias, 57, 180.
Brazil): confirmation of rapid icehouse to greenhouse transition in Zaine, M. F. & Fairchild, T. R. 1987. Novas considerações sobre os
snowball Earth. Geology, 31, 613– 616. fósseis da Formação Tamengo, Grupo Corumbá, SW do Brasil. In:
Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V., Anais Xº Congresso Brasileiro de Paleontologia, Rio de Janeiro,
Trindade, R. I. F. & Fairchild, T. R. 2007. Carbon and strontium Brazil, 797– 806.
Chapter 46

The Serra Azul Formation, Paraguay Belt, Brazil

MILENE F. FIGUEIREDO1 *, MARLY BABINSKI1 & CARLOS J. S. ALVARENGA2


1
Instituto de Geociências, Universidade de São Paulo, São Paulo-SP, 05508-080, Brazil
2
Instituto de Geociências, Universidade de Brası́lia, Brası́lia, DF, 70910-900, Brazil
*Corresponding author (e-mail: milene.figueiredo@gmail.com)

Abstract: A new succession of diamictites and siltstones has been found within the Araras Group and is interpreted to record Ediacaran
glaciation in the northern Paraguay Belt, Brazil. This discontinuous stratigraphic unit, named the Serra Azul Formation (Fm.) (Figueiredo
et al. 2005; Alvarenga et al. 2007), is up to 300 m thick. It lies above dolomites of the Nobres Fm. and below sandstones of the Raizama
Fm. At the stratotype section, the Serra Azul Fm. comprises c. 70 m of massive glacial diamictites, overlain by 200 m of laminated
siltstones and rhythmites that contain sparse intercalations of very fine sandstone lenses.

The Serra Azul Fm. lies between carbonates of the upper Araras stromatolitic dolostones, mud cracks and dolomitized evaporites
Group and the siliciclastic lower Alto Paraguai Group in the north- (Alvarenga et al. 2004; Figueiredo 2006; Nogueira et al. 2007).
ern Paraguay Belt, southeastern Amazon craton (Fig. 46.1). The Zaine (1991) and Nogueira et al. (2007) interpreted the deposi-
Serra Azul Fm. was first described (Figueiredo et al. 2004) as a tional environment of the uppermost part of the Nobres Fm. as
new diamictite level associated with clayey-siltstones and was peritidal, arid tidal and sabkha.
recently elevated to formation status (Alvarenga et al. 2007). The Serra Azul Fm. ranges between 250 and 300 m thick in the
These rocks are poorly exposed because they are generally either region of its type section where it is completely exposed. It is subdi-
covered by talus or strongly weathered. The formation occurs in vided into two informal members: a lower unit comprising diamic-
elongated areas of low relief between ranges of Raizama Fm. sand- tites and an upper unit of pelites with increasing fine sand upsection
stones and hills comprising the Araras Fm. carbonates. The best (Figueiredo 2006; Alvarenga et al. 2007). The diamictite member is
exposures are found in rills, quarries and road cuts. The stratotype exposed only around the Serra Azul Syncline and near the town of
of the Serra Azul Fm. is exposed on the east –west-trending Serra Nobres. The clayey-siltstone member occurs over hundreds of kilo-
Azul Syncline (Fig. 46.1) in Mato Grosso State. Other exposures metres in the northern part of the Paraguay Belt (Fig. 46.1). Lateral
along the belt are stratigraphically incomplete, typically preser- variations in the thickness of the Serra Azul Fm. result mainly from
ving only the upper part of the formation. variations in the thickness of the diamictite member, which does
not crop out in some areas due to non-deposition or erosion prior
to deposition of the upper member. Where the Serra Azul Fm. is
Structural framework absent, stromatolitic dolomites from the upper Nobres Fm. lie dis-
conformably beneath sandstones of the Raizama Fm.
The Serra Azul Fm. was deposited on a passive margin along The Raizama Fm. is c. 1200 m thick (Hennies 1966) and is com-
the border of the Amazon Craton, which was folded by the posed of conglomerates, sandstones and arkoses interbedded with
Brasiliano – Pan-African Orogeny during the Cambrian Period siltstones and shales (Hennies 1966; Ribeiro Filho et al. 1975).
(Alvarenga & Trompette 1993; Trompette 1994; Trompette et al. Shales are more abundant in the lower part of the Raizama Fm.
1998; Trindade et al. 2003; Tohver et al. 2010). The stratotype Structures suggesting a shallow to emergent depositional environ-
of the formation occurs along the Serra Azul (Azul Range) in the ment, such as cross-stratification in sandstones, ripple-waves in silt-
northern Paraguay Belt on the southern limb of an openly stones and mud cracks in claystones, are common (Ribeiro Filho
folded, east – west-trending asymmetric syncline (Fig. 46.1; Alvar- et al. 1975; Zaine 1991). Zaine (1991) found a low-diversity ichno-
enga et al. 2007), located in the external tectonic zone of the fossil assemblage in the sandstones from the basal portion of this
fold-and-thrust belt, where the metamorphism is minimal and the formation: (i) Cochlichnus (simple and sinuous caving) between
sedimentary structures are well preserved. the crests of ripples and (ii) Lockeia (resting mark) associated with
(iii) Planolites (trackway), all in the same stratum. The contact
with the Serra Azul Fm. is gradational in the Azul Range region
Stratigraphy
and erosive in other localities (Figueiredo et al. 2008), where the
Nobres Fm. is in sharp contact with the Raizama Fm. (Hennies
The Serra Azul Fm. occurs between the Araras Group below and
1966; Ribeiro Filho et al. 1975; Nogueira et al. 2007). The basal
the Alto Paraguai Group above (Figs 46.1 & 46.2). It is underlain
part of the Raizama Fm. has been interpreted as transgressive depos-
by dolomites of the predominantly shallow marine Nobres Fm.
its with marine and fluvial facies (Nogueira et al. 2007).
(Araras Group) and is overlain by sandstones of the tidally influ-
enced Raizama Fm. (Alto Paraguay Group).
Dolostones of the Nobres Fm. were first documented by Castel-
neau (1850) and then by Almeida (1964), who attributed them to Glaciogenic deposits and associated strata
the upper Araras Group. It is c. 500 m thick (Hennies 1966) and
overlies limestones of the Guia Fm. (Alvarenga et al. 2011). The Serra Azul Fm.
Nobres Fm., from bottom to top, comprises dolosparite with two
interbedded layers of arkoses, a thick silicified carbonate level, The diamictite unit is c. 70 m thick and was deposited on top of the
impure dolograinstone with cross-bedding, peloids and upward Nobres Fm., but this contact is not exposed (Fig. 46.2). It is com-
increasing intraclasts and oolites, sometimes interbedded with thin posed of massive to crudely stratified diamictites with an abundant
layers of primary dolomudstone. The succession ends with sandy-clayey-silty matrix and sparse clasts of varied composition,

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 499– 502. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.46
500 M. F. FIGUEIREDO ET AL.

Fig. 46.1. Geological maps showing the location of Paraguay Belt on the Amazon craton and the occurrence of the Serra Azul Fm. in the northern part of the
Paraguay Belt.

size and shape (Alvarenga et al. 2007). The matrix is commonly indicate a glacial transport direction from the Amazon Craton
reddish and massive, becoming yellowish and stratified in the (WNW) towards the Paraguay Belt (ESE).
uppermost 60 cm, immediately below the siltstone unit. The siltstone unit is about 25 m thick and sharply overlies the
Clast lithologies include sandstone, arkose, chert, quartz, quart- diamictite level. This unit comprises a reddish laminated siltstone
zite, meta-conglomerate, weathered carbonate, claystone, diabase, that in places contains sparse granules (2 –3 mm) of quartzite and
granite, gneiss, basalt and rhyolite. The clast size ranges from quartz (Figueiredo et al. 2008). Towards the top, this reddish silt-
granules, pebbles (c. 5%) and cobbles (c. 1%) to boulders (1%), stone grades into rhythmites that are about 150 m thick (Figueiredo
with a few reaching 30 cm in diameter (Figueiredo et al. 2008). et al. 2008). At the base of this succession, there are alternating
Their shape also varies from well rounded to highly angular; clayey and silty plane-parallel laminae with intercalations of sand-
some are faceted, polished and striated. Some rounded and stone layers (5 –10 cm thick) showing small hummocky lami-
polished clasts have impact marks preserved on the non-striated nation that indicates oscillatory flow events. In the Nobres town
surface, while some oblate clasts have parallel sharply abraded region, there is a lens (12 m) of limestone in the basal portion of
and striated surfaces, suggesting a reworking of fluvial pebbles this rhythmite succession (Alvarenga et al. 2007). The textures
by glaciers. Other clasts are bullet-shaped; the major axes of one of the limestone, from bottom to top, include laminated mudstone,
large cobble and one boulder have a WNW trend that could massive mudstone with small-scale cut-and-fill channels filled by
wackestone, laminated mudstone with clotted texture, mudstone
with truncated laminae, and cyclic successions of breccias with
slumping deformation and plane-parallel lamination. Red clayey
films occur between limestone laminae, the upwards increase of
which causes a nodular texture due to the progressive deformation
of the limestone by compaction (Möller & Kvingan 1988).
In the intermediary portion of the rhythmite succession, there
is a constant input of sandy, alternating clayey (2 mm to 3 cm)
and very fine sandy (1 –5 mm) laminae overlain by a massive
medium sandstone layer c. 15 cm thick with very angular quartz
grains. Above this layer, the rhythmites continue, although with
more contribution from very fine sand (laminae of 2 mm to
2 cm) in relation to silt (laminae of 1–10 mm).
The rhythmites grade into a heterolithic succession c. 40 m thick
that is composed of inter-bedded siltstones and sandstones (Figueir-
edo et al. 2008). The basal part has intercalated lenses (0.5–2 cm
thick) of cross-laminated very fine sandstone within plane-parallel
clayey-siltstones. The sandstone lenses become thicker (from
laminae to metric to decimetric-scale beds) and coarser towards
the top, while siltstone beds decrease upwards. At the top of the suc-
cession, the sandstone facies dominates in metre-scale sigmoidal
beds (lobes) interbedded in a heterolithic bed. The first thick sand-
stone bed is parallel laminated, and the upper beds display
massive bedding, cross-stratification and hummocky structures.

Boundary relations with overlying and underlying


non-glacial units

Fig. 46.2. Generalized Paraguay Belt stratigraphy and a detailed stratigraphic In the Azul Range region (Fig. 46.1), the basal contact of diamic-
log of the shelfal Serra Azul Fm. in the northern part of the belt. tites from the Serra Azul Fm. with the dolomites of the underlying
THE SERRA AZUL FORMATION, PARAGUAY BELT 501

Nobres Fm. is not exposed. However, the presence of brecciated, For the last 60 cm of the diamictite unit, a crude stratification is
silicified, stromatolitic and nodular carbonate clasts with charac- apparent, and the clasts become smaller, suggesting retreat of the
teristics similar to those of the Nobres Fm. suggests an erosive glacier. The laminated siltstones deposited directly on the diamic-
contact between these two units. tites could represent further retreat of the glacier, producing distal
The upper contact between the diamictites and the laminated facies deposits that are less glacially influenced.
siltstones is sharp and conformable (Figueiredo 2006; Alvarenga A calcareous lens deposited above the reddish laminated silt-
et al. 2007). However, it is difficult to define where the last evi- stone suggests an interruption of clastic deposition and could
dence of glacial influence occurs within the above siltstone mark a maximum flooding surface. The rhythmite (sand and
member (Figueiredo et al. 2008). The upper silstones grade into clay) deposited on top of the carbonate represents a new sequence
rhythmites. Towards the top of the formation, sandstone inter- of clastic sedimentation. This sequence began as cyclic (rhythmite)
bedded with these rhythmites grades into heterolithic sediments. with storm events (hummocky cross stratification). Upsection,
The sandstones are cross-stratified and become thicker and sandstone layers become thicker and coarser and contain tidal
coarser towards the top of the Serra Azul Fm., marking a transi- bundles. This progression of facies suggests a progradational
tional contact with the Raizama Fm. deltaic sequence.
The dolomite –siliciclastic succession of facies in the under-
lying Nobres Fm. is consistent with a shallowing-upward platform
Chemostratigraphy succession (Almeida 1964; Alvarenga et al. 2004; Nogueira et al.
2007) with evolution from a slope and moderately deep platform to
No geochemical data are available for the Serra Azul Fm. Stable transitional and supratidal environments like arid tidal flats and
and radiogenic isotope data from the underlying Araras Group sabkhas (Nogueira et al. 2007). Thus, Serra Azul glaciers trans-
are reviewed elsewhere in this volume (Alvarenga et al. 2011). gressed an Ediacaran-aged carbonate platform.
To summarize, d13C (carbonate) values in the Nobres Group The siliciclastic Raizama Fm. that overlies the Serra Azul Fm.
vary between c. –2 and 0‰ and the least radiogenic 87Sr/86Sr suggests deposition in an epeiric environment during slow and pro-
ratios in the upper Guia Fm. limestones and Nobres Fm. range gressive basin subsidence (Almeida 1964; Nogueira et al. 2007).
from 0.7078 and 0.7081 (Nogueira et al. 2007; Alvarenga The conglomerates record development of fluvial channels with
et al. 2008). subaerial exposure, and the sedimentary structures and textures
of the sandstones record a tidally influenced platform (Almeida
1964; Nogueira et al. 2007). These characteristics indicate that
Geochronological constraints the basal Raizama Fm. was deposited from proximal to distal
facies in a transgressive regime (Nogueira et al. 2007).
No direct radiometric ages or fossils are available for the Serra
Azula Fm., and precise ages are notably sparse in the Paraguay We thank the Conselho Nacional de Desenvolvimento Cientı́fico e Tecnológico
Belt. A Pb/Pb whole-rock isochron age of 633 + 25 Ma (Babisnki (CNPq) for a research grant (Proc. 473614/2004-9) and the PhD scholarship
et al. 2006) was determined on carbonates of the Mirassol d’Oeste awarded to the senior author. M. F. Figueiredo thanks the Fundação de Amparo
Fm. at the base of the Araras Group, providing a maximum age à Pesquisa do Estado de São Paulo (FAPESP) for a Master’s scholarship (Proc.
constraint on the Serra Azul Fm. that implies a mid-Ediacaran 04/06225-5). M. Babinski and C. J. S. Alvarenga are CNPq Research Fellows.
age. A minimum age constraint is provided by an encapsulation This represents a contribution of the IUGS- and UNESCO-funded IGCP (Inter-
40Ar/39Ar metamorphic age of 528 + 26 Ma obtained on clay national Geoscience Programme) project #512.
fractions from multiple stratigraphic levels within the Paraguay
Belt (Tohver et al. 2010). Additional radiometric data include References
K –Ar ages of c. 730 Ma obtained on volcanic clasts recovered
from the diamictites, and Sm– Nd TDM ages ranging from 1.9 to Almeida, F. F. M. 1964. Geologia do Centro-Oeste Matogrossense. Min-
1.6 Ga obtained on the matrix of the diamictites and pelites of istério de Minas e Energia, DNPM, Boletim da Divisão de Geologia
the Serra Azul Fm. (Figueiredo, unpublished data). These TDM Mineral, 215, 1– 137.
ages suggest that the rocks of the Ventuari-Tapajós and Sunsás Alvarenga, C. J. S. & Trompette, R. 1993. Evolução Tectônica Brasili-
provinces (Tassinari & Macambira 1999) from the Amazon ana da Faixa Paraguai na região de Cuiabá. Revista Brasileira de
craton could be the main source of sediments to the Serra Geociências, 23, 18 –30.
Azul deposits. The source area of the younger volcanic clasts Alvarenga, C. J. S., Santos, R. V. & Dantas, E. L. 2004. C-O-Sr iso-
(c. 730 Ma), however, has not yet been identified. topic stratigraphy of cap carbonates overlying Marinoan-age glacial
diamictites in the Paraguay Belt, Brazil. Precambrian Research,
131, 1– 21.
Discussion Alvarenga, C. J. S., Figueiredo, M. F., Babinski, M. & Pinho, F. E. C.
2007. Glacial diamictites of Serra Azul Formation (Ediacaran, Para-
The massive and poorly sorted characteristics of the Serra Azul guay Belt): evidence of the Gaskiers glacial event in Brazil. Journal
Fm. diamictite indicate deposition from high density flows of South American Earth Science, 23, 236– 241.
(Edwards 1978) without reworking by currents. Furthermore, the Alvarenga, C. J. S., Boggiani, P. C. et al. 2011. Glacially-influenced
large variety of clast compositions requires a transportation sedimentation of the Puga Fm., Cuiabá Group and Jacadigo
Group, and associated carbonates of the Araras and Corumbá
agent capable of scouring huge areas and mixing many rock
groups, Paraguay Belt, Brazil. In: Arnaud, E., Halverson, G. P.
types (Eyles & Miall 1984). In addition, large quantities of striated,
& Shields-Zhou, G. (eds) The Geological Record of Neo-
faceted and polished clasts were recovered from the diamictites, proterozoic Glaciations. Geological Society, London, Memoirs,
which normally indicates a glacial environment (Eyles 1993). 36, 487– 497.
Two bullet-shaped boulders both orientated along the same trend Babinski, M., Trindade, R. I. F., Alvarenga, C. J. S., Boggiani, P. C.,
were also found that suggest east – SE ice movement, that is, Liu, D., Santos, R. V. & Brito Neves, B. B. 2006. Chronology
from the palaeocontinent (Amazon craton) to the ocean (Paraguay of Neoproterozoic ice ages in central Brazil. In: V South American
Belt). However, these two boulders alone are insufficient evidence Symposium on Isotope Geology, 2006, Punta del Este, Uruguay,
to classify this diamictite as an ice contact deposit. Other charac- Short Papers, 1, 303–306.
teristic structures, such as deformed or sheared substrate, eroded Castelneau, F. 1850. Expedition dans les parties centrales de l’Améri-
basement and striated pavement, are necessary for such a classifi- que du Sud, de Rio de Janeiro a Lima, et de Lima au Pará. Histoire
cation, but these were not observed (Figueiredo et al. 2008). du Voyage. Librairie Editeur, Paris, Tomo II.
502 M. F. FIGUEIREDO ET AL.

Edwards, M. B. 1978. Glacial environments. In: Reading, H. G. (ed.) Nogueira, C. R. A., Riccomini, C., Sial, A. N., Moura, C. A. V., Trin-
Sedimentary Environments and Facies. Elsevier, New York, dade, R. I. F. & Fairchild, T. R. 2007. Carbon and strontium isotope
416– 438. fluctuations and paleoceanographic changes in the late Neoprotero-
Eyles, N. 1993. Earth’s glacial record and its tectonic setting. Earth zoic Araras carbonate platform, southern Amazon Craton, Brazil.
Science Reviews, 35, 1– 248. Chemical Geology, 237, 168– 190.
Eyles, N. & Miall, A. D. 1984. Glacial facies. In: Walker, R. G. (ed.) Ribeiro Filho, W., Luz, J. S. & Abreu Filho, W. 1975. Relatório Final
Facies Models, 2nd edn. Geoscience Canada, Reprint Series, 1, do Projeto Serra Azul. Ministério de Minas e Energia, DNPM/
15 –38. CPRM, 1, 1 –104.
Figueiredo, M. F. 2006. Quimioestratigrafia das rochas ediacarianas do Tassinari, C. C. G. & Macambira, M. J. B. 1999. Geochronological pro-
extremo norte da Faixa Paraguai, Mato Grosso. MS thesis, Instituto vinces of the Amazonian Craton. Episodes, 22, 174–182.
de Geociências, Universidade de São Paulo. Tohver, E., Trindade, R. I. F., Solum, J. G., Hall, C. M., Riccomini, C.
Figueiredo, M. F., Babinski, M., Alvarenga, C. J. S. & Pinho, F. E. C. & Nogueira, A. C. 2010. Closing the Clymene ocean and bending a
2004. Diamictites overlying Marinoan-age carbonates of Araras Brasiliano belt: evidence for the Cambrian formation of Gondwana,
Formation, Paraguay Belt, Brazil: evidence of a new glaciation? southeast Amazon craton. Geology, 38, 267– 270.
In: Symposium on Neoproterozoic-Early Paleozoic Events in SW- Trindade, R. I. F., Font, E., D’Agrela Filho, M. S., Nogueira, A. C. &
Gondwana, IGCP-478, Second Meeting, Brazil, 18– 19. Riccomini, C. 2003. Low-latitude and multiple geomagnetic rever-
Figueiredo, M. F., Babinski, M., Alvarenga, C. J. S. & Pinho, F. E. C. sals in the Neoproterozoic Puga cap carbonate, Amazon craton.
2005. Nova unidade litoestratigráfica: Formação Serra Azul, Faixa Terra Nova, 15, 441–446.
Paraguai, Mato Grosso. IX Simpósio de Geologia do Centro-Oeste, Trompette, R. 1994. Geology of Western Gondwana (2000– 500 Ma).
Brazil, 23 –25. Pan-African –Brasiliano Aggregation of South America and Africa.
Figueiredo, M. F., Babinski, M., Alvarenga, C. J. S. & Pinho, F. E. C. Balkema.
2008. Nova unidade litoestratigráfica registra glaciação ediacarana Trompette, R., Alvarenga, C. J. S. & Walde, D. 1998. Geological evol-
em Mato Grosso: Formação Serra Azul. Geologia USP, 8, 65 –75. ution of the Neoproterozoic Corumbá graben system (Brazil).
Hennies, W. T. 1966. Geologia do Centro-Norte Mato-Grossense. Depositional context of the stratified Fe and Mn ores of Jacadigo
Unpublished PhD dissertation, Escola Politécnica, Universidade de Group. Journal of South American Earth Science, 11, 587– 597.
São Paulo. Zaine, M. F. 1991. Análise dos fósseis de parte da Faixa Paraguai
Möller, N. K. & Kvingan, K. 1988. The genesis of nodular limestone in (MS, MT) e seu contexto temporal e paleoambiental. Unpublished
the Ordovician and Silurian of the Oslo Region (Norway). Sedimen- PhD dissertation, Instituto de Geociências, Universidade de São
tology, 35, 405– 420. Paulo.
Chapter 47

The Bebedouro Formation, Una Group, Bahia (Brazil)

J. T. GUIMARÃES1, A. MISI2*, A. J. PEDREIRA1 & J. M. L. DOMINGUEZ2


1
Geological Survey of Brazil (CPRM) Av. Ulysses Guimarães, 2862 – Sussuarana – Centro Administrativo
da Bahia 41213-000 Salvador-Bahia, Brazil
2
Universidade Federal da Bahia, Centro de Pesquisa em Geofı́sica e Geologia, Instituto de Geociências,
Campus da Federação 40170-290 Salvador-Bahia, Brazil
*Corresponding author (e-mail: aroldo.misi@gmail.com)

Abstract: The Bebedouro Formation (Fm.) is a Neoproterozoic glaciogenic succession at the base of the Una Group on the São
Francisco Craton. The glaciogenic sequence is composed of diamictites, pelites and sandstones with a variety of lithofacies that are
grouped into four associations: (i) ice-contact, (ii) pro-glacial, (iii) ice-rafted and (iv) aeolian (extraglacial). Thus, the Bebedouro Fm.
is interpreted to have been deposited in a shelf marine environment where glacio-proximal sedimentation dominated. It variably overlies
Palaeoproterozoic/Archaean basement and Mesoproterozoic metasedimentary rocks of the Chapada Diamantina Group.
The overlying carbonate succession (Salitre Fm.) is informally subdivided into five mappable units. Unit C, at the base, is composed of
red argillaceous dolostone with typical negative d13C signatures averaging – 5.1‰ (VPDB). Unit B consists of grey laminated limestones
that grade upward into grey dolostone with tepee structures of Unit B1; both have d13C values around 0‰. The overlying Unit A consists
of interbeds of greyish marl, shales and siltstones grading upward into massive black organic-rich limestone with oolitic and pisolitic beds
of Unit A1. Positive d13C signatures averaging þ8.5‰ VPDB characterize the black limestones of Unit A1. 87Sr/86Sr ratios of well-
preserved samples of these carbonate units show values of 0.70745 and 0.70765. These units can be correlated litho- and chemostrati-
graphically with formations of the Bambuı́ Group (São Francisco Basin) to the west.

The Bebedouro Formation crops out in the central part of Bahia epicontinental and passive margin Neoproterozoic basins were
state, Brazil, where it occurs in narrow and apparently discontinu- formed over and around the 1.8 Ga stabilized terrains of the São
ous north – south strips that span a length of 400 km and cover an Francisco Craton (Almeida 1977). They were intensely deformed
area of more than 40 000 km2 (Fig. 47.1). This formation occurs and faulted by the events of the Brasiliano orogeny, during which
in several isolated sub-basins (Fig. 47.1) but was probably depos- the most important movements occurred between 650 and 500 Ma.
ited within a single, large, precursor basin that was subsequently The present configuration of the craton separates the Una Group
segmented during tectonic events related to the assembly of into four geographically isolated basins on the stabilized eastern
western Gondwana (Teixeira et al. 2007). part of the São Francisco craton: the Salitre, Irecê, Una-Utinga
The name ‘Bebedouro’ was introduced by Oliveira & Leonardos and Ituaçu sub-basins (Fig. 47.1). Presumably, these sub-basins
(1940) to designate conglomerates occurring in the eastern border were originally connected, forming a single, large basin that pre-
of the Mesoproterozoic basin of the Chapada Diamantina, in Bahia dated the Brasiliano orogeny. Deformed equivalents of the Una
state. The conglomerates were first described by Derby (1905) but Group occur in fold belts along the north and east borders of the
their glacial origin was proposed by A. I. Oliveira in 1921, as cratonic area within the Rio Preto, Riacho do Pontal, Vasa
reported by Moraes Rego (1930), and Williams (1930). Sub- Barris/Miaba and Rio Pardo groups. Metamorphic grade is very
sequent work by several authors demonstrated their glacial low or absent in the intra-cratonic sub-basins and low (up to green-
origin (Mello Jr 1938; Kegel 1959; Brito Neves 1967; Söfner shist facies) in the adjacent fold belts.
1973; Montes 1977; Rocha Campos & Hasui 1981; Dominguez Structural lineaments (fractures and faults) trending NNW –
1993). Montes (1977) noted the presence of striated pavements SSE, probably related to extensional movements, cross both
and dropstones, and proposed a continental glaciation for the the basement (Palaeoproterozoic and Mesoproterozoic) and the
origin of the Bebedouro Fm. Glaciomarine sequences and ‘conti- Neoproterozoic sedimentary cover, suggesting that the extensional
nental tillites’ were described by Söfner (1973), Montes (1977) tectonic regime continued during the sedimentation of the Una
and Karfunkel & Hoppe (1988), who showed evidence of ice Group (Misi et al. 2005). Later orthogonal compression (north –
flow from east to west. Subsequently, Guimarães (1996) proposed south and east –west) generated a basin and dome structure
that the deposition of the Bebedouro Fm. occurred in a glaciomar- pattern in the sedimentary cover. In some places, as in the
ine environment by sub-aqueous debris-flow, turbidity currents, western border of the Irecê sub-basin, the Mesoproterozoic base-
iceberg melting and locally aeolian processes. ment is exposed.
The Bebedouro Fm. correlates with the Jequitaı́ Fm. of the São
Francisco Basin (Bambuı́ Group) to the west and SW (Uhlein et al.
2011), which is also referred to as the Macaubas Group by Stratigraphy
some authors.
The Una Group carbonates were deposited in a series of sub-basins
in Bahia state, northeastern São Francisco craton, whereas the
Structural framework equivalent Bambuı́ Group was deposited in the São Francisco
Basin, located in the states of Minas Gerais, Goiás and along the
The Bebedouro Fm. and the overlying carbonate succession of the western border of Bahia state (Fig. 47.1). The Bebedouro Fm.,
Una Group were likely deposited in a basin formed by extension which is less than 200 m thick, lies unconformably over either
related to the fragmentation of the Rodinia supercontinent, starting Mesoproterozoic metasedimentary rocks of the Chapada Diaman-
c. 950 Ma and continuing until c. 600 Ma (Condie 2002). These tina Group or a Palaeoproterozoic/Archaean medium to high-
events occurred synchronous with the onset of Gondwana assem- grade metamorphic complex. It is overlain by the carbonate
bly (Condie 2002; Cordani et al. 2003; Teixeira et al. 2007). The platform succession of the Salitre Formation (Fig. 47.2).

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 503– 508. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.47
504 J. T. GUIMARÃES ET AL.

Localization 43º15’00’’W 40º15’00’’W


65º00’00’W 9º30’00’’S 9º30’00’’S
NOVA
REMANSO

42º00’00’’W

2º00’00’’N 2º00’00’’N

Study
BRAZIL Area
S. Francisco Basin
(Bambuí Group)

São BAHIA
Francisco XIQUE-
Craton XIQUE
Salitre
21º00’00’’S 21º00’00’’S
Sub-basin

42º00’00’’W
IRECÊ

MORRO
Irecê DO CHAPEU
Sub-basin
500 km 0 500 1.000 1.500 km

BA
44º00’00’’S 44º00’00´´S 052

SEABRA
BR
242
65º00’00’’W

Town
Una-Utinga
Syncline ANDARAI Sub-basin
Brachysyncline ITAETÊ
MACAUBAS
Phanerozoic Cover

São Francisco Supergroup


Salitre Formation

Bebedouro Formation

Glacial contact deposits

Proglacial and ice-rafted debris

Aeolian extraglacial deposits Ituaçu


Sub-basin
Espinhaço Supergroup
ITUAÇU
14º10’00’’S 14º10’00’’S
Basement older than 1.8 Ga 43º15’00’’W 40º15’00’’W

Direction of glaciers movement


50 km 25 0 50 km

Fig. 47.1. Geological map showing the Bebedouro Fm. in the central area of Bahia state, Brazil (modified from Guimarães 1996).

Misi & Souto (1975) identified two members in the Bebedouro The Bebedouro Fm.
Fm., within the Irecê Sub-basin: the Lower Member is composed of
green and reddish meta-siltstones and slates with chert and lime- The Bebedouro Fm. is composed of diamictites, sandstones and
stone lenses, with lithic metagreywacke forming the matrix pelites. These lithofacies show sharp, gradational or erosive con-
above the diamicitites (total thickness c. 100 m); the Upper tacts and are generally texturally and mineralogically immature.
Member is a feldspatic sandstone (meta-arkose), up to 80 m The diamictites occur mainly in the eastern margins of the
thick. A more detailed study by Guimarães (1996) that included Una-Utinga and Campinas sub-basins (Fig. 47.1). They contain
the other sub-basins showed that the diamicites are in fact interca- poorly sorted, heterolithic clasts. At least four types of coarse-clast
lated with arkose, greywackes and lithic greywacke. These lithofa- bearing facies have been identified by Guimarães (1996): (i)
cies are described in the following section. matrix-supported massive diamictite, (ii) matrix-supported dia-
The Salitre Fm., overlying the Bebedouro Fm., is more than mictite with parallel stratification, (iii) matrix-supported diamic-
500 m thick and is composed of five informal units (Misi & tites with sigmoidal cross-bedding and (iv) massive clast-
Souto 1975; Misi 1978) (Fig. 47.2), briefly summarized in the supported conglomerate/breccia. In all facies, clasts are domi-
next section. These units are correlative with the Bambuı́ Group nantly but not exclusively angular and range in size from granules
formations of the São Francisco Basin, as shown in Table 47.1. to blocks .2 m in diameter. Clast compositions include granite,
gneiss, pegmatite, schist, phyllite, basic/ultrabasic rocks, volca-
nics, calc-silicates, quartz, grey and green quartzites, sandstones,
Glaciogenic deposits and associated strata argillites, carbonates and chert. Faceted and polished surfaces
are common, and striated clasts have been found.
The descriptions that follow are essentially based on Misi & Souto The matrix is composed of greywacke, lithic greywacke, lithic
(1975), Misi (1978) and Guimarães (1996). sandstone, arkose, sub-arkose and argillite with quartz and rock
THE BEBEDOURO FORMATION, BRAZIL 505

Fig. 47.2. Lithostratigraphic successions of


the Una Group in the Irecê sub-basin
(Source: Misi & Souto 1975; Misi 1978).

fragments. Clay material and impregnated iron oxides are impor- of the sandstones includes spherical to more commonly elongate
tant components of the matrix and clasts, suggesting that the rock fragments and granules and clasts of quartz and feldspar.
matrix component is at least partially derived from the comminu- Guimarães (1996) recognized parallel-bedded reddish (Fe-rich)
tion of larger fragments. Generally, the matrix is poorly sorted and sandstones with and without dropstones, hummocky cross-bedding
contains fragments of variable sizes and origins. The larger frag- and large-scale tangential cross-bedding. Asymmetric ripples and
ments correspond are angular. normal gradating are also observed in these lithofacies. Lone-
The diamictite bodies are lenticular, lobate or tabular at the stones, observed in some places, are gravel to block-size clasts
outcrop scale and commonly show reworking by storm waves. with variable composition and shape. These sandstones represent
Internally they are variably massive or stratified, and sometimes wave-reworked fluidized/liquefied and confined gravitational
exhibit slump structures. Normal and inverse grading, clast imbri- flows, which evolved to high-concentration turbidity currents as
cation, plane-parallel bedding and sigmoidal cross-bedding occur a result of iceberg melting.
within stratified bodies. Bimodal, medium to coarse-grained sandstones with large-
Guimarães (1996) has suggested that the diamictites of the scale tangential and trough cross-bedding interpreted to represent
Bebedouro Fm. were derived from rapid deposition with limited aeolian deposition occur in isolation on the eastern margin of the
transport in confined debris flows (cohesive and non-cohesive), Una-Utinga sub-basin, close to the town of Itaetê (Fig. 47.1).
high concentration turbidity currents, and grain falls in a subaqu- The pelites include variably silicified grey-greenish, red-ochre
eous environment (Fig. 47.3). and purplish siltstone, greywacke and arkose. These lithotypes
The Bebedouro sandstones, although widespread within the are widespread, found mainly in the Irecê and Ituaçu sub-basins
Neoproterozoic basins, are subordinate lithotypes. They include as well as in the western margins of the Campinas and Una-Utinga
ochre, brown and grey-greenish arkose, sub-arkose, greywacke, sub-basins (Fig. 47.1). Laminated and massive pelites are present
lithic greywacke, lithic sandstone and quartz arenite. Generally
they are interbedded with diamictites and pelites. The framework
WEST EAST

BUOYANT SEDIMENTS PLUME


Table 47.1. Lithostratigraphic correlation between the Bambuı́ and
ICEBERGS
Una groups
SL ICE SHEET

Lithotypes Bambuı́ Group Una Group (units


(formations) or formations)

Arkose, siltstone Três Marias –


Siltstone, pelites Serra da Saudade –
Black oolitic limestones Lagoa do Jacaré Unit A1 Log A

Marl, shales Serra Santa Helena Unit A Log B


Dolostone laminated Sete Lagoas Unit B1 Log C
limestones, red dolostone Unit B
Unit C
Diamictite, arkose, Jequitaı́, Macaubas Bebedouro Fig. 47.3. Ice-contact glaciomarine system, Bebedouro Fm. See Figure 47.4
pelitic rocks for schematic representation of logs A, B and C as indicated in this figure
(Source: Guimarães 1996).
506 J. T. GUIMARÃES ET AL.

(a) (b) (c)

SUSPENSION
SUSPENSION DERIVED SEDIMENTS
DERIVED SEDIMENTS

DEBRIS FLOW

HIGH-DENSITY
DEBRIS FALL TURBIDITE SUSPENSION

DERIVED
DEBRIS FLOW DEBRIS FLOW SEDIMENTS

HIGH-DENSITY
TURBIDITE
DEBRIS FALL Vertical Scale
0

DEBRIS FLOW SUSPENSION


DERIVED SEDIMENTS 2m
Fig. 47.4. Schematic representation of
glacial lithofacies associations within the
DEBRIS FLOW
Bebedouro Fm. (logs A, B and C, Fig. 47.4)
(Source: Guimarães 1996). (a) Ice – contact
diamictites; (b) Ice – marginal fan; and (c)
Iceberg melt.

with or without dropstones. Pelite packages range in thickness exposed in all of the studied sub-basins. It is informally subdivided
from centimetres to 20 m. Reworking by storm waves is com- into five mappable units within the Una Group. The following
monly observed. Compositionally, the pelites contain clay- lithofacies, briefly described bottom to top, were originally ident-
minerals, micro-crystalline quartz, iron oxides and carbonate ified in the Irecê sub-basin (Misi & Souto 1975) and later in the
cement. The larger grains are spherical to elongated and either other sub-basins (Misi 1979): red argillaceous dolostones (Unit
angular or sub-rounded. They are composed of quartz, feldspar, C), laminated limestone (Unit B), evaporitic dolostone (Unit
mica and rock fragments. Interspersed dropstones are of diverse B1), marl (Unit A) and black oolitic limestone (Unit A1).
shape, size and composition. The red dolostone of Unit C overlies the Bebedouro Fm.
The pelitic rocks are believed to be resedimented sub-aqueous (Fig. 47.2) and has a maximum thickness of 70 m. It is overlain
deposits, formed by cohesive and unconfined gravity flows, low- by c. 180 m of grey laminated limestone with rhythmic successions
concentration turbidity currents, and suspension fallout from of centimetre-thick beds of limestone or dolomitic limestone and
icebergs and hypopycnal plumes (Figs 47.3 & 47.4). argillaceous material (Unit B). This facies grades upward into
dolomitic layers with replaced nodular anhydrite, tepee structures,
microbialaminites and columnar stromatolites that comprise Unit
The Salitre Fm. B1, which has a total thickness of c. 50 m and was observed in
its entirety in drill hole IL 53 (Misi & Kyle 1994). The overlying
The carbonate succession overlying the Bebedouro Fm., referred unit A is composed of marl, argillites and siltstones and varies
to as the Salitre Fm. in the available geological maps, is well from 0 to 100 m in thickness. This unit is followed above by
more than 150 m of black, organic-rich oolitic and pisololitic
13
δ C(%0VPDB)
limestone of Unit A1 (Fig. 47.2).
-15 -10 -5 0 +5 +10 +15 At least two major sedimentary cycles can be identified in the
Salitre Fm. (Misi et al. 2005, 2007). Cycle 1 starts at the base of
87 86
Sr/ S r the Salitre Fm. (Units C and B) and ends in the dolomites with
0.70745
tepee structures (Unit B1) that representing the exposure surface
Unit A 1 of a typical shallowing-upward sequence. Cycle 2 starts with a
transgressive event represented by the marl and shale (Unit A)
Olithic, black
400 m limestone immediately above the dolomitic facies of Unit B1, passing to
Unit A oolitic and pisolitic black limestones with trough cross-bedding,
Marl
Unit B 1 Dolostone indicating shallow, high-energy sedimentation (Unit A1).
Salitre Fm.

Laminated
0.70765 limestone
Pink dolostone
Unit B Boundary relations with overlying and underlying
Diamictite
0.70780 non-glacial units
200 m

In most locations, the Bebedouro Fm. lies unconformably over


Mesoproterozoic sedimentary rocks of the Morro do Chapeu Fm.
of the Chapada Diamantina Group. On the eastern side of the
Unit C
Una-Utinga sub-basin (Fig. 47.1), relatively small outcrops of the
Bebedouro Fm. directly overlie the Archaean/Palaeoproterozoic
Bebedouro Fm.
terrains of the São Francisco Craton, which there is composed of
0m
gneisses and migmatites.
Fig. 47.5. Stratigraphic variation of d13C and 87Sr/86Sr in carbonates of the The Bebedouro Fm. is typically overlain by red argillaceous
Una Group, Irecê Basin (Isotopic data from Torquato & Misi 1977; and Misi & dolostones of Unit C of the Salitre Fm., which is widely occurring
Veizer 1998). but absent on the western margin of several sub-basins. In these
THE BEBEDOURO FORMATION, BRAZIL 507

locations, the Bebedouro Fm. is overlain by laminated limestones Evidence for the glacial origin of the Bebedouro Fm., as estab-
of Unit B. lished by many subsequent studies, may be summarized as follows
(Montes 1977; Montes et al. 1985; Guimarães 1996):
Chemostratigraphy † the presence of clasts from different sources and variable forms
and sizes, within pelitic rocks, which are angular to sub-angular
Chemostratigraphic studies of the carbonate successions of the as well as sub-rounded and rounded, some showing striae and
Una Group in the Irecê sub-basin (Torquato & Misi 1977; Misi preferential orientation;
& Veizer 1998) show remarkable stratigraphic variation of Sr † the presence of striated pavements;
and C isotopes. These trends are summarized in Figure 47.5. † the presence of fragmented and angular forms of minerals like
In summary, 87Sr/86Sr ratios vary between 0.70745 and 0.71776, quartz, microcline, oligoclase and andesine, as observed in thin
but the best-preserved samples, indicated by the lowest Mn –Sr sections (garnet crystals with chatter marks were also observed
ratios (,0.03) and highest total Sr contents (.300 to by Montes et al. 1985);
c. 2300 ppm) define a much narrower range of 0.70745– † the widespread distribution of the Bebedouro Fm. in all the
0.70769. The d13C signatures show low values of c. –5‰ studied sub-basins.
(n ¼ 12) in Unit C (Torquato & Misi 1977), values varying
around 0‰ through Units B and B1, and a positive excursion of The lithofacies of the Bebedouro Fm. are thought to be the pro-
up to 9.4‰ in the black and organic-rich carbonates of Unit A1. ducts of debris flows, turbidity currents, ice-rafting and suspension
Powis (2006), Misi et al. (2007) and Vieira et al. (2007) report mechanisms. The palaeo-flow directions of sediment transport
similar isotopic variation in the carbonate successions of the were from east to west. Locally, there is evidence of sediment
Bambuı́ Group, in the São Francisco Basin. Negative anomalies reworking by aeolian processes. Guimarães (1996) grouped these
like that found in Unit C are characteristic of many carbonate suc- lithofacies into four genetic associations: glacier contact, pro-
cessions immediately above glaciogenic lithotypes around the glacial, iceberg meltout and aeolian extraglacial, predominantly
world (Knoll et al. 1986; Kaufman et al. 1991; Halverson et al. in a proximal glaciomarine environment.
2005; Shields 2005). Misi et al. (2007) pointed out that ‘glaciogenic diamictite and
associated lithologies are known to occur in two discrete strati-
graphic positions within Neoproterozoic successions of South
Geochronological constraints America, and these most-likely correspond to Sturtian and Mar-
inoan (or Varanger) ice ages worldwide’. The Bebedouro Fm.
The ages of the Neoproterozoic successions of the São Francisco and the correlated units of the Bambuı́ Group are thought by
Craton, including the Bebedouro Fm. in Bahia state, are poorly many authors to be Sturtian in age (Babinski et al. 2007), that is,
resolved due to the lack of datable volcanic rocks. Metamorphism belonging to the older of two Cryogenian glacial epochs. Direct
of the rocks, although of low grade, has contributed to a general evidence of a younger (end-Cyrogenian or Marinoan) glaciation
lack in confidence of the Rb –Sr and Pb/Pb age determinations within the Bambuı́ Group or Una Group has remained elusive,
(Macedo 1982; Macedo & Bonhomme 1984; Trindade et al. but may be cryptically expressed by the negative d13C anomaly
2004). The most widely accepted radiometric determination is a in dolostones in Unit B1 (Fig. 47.5) and corresponding to the top
Pb/Pb isochron age of 740 + 27 Ma from six samples of the of the first of two cycles within the Salitre Fm.
basal Sete Lagoas Fm. (Bambuı́ Group) (Babinski et al. 2007).
Pelitic units of the Mesoproterozoic Chapada Diamantina Group The authors are grateful to the Geological Survey of Brazil (CPRM) and to the
underlying the Bebedouro Fm., dated by Babinski et al. (1993), National Research Council of Brazil (CNPq), which provided funding for the
yielded a Pb/Pb age of 1140 + 140 Ma, whereas intrusive mafic research over the last 10 years. Currently, A. M. receives support from project
rocks within the Chapada Diamantina Group were dated by U – no. 486416/2006-2 from CNPq. This represents a contribution of the IUGS-
Pb zircon at 1496 + 3.2 Ma (Guimarães et al. 2005). and UNESCO-funded IGCP (International Geoscience Programme) Project #512.

Mineralization
References
In the Irecê sub-basin, Unit B1 hosts massive stratiform and strata-
bound pyrite, sphalerite and galena deposits associated with Almeida, F. F. M. 1977. O Cráton do São Francisco. Revista Brasileira de
Geociências, 1, 13 –21.
medium grey tepee dolostone containing nodules of calcite,
Babinski, M., Van Schmus, W. R., Chemale, F. Jr, Neves, B. B. B. &
silica and barite (Misi & Souto 1975; Misi et al. 2005). Galena Rocha, A. J. D. 1993. Idade isocrônica Pb –Pb em rochas carbonáti-
and sphalerite mineralization also occurs in the Una-Utinga sub- cas da Formação Caboclo em Morro do Chapéu, BA. II Simpósio
basin. The sulphide concentrations in the Una Group are associated sobre o Cráton do São Francisco, Sociedade Brasileira de Geologia,
NNE – SSW faults (Misi et al. 2005). Economic phosphate deposits 2, Salvador (BA), Anais, 160– 163.
intimately associated with columnar and laminar stromatolites Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating of
occur within massive grey to reddish-grey dolostone in a lower the Sete Lagoas cap carbonate (Bambuı́ Group), Brazil and impli-
stratigraphic position of the Unit B1 in the Irecê sub-basin (Misi cations for the Neoproterozoic glacial events. Terra Nova, 19,
& Kyle 1994). 401– 406.
Brito Neves, B. B. 1967. Geologia das Folhas de Upamirim e Morro do
Chapéu-BA. CONESP/SUDENE. Relatório, 17, 35.
Discussion
Condie, K. C. 2002. The supercontinent cycle: are there two patterns of
cyclicity? Journal of African Earth Sciences, 35, 179– 183.
A glacial origin for the Bebedouro Fm. was originally proposed Cordani, U. G., Brito Neves, B. B. & D’Agrella Filho, M. S. D. 2003.
by Williams (1930), who described conglomerates and shales From Rodinia to Gondwana: a review of the available evidence from
cropping out in the northern Una-Utinga sub-basin. This author South America. Gondwana Research, 6, 265– 273.
stated that ‘it is difficult to explain the presence of a peculiar Derby, O. A. 1905. Notas geológicas sobre o Estado da Bahia. Secretaria
kind of clasts within the pelitic rocks. In this sense, we do not hes- de Agricultura, Viaçăo, Industria e Obras. Públicas Boletim, 7, 12–31.
itate in suggesting a glacial origin [for the conglomerate] which Dominguez, J. M. L. 1993. As coberturas do Cráton do Francisco: uma
was probably deposited in calm and deep water within a wide- abordagem do ponto de vista de análise de bacias. In: Dominguez,
spread sea, on which icebergs were floating and liberating clasts J. M. L. & Misi, A. (eds) O Cráton do São Francisco. Sociedade
[to the deep sea sediments]’. Brasileira de Geologia, Núcleo Bahia-Sergipe, 137– 159.
508 J. T. GUIMARÃES ET AL.

Guimarães, J. T. 1996. A Formação Bebedouro no Estado da Bahia; sequences, São Francisco Craton, Brazil: a review and a
faciologia, estratigrafia e ambientes de sedimentação. Dissertação possible metallogenic evolution model. Ore Geology Reviews, 26,
(Mestrado em Geologia), Instituto de Geociências, Universidade 263– 304.
Federal da Bahia (UFBA), Brasil. Misi, A., Kaufman, A. J. et al. 2007. Chemostratigraphic correlation of
Guimarães, J. T., Martins, A. A. M. et al. 2005. Projeto Ibitiara-Rio de Neoproterozoic successions in South America. Chemical Geology,
Contas, Ba. CPRM. 237, 143– 167.
Halverson, G. P., Hoffman, P. R., Schrag, D. P., Maloof, A. C. & Montes, A. S. L. 1977. O contexto estratigráfico e sedimentológico da
Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon- Formação Bebedouro na Bahia. Brası́lia. Tese (Mestrado) Universi-
isotope record. GSA Bulletin, 117, 1181–1207. dade de Brası́lia (UNB).
Karfunkel, J. & Hoppe, A. 1988. Late Proterozoic glaciation in Montes, A. S. L., Gravenor, C. P. & Montes, M. L. 1985. Glacial sedi-
Central-Eastern Brazil: synthesis and model. Palaeogeography, mentation in the Late Precambrian Bebedouro Formation, Bahia,
Palaeoclimatology, Palaeoecology, 65, 1 –21. Brazil. Sedimentary Geology, 44, 349–358.
Kaufman, A. J., Hayes, J. M., Knoll, A. H. & Germs, G. J. B. 1991. Moraes Rego, L. F. 1930. Glaciação eopaleozóica no centro do Brasil.
Isotopic composition of carbonates and organic carbon from upper Anais da Academia Brasileira de Ciências, 2, 109– 112.
Proterozoic successions in Namibia: stratigraphic variations and the Oliveira, A. I. & Leonardos, O. H. 1940. Geologia do Brasil. Comissão
effects of diagenesis and metamorphism. Precambrian Research, Brasileira dos Centenários de Portugal.
49, 301– 327. Powis, K. 2006. Stable isotope geochemistry of the Neoproterozoic
Kegel, W. 1959. Estudos Geológicos na Zona Central da Bahia. Bambuı́ Group at Serra do Ramalho, Bajia, Brazil. PhD
Divisão de Geologia e Mineralogia, Departamento Nacional da thesis, Ottawa-Carleton Geoscience Center, University of Ottawa,
Produção Mineral (DGM-DNPM), Boletim 198, Rio de Janeiro. Canada.
Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K. & Lambert, L. Rocha Campos, A. C. & Hasui, Y. 1981. The Late Precambrian Bebe-
B. 1986. Secular variation in carbon isotope ratios from upper Proter- douro Formation, Bahia, Brazil. In: Hambrey, M. J. & Harland,
ozoic successions of Svalbard and East Greenland. Nature, 321, W. B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge
832– 838. University Press, Cambridge. (IGCP Project, 38: Pre-Pleistocene
Macedo, M. H. F. 1982. Les sistémes isotopiques rubidium-strontium et Tillites.)
potassium-argon dans lês argiles extraités de sediments carbonatés. Shields, G. 2005. Neoproterozoic cap carbonates: a critical appraisal of
Application à la datation du protérozoique sedimentary du Brésil existing models and the plumeworld hypothesis. Terra Nova, 17,
dans les Etáts de Bahia et Santa Catarina. Thése presentée a 299– 310.
l’Université L. Pasteur pour obtenir le titre de Docteur-Ingénieur, Söfner, B. 1973. Observações sobre a estratigrafia do Pré-Cambriano
Strasbourg, France, 119. da Chapada Diamantina Sudeste e da área contı́gua. In: Sociedade
Macedo, M. H. F. & Bonhomme, M. G. 1984. Contribuição à cronoestra- Brasiliera de Geologia, Congresso Brasiliera, 27, Aracajú, Anais,
tigrafia das formações Caboclo, Bebedouro e Salitre na Chapada 1, 23 – 33.
Diamantina (BA) pelos métodos Rb –Sr e K –Ar. Revista Brasileira Teixeira, J. B. G., Misi, A. & Silva, M. G. 2007. Supercontinent evol-
de Geociências, 153– 163. ution and the Proterozoic metallogeny of South America. Gondwana
Mello, J. L. Jr. 1938. Geologia e Hidrologia do Noroeste da Bahia. Research, 11, 346– 361.
Serviço Geológico e Mineralógico (SGM), Boletim 90, Rio de Torquato, J. R. F. & Misi, A. 1977. Medidas isotópicas de carbono e
Janeiro. oxigênio em carbonatos do Grupo Bambuı́ na região centro-norte
Misi, A. 1978. Ciclos de sedimentaupoçao e mineralċzçŏes de Pb –Zn nas do Estado da Bahia. Revista Brasileira de Geociências, 7, 14 –24.
sequências Bambui (Supergrupo São Francisco), Estado da Bahia. Trindade, R. I. F., D’agrella Filho, M. S., Babinski, M., Font, E. &
XXX Congresso Brasileriro de Geologia, Sociedae Brasileira de Geo- Neves, B. B. B. 2004. Paleomagnetism and geochronology of the
logia, Anais, 4, 2548–2561. Bebedouro cap carbonate: evidence for continental-scale Cambrian
Misi, A. 1979. O Grupo Bambuı́ no Estado da Bahia. In: Inda, H. A. V. remagnetization in the São Francisco craton, Brazil. Precambrian
(ed.) Geologia e Recursos Minerais do Estado da Bahia. Textos Research, 128, 83– 103.
Básicos. Salvador, SME/COM, 1, 120– 154. Uhlein, A., Alvarenga, C. J. S., Dardenne, M. A. & Trompette, R. R.
Misi, A. & Souto, P. G. 1975. Controle estratigráfico das mineralizações 2011. The glaciogenic Jequitaı́ Formation, southeastern Brazil.
de Pb –Zn –F –Ba no Grupo Bambuı́, parte leste da Chapada de Irecê, In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The
BA. Revista Brasileira de Geociências, 5, 30 –45. Geological Record of Neoproterozoic Glaciations. Geological
Misi, A. & Kyle, J. R. 1994. Upper Proterozoic carbonate stratig- Society, London, Memoirs, 36, 541–546.
raphy, diagenesis and stromatolitic phosphorite formation, Irecê Vieira, L. C., Trindade, R. I. F., Nogueira, A. C. R. & Ader, M. 2007.
Basin Bahia, Brazil. Journal of Sedimentary Research, 64, 299– 310. Identification of a Sturtian cap carbonate in the Neoproterozoic Sete
Misi, A. & Veizer, J. 1998. Neoproterozoic carbonate sequences of the Lagoas carbonate platform, Bambui Group, Brazil. Comptes Rendus
Una Group, Irecê Basin, Brazil: Chemostratigraphy, age and corre- Geoscience, 339, 240–258.
lations. Precambrian Research, 89, 87– 100. Williams, H. E. 1930. Estudos geológicos na Chapada Diamantina,
Misi, A., Iyer, S. S. S. et al. 2005. Sediment-hosted lead-zinc Estado da Bahia. Boletim do Serviço Geológico e Mineralógico,
deposits of the Neoproterozoic Bambui Group and correlative Brasil, 44.
Chapter 48

Neoproterozoic successions of the São Francisco Craton, Brazil: the Bambuı́, Una,
Vazante and Vaza Barris/Miaba groups and their glaciogenic deposits

AROLDO MISI1*, ALAN J. KAUFMAN2, KAREM AZMY3, MARCEL AUGUSTE DARDENNE4,


ALCIDES NÓBREGA SIAL5 & TOLENTINO FLÁVIO DE OLIVEIRA6
1
Metallogenesis Group, Department of Geology and Research Center on Geophysics and Geology (CPGG),
Federal University of Bahia, Instituto de Geociências, Campus da Federação 40170-290, Salvador, Bahia, Brazil
2
Departments of Geology and ESSIC, University of Maryland, College Park, Maryland, USA
3
Department of Geology, Memorial University of Newfoundland, St. John’s, Canada
4
Institute of Geosciences, University of Brasilia, Brasilia (DF), Brazil
5
NEG-LABISE, Department of Geology, Federal University of Pernambuco, Recife (PE), 50670-000, Brazil
6
Votorantim Metais, P.O. Box 03, 38780-000, Vazante, MG, Brazil
*Corresponding author (e-mail: aroldo.misi@gmail.com)

Abstract: The Neoproterozoic successions of the São Francisco Craton are primarily represented by the Bambuı́ and Una groups,
deposited in cratonic epicontinental basins, and by the Vazante and Vaza Barris/Miaba groups, which accumulated on passive
margins on the edges of the craton. The epicontinental basins comprise three megasequences: glaciogenic, carbonate platform
(marine) and dominantly continental siliciclastics. Possible correlative sequences are observed in the passive margin deposits.
At least two major transgressive–regressive sea-level cycles occurred during the evolution of the carbonate megasequence, which
lies above glaciomarine diamictites of probable early Cryogenian (i.e. Sturtian) age. C, O, Sr and S isotope trends from analyses of
well-preserved samples, together with lithostratigraphic observations, provide reasonable correlations for most of the Neoproterozoic suc-
cessions of the São Francisco Craton. The 87Sr/86Sr record of these successions, ranging from 0.70769 to 0.70780, supports the proposed
correlation with the Bambuı́, Una and Vaza/Barris successions, and with the basal units of the Vazante Group. In addition, C-isotope posi-
tive excursions ranging from þ8.7 to þ14‰ and negative excursions from –5.7 to – 7‰ VPDB in the Bambuı́, Una and Vaza-Barris suc-
cessions provide key markers for correlations. The precise ages of the sedimentation in these successions remains a matter of debate, but
organic shales of two units of the Vazante Group have been dated by Re–Os techniques in two different laboratories, both yielding
Mesoproterozoic ages. The Neoproterozoic and Mesoproterozoic successions preserve significant glaciogenic deposits.

The global occurrences of low-latitude Neoproterozoic glacial With these caveats in mind, aspects of the litho- and
sediments and the texturally and isotopically anomalous post- sequence-stratigraphy, chemostratigraphy and radiometric ages
glacial carbonates deposited during potential ‘Snowball Earth’ of the most representative glaciogenic and carbonate-rich succes-
events (Hoffman et al. 1998a; Hyde et al. 2000; Hoffman & sions of the São Francisco Craton, and their possible global corre-
Schrag 2002) provide stratigraphic markers that have been used lation, are discussed in this chapter.
for inter-continental correlations. However, the timing and
duration of the Neoproterozoic ice ages are in most cases poorly
Structural and geotectonic framework
constrained, due to the lack of absolute ages from appropriate
lithologies. This is the case for the carbonate-dominated
Most of the glaciogenic successions of the São Francisco Craton
Vazante, Bambuı́ and Una groups (cf. Azmy et al. 2001, 2006;
were likely deposited during extensional events related to the frag-
Misi et al. 2007) on the São Francisco Craton in east central
mentation of the Rodinia supercontinent starting around 950 Ma
Brazil (Fig. 48.1), although new chronometric techniques are
and extending until c. 600 Ma (Condie 2002). These extensional
revealing surprising results (cf. Babinski et al. 2007).
events were episodic during the closure of the Pan-African-Brasi-
In the absence of radiometric or biostratigraphic tie-lines in
liano rift (Porada 1989; Brito-Neves et al. 1999; Condie 2002;
these Neoproterozoic successions, time-series trends in stable
Cordani et al. 2003). A relationship between Neoproterozoic gla-
isotope signatures have been used as long-distance correlation
ciation and extensional tectonics has been previously postulated
tools for the cap carbonates and other carbonate lithofacies
(e.g. Young 1995).
throughout the marine sequences (e.g. Knoll et al. 1986; Derry
These sedimentary successions deposited on the São Francisco
et al. 1989; Fairchild et al. 1990; Kaufman et al. 1991; Kaufman
Craton were distributed in the following geotectonic settings
& Knoll 1995; Jacobson & Kaufman 1999; Brasier & Shields
(Misi et al. 2007) (Fig. 48.1):
2000; Azmy et al. 2001; Cozzi et al. 2004, Halverson et al.
2005; Azmy et al. 2006). However, a priori assumptions of the † mixed carbonate and siliciclastic strata deposited on tectoni-
ages of some of the Neoproterozoic glacial deposits (i.e. Sturtian cally stable cratons, including the Bambuı́ and Una groups in
v. Marinoan) in the absence of direct radiometric constraints the São Francisco Basin and in the Irecê, Campinas, Una-
(Hoffman & Schrag 2002), and the likelihood of depositional Utinga and Ituaçú sub-basins, respectively;
(and diagenetic) variations in C- and Sr-isotope compositions † intensely deformed mixed carbonate and siliciclastic strata in
complicate the use of chemical stratigraphy for basin-to-basin passive-margin basins surrounding the stable cratons, including
comparisons. In addition, correlations need to take into account part of the Vazante group (Brası́lia Fold Belt) and the Vaza
the possibility of structural complications due to the tectonic Barris/Miaba group (Sergipano Fold Belt), among others sur-
setting during and after sedimentation. rounding the cratonic area.

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 509– 522. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.48
510 A. MISI ET AL.

Fig. 48.1. The São Francisco Craton in


NE Brazil with the Neoproterozoic cover
(São Francisco Supergroup) and the
folded belts surrounding the
cratonic area.

Some authors (Chang et al. 1988; Thomaz Filho et al. 1998; into three megasequences (glaciogenic, carbonate platform and
Dardenne 2001; Pimentel et al. 2001) have postulated that the Neo- dominantly continental siliciclastic units) separated by erosional
proterozoic sediments on the São Francisco Craton accumulated in unconformities (Misi 2001; Misi et al. 2007). Each megasequence
foreland basins during compressive events related to the Brasi- may have other smaller sequence boundaries that, when associated
liano–Pan African orogeny. On the other hand, D’el Rey Silva with high-resolution chemostratigraphic data, may provide key
(1999) demonstrated that sedimentation of the Vaza Baris/Miaba event markers for stratigraphic correlations. These megasequences
domains in the Sergipano Belt (NE Brazil) occurred on a passive are briefly described below (Fig. 48.2).
continental margin. The similarity of cyclic sedimentation in the
marginal and epicontinental marine basins suggests that sediments Glaciogenic megasequence
originally spanned the craton. Extensional tectonism associated
with the glacial deposits ‘is indicated by the presence of large poly- Siliciclastic rocks at the base of the successions are dominated
lithic clasts of both basement and cover in the diamictites . . . point- by conglomerate, metagreywacke, diamictite, pelite and quartzite.
ing to uplift-erosion of footwall and possibly hanging wall blocks These deposits are interpreted as having a continental glacial to
twice in the evolution of the basin’ (Dardenne 1978a, 1979, 1981; glacial –marine origin. Striated pavements and dropstones as
D’el Rey Silva 1999, p. 463). Alkmim (1996) and Misi (1999) well as faceted and striated clasts in most of the lithofacies have
demonstrated the existence of widespread aligned structures with been described (Söfner 1973; Karfunkel & Hoppe 1988; Montes
orientation 108N–208W intersecting both the older Proterozoic 1977; Guimarães 1996; Uhlein et al. 2004). These lithofacies
basement and the Neoproterozoic sedimentary cover (Fig. 48.1). unconformably overlie the Palaeoproterozoic/Archaean basement
complex and the Mesoproterozoic Espinhaço Supergroup. They
Stratigraphy constitute the Jequitaı́ Fm. in the Bambuı́ Group (states of Minas
Gerais, Bahia and Goiás), and the Bebedouro Fm. in the Una
The Neoproterozoic successions of the São Francisco Craton are Group (state of Bahia). In the passive margin basins, glacial dia-
part of the São Francisco Supergroup. They may be subdivided mictite is present in the Panelinha Fm. (Rio Pardo Group) and
NEOPROTEROZOIC SUCCESSIONS OF THE SÃO FRANCISCO CRATON 511

Fig. 48.2. Megasequences of the


Neoproterozoic successions of the São
Francisco Supergroup (modified from Misi
et al. 2007).

the Macaúbas Group, both of the Araçuaı́ Fold Belt, the Palestina (northern Bahia) and by part of the Vazante Group (western
Fm. in the Vaza Barris/Miaba Group of the Sergipano Fold Belt, Minas Gerais), include possibly equivalent strata, although the tec-
and the Santo Antonio do Bonito Fm. in the Vazante Group of tonic setting differs in each of these regions. The stratigraphic
the Brasilia Fold Belt (Fig. 48.2). thicknesses in these marginal basins are greater than in the intra-
In addition, Brody et al. (2004), Olcott et al. (2005) and Azmy cratonic basins, and parts were intensely deformed during the
et al. (2006) document the presence of abundant dropstones in the Brasiliano – Pan African orogeny.
black shale of the Serra do Poço Verde and Lapa formations higher At least two transgressive –regressive cycles are clearly rep-
up in the Vazante Group. The Serra do Poço Verde shale is sand- resented in the carbonate platform sediments deposited in the intra-
wiched between carbonate breccia, which may also be glacial in cratonic basins (Dardenne 1978b, 1979, 2000; Misi 1978; Misi
origin (Olcott et al. 2005), and contains glendonite, a carbonate et al. 2007). The first cycle ends with an extensive sub-aerial
mineral after ikaite that only forms in sediments at frigid tempera- exposure surface associated with tepee structures, replaced nodular
tures. Based on recent geochronological data (Re – Os in black anhydrite, dissolution breccias, and laminated and columnar stro-
shales, Azmy et al. 2008; Geboy et al. 2009), this possible glacio- matolites (Fig. 48.2).
genic event is considered by some authors as Mesoproterozoic in
age, as will be discussed later.

Table 48.1. Correlations between the Bambuı́ formations and the Una
Carbonate platform megasequence (marine) Group units

The carbonate platform facies are composed of limestone, dolo-


Group
stone, marl, shale and siltstone. Internal to the São Francisco Lithotypes BAMBUÍ UNA ,
craton, the Bambuı́ Group (Dardenne 1978b, 1979, 2000) is subdi- ,
(Formations)* † (Units) ( Form.)‡ §
vided into the following formations: Sete Lagoas (dolarenite with
teepee structures at the top, laminated limestones and red argillac- Arkose, siltstone Três Marias –
eous dolostone at the base), Serra de Santa Helena (marl and inter-
calated limestone), Lagoa do Jacaré (black, organic-rich oolitic Siltstone, pelite and intercalated limestone Serra da Saudade –

limestone) and Serra da Saudade (siltstone, pelites and intercalated


limestone). In the Irecê and Una-Utinga basins of the Una Group, Black organic -rich calcarenite and calcilutite Lagoa do Jacaré Unit A1
with oolitic and pisolitic limestones with
the dominant carbonate succession has been subdivided into five
Salitre Formation

interbedded pelite and marl


informal units that correlate with those of the Bambuı́ Group
(Misi & Souto 1975; Misi 1978; Misi & Veizer 1998): Unit C Grey marl, pelite and siltstone with black Serra de Santa Unit A
limestones Helena
(base of Sete Lagoas Fm.), Unit B (mid section of Sete Lagoas
Fm.), Unit B1 (top of Sete Lagoas Fm.), Unit A (Serra de Sta. Clear grey dolarenite w/stromatolites and Sete Lagoas 3 Unit B1
tepee structures (Upper)
Helena Fm.) and Unit A1 (Lagoa do Jacaré Fm.) See Table 48.1 Sete Lagoas 2 Unit B
Laminated limestone and shale (Middle)
for a brief description of these facies.
Red argillaceous dolostone (Lower) Sete Lagoas 1 Unit C
The passive margin successions bordering the craton, which are
represented by the Vaza Barris/Miaba Group (NE of the São
Diamictite, arkose, pelitic rocks Jequitaí Bebedouro
Francisco Craton, in Bahia and Sergipe states), the Rio Pardo
*Dardenne (1978), †Misi (1978), ‡Misi & Souto (1975) and §Branner (1911).
Group (Bahia), the Rio Preto and Riacho do Pontal groups
512 A. MISI ET AL.

Dominant continental siliciclastic megasequence (molasses) Vazante Group

Continental siliciclastic rocks composed of arkose, siltstone, phyl- The Vazante Group (Dardenne 2000, 2001) consists mainly of stro-
lite and conglomerate constitute the Três Marias Fm. of the matolitic carbonate muds and rhythmites deposited on a shallow
Bambuı́ Group. The origin of these deposits is related to post- marine platform (Dardenne 2001) that extended more than
tectonic molassic sedimentation along the borders of the cratonic 300 km north–south along the western margin of the craton. The
area (Brito Neves & Cordani 1991). Whereas Brito Neves & succession now lies in the external zone of the Brası́lia Fold Belt
Cordani (1991) claimed that the contact between the Três Marias (Fig. 48.3) in the São Francisco Basin (SFB). The stratigraphy of
Fm. and the Bambuı́ Group is an erosional unconformity, others the marginal marine sediments of the Vazante Group (Fig. 48.4)
have postulated a gradational contact between the Três Marias has been studied in detail and refined by several authors (e.g. Dard-
Fm. and the Serra da Saudade Fm. (Dardenne 1978b, 1979, enne 1978, 1979, 2000, 2001; Madalosso 1979; Karfunkel & Hoppe
2000; Martins-Neto, pers. comm; cf. Misi 2001). 1988; Fairchild et al. 1996; Azmy et al. 2001; Misi 2001; Misi et al.
2007). The Vazante sequence, from bottom to top, includes the
Santo Antonio do Bonito, Rocinha, Lagamar, Serra do Garrote,
Characteristics and distribution of the glaciogenic Serra do Poço Verde, Serra do Calcario and Serra da Lapa-
Velosinho formations (Fig. 48.4). In the eastern part of the basin,
deposits and associated strata carbonate, diamictite and shale of the Vazante Group are generally
well preserved and little metamorphosed. To the west near the Bra-
The Neoproterozoic glaciogenic deposits of the São Francico Craton silia Fold Belt, however, the sediments are highly deformed and
are widespread and found in the basal sections of the epicontinental have experienced amphibolite to granulite facies metamorphism
basins of the Bambuı́ Group (São Francisco Basin; see also Uhlein (Dardenne 1978a; Fuck et al. 1994). Earlier studies indicated that
et al. 2011) and the Una Group (Irecê, Una-Utinga and Campinas the Vazante Group sediments accumulated in a passive margin
sub-basins; see also Guimarães et al. 2011) and in the basal and setting (e.g. Dardenne 1979; Campos Neto 1984a, b; Fuck et al.
middle parts of the passive margin basin of the Vaza Barris/ 1994), which is consistent with the lack of volcanic input to the
Miaba Group (Sergipano Fold Belt). The glaciogenic deposits of sequence. The northern and southern segments were separated by
the Vazante Group (Brasilia Fold Belt) occur in the basal and a regional WSW–ESE lineament situated on the same latitude of
upper parts of the sequence, but their ages are still a matter of the Federal District (DF, Fig. 48.1). They show distinct geotectonic
debate, as will be discussed in subsequent sections. Figure 48.1 characteristics: intensely deformed in the southern segments with
shows the distribution of these basins in the studied area. tectonic transport of great magnitude (from west to east), while in
the northern segment the sedimentary units are less deformed and
the stratigraphic organization remains well preserved, probably
Bambuı́ and Una Groups due to the location of this segment at an upper crustal level of
the granite-gneiss basement, ‘which acted as a rigid block in front
The Bambuı́ Group in the São Francisco Basin and the Una Group of the compressive trend of the Brasiliano Cycle’ (Dardenne
in the Irecê, Una-Utinga and Campinas sub-basins (Fig. 48.1) were 2000, p. 234).
deposited in epicontinental, shallow-marine environments. Recent geochronological studies suggest that at least part of
Remarkable lithostratigraphic and chemostratigraphic similarities what is referred to as the Vazante Group may predate the Bambuı́
support detailed correlation between these isolated sub-basins, Group (Geboy 2006; Rodrigues 2007; Azmy et al. 2008; Rodri-
despite their present discontinuity. gues et al. 2008).
Deposition in this likely once contiguous basin began with The Vazante Group rests on a glaciogenic unit (D), which con-
dominantly glacial and glaciomarine sedimentation in the Jequitaı́ stitutes the uppermost part of the Santo Antonio do Bonito Fm.
Fm. (Bambuı́ Group) and the Bebedouro Fm. (Una Group), as (Fig. 48.4). This diamictite contains faceted and striated cobbles
represented by glacial diamictites, which contain faceted and and is overlain by the Cubatão Fm., a thick stromatolitic dolostone.
striated clasts, as well as dropstones in finely laminated pelites. New observations of dropstones in organic-rich shale sand-
In some regions, this diamictite contains basement clasts in a wiched between previously recognized dolomite breccia in the
fine-grained Fe-rich mudstone, and in the Sete Lagoas region, Serra do Poço Verde Fm. suggest the presence of a second glacial
it is overlain by a carbonate lithofacies containing centimetre- horizon (D II) near the top of the succession (Olcott et al. 2005;
scale sea-floor precipitates. Discontinuous aeolian sandstone with Azmy et al. 2006). The carbonate diamictites were found in
variable thickness occurs above the Bebedouro Fm. diamictite exploration drill holes in a known geographic extension of over
(Guimarães 1996). These units were deposited unconformably 150 km in an approximate north–south direction, between the
over the Palaeoproterozoic and Archaean basement, as well as cities of Unai and Lagamar. In most of the area they are associated
the Mesoproterozoic sedimentary rocks of the Espinhaço Super- with black shales and debris flows with fist-sized dropstones. The
group (Fig. 48.1). preserved thickness of this unit ranges from 20 m to .100 m, and
The carbonate platform marine sedimentation that follows was dolostone clasts vary from from millimetres to several metres
deposited unconformably on the glaciogenic and glaciomarine in diameter.
lithotypes. The lithostratigraphic units from the top to bottom This unit can be broken down into two distinct levels: one in the
include formations with similar characteristics in the Bambuı́ upper Morro do Calcário Formation, frequently in unconformable
and Una groups, as noted by Dardenne (1978b, 1979, 2000) and contact with the Lapa Fm., and a lower one in the Lower Poço
Misi et al. (2007) (Table 48.1). Verde Fm. (DNPM hole 1-PSB01 from 275.50 to 304.50 m).
The thickness of the glaciogenic megasequence ranges from 0 to Thin section observations revealed the presence of the unusual
150 m, while the thickness of the carbonate-rich interval is variable, cold-water carbonate ‘glendonite’ in some of the shale samples
probably controlled by a basement fault system active during sedi- (Olcott et al. 2005), which may support the syn-glacial interpret-
mentation (Dardenne 1978, 1979, 2000; Misi et al. 2007). Seismic ation of this unit. A sharp drop in d13C values is recorded in
surveys from the central part of the São Francisco Basin indicate bedded dolomite above the shale horizon in one core (Azmy
that the carbonate platform is up to 1000 m thick (Teixeira et al. et al. 2001), and possible cap-carbonate lithofacies (red dolo-
1993). In contrast, it reaches only 600 m in thickness in the Sete stones) have been recognized in field exposures near Vazante
Lagoas area (Pedrosa Soares et al. 1994), 400 m in the Serra do mine. A regional unconformity separating the Morro do Calcário
Ramalho and Januaria-Itacarambi areas (Dardenne 1978b, 1979, Fm. from the overlying Lapa Fm. occurs at the top of the upper
2000; Misi 1979), and 600 m in the Irecê Basin (Misi 1979). diamictite (D II) (Misi et al. 2005).
NEOPROTEROZOIC SUCCESSIONS OF THE SÃO FRANCISCO CRATON 513

Fig. 48.3. Simplified regional geological


map of the Brasilia Belt showing the
distribution of the Vazante Group, based on
Marini et al. (1984), Dardenne (2000),
Pimentel et al. (2001) and Valeriano et al.
(2004).

Coupled with a negative d13C excursion in overlying carbonate represented by a lower siliciclastic megasequence (Juetê, Itabaiana
and marl, Brody et al. (2004) and Azmy et al. (2006) interpret basal and Ribeirópolis formations) and an upper carbonate megase-
Lapa sediments as a post-glacial cap-carbonate lithofacies. Alter- quence (Acauã and Jacoca formations). According to D’el
natively, the Morro do Calcário interval of carbonate breccia has Rey Silva (1999), ‘the lateral correlation between the formations
been interpreted as debris-flow deposit along the western flank of each megasequence in Cycle 1 is constrained by their same
of a stromatolitic platform, and thus may be intra-formational in stratigraphic position, by the strong similarity of the rock types,
origin and not directly related to glaciation (Dardenne 1979, and by the fact that the Simão Dias Group forms a blanket of sili-
2000, 2006, 2007). ciclastics that is continuous across the craton-fold belt boundary’
(Fig. 48.6). Cycle 2 is similarly composed of a lower siliciclastic
interval, including the Simão Dias Group and the Palestina Fm.
Vaza Barris/Miaba Group and an upper carbonate interval, represented by the Olhos
d’Água Fm.
The Sergipano belt was divided by D’el Rey Silva (1999) into three Carbonates of the Jacoca Fm., at the basal section of the Vaza
tectonic domains: the cratonic Estância domain, the external Vaza Barris/Miaba Group, are in sharp contact with the Ribeiropolis
Barris domain and the intermediate Macururé domain. These diamictite below. This unit is 0– 300 m thick and locally reaches
domains form a continuous wedge of craton, platform and basinal thicknesses of c. 500 m. It consists of diamictite with dropstones,
deposits developed along a continental margin (Fig. 48.5). The metagreywacke and quartz-sericite phyllite overlying quartzites
intermediate Macururé domain (max. thickness, 13 km) is com- of the Itabaiana Fm. D’el Rey Silva (1995, 1999) has reported
posed of carbonate and siliciclastic sedimentary rocks with associ- volcanics within this formation. A typical section of the Jacoca
ated metavolcanics. The Vaza Barris domain (1– 4 km thick) Fm. starts with a thick layer of laminated grey to pink dolostone
includes marine platform siliciclastic and carbonate sediments containing pyrite and chalcopyrite, followed up-section by a 3–
with minor volcanic horizons, comprising the Miaba, Simão 15-m-thick layer of laminated dolostone, and dark grey to black
Dias and Vaza Barris groups. The Estancia domain (1 –3 km phyllites, followed by c. 10-m-thick massive dolostone (Sial
thick) is composed of continental to shallow marine sediments. et al. 2009). This unit is overlain by a 40-m-thick heterolithic
D’el Rey Silva (1999) described two sedimentary cycles in these sequence of grey limestones and dark to black phyllites and
domains, each including two megasequences: Cycle 1 is finally by grey dolostones.
514 A. MISI ET AL.

(Bambuı́ Group) formations on the São Francisco craton

Formation
(Tables 48.1 & 48.2).

(m ne x.
Group

) ss
ic pro
Member

Th Ap
k
Chemostratigraphy

High-resolution chemostratigraphic studies have been conducted


Serra da Lapa Velosinho
Lapa

650 in the Irecê Sub-basin (Fig. 48.1), Una Group (Torquato & Misi
organic-rich shales 1977; Misi & Veizer 1998), in the São Francisco Basin, Serra do
D II Ramalho area (Powis 2006; Misi et al. 2007), in the Brasilia
Fold Belt, Vazante Group (Azmy et al. 2001, 2006) and in the Ser-
Morro do
Calcário

gipano Fold Belt (Sial et al. 2000, 2003, 2005, 2006a, b). Kawa-
Upper Pamplona shita (1998) analysed Sr, C and O isotopes in the carbonate
300
sections at a lower stratigraphic resolution. The following review
is mostly based on the recent synthesis by Misi et al. (2007).

Irecê Basin
Rocinha Lagamar Serra do Garrote Serra do Poço Verde

Middle Pamplona 400


One of the earliest isotopic studies of a post-glacial cap carbonate
was performed in the Irecê Basin (Una Group) by Torquato & Misi
Lower Pamplona 200 (1977), although at the time the unit was considered lacustrine in
V A Z A N T E

origin. These authors analysed 85 carbonate samples for C and


Upper Morro do Pinheiro 500 O isotopes through two complete sections of the Una Group. The
results reveal a d13C trend from consistently negative values in
Lower Morro do Pinheiro 500 the red argillaceous dolostone (Unit C, c. –6‰) immediately
above diamictite of the Bebedouro Fm. to values c. 0‰ up-section
(Units B and B1) and continuing to highly positive values (þ8‰,
Unit A1). The negative values in the dolostone overlying the Bebe-
douro Fm. diamictite are similar to post-glacial trends worldwide
and may support the glacial interpretation for these deposits.
>1000 This trend has been reproduced by Misi & Veizer (1998), who ana-
lysed well-preserved micro-samples that were first characterized
by petrographic and geochemical techniques. In addition, these
authors determined 87Sr/86Sr ratios on the best-preserved carbon-
ates. The least radiogenic 87Sr/86Sr value of 0.70780 from Unit B
Sumidouro is consistent with a sedimentation age between 750 and 600 Ma
250 (cf. Jacobsen & Kaufman 1999) (Fig. 48.7).
Arrependido
São Francisco Basin
1000
The carbonates of the Bambuı́ Group in the São Francisco Basin
have been the subject of many chemostratigraphic studies (e.g.
arkosic sandstone Iyer et al. 1995; Kiang 1997; Kawashita 1998; Santos et al.
St. Antonio
do Bonito

D 2000, 2004; Kaufman et al. 2001). The most detailed chemostrati-


250 graphic study developed in the Bambuı́ Group thus far comes from
high-resolution profiles from the Serra do Ramalho area, western
Bahia (Powis 2006; Misi et al. 2007). Chemostratigraphic
studies have also been conducted by Iyer et al. (1995), Kaufman
Fig. 48.4. Stratigraphic units of the Vazante Group. et al. (2001) and Vieira et al. (2007) in the Sete Lagoas area of
the São Francisco Basin. As in the Irecê Basin, the d13C compo-
sition of carbonate rocks in the Serra do Ramalho and in the Sete
The Olhos d’Água Fm. (200 – 1300 m) is composed of inter- Lagoas area increase up-section from c. –4‰ at the base of the
bedded limestones that become organic-rich towards the top. Sete Lagoas Fm. overlying glacial diamictite of the Jequitaı́ Fm.
The upper section consists of c. 50 m of limestone –pelite inter- (assumed to be correlative to the Bebedouro Fm.) and rising stea-
calations that are overlain by laminated limestones (40 m thick) dily up to c. þ14‰ near the top of the succession in Lagoa do
culminating with a 30-m-thick layer of organic-rich limestone. Jacaré Fm. The best-preserved samples (Mn/Sr  0.02) show a
The lower section of laminated limestones is in sharp contact narrow range of 87Sr/86Sr values between 0.7074 and 0.7075
with the diamictites of the Palestina Fm., which is composed of (Fig. 48.7).
metagreywackes with clasts up to 30 cm in diameter. The clasts
are of gneiss, granite, quartz, quartzite, phyllite and metacarbonate
supported in a quartz –sericite matrix. Lenses of Fe-cemented Brasilia Fold Belt (Vazante Group)
quartzite unconformably overlie the siliciclastic sediments of the
Simão Dias Group (Fig. 48.6). Petrographic investigations showed that some of the Vazante
The Palestina Fm. is interpreted as a glaciogenic deposit. At a carbonates were affected by dolomitization, but examination of
lower stratigraphic level, the equivalent Juetê and Ribeirópolis for- thin sections shows that these rocks have very fine-grained and
mations (D’el Rey Silva 1999) are also interpreted as glaciogenic fabric-retentive dolomicrites (Azmy et al. 2001, 2006). Three gen-
and may correlate with the Bebedouro (Una Group) and Jequitaı́ erations of secondary dolomitic cements have been observed,
NEOPROTEROZOIC SUCCESSIONS OF THE SÃO FRANCISCO CRATON 515

Fig. 48.5. Simplified geological map of


the Sergipano Belt, modified from D’el
Rey Silva (1999). Basement domes:
1, Itabaiana; 2, Simão Dias; 3, Jirau do
Ponciano. Main thrust and strike–slip
fault: MF, Macururé fault; BMJF, Belo
Monte-Jeremoabo fault; SMAF, São
Miguel do Aleixo fault; IF, Itaporangã
fault.

including fibrous, equant and late fracture-filling phases (cf. Azmy Sergipano Fold Belt (Vaza Barris Group)
et al. 2001). No significant increase in crystal size was observed
associated with dolomitization, suggesting that the original sedi- Chemostratigraphic studies have been performed on some carbon-
ment did not suffer from extensive and/or repeated meteoric ate units of the Vaza Barris Group by Kawashita (1998) and by Sial
alteration. et al. (2000, 2003, 2005, 2006a, b). The first author determined
The trace element compositions of the Vazante carbonates, least-altered 87Sr/86Sr ratios of 0.7078 in black limestones of the
particularly those of the fabric retentive micrites, suggest signifi- Olhos d’Água Fm. In addition Sial et al. (2000) report d13C low
cant geochemical preservation (Azmy et al. 2001, 2006). The values of – 4.7‰ immediately above diamictite of the Palestina
C-isotope profiles reveal a negative shift up to 8‰ in dolomite Fm. These numbers increase up-section to values as high as
beds above the dropstone-laden shale of the Serra do Poço þ10‰ in the upper Olhos d’Água Fm. Sial et al. (2006a) also
Verde Fm. (Fig. 48.8) and near the base of the Lapa Fm. report negative excursions as low as – 5% in the lower carbonates
(cf. Azmy et al. 2006; Misi et al. 2007). The least radiogenic of the Acauã Fm. (equivalent to the Jacoca Fm.; Fig. 48.9) immedi-
Sr-isotope values associated with the negative d13C shifts are ately above the diamictites of the Ribeirópolis Fm., suggesting that
0.7061 in the Serra do Poço Verde Fm. and 0.7068 in the Lapa at least two important glaciogenic events are represented in the
Fm. (Fig. 48.8). Vaza Barris Group. The results support a possible correlation of

Fig. 48.6. Summary stratigraphy of the


Itabaiana dome area, based on D’el Rey
Silva (1999). Megasequences: LSM, lower
siliciclastic; LCM, lower carbonate
(megasequences of Cycle I); USM, upper
siliciclastic; UCM, upper carbonate
(megasequences of Cycle II). Formations:
JU, Juetê A, Acauã; I, Itabaiana; R,
Ribeirópolis; JC, Jacoca; LP, Laegarto-
Palmares; Ja, Jacaré; FP, Frei Paulo; P,
Palestina; OA, Olthos d’Água Formation.
516 A. MISI ET AL.

Table 48.2. Correlations between intracratonic and passive-margin basins Geochronological constraints
Brasiliano Fold Belt Intracratonic Basins Sergipano Fold Belt
Passive-margin Basin Passive-margin Basin Despite the absence of reliable absolute geochronological age
Vazante group Bambuí group Una group Vaza barrís/miaba gr.
estimates in the platform carbonate successions of the São Fran-
(Formations) (Formations) (Units) (Form.) (Formations) cisco craton, there is a general agreement in considering a
Três Marias – –
middle Cryogenian age for the Bambuı́ and Una groups. The
Mesoproterozoic Units

Serra da Lapa
least radiogenic Sr isotope values of 0.7074– 0.7077 and S (from
(overthrusted terrain)

Morro do Calcário Serra sa Saudade – – stratform barite), with a mean value of c. þ30 per mil CDT in
Serra do Poço Verde Lagoa do Jacaré A1 the Una Group, indicate that sedimentation likely occurred

Salitre Formation
Olhos d’Água
Serra de Santa Helena A between c. 780 Ma and 600 Ma (Misi & Veizer 1998; Misi et al.
Serra da Garrote
Sete Lagoas 3 B1 Palestina 1999). More recently, Babinski & Kaufman (2003) and Babinski
Simão Dias Group
Lagamar
Sete Lagoas 2 B et al. (2007) determined new Pb/Pb carbonate ages from well-
Jacoca
Rocinha Sete Lagoas 1 C preserved post-glacial sea-floor cements in the Sete Lagoas
Santo Antônio do Bonito Jequitaí Bebedouro Ribeirópolis Fm. of the Bambuı́ Group, suggest a depositional age of
740 + 22 Ma (MSDW ¼ 0.66). Notably, almost the same age
(730 Ma, U –Pb in zircon) has been recorded by Brito-Neves
(pers comm.) in acidic tuffs intercalated in the Ribeirópolis
the Vaza Barris Group with the carbonate platform successions of diamictites of the Vaza Barris/Miaba Group, in the Sergipano
the Bambuı́ and Una Groups, as sustained by some authors (e.g. Belt. According to D’el Rey Silva (1999), sedimentation in the Ser-
Misi 1979; Teixeira & Figueiredo 1991; Trompette 1994; D’el gipano Belt lasted from c. 1.0 Ga (U –Pb zircon in metavolcanics,
Rey Silva 1999). The diamictites of the Ribeirópolis Fm. may be Brito Neves et al. 1993) to possibly 0.65 Ga.
equivalents of the Bebedouro and the Jequitaı́ formations (respect- On the other hand, the global stratigraphic correlation of the
ively, from the Una and Bambuı́ groups), but no equivalent glacial Vazante Group in the São Francisco Basin has been a challenging
deposits correlating with the Palestina Fm. have been clearly issue. The resetting of the radiogenic clock by the major
demonstrated from these epicontinental successions (Fig. 48.9 & Brası́lian –Pan African orogeny (c. 550– 600 Ma) (Trindade
Table 48.2). et al. 2004) and the scarcity of volcanic ash layers makes absolute

Fig. 48.7. Proposed lithostratigraphic and


chemostratigraphic correlation between the
Bambuı́ and Una Groups, respectively in the
São Francisco Basin (Serra do Ramalho
area) and in the Irecê Basin (Misi & Veizer
1998; Powis 2006; Misi et al. 2007).
NEOPROTEROZOIC SUCCESSIONS OF THE SÃO FRANCISCO CRATON 517

the basal portion of the Lagamar Fm. with a maximum age of


1.95 Ga. consistent with the c. 1.35 Ga Re – Os ages obtained by
Geboy (2006) from organic shales of the same unit.

Mineralization

Mineral deposits associated with Late Precambrian glacial and


carbonate sediments of the São Francisco craton include the
following:
† Fe – Mn deposits near Porteirinha (Minas Gerais state), which
are intimately associated with glacial sedimentation of the
Macaúbas Group in the Araçuaı́ belt, and likely related to exha-
lative processes (Dardenne & Schobbenhaus 2001).
† Phosphorite deposits interbedded with carbonates in the lower
Vazante Group at Coromandel, Rocinha and Lagamar (Dard-
enne et al. 1986, 1997; Da Rocha-Araújo et al. 1992; Sanches
et al. 2007), at the base of the Sete Lagoas Fm. in the
Bambuı́ Group at Campos Belos (Dardenne et al. 1986), in
the lower half of the Serra da Saudade Fm. in the Bambuı́
Group at Cedro do Abaeté (Dardenne et al. 1986; Lima et al.
2005), and in stromatolitic beds of the Salitre Fm. in the Irecê
basin (Misi & Kyle 1994).
† Fluorite, barite and Pb –Zn deposits associated with the upper
part of the first carbonate megacycle of the Bambuı́ Group
near Januária-Itacarambi, Montalvânia and Serra do Ramalho
(Dardenne 1979; Misi 1979; Dardenne & Freitas-Silva 1999).
† Pb– Zn sulphide deposits associated with brecciated dolomite
in the middle Vazante Group at Morro Agudo mine (Hitzman
et al. 1995; Misi et al. 1998, 1999; Dardenne & Freitas-Silva
1999; Bettencourt et al. 2001; Monteiro 2002).
† Zn silicate deposits associated with major fault structures at the
Vazante mine, Minas Gerais (Dardenne & Freitas-Silva 1999;
Bettencourt et al. 2001; Monteiro 2002; Misi et al. 2005).
Fig. 48.8. Chemostratigraphy of the Vazante Group.

geochronological age control of this succession problematic. The Discussion


abundance of Conophyton metula Kirichenko in the Lagamar
Fm. (Fig. 48.4) suggests an age estimate between 1350 Ma and Broadly speaking, the Neoproterozoic Era experienced at least three
950 Ma for the sedimentation of this part of the Vazante Group widespread ice ages, commonly referred to as the Gaskiers, Mar-
(Cloud & Dardenne 1973). Nevertheless, Misi et al. (1997) and inoan and Sturtian events, which are radiometrically constrained
Sanches et al. (2007) analysed carbonate fluorapatite and fine- to be c. 585 Ma (Newfoundland: Bowring et al. 2003; Halverson
grained organic-rich limestone (micrite) from the lower Rocinha et al. 2005), c. 630 Ma (South China: Condon et al. 2005; Zhang
Formation for their Sr-isotope composition. 87Sr/86Sr values et al. 2005) and c. 750 Ma (Namibia: Hoffman et al. 1996), respect-
range from 0.70760 to 0.70790, similar to those reported for the ively. Some researchers suggest that these singular glacial events
Bambuı́, Una and Vaza Barris/Miaba Groups. All seven samples may have been diachronous, spanning over tens of millions of
had high total Sr (.1300 ppm) and low Mn– Sr (0.1 –0.01), indi- years, or alternatively split into several discrete ice ages (e.g.
cating a high degree of preservation of the original seawater signal. Kaufman et al. 1997; Kendall et al. 2006). In most cases, the
Sm –Nd studies of provenance (Pimentel et al. 2001) on fine- glacial diamictites are overlain by texturally and isotopically enig-
grained sediments of the Bambuı́, Vazante and Paranoá groups matic cap carbonates with strong negative C-isotope anomalies
have furnished TDM model age intervals of 1.9 –1.4 Ga, 2.3– (e.g. Kaufman et al. 1997; Hoffman et al. 1998b), which have
1.7 Ga and 2.3 –1.9 Ga for the respective groups. According to also been recorded in syn- and immediately pre-glacial carbonates
these authors, the provenance results suggest that the Vazante (e.g. Kaufman et al. 1997; Kennedy et al. 2001; Halverson et al.
Group constitutes a transitional sequence between the Paranoá 2002). In contrast, S-isotope compositions in the cap carbonates pre-
and Bambuı́ groups. serve remarkably positive d34S anomalies (e.g. Hurtgen et al. 2002).
Recent Re – Os investigations of organic-rich shales of the upper In all the studied Neoproterozoic successions from the São Fran-
Vazante Group (Geboy 2006; Azmy et al. 2008) suggest a signifi- cisco craton, at least one horizon of glacial diamictite (generally
cantly older, Mesoproterozoic age for the Morro do Calcário and recognized on the basis of faceted and striated clasts of hetero-
Serra do Garrote units, with ages of c. 1 Ga and older. However, geneous composition in a fine-grained matrix) has been recog-
U –Pb determinations on detrital zircons (Dardenne et al. 2003, nized. Most often, these glacial horizons occur at the base of the
2005; Dardenne 2007a; Rodrigues 2008; Rodrigues et al. 2008) succession. These ice age deposits are overlain by cap carbonates,
indicated maximum depositional ages for the Santo Antonio do although in some cases a thin siliciclastic wedge, perhaps record-
Bonito Fm. (diamictites at the basal section of the Vazante ing glacial outwash facies during meltback (Bailey & Peters 1998),
Group, Fig. 48.4) of 988 + 15 Ma. Rodrigues et al. (2008) and separates the diamictite from carbonate lithofacies. Chemostrati-
Rodrigues (2008) also report a significant Neoproterozoic zircon graphic studies (including C, S and Sr isotopes) and Pb/Pb
population from the Rocinha Fm. in which the youngest grains dating of carbonates from these glaciogenic successions provide
have been dated c. 925 Ma. Nonetheless, in the same study the tools for possible correlations with Neoproterozoic counterparts
authors identified provenance patterns for the conglomerates at around the world.
518 A. MISI ET AL.

Fig. 48.9. Chemostratigraphy of the Vaza


Barris/Miaba Groups (modified from Sial
et al. 2000).

Perhaps of greatest interest are the correlation of Bambuı́, diamictite with the Otavi Group is missing and we regard it
Una, Vazante and Vaza Barris/Miaba group deposits on the São is possible that these glacial deposits instead correlate with
Francisco Craton with Neoproterozoic equivalents on the Congo possibly younger glacial deposits of the Witvlei Group on the
Craton in Namibia. If we accept that throughout much of the Kalahari Craton in southern Namibia (Kaufman et al. 1997;
Mesoproterozoic and Neoproterozoic eras these cratons were Saylor et al. 1998).
joined, the broadly equivalent sedimentary successions should On the western margin of the Sao Francisco Craton in Brazil,
reveal striking similarities. However, whereas the well-studied the basal glacial deposit of the Vazante Group (Santo Antonio
Otavi Group, in Namibia, hosts at least two cap carbonates atop do Bonito Fm.) has long been considered as an equivalent to the
glacial sediments (Kaufman et al. 1991; Hoffmann & Prave basal diamictite of the Bambuı́ Group beneath the Sete Lagoas
1996; Hoffman et al. 1998b), only one ice age deposit is currently Fm. Overlying strata in both successions are variably mineralized
recognized in the Bambuı́ and equivalent Una groups. Diamictites (phosphate and Pb– Zn) and contain thick carbonate intervals.
from these units containing basement clasts in a fine-grained However, C-isotope compositions of the carbonates differ signifi-
Fe-rich mudstone and overlained by a cap-carbonate lithofacies cantly. The d13C profile for Vazante carbonates is generally near
with centimetre-scale sea-floor precipitates, are strikingly similar zero, and shows only minor excursions to negative values that
to those of the younger Maieberg Fm. cap carbonate in Namibia. are interpreted to reflect cap-carbonate lithofacies (Azmy et al.
The Sete Lagoas precipitates are preserved in limestone with nega- 2001, 2006). In contrast, the strongly negative d13C values of the
tive d13C values and high Sr abundances. Sr-isotope compositions Sete Lagoas cap carbonate are followed by remarkably positive
of these lithologies are consistently around 0.7073, which is an values (upwards to þ16‰) in overlying strata (Misi et al. 2007).
exact match with the precipitate interval in the Maieberg Fm., sup- Thus the equivalence of these successions is suspect, although
porting this intracontinental correlation. Notably, Pb/Pb dating of Sr-isotope compositions of lower Vazante phosphorites are quite
the Sete Lagoas precipitates yielded an age of 740 + 22 Ma similar to those in well-preserved limestones of the Bambuı́ and
(11-point isochron; MSWD ¼ 0.66; Babinski et al. 2007). Accep- Una groups (Misi & Veizer 1998). Further complicating the inter-
tance of this date and the correlation with the precipitates in the basinal correlation is the presence of a second Vazante Group dia-
Maieberg Fm. associates both of these post-glacial deposits with mictite (D II) in the Serra do Poço Verde Fm. (Fig. 48.2: Olcott
the Sturtian ice age. This assignment is at odds with the view of et al. 2005; Azmy et al. 2006), and the possibility that the over-
the Maieberg Fm. as the archetypical Marinoan cap carbonate lying Lapa Fm. was deposited during oceanic transgression after
(Hoffman & Schrag 2002), and the c. 630 Ma age assigned to a third possible glacial event (Brody et al. 2004; Azmy et al. 2006).
the underlying Ghaub diamictite by the correlation of this unit It has been suggested previously that the shale and marly car-
with an ash-bearing diamictite in the deformed Swakop Group of bonate of the Lapa Fm. may be equivalent to the Rasthof Fm. in
central Namibia (Hoffmann et al. 2004). Independent chemostrati- northern Namibia (the older of the two Otavi Group cap carbon-
graphic evidence for the equivalence of the dated Swakop Group ates). This assignment, however, was based on comparable
NEOPROTEROZOIC SUCCESSIONS OF THE SÃO FRANCISCO CRATON 519

negative d13C excursions in both units, and on a single Sr-isotope Babinski, M. & Kaufman, A. J. 2003. First direct dating of a Neoproter-
value (c. 0.7068) from a dolomite cement in the basal Lapa For- ozoic post-glaciogenic cap carbonate. IV South American Symposium
mation (Azmy et al. 2001, 2006) that matches with the lowest on Isotope Geology, Short Papers 1, 321– 323.
values recorded in Sr-rich Rasthof limestones. However, the Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating of
Sr-isotope equivalence of these units is problematic and, further- the Sete Lagoas cap carbonate (Bambuı́ Group, Brazil) and impli-
more, new Re –Os age constraints for shale of the Lapa and cations for the Neoproteozoic glacial events. Terra Nova, 19,
older organic-rich lithologies in the Vazante Group suggest that 401– 406.
this succession may be far older than previously considered. Bailey, C. M. & Peters, S. E. 1998. Glacially influenced sedimentation in
the late Neoproterozoic Mechum River Formation, Blue Ridge pro-
Re –Os age determinations on Lapa Fm. shale suggest an age of
vince, Virginia. Geology, 26, 623– 626.
close to one billion years (Azmy et al. 2008), and similar measure- Bettencourt, J. S., Monteiro, L. V. S., Bello, R. M. S., Oliveira, T.
ments of organic-rich shale in the underlying Serra do Poço Verde F. & Juliani, C. 2001. Metalogênese do Zinco e chumbo na
and Serra do Garrote formations yield ages of c. 1.1 and 1.3 Ga região de Vazante-Paracatu, Minas Gerais. In: Pinto, C. P. &
(Geboy 2006; Geboy et al. 2006, 2009). A Mesoproterozoic Martins-Neto, M. A. (eds) Bacia do São Francisco: Geologia e
age for this succession is consistent with the relative invariance recursos minerais. SBG-Núcleo de Minas Gerais, Belo Horizonte,
of C-isotope trends in the Vazante Group, which are comparable 161– 198.
with the Paranoá Group stratigraphically beneath Bambuı́ Bowring, S., Myrow, P., Landing, E., Ramezani, J. & Grotzinger, J.
sediments in central and eastern Brazil (Santos et al. 2000). 2003. Geochronological constraints on terminal Proterozoic events
This older age is similarly suggested from Nd-isotope studies of and the rise of metazoans. Geophysical Research Abstracts (EGS,
detrital grains in the Vazante Group (Pimentel et al. 2001) and Nice), 5, 219.
the appearance of Conophyton metula Kirichenko stromatolites Brasier, M. D. & Shields, G. 2000. Neoproterozoic chemostratigraphy
(Cloud & Dardenne 1973), which are also known from the and correlation of the Port Askaig glaciation, Dalradian Supergroup
Paranoá Group. of Scotland. Journal of the Geological Society of London, 157,
The Re –Os ages provided by two different laboratories would 909– 914.
push the known record of Late Precambrian glaciation into the Brito Neves, B. B., Van Schmus, W. R., Babinski, M. & Sabin, T. 1993.
Mesoproterozoic Era, some 500 million years earlier than pre- O evento de magmatismo de 1,0 Ga nas faixas mó veis ao norte do
viously considered. Further tests of the Re – Os ages, as well as Craáton São Francisco. Sociedade Brasileira de Geologia, Simpósio
time-series elemental, isotopic and biomarker studies (cf. Olcott o Cráton do São Francisco 2, Salvador, Anais, 243– 245.
et al. 2005) of the Vazante Group, which are currently in progress, Brito-Neves, B. B., Campos-Neto, M. C. & Fuck, R. A. 1999.
From Rodinia to Western Gondwana: an approach to the Brasiliano–
should shed light on the depositional environment of the glacio-
Pan African Cycle and orogenic collage. Episodes, 22, 155– 166.
genic Vazante Group and provide better resolution of these contro- Brody, K. B., Kaufman, A. J., Eigenbrode, J .L. & Cody, G. D. 2004.
versial age estimates. Biomarker geochemistry of a post-glacial Neoproterozoic succession
in Brazil (Abstract), Geological Society of America, Abstracts with
During the last ten years CNPq has provided funds for research on the Neoproter- Programs, 36, 477.
ozoic basins of the São Francisco Craton. Currently, A. M. receives support from Campos-Neto, M. C. 1984a. Litoestratigrafia, relações estratigráficas e
project no. 486416/2006-2. A. J. K. acknowledges the NSF and NASA for evolução paleogeográfica dos grupos Canastra e Paranoá (região
funding for work in the Proterozoic successions of Brazil, as well as the Deutsche Vazante-Lagamar, MG). Revista Brasileira de Geociências, 14,
Forschungsgemeinschaft and Westfälische Wilhelms-Universität Münster for 81– 91.
sabbatical funding in 2007–2008. A. N. S acknowledges support from the CNPq Campos Neto, M. C. 1984b. Geometria e fases de dobramentos brasilia-
(grants PROSUL/CNPq 490136/2006-0 and 475657/2006-3) and PRONEX nos superpostos no oeste de Minas Gerais. Revista Brasileira de Geo-
APQ-047-1.07/06 that have defrayed costs with fieldwork in Sergipe and ciência, 14, 60– 68.
Bahia. We are grateful to the directors of Votorantim Metais for supporting Chang, H. K., Miranda, F. P. & Alkmim, F. F. 1988. Considerações
fieldwork at the Vazante area. This represents a contribution of the IUGS- and sobre a evolução tectônica da Bacia do São Francisco. In: 35th Con-
UNESCO-funded IGCP (International Geoscience Programme) Project #512. gresso Brasileiro de Geologia (Belém), Anais, 5, 2076–2090.
Cloud, P. E. & Dardenne, M. A. 1973. Proterozoic age of the Bambuı́
Group in Brazil. Geological Society of America Bulletin, 84,
References 1673–1676.
Condie, K. C. 2002. The supercontinent cycle: are there two patterns of
Alkmim, F. F., Brito Neves, B. B. & Alves, J. A. C. 1993. Arcabouço cyclicity? Journal of African Earth Sciences, 35, 179– 183.
Tectônico do Cráton do São Francisco: uma revisão. In: Dominguez, Cordani, U. G., Brito-Neves, B. B. & D’Agrella-Filho, M. S. 2003.
J. M. L. & Misi, A. (eds) O Craton do São Francisco SBG, SGM, From Rodinia to Gondwana: a review of the available evidence
CNPq, 45 –62. from South América. Gondwana Research, 6, 275– 283.
Alkmim, F. F., Chemale, F., Jr. & Endo, I. 1996. A deformação das Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005.
coberturas proterozóicas do Craton do São Francisco e o seu signifi- U–Pb ages from the Neoproterozoic Doushantuo Formation, China.
cado tectônico. Revista da Escola de Minas de Ouro Preto, 49, Science, 308, 95– 97.
22 – 38. Cozzi, A., Allen, P. & Grotzinger, J. 2004. Understanding carbonate
Amaral, G. & Kawashita, K. 1967. Determinações da idade do Grupo ramp dynamics using d13C profiles: examples from the Neoprotero-
Bambuı́ pelo método Rb– Sr. 21st Congresso Brasileiro de Geologia. zoic Buah Formation of Oman. Terra Nova, 16, 62 –67.
Curitiba, Anais, 214– 217. DA ROCHA ARAUJO, P. R., FLICOTEAUX, R., PARRON, C. & TROMPETTE, R.
Azmy, K., Veizer, J., Misi, A., Oliveira, T. F., Sanches, A. L. & Dard- 1992. Phosphorites of Rocinha mine, Patos de Minas (Minas Gerais
enne, M. A. 2001. Dolomitization and isotope stratigraphy of the Brasil): genesis and evolution of a Middle Proterozoic Deposit tecto-
Vazante Formation, São Francisco Basin, Brazil. Precambrian nized by the Brasiliano Orogeny. Economic Geology, 87, 332– 351.
Research, 112, 303– 329. Dardenne, M. A. 1978a. Zonação tectônica na borda ocidental do Cráton
Azmy, K., Kaufman, A. J., Misi, A. & Oliveira, T. F. 2006. Isotope São Francisco. 30th Congresso Brasileiro de Geologia, Recife, SBG,
stratigraphy of the Lapa Formation, São Francisco Basin, Brazil: 1, 299– 308.
implications for Late Neoproterozoic glacial events in South Dardenne, M. A. 1978b. Sı́ntese sobre a estratigrafia do Grupo Bambuı́
America. Precambrian Research, 149, 231– 248. no Brasil Central. 30th Congresso Brasileiro de Geologia, SBG, 2,
Azmy, K., Kendall, B., Creaser, R. A., Heaman, L. & de Oliveira, 597– 610.
T. F. 2008. Global correlation of the Vazante Group São Francisco Dardenne, M. A. 1979. Les minéralisations de plomb, zinc, fluor du Pro-
Basin, Brazil: Re– Os and U– Pb radiometric age constraints. Pre- térozoı̈que Supérieur dans le Brésil Central. Thèse de Doctorat
cambrian Research, 164, 160172. d’État, Université de Paris VI. Coromandel. Dissertação de Mestrado.
520 A. MISI ET AL.

Dardenne, M. A. 1981. Os grupos Paranoá e Bambuı́ na Faixa Dobrada Guimarães, J. T. 1996. A Formação Bebedouro no Estado da Bahia;
Brası́lia. An. Simp. Cráton São Francisco e suas Faixas Marginais, faciologia, estratigrafia e ambientes de sedimentação. Dissertação
Salvador, SBG, 140–157. (Mestrado em Geologia), Instituto de Geociências, Universidade
Dardenne, M. A. 2000. The Brası́lia Fold Belt. In: Cordani, U. G., Federal da Bahia (UFBA), Brasil.
Milani, E. J., Thomaz-Filho, A. & Campos, D. A. (eds) Tectonic Guimarães, J. T., Misi, A., Pedreira, A. J. & Dominguez, J. M. L. 2011.
Evolution of South America. International Geological Congress, 31, The Bebedouro Formation, Una Group, Bahia (Brazil). In: Arnaud,
Rio de Janeiro, 231– 263. E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
Dardenne, M. A. 2001. Lithostratigraphic sedimentary sequence of the Record of Neoproterozoic Glaciations. Geological Society, London,
Vazante Group. In: Misi, A. & Teixeira, J. B. G. (eds) Proterozoic Memoirs, 36, 503–508.
Base Metal Deposits of Africa and South America. Proceedings of Halverson, G. P., Hoffman, P. F., Schrag, D. P. & Kaufman, A. J.
the 1st Field Workshop IGCP 450, CNPq/UNESCO/IUGS, Belo 2002. A major perturbation of the carbon cycle before the Ghaub
Horizonte and Paracatu (MG), Brazil, 48 –50. glaciation (Neoproterozoic) in Namibia: prelude to snowball Earth.
Dardenne, M. A. 2006. A glaciação neoproterozóica do Grupo Vazante Geophysics, Geochemistry, Geosystems, 3, doi: 10.1029/2001G
na Faixa Brası́lia: discussão e alternativas. Congresso Brasileiro de C000244.
Geologia, 43, SBG, Aracaju, Resumos. Halverson, G. P., Hoffman, P. F. & Schrag, D. P. 2005. Toward a
Dardenne, M. A. 2007a. The pseudo-diamictites of the Vazante Group. Neoproterozoic composite carbon isotope record. Geological
Symposium IGCP 478, Stellenbosch, Abstracts. Society of America Bulletin, 117, 1181– 1207.
Dardenne, M. A. 2007b. Lithostratigraphy of the Vazante and Bambuı́ Hitzmann, M. W., Thorman, C. H., Romagna, G., Oliveira, T. F.,
groups in the São Francisco Craton and the Brası́lia Fold Belt. Dardenne, M. A. & Drew, L. J. 1995. The Morro Agudo Zn –Pb
IGCP 478, Stellenbosch, Abstracts. deposit, Minas Gerais, Brazil: a proterozoic Irish-type carbonate-
Dardenne, M. A. & Freitas-Silva, F. H. 1999. Modelos genéticos dos hosted SEDEX – replacement deposit. Annual Meeting, New
depósitos Pb-Zn nos grupos Bambuı́ e Vazante. Workshop: Depósitos Orleans, GSA, Abstract, 408.
Minerais Brasileiros de Metais-Base. Salvador, CAPES-PADCT- Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
ADIMB, 86– 93. testing the limits of global change. Terra Nova, 14, 129–155.
Dardenne, M. A. & Schobbenhaus, C. 2001. Metalogênese do Brasil. Hoffman, P. F., Hawkins, D. P., Isachsen, C. E. & Bowring, S. A. 1996.
Editora UnB, Brası́lia, 392. Precise U–Pb zircon ages for early Damaran magmatism in the
Dardenne, M. A., Trompette, R., Magalhães, L. F. & Soares, L. A. Summas Mountains and Welwitschia inlier, northern Damara belt,
1986. Proterozoic and Cambrian phosphorites-regional review: Namibia. Geological Survey of Namibia Communications, 11,
Brazil. In: Cook, P. J. & Shergold, J. H. (eds) Phosphate Deposit 47 – 52.
of the World. Proterozoic and Cambrian Phosphorites. Cambridge Hoffman, P. F., Kaufman, A. J., Halverson, G. P. & Schrag, D. P.
University Press, 116–131. 1998a. A Neoproterozoic snowball Earth. Science, 281, 1342– 1346.
Dardenne, M. A., Freitas-Silva, F. H., Nogueira, G. M. S. & Souza, Hoffman, P. F., Kaufman, A. J. & Halverson, G. P. 1998b. Comings
J. F. C. 1997. Depósitos de fosfato de Rocinha e Lagamar, Minas and goings of global glaciation on a Neoproterozoic tropical platform
Gerais. In: Schobbenhaus, C., Queiroz, E. T. & Coelho, in Namibia. GSA Today, 8, 1– 9.
C. E. S. (eds) Principais Depósitos Minerais do Brasil. DNPM/ Hoffmann, K.-H. & Prave, A. R. 1996. A preliminary note on a revised
CPRM, IVC, 113– 122. subdivision and regional correlation of the Otavi Group based on gla-
Dardenne, M. A., Pimentel, M. M. & Alvarenga, C. J. S. 2003. ciogenic diamictites and associated cap dolostones. Communications
Provenance of conglomerates of the Bambuı́, Jequitaı́, Vazante and of the Geological Survey of Namibia, 11, 77 – 82.
Ibiá groups, implications for the evolution of the Brası́lia Belt. Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
In: Sociedade Brasileira de Geologia, Simpósı́o Nacional de Estudos 2004. A U– Pb zircon date from the Neoproterozoic Ghaub For-
Tectônicos, 9, International Symposium on Tectonics, 3, Armação de mation, Namibia: Constraints on Marinoan glaciation. Geology, 32,
Búzios, 2003. Boletim de Resumos, 47–50. 817– 820.
D’el-Rey Silva, L. J. H. 1995. Tectonic evolution of the Sergipano Belt, Hurtgen, M. T., Arthur, M. A., Suits, N. S. & Kaufman, A. J. 2002.
northeastern Brazil. Revista Brasileira de Geociências, 25, 315– 332. The sulfur isotopic composition of Neoproterozoic seawater sulfate:
D’el Rey Silva, L. J. H. 1999. Basin infilling in the southern-central implications for a snowball earth? Earth Planet. Sci. Lett. 203,
part of the Sergipano Belt (NE Brazil) and implications for the 413– 430.
evolution of Pan-African/Brasiliano cratons and Neoproterozoic Hyde, W. T., Crowley, T. J., Baum, S. K. & Peltier, W. R. 2000.
sedimentary cover. Journal of South American Earth Sciences, 12, Neoproterozoic ‘snowball Earth’ simulations with a coupled
453– 470. climate/icesheet model. Nature, 405, 425– 429.
Derry, L. A., Keto, L. S., Jacobsen, S. B., Knoll, A. H. & Swett, K. Iyer, S. S., Babinski, M., Krouse, H. R. & Chemale, F., Jr. 1995. Highly
13
1989. Sr isotopic variations of upper Proterozoic carbonates from C enriched carbonate and organic matter in the Neoproterozoic
East Greenland and Svalbard. Geochimica Cosmochimica Acta, 53, sediments of the Bambuı́ Group, Brazil. Precambrian Research, 73,
2331– 2339. 271– 282.
Fairchild, I. J., Marshall, J. D. & Bertrand-Safati, J. 1990. Strati- Jacobsen, S. B. & Kaufman, A. J. 1999. The Sr, C and O
graphic shifts in carbon isotopes from Proterozoic stromatolitic isotope evolution of Neoproterozoic seawater. Chemical Geology,
carbonates (Mauritania): influence of primary mineralogy and dia- 161, 37– 57.
genesis. American Journal of Science, 290A, 46 –79. Karfunkel, J. & Hoppe, A. 1988. Late Proterozoic glaciation in
Fairchild, T. R., Schopf, J. W. et al. 1996. Recent discoveries of Central-Eastern Brazil: synthesis and model. Paleogeography, Paleo-
Proterozoic microfossils in south-central Brazil. Precambrian climatology, Paleoecology, 65, 1– 21.
Research, 80, 125– 152. Kaufman, A. J. & Knoll, A. H. 1995. Neoproterozic variations in the C-
Fuck, R. A., Pimentel, M. M. & Silva, J. H. D. 1994. Compartimentação isotopic composition of seawater: stratigraphic and biogeochemical
tectônica na porção oriental da Provı́ncia Tocantins. Anais 388 implications. Precambrian Research, 73, 27 – 49.
Congresso Brasileiro de Geologia, Camboriú, SBG, 1, 215– 216. Kaufman, A. J., Hayes, J. M., Knoll, A. H. & Germs, G. J. B. 1991.
Geboy, N. J. 2006. Rhenium– osmium age determinations of glaciogenic Isotopic compositions of carbonates and organic carbon from upper
shales from the Mesoproterozoic Vazante Formation, Brazil. MS Proterozoic successions in Namibia: stratigraphic variation and the
thesis, University of Maryland, 99. effect of diagenesis and metamorphism. Precambrian Research, 49,
Geboy, N. J., Kaufman, A. J. & Walker, R. J. 2006. A stable isotope and 301– 327.
Re –Os study of organic-rich mudstones across a Proterozoic glacial Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes, ice
cycle in Brazil. Geological Society of America Annual Meeting ages, and terminal Proterozoic earth history. Proceedings of the
(Philadelphia), Abstracts, 38, 125. National Academy of Science USA, 94, 6600–6605.
Geboy, N. J., Kaufman, A. J. et al. 2009. Evidence of Mesoproterozoic Kaufman, A. J., Varni, M. A., Misi, A. & Brito-Neves, B. B. 2001.
Ice Ages in Brazil. Internal report, unpublished. Anomalous d34S signatures in trace sulfate from a potential cap
NEOPROTEROZOIC SUCCESSIONS OF THE SÃO FRANCISCO CRATON 521

carbonate in the Neoproterozoic Bambuı́ Group, Brazil. In: Misi, A. Francisco Craton, Brazil: a review and a possible metallogenic evol-
& Teixeira, J. B. G. (organizers) Proterozoic Sediment-Hosted ution model. Ore Geology Reviews, 3, 263–304.
Base Metal Deposits of Western Gondwana. I Field Workshop, Misi, A., Kaufman, A. J. et al. 2007. Chemostratigraphic correlation of
Beolo Horizonte and Paracatu (Minas Gerais), Brazil, 62 –65. Neoproterozoic successions in South America. Chemical Geology,
Kawashita, K. 1998. Rochas carbonáticas neoproterozóicas da América 237, 161–185.
do Sul: idades e inferências quimioestratigráficas. Full professor Montes, A. S. L. 1997. O contexto estratigráfico e sedimentológico da
thesis, University of São Paulo, Brazil, 126. Formação Bebedouro na Bahia. Um possı́vel portador de diamantes.
Kiang, C. H. 1997. Isótopos estáveis (C, H, O) e 87Sr/86Sr: implicações na MSc thesis, University of Brası́lia, Brazil.
estratigrafia e na paleo-circulação de fluidos na Bacia do São Fran- Monteiro, L. V. S. 2002. Modelamento metalogenético dos depósitos
cisco. Full Professor Thesis, UNESP, São Paulo, Brazil. de Vazante, Fagundes e Ambrósia associados ao Grupo Vazante,
Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K. & Lambert, L. Minas Gerais. Tese Doutorado, Universidade de São Paulo-
B. 1986. Secular variation in carbon isotope ratios from upper USP, 317.
Proterozoic successions of Svalbard and East Greenland. Nature, Olcott, A. N., Sessions, A. L., Corsetti, F. A., Kaufman, A. J. & de
321, 832– 838. Oliviera, T. F. 2005. Direct evidence for significant primary
Kendall, B., Creaser, R. A. & Selby, D. 2006. Re– Os geochronology production during widespread Neoproterozoic glaciation. Science,
of post-glacial black shales in Australia: constraints on the timing of 310, 471–474.
‘Sturtian’ glaciation. Geology, 34, 729–732. Pedrosa-Soares, A. C., Dardenne, M. A., Hasui, Y., Castro, F. D. C.,
Kennedy, M. J., Christie-Blick, N. & Sohl, L. E. 2001. Are Proterozoic Carvalho, M. V. A. & Reis, A. C. 1994. Mapa Gológico do Estdo de
cap carbonates and isotopic excursions a record of gas hydrate Minas gerais e Nota Explicativa. Secretaria de Recursos Minerais,
destabilization following Earth’s coldest intervals? Geology, 29, Hı́dricos e Energéticos, Companhia Mineradora de Minas Gerais
443– 446. (COMIG), mapas e texto, 97.
Lima, O. N. B., Uhlein, A. & Britto, W. 2005. Geologia dos depósitos Pimentel, M. M., Dardenne, M. A., Viana, M. G., Costa, S. M., Gioia,
fosfáticos do Grupo Bambuı́ na Serra da Saudade, Cedro do Abaeté, L., Junges, S. & Seer, H. J. 2001. Nd isotopes and the provenance
Minas Gerais. In: Short Papers-III Simpósio sobre o Craton do São of sediments from the Neoproterozoic Brası́lia Belt, Central Brazil:
Francisco, Salvador, SBG, 320–323. geodynamic implications. Journal of South American Earth Sciences,
Madalosso, A. 1979. Stratigraphy and sedimentation of the Bambuı́ 14, 571– 585.
Group in Paracatu region, minas Gerais, Brazil. MA thesis, Univer- Porada, H. 1989. Pan-African rifting and orogenesis in southern to
sity of Missouri, USA, 127. equatorial Africa and Eastern Brazil. Precambrian Research, 44,
Marini, O. J., Fuck, R. A., Danni, J. C., Dardenne, M. A., Loguercio, 103– 136.
S. O. C. & Ramalho, R. 1984. As faixas de dobramentos Brası́lia, Powis, K. 2006. Stable isotope geochemistry of the Neoproterozoic
Uruaçu e Paraguai-Araguaia e o Maciço Mediano de Goiás. In: Bambuı́ Group at Serra do Ramalho, Bahia, Brazil. PhD thesis,
Schobbenhaus, C., Campos, D. A., Derze, G. R. & Asmus, H. E. Ottawa-Carleton Geoscience Center, University of Ottawa, Canada,
(eds) Geologia do Brasil. Texto Explicativo do Mapa Geológico do 197.
Brasil e da Área Oceânica Adjacente Incluindo Depósitos Minerais. Rodrigues, J. B. 2007. Estudos de proveniência de sedimentos da região
Ministério da Minas e Energias/Departamento Nacional da Centro-Oeste da Faixa Brası́lia-uma abordagem geocronológica.
Produção Mineral, Brası́lia, 251– 303. Exame de Qualificação Doutorado, Universidade de Brası́lia-UnB,
Misi, A. 1978. Ciclos de sedimentação e mineralizações de Pb –Zn nas Brası́lia, 50.
sequências Bambui (Supergrupo São Francisco), Estado de Bahia. Rodrigues, J. B. 2008. Proveniência de sedimentos dos grupos Canastra,
XXX Congresso Brasileiro de Geologia. Sociedade Brasileira de Geo- lbiá, Vazante e Bambuı́, Um estudo de zircões detrı́ticos e ldades
logia, Anais, 4, 2548–2561. Modelo Sm –Nd. Dr. Thesis, University of Brasilia, Brasilia DF.
Misi, A. 1979. O Grupo Bambuı́ no Estado da Bahia. In: Inda, H. V. (ed.) Rodrigues, J. B., Pimentel, M. M., Buhn, B., Dardenne, M. A. &
Geologia e Recursos Minerais do Estato da Bahia. Textos Básicos, Alvarenga, C. J. S. 2008. Provenance of the Vazante Group–
SME/CPM, Salvador, 1, 120– 154. Preliminary data. VI South American Symposium on Isotope
Misi, A. 1999. Um modelo de evolução metalogenética para os depósitos Geology, Bariloche, Argentina. Extented Abstracts, CD-ROM.
de zinco e chumbo hospedados em sedimentos proterozóicos de Sanches, A. L., Misi, A., Kaufman, A. J. & Azmy, K. 2007. As sucessões
cobertura do Craton do São Francisco. Full Professror Thesis, carbonáticas neoproterozóicas do Craton do São Francisco e os
Federal University of Bahia (Brazil). depósitos de fosfato: correlações e fosfogênese, Revista Brasileira
Misi, A. 2001. Estratigrafia isotópica das sequências do Supergrupo São de Geociências, 37, 1034–1046.
Francisco, coberturas neoproterozóicas do Cráton do São Francisco. Santos, R. V., de Alvarenga, C. J. S., Dardenne, M. A., Sial, A. N. &
Idade e correlações. In: Pinto, C. P. & Martins-Neto, M. A. Ferreira, V. F. 2000. Carbon and oxygen isotopes across
(eds) Bacia do São Francisco, Geologia e Recursos Naturais, SBG/ Meso-Neoproterozoic limestones from Central Brazil: Bambuı́ and
MG, Belo Horizonte, 67– 92. Paranoá groups. Precambrian Research, 104, 107–122.
Misi, A. & Kyle, J. R. 1994. Upper Proterozoic carbonate Santos, R. V., Alvarenga, C. J. S. et al. 2004. Carbon isotopes of
stratigraphy, diagenesis, and stromatolitic phosphorite formation, Mesoproterozoic-Neoproterozoic sequences from Southern São
Irecê Basin, Bahia, Brazil. Journal of Sedimentary Research, 64, Francisco Craton and Araçuaı́ Belt, Brazil: palaeogeographic impli-
299– 310. cations. Journal of South American Earth Sciences, 18, 27 –39.
Misi, A. & Souto, P. G. 1975. Controle estratigráfico das mineralizações Saylor, B. Z., Kaufman, A. J., Grotzinger, J. P. & Urban, F. 1998. A
de Pb –Zn –F –Ba do Grupo Bambui, parte leste da Chapada de Irecê composite reference section for terminal Proterozoic strata of
(Bahia). Revista Brasileria de Geosciências, 5, 30 – 45. southern Namibia. Journal of Sedimentary Research, 68, 1223– 1235.
Misi, A. & Veizer, J. 1998. Neoproterozoic carbonate sequences of the Sial, A. N., Ferreira, V. P., Almeida, A. R., Romano, A. W., Parente,
Una Group, Irecê Basin, Brasil: chemostratigraphy, age and corre- C., Da Costa, M. L. & Santos, V. H. 2000. Carbon isotope
lations. Precambrian Research, 89, 87 –100. fluctuations in Precambrian carbonate sequences of several
Misi, A., Veizer, J., Kawashita, K. & Dardenne, M. A. 1997. The age localities in Brazil. Anais Academia Brasileira de Ciências, 72,
of the Neoproterozoic carbonate platform sedimentation based on 540– 557.
87
Sr/86Sr determinations, Bambuı́ and Una Groups, Brazil. I South Sial, A. N., Ferreira, V. P., Moura, C. V. A. & Santos, V. H. 2003.
American Symposium on Isotope Geology, Campos do Jordão, São C-, O- and Sr-isotope stratigraphy of the Sturtian Jacoca and Olho
Paulo, Brazil. Extented Abstracts, 199– 200. D’Agua Formations, state of Sergipe, Northeastern Brazil. Short
Misi, A., Iyer, S. S. S. et al. 1999. Geological and isotopic constraints papers. IV South American Symposium on Isotope Geology. Salvador,
on the metallogenic evolution of the Proterozoic sediment-hosted Bahia, 394–397.
Pb – Zn (Ag) deposits of Brazil. Gondwana Research, 2, 47 – 65. Sial, A. N., Ferreira, V. P. et al. 2005. Two Neoproterozoic cap carbon-
Misi, A., Iyer, S. S. S. et al. 2005. Sediment-hosted lead– zinc deposits of ates in the states of Sergipe and Bahia, Northeastern Brazil: C- and
the Neoproterozoic Bambui Group and correlative sequences, São Sr-isotopes and mercury as paleoclimatic tracer. X Congresso
522 A. MISI ET AL.

Brasileiro de Geoquı́mica e II Simpósio de Geoquı́mica de Paı́ses do Trindade, R. I. F., D’Agrella-Filho, M., Babinski, M. & Brito Neves,
Mercosul, Porto de Galinhas. Short paper CD-ROM. B. B. 2004. Paleomagnetism and geochronology of the Bebedouro
Sial, A. N., Ferreira, V. P. et al. 2006a. Chemostratigraphy cap carbonate: evidence for continental-scale Cambrian remagnetiza-
of two Neoproterozoic cap carbonates from the Sergipano tion in the São Francisco craton, Brazil. Precambrian Research, 128,
belt (northeastern Brazil). Short Papers, V South American 83 – 103.
Symposium on Isotope Geology (V SSAGI). Punta del Este, Trompette, R. 1994. Geology of Western Gondwana (2000– 500 Ma).
Uruguay, 314–317. A.A. Balkema, Rotterdam, 350.
Sial, A. N., Ferreira, V. P. et al. 2006b. C- and Sr-isotopes and mercury Uhlein, A., Alvarenga, C. J. S., Trompette, R., Dupont, H. S. J. B.,
as paleoclimatic tracer in two Neoproterozoic cap carbonates in Egydio-Silva, M., Cukrov, N. & Lima, O. N. B. 2004. Glaciação
northeastern Brazil. Snowball Earth Conference. Ascona, Suiça, neoproterozóica sobre o Craton do São Francisco e faixas dobradas
100– 101. adjacentes. In: Mantesso-Neto, V., Bartorelli, A., Carneiro,
Sial, A. N., Dardenne, M. A. et al. 2009. The São Francisco Paleocon- D. R. C. & Brito Neves, B. B. (eds) Geologia do Continente
tinent. In: Neoproterozoic-Cambrian Tectonics, Global Change and Sul-Americano: Evolução da obra de Fernando Flávio de Almeida.
Evolution: A Focus on Southwestern Gondwana. Developments in Beca Ed., São Paulo, 539– 553.
Precambrian Geology, 16, 31– 69. Uhlein, A., de Alvarenga, C. J. S., Dardenne, M. A. & Trompette, R.
Söfner, B. 1973. Observações sobre a estratigrafia do Pré-Cambriano R. 2011. The glaciogenic Jequitaı́ Formation, southeastern Brazil.
da Chapada Diamantina Sudeste e da área contı́gua. In: SBG, 27th In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The
Congresso Brasileiro de Geologia. Aracajú, Anais, 1, 23 –33. Geological Record of Neoproterozoic Glaciations. Geological
Souza, J. C. F. 1997. Litoestratigrafia e sedimentologia da Formação Society, London, Memoirs, 36, 541–546.
Vazante na região de Coromandel. MSc thesis, Universidade de Valeriano, C. M., Dardenne, M. A., Fonseca, M. A., Simões, L. S. A.
Brası́lia. & Seer, H. J. 2004. A Evolução Tectônica da Faixa Brası́lia. In:
Teixeira, L. B., Martins, M. & Braun, O. P. G. 1993. Evolução geo- Mantesso-Neto, V., Bartorelli, A., Carneiro, D. R. C. &
lógica da Bacia São Francisco com base em sı́smica de reflexão e Brito Neves, B. B. (eds) Geologia do Continente Sul-Americano:
métodos potenciais. II Simpósio sobre o Craton do São Francisco. Evolução da obra de Fernando Flávio de Almeida. Beca Edta, São
SBG, Núcleo Bahia, Anais, 179–181. Paulo, 575–593.
Teixeira, W. & Figueiredo, M. C. H. 1991. An outline of Early Proter- Viera, L. C., Almeida, R. P., Trindade, R. I. F., Nogueira, A. C. R. &
ozoic crustal evolution in the São Francisco Craton, Brazil: a review. Janikian, L. 2007. A Formação Sete Lagoas em sua área tipo: fácies,
Precambrian Research, 53, 1 –22. estratigrafia e sistemas deposicionais. Revista Brasileira de Geo-
Thomaz-Filho, A., Kawashita, K. & Cordani, U. G. 1998. A sciências, 37, 1020– 1033.
origem do Grupo Bambuı́ no contexto da evolução geotectônica e Young, G. M. 1995. Are Neoproterozoic glacial deposits preserved on the
de idades radiométricas. Anais Academia Brasileira de Ciências, margins of Laurentia related to the fragmentation of two superconti-
70, 527– 548. nents? Geology, 23, 153– 156.
Torquato, J. R. F. & Misi, A. 1977. Medidas isotópicas de carbono Zhang, S., Jiang, G., Zhang, J., Song, B., Kennedy, M. J. & Christie-
e oxigênio em carbonatos do Grupo Bambui na região centro- Blick, N. 2005. U– Pb sensitive high-resolution ion microprobe ages
norte do Estado da Bahia. Revista Brasileira de Geociências, 7, from the Doushantuo Formation in south China: constraints on late
14 –24. Neoproterozoic glaciations: Geology, 33, 473–476.
Chapter 49

The Neoproterozoic Macaúbas Group, Araçuaı́ orogen, SE Brazil

A. C. PEDROSA-SOARES1*, MARLY BABINSKI2, CARLOS NOCE1, MAXIMILIANO MARTINS1,


GLÁUCIA QUEIROGA1 & FRANCISCO VILELA1
1
Universidade Federal de Minas Gerais, Instituto de Geociências-CPMTC, Campus Pampulha,
31270-901 Belo Horizonte, MG, Brazil
2
Universidade de São Paulo, Instituto de Geociências, Rua do Lago 562, Cidade Universitária 05508-080 São Paulo, SP, Brazil
*Corresponding author (e-mail: pedrosa@pesquisador.cnpq.br)

Abstract: The Neoproterozoic Macaúbas Group records accumulation in the precursor basin of the Araçuaı́ orogen, located on the
eastern margin of the São Francisco craton (SE Brazil). The Macaúbas basin evolved from a late Tonian continental rift to a passive
margin that lasted at least until c. 660 Ma. It was orogenically inverted during the late Neoproterozoic Brasiliano event. The Macaúbas
Group includes the pre-glacial Matão, Duas Barras and Rio Peixe Bravo formations, the glaciogenic Serra do Catuni, Nova Aurora and
Lower Chapada Acauã formations, and the post-glacial Upper Chapada Acauã and Ribeirão da Folha formations. In the central sector of
the Araçuaı́ orogen, the oldest glaciogenic unit of the group, the Serra do Catuni Formation, overlies the sandstone-conglomerate package
of the Duas Barras Formation. The Serra do Catuni Formation consists of massive diamictite with minor sandstone and rare pelite, depos-
ited mostly in a proximal glaciomarine environment. This unit passes upward and eastward into the Lower Chapada Acauã Formation, a
thick succession of stratified diamictite, graded sandstone, pelite, transitional basalt and rare carbonate. This distal glaciomarine unit is
covered by the diamictite-free Upper Chapada Acauã Formation, which passes eastward into the Ribeirão da Folha Formation, which
includes fine-grained turbidite, pelite and ocean-floor rocks. In the northern sector of the Araçuaı́ orogen, the sandstone-pelite succession
of the pre-glacial Rio Peixe Bravo Formation is covered by the Nova Aurora Formation, the glaciomarine unit rich in diamictite and Fe-
rich diamictite, with minor graded sandstone and rare pelite. The Nova Aurora Formation is covered by the sandstone-pelite package
of the Upper Chapada Acauã Formation. The pre-glacial and glaciogenic successions record the continental rift to transitional stages
of the Macaúbas basin. The post-glacial succession represents proximal and distal passive margin to ocean floor environments. The
Serra do Catuni Formation seems to be a proximal glaciomarine equivalent of the Jequitaı́ glacio-terrestrial deposits located on the
São Francisco craton.

The Macaúbas Group represents the precursor basin of the was thrust over the pelite-carbonate cratonic cover (Bambuı́
Neoproterozoic Araçuaı́ orogen, located in southeastern Brazil Group) and does not show any direct field relation with the Jequitaı́
(Fig. 49.1a). This orogen extends from the eastern border of the and Carrancas diamictites located on the southern São Francisco
São Francisco craton to the Atlantic Ocean (Pedrosa-Soares craton (Figs 49.1a & 49.2), although their correlation is often
et al. 2001, 2008). assumed (Uhlein et al. 2011).
Since the beginning of the 1930s the name Macaúbas has Several names such as conglomeratic phyllite, conglomeratic
appeared in the international geological literature to refer to a sedi- phyllonite, paraconglomerate, conglomeratic greywacke, mixtite
mentary unit that includes Precambrian glaciogenic rocks (Moraes and tillite were applied to the Macaúbas metadiamictites until
& Guimarães 1931). The first field studies were published by the 1970s, reflecting disagreement (and misunderstandings) over
Moraes (1929), who described the succession of ‘conglomeratic the glaciogenic interpretation, evolution of geological knowledge,
phyllites, phyllites, quartzites and mica schists cropping out in and standardization of nomenclature. Indeed, the diamictite-
the Macaúbas River valley and Catuni Ridge’, located to the bearing units preserved on the São Francisco craton record the
north of the city of Diamantina (Fig. 49.2). Moraes & Guimarães best evidence of glaciation, despite the relatively small area over
(1930) named that succession the Macaúbas Fm. and interpreted which they occur. The Jequitaı́ Fm. includes glacio-terrestrial sedi-
the ‘conglomeratic phyllites’ as metamorphosed glaciogenic ments and structures preserved from orogenic deformation (e.g.
rocks. In 1932, Moraes noted the very extensive distribution of Isotta et al. 1969; Viveiros & Walde 1976; Walde 1976; Rocha-
the Macaúbas Fm. after identifying a thick pile of ‘highly meta- Campos & Hasui 1981; Gravenor & Monteiro 1983; Karfunkel &
morphosed conglomeratic phyllonites’ associated with quartzite, Hoppe 1988; Martins-Neto et al. 1999; Uhlein et al. 1999, 2011;
schist and rare carbonate layers along the western valleys of the Karfunkel et al. 2002; Martins-Neto & Hercos 2002), and has a
Jequitinhonha and Rio Pardo rivers, and many of their tributaries maximum sedimentation age around 880 Ma (U – Pb LA-ICPMS,
(Fig. 49.2). Moraes (1932) was also the first to suggest a correlation detrital zircon; Rodrigues 2008). The Carrancas diamictite
between the Macaúbas Fm. and the glaciogenic Jequitaı́ Fm. occurs in a few outcrops that lack clear evidence of glacial
located on the São Francisco craton (Fig. 49.1a). Since Schöll origin, but it is overlain by the c. 740 Ma carbonate typical of
(1972) and Pflug & Schöll (1975), the designation Macaúbas post-glacial rocks within the cratonic cover (Schöll 1972; Babinski
Group (instead of formation) has been used. Almeida (1977) con- et al. 2007).
sidered the Macaúbas Group as the main fill of the ‘Alpine-type On the other hand, most of the Macaúbas metadiamictites rep-
geosyncline’ that became the Araçuaı́ fold belt after the Brasiliano resent sub-aqueous deposits that underwent strong deformation
orogenic event in late Neoproterozoic time. and metamorphism, hampering the recognition of solid evidence
Based on its first definition (Moraes 1929, 1932; Moraes & Gui- of glaciation. Nevertheless, the areas of occurrence, regional stra-
marães 1930, 1931), the name Macaúbas Group is here exclusively tigraphy, Neoproterozoic age and glaciogenic nature of the Macaú-
applied to refer to the extensive unit with metadiamictite-bearing bas metadiamictites seem to be definitively demonstrated by the
formations that underwent regional deformation and metamorph- data and correlations presented by Hettich (1975, 1977), Karfunkel
ism within the Araçuaı́ orogen (Figs 49.2 & 49.3). This group & Karfunkel (1976, 1977), Hettich & Karfunkel (1978), Viveiros

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 523– 534. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.49
524 A. C. PEDROSA-SOARES ET AL.

Structural framework

The correlation between the Araçuaı́ orogen and its counterpart


located in southwestern Africa, the West Congo belt (Fig. 49.1b),
is well-established in the geological literature (e.g. Porada 1989;
Pedrosa-Soares et al. 1992, 2008; Trompette 1994; Cordani
et al. 2003; Alkmim et al. 2006). The Araçuaı́– West Congo oro-
genic system, as well as its precursor basin, developed inside an
embayment of the São Francisco –Congo craton (Fig. 49.1b), char-
acterizing a Neoproterozoic confined orogen (Pedrosa-Soares et al.
2001, 2008; Rogers & Santosh 2004). That precursor basin
evolved from a Tonian continental rift to an inland-sea basin (a
large gulf) that was partially floored by Cryogenian oceanic crust
along its southern sector, and was inverted during the late Neo-
proterozoic Brasiliano –Pan African orogenesis, resulting in the
Araçuaı́ and West Congo belts (Pedrosa-Soares et al. 2001,
2007, 2008; Tack et al. 2001; Alkmim et al. 2006, 2007).
This system was split up in two quite different but complemen-
tary counterparts, after the Atlantic Ocean opening in Cretaceous
time (Pedrosa-Soares et al. 2008). The West Congo belt inherited
the thick package of bimodal volcanic rocks of the early Tonian
continental rift, as well as rift and passive margin successions
from Tonian to late Cryogenian age, but no Neoproterozoic ophio-
lite sliver (Tack et al. 2001; Frimmel et al. 2006; Pedrosa-Soares
et al. 2008). The Araçuaı́ orogen inherited the basin sector,
lacking in rift-related magmatic rocks, but including the remnants
of Neoproterozoic oceanic lithosphere and the extensive Macaúbas
Group that includes rift, transitional and passive margin sediments
from late Tonian to late Cryogenian age (Pedrosa-Soares et al.
2007, 2008). Neoproterozoic diamictites occur in both counter-
parts of the Araçuaı́ –West Congo orogenic system; those
located in the West Congo belt are reviewed in Tait et al. (2011).
The Macaúbas Group is the most complete record of the precur-
sor basin within the Araçuaı́ orogen because it shows evidence of
the rift and passive margin stages. Normal faults and clastic dykes
associated with the sedimentation of the lower Macaúbas Group,
as well as late Tonian mafic dykes and anorogenic granites, the
main evidence of the rift-related extensional tectonics (Karfunkel
et al. 1976, 1977, 1984; Machado et al. 1989; Guimarães et al.
1993; Grossi-Sad et al. 1997a; Uhlein et al. 1988, 1999; Martins-
Neto et al. 2001; Martins-Neto & Hercos 2002; Knauer et al. 2006;
Martins 2006; Martins et al. 2008; Silva et al. 2008). Transitional
mafic volcanism occurred in the late continental rift stage (Gradim
et al. 2005). Transgressive sedimentary successions free of
diamictite and ophiolite slivers record the passive margin stage
and associated oceanic spreading that lasted at least until
c. 660 Ma (Pedrosa-Soares et al. 1998, 2001, 2008; Queiroga
et al. 2007).
The Macaúbas Group was deformed, metamorphosed and
intruded by orogenic granites during the Brasiliano orogeny that
lasted from c. 630 Ma (onset of the pre-collisional stage) to c.
490 Ma (end of the post-collisional stage) in the Araçuaı́ orogen
(Pedrosa-Soares et al. 2001, 2008; Silva et al. 2005).
The Macaúbas Group occurs in the west-verging fold-and-thrust
belt that characterizes the external (western) tectonic domain of the
Fig. 49.1. (a) Location of the Araçuaı́ orogen in relation to the São Francisco
Araçuaı́ orogen (Figs 49.1 & 49.2). Along that belt, the major
craton, showing the distribution of the Macaúbas Group and Neoproterozoic structures of the first deformational phase are thrust faults and
cratonic covers. (b) The Araçuaı́–West Congo orogenic system in relation to ductile shear zones, some of them representing the tectonic inver-
the São Francisco-Congo craton. sion of normal faults related to the rift stage of the Macaúbas basin,
and west-verging tight asymmetric folds. These structures are
associated with an east-dipping regional foliation and a stretching
et al. (1979), Karfunkel et al. (1984, 1985), Karfunkel & Hoppe lineation (very well marked by stretched clasts in metadiamictites).
(1988), Mourão & Pedrosa-Soares (1992), Pedrosa-Soares et al. The kinematic indicators, such as fold vergence, S-C foliation and
(1992, 1998, 2000, 2008), Grossi-Sad et al. (1997a), Noce et al. asymmetric sigmoidal features (e.g. stretched sigma clasts in dia-
(1997), Uhlein et al. (1998, 1999, 2007), Babinski et al. (2005, mictites) of this first deformational phase, record tectonic transport
2011), Gradim et al. (2005) and Martins (2006). towards the São Francisco craton. The regional metamorphism
In this chapter, we present an updated synthesis on the Macaú- related to the first deformational phase increases from west to
bas Group with special emphasis on the glaciogenic units that are east so that the Macaúbas Group records the chlorite, biotite,
considered to represent at least one Cryogenian glacial event. garnet, staurolite, kyanite and sillimanite zones of an intermediate
NEOPROTEROZOIC MACAÚBAS GROUP 525

Fig. 49.2. Simplified geological map of the Araçuaı́ orogen (modified from Pedrosa-Soares et al. 2007, 2008). Section A– B shows structural and metamorphic features
of the Macaúbas Group.
526 A. C. PEDROSA-SOARES ET AL.

Fig. 49.3. Sketched map showing the distribution of the different formations of the Macaúbas Group in the external (western) tectonic domain of the Araçuaı́ orogen,
along with the regional stratigraphic scheme (modified from Pedrosa-Soares et al. 2007).

pressure metamorphic regime (Dossin & Dardenne 1984a; Tupinambá & Grossi-Sad 1997; Alkmim et al. 2006). These struc-
Pedrosa-Soares et al. 1984, 1992, 1993, 1996, 2001, 2008; Alkmim tures have been associated with the gravitational collapse of the
et al. 2006; Uhlein & Chaves 1989; Grossi-Sad et al. 1997a; Araçuaı́ orogen (Marshak et al. 2006). Two NE- and NW-trending,
Uhlein et al. 1998; Marshak et al. 2006). high-angle-dip regional fracture systems associated with open
A west-dipping crenulation cleavage associated with asym- megafolds (flexures) characterize the third deformational
metric folds and normal faults characterizes the second deforma- phase (Pedrosa-Soares et al. 1992; Grossi-Sad et al. 1997a),
tional phase (Pedrosa-Soares et al. 1993; Mourão & Grossi-Sad which may also be related to the gravitational collapse of
1997; Oliveira et al. 1997a; Pedrosa-Soares & Grossi-Sad 1997; the orogen.
NEOPROTEROZOIC MACAÚBAS GROUP 527

of the Guanhães block, only the Ribeirão da Folha Fm. occurs


(Figs 49.3 & 49.4).
The Matão, Duas Barras and Rio Peixe Bravo formations rep-
resent the oldest deposits of the Macaúbas Group (Figs 49.3 &
49.4). The Matão Fm. consists of breccia and conglomerate rich
in sandstone pebbles and cobbles, covered by sandstone with con-
glomerate lenses. Its maximum thickness is 200 m. An erosive
unconformity and normal faults define the contact between the
Matão Fm. and the basement (locally represented by eolian sand-
stone of the Statherian Espinhaço Supergroup). The Matão Fm.
records sedimentation under unstable tectonic conditions related
to the early rift stage of the Macaúbas basin (Martins 2006;
Martins et al. 2008). The Duas Barras Fm. consists of sandstone
and conglomeratic sandstone with variable contents of mica, feld-
spar, iron oxide and/or lithic fragments, quartz sandstone, con-
glomerate and rare pelite. It shows fluvial and shallow marine
sedimentary structures and bimodal (NW –SE and SE –NW)
palaeocurrent sets. The maximum thickness is approximately
100 m (Noce 1997a; Grossi-Sad et al. 1997b; Martins 2006).
The Rio Peixe Bravo Fm. consists of micaceous, ferruginous
Fig. 49.4. General stratigraphic scheme for the central sector (17–188S) of the and/or feldspathic sandstone, pelite locally rich in hematite and/
Macaúbas Group (not to scale; estimated thicknesses are referred to in the text or graphite, and rare conglomerate, with a maximum thickness
and Fig. 49.5; modified from Pedrosa-Soares et al. 2008). of around 700 m (Viveiros et al. 1979; Mourão & Grossi-Sad
1997; Oliveira et al. 1997a; Roque et al. 1997; Knauer et al.
2006). The Duas Barras and Rio Peixe Bravo formations, which
In fact, all primary rocks of the Macaúbas Group were modified lack evidence of glacial influence, represent fluvial to marine sedi-
by regional deformation and metamorphism of variable intensity. mentation during the continental rift stage of the Macaúbas basin
However, the general low metamorphic grade (greenschist to inter- (Noce et al. 1997; Uhlein et al. 1998, 2007; Martins-Neto et al.
mediate amphibolite facies) and the preservation of sedimentary 2001; Martins 2006; Pedrosa-Soares et al. 2007, 2008).
and volcanic structures in low strain zones allow the interpretation The Serra do Catuni Fm., the oldest diamictitic unit of the
of the pre-metamorphic lithology and other features, such that the Macaúbas Group, is a package of massive diamictite with minor
protoliths can be confidently described. sandstone and rare pelitic rhythmite lenses (Figs 49.3 & 49.4).
This formation is locally rich in chaotic boulders, and contains
Stratigraphy faceted and striated flat-iron-shaped cobbles and pebbles (Karfun-
kel & Karfunkel 1976, 1977; Karfunkel et al. 1984; Karfunkel &
The regional stratigraphic scheme adopted here resulted from the Hoppe 1988; Pedrosa-Soares et al. 1992, 2008; Noce et al. 1997;
Espinhaço Project (a 1:100 000 scale geological mapping pro- Uhlein et al. 1999, 2007; Martins-Neto et al. 2001; Martins-Neto &
gramme that largely covered the Macaúbas Group; Grossi-Sad Hercos 2002; Martins 2006).
et al. 1997a; Noce et al. 1997), with modifications inserted by The Nova Aurora Fm. mainly comprises stratified diamictite,
Lima et al. (2002), Pedrosa-Soares et al. (2007, 2008) and ferruginous diamictite-bearing and diamictitic Fe formation, with
Martins et al. (2008). The Macaúbas Group is subdivided into minor intercalations of sandstone with fining-up graded bedding
seven formations, from oldest to youngest Matão, Duas Barras, and rare pelite (Viveiros et al. 1979; Pedrosa-Soares et al. 1992;
Rio Peixe Bravo, Serra do Catuni, Nova Aurora, Chapada Acauã Mourão & Grossi-Sad 1997; Noce et al. 1997; Oliveira et al.
and Ribeirão da Folha (Fig. 49.3). 1997a; Pedrosa-Soares & Oliveira 1997; Roque et al. 1997;
Two uplifted basement blocks divide the distribution of the Uhlein et al. 1998, 1999; Drumond 2000).
Macaúbas Group into three sectors with variably complete strati- The Chapada Acauã Fm. includes a lower diamictitic unit and
graphic records: the northern sector, related to the Porteirinha an upper diamictite-free unit (Figs 49.3 & 49.4). The Lower
block; the central sector, located between 178 and 188S; and the Chapada Acauã Fm. essentially consists of stratified diamictite,
southern sector, related to the Guanhães block (Fig. 49.3). graded sandstone, pelite and mafic volcanic rocks (Hettich 1977;
The most complete package of the Macaúbas Group occurs in Mourão & Pedrosa-Soares 1992; Pedrosa-Soares et al. 1992,
the central sector where the lowermost units are the diamictite-free 2007, 2008; Mourão & Grossi-Sad 1997; Oliveira et al. 1997a, b;
Matão and Duas Barras formations that are covered by the most Tupinambá & Grossi-Sad 1997; Uhlein et al. 1999; Gradim et al.
proximal and oldest diamictite-bearing unit of the group, the 2005; Martins 2006). At least one thick carbonate lens (c. 20 m)
Serra do Catuni Fm. (Figs 49.3 & 49.4). This formation passes occurs intercalated in a pelite layer with isolated outsized clasts at
upward and eastward into the diamictite-bearing Lower Chapada the top of the Lower Chapada Acauã Fm. (Pedrosa-Soares & Grossi-
Acauã Fm., which, in turn, is covered by the diamictite-free Sad 1997). The mafic volcanic rocks, metamorphosed to green-
Upper Chapada Acauã Fm. This unit passes eastwards into the schists, show pillow structures and other features of sub-aqueous
diamictite-free Ribeirão da Folha Fm. (Fig. 49.4). flows (Schrank et al. 1978; Chula et al. 1996; Uhlein et al. 1998;
In the northern sector, only the Serra do Catuni Fm. occurs Gradim et al. 2005). They have tholeiitic basalt protoliths with a
between the São Francisco craton boundary and the western edge dominant within-plate signature (Gradim et al. 2005), Sm –Nd
of the Porteirinha block (Fig. 49.3). To the east of the Porteirinha TDM model age of c. 1.5 Ga, and inherited zircon grains with det-
block the diamictite-free Rio Peixe Bravo Fm. is overlain by the rital features and U – Pb SHRIMP ages from Archaean to Late
diamictite-bearing Nova Aurora Fm., which is partially capped Mesoproterozoic (Babinski et al. 2005). However, samples with
by the Upper Chapada Acauã Fm. oceanic signature and slightly positive 1Nd(900 Ma) of þ0.23,
In the southern sector, the stratigraphic record of the Macaúbas together with the inherited zircon grains, suggest a transitional
Group is very incomplete. The Serra do Catuni Formation, mafic magma that migrated through a thinned continental crust
locally overlain by strata that probably correlate to the Chapada (Babinski et al. 2005; Gradim et al. 2005). Accordingly, these
Acauã Formation, occurs along a narrow zone in the western greenschists provide strong evidence of volcanism in an exten-
thrust front of the Araçuaı́ orogen (Figs 49.2 & 49.3). To the east sional marine basin floored by thin continental crust, during the
528 A. C. PEDROSA-SOARES ET AL.

transitional phase from the late rift to early passive margin stages
of the Macaúbas basin.
Both the Nova Aurora and Lower Chapada Acauã formations
show cyclic deposition from coarse- to fine-grained sediments,
fining-up graded bedding, erosional contacts between cycles and
load structures (Mourão & Pedrosa-Soares 1992; Pedrosa-Soares
et al. 1992; Noce et al. 1997; Pedrosa-Soares & Oliveira 1997;
Uhlein et al. 1999, 2007; Martins-Neto et al. 2001). The features
of these formations, as well as of the Serra do Catuni Fm., are
described and interpreted in subsequent sections.
The Upper Chapada Acauã Fm. is a succession of interbedded
sandstone and pelite, free of diamictite, interpreted as a post-
glacial unit deposited in a shelf environment during the passive
margin stage of the Macaúbas basin (Pedrosa-Soares et al.
1992; Mourão & Grossi-Sad 1997; Noce et al. 1997; Oliveira
et al. 1997a; Pedrosa-Soares & Grossi-Sad 1997; Martins-Neto
et al. 2001).
The Ribeirão da Folha Fm. is free of diamictite and includes
distal passive margin and ocean floor deposits (Figs 49.2, 49.3
& 49.4). The proximal part of the Ribeirão da Folha Fm. is a suc-
cession of fine-grained turbidites and very rare limestone lenses.
It overlies the diamictite-turbidite unit of the Lower Chapada
Acauã Fm. and is interpreted as a deep-water correlative of the
shelf sandstone-pelite pile of the Upper Chapada Acauã Fm.
(Pedrosa-Soares et al. 1992, 2007, 2008; Pedrosa-Soares &
Grossi-Sad 1997). The distal part of the Ribeirão da Folha For-
mation is locally associated with thrust slices of oceanic meta-
mafic and meta-ultramafic rocks, constituting a tectonically
dismembered ophiolite complex (Pedrosa-Soares et al. 1992,
1998, 2001, 2008; Queiroga et al. 2007). The distal Ribeirão
da Folha Fm. is rich in peraluminous and carbonaceous pelites
with intercalations of volcanic-exhalative deposits (sulphide-
bearing cherts, massive sulphide bodies, and banded iron for-
mations of oxide, silicate and sulphide types) and mafic volcanic
rocks with ocean-floor geochemical and isotopic (1Nd(660 Ma) of
c. þ4) signatures (Pedrosa-Soares et al. 1998, 2001, 2008; Queir-
oga et al. 2006, 2007).

Glaciogenic deposits and associated strata

The Macaúbas Group includes three diamictite-bearing units that


have been interpreted as glaciogenic deposits: the Serra do
Catuni, Nova Aurora and Lower Chapada Acauã formations
(Figs 49.3, 49.4 & 49.5). The rock-assemblage included in the
Serra do Catuni Fm. was mapped and/or studied in detail by Kar-
funkel & Karfunkel (1976, 1977), Dossin & Dardenne (1984a),
Karfunkel et al. (1984), Pedrosa-Soares et al. (1984, 1992, 2007,
2008), Uhlein & Chaves (1989), Fogaça (1997), Grossi-Sad et al.
(1997b), Guimarães (1997), Guimarães et al. (1997), Knauer &
Grossi-Sad (1997), Mourão et al. (1997), Noce (1997a, b), Oliveira
et al. (1997b), Tupinambá & Grossi-Sad (1997), Uhlein et al. (1998,
1999, 2007), Knauer et al. (2006) and Martins (2006). The Serra do
Catuni Fm. is a very extensive and rather homogeneous unit that
persistently occurs for over than 400 km along the western
border of the Araçuaı́ orogen (Fig. 49.3). This formation is rich
in massive diamictite, the thickness of which ranges from a few
tens of metres up to 300 m (Fig. 49.5). The following description
is based on sections in the Macaúbas River valley (Fig. 49.2),
where the Serra do Catuni Fm. is reasonably well preserved
Fig. 49.5. Composite stratigraphic columns for the Serra do Catuni, Nova
from deformation and erosion.
Aurora and Lower Chapada Acauã formations (Macaúbas Group). The Serra do
The Serra do Catuni Fm. overlies the Duas Barras Formation
Catuni column is based on sections described by Martins (2006) along the and contains a 250-m-thick pile of massive diamictite with
Macaúbas River valley (Fig. 49.2). The Nova Aurora column synthesizes lenses of massive sandstone (Martins 2006). The diamictite has
field and drill holes data from the region limited by a white rectangle in an abundant (60 –80% in volume), poorly sorted, muddy-
Figure 49.3. The column for the Lower Chapada Acauã Fm. is based on field sandstone matrix composed of detrital quartz, K-feldspar and
data from the eastern escarpments of the Acauã Plateau (Fig. 49.3) and medium carbonate, with a metamorphic foliation marked mainly by fine-
valley of the Araçuaı́ River (Fig. 49.2), as well as on data from the Macaúbas grained muscovite. The clasts, ranging in size from granules to
River valley. boulders, are also poorly sorted in texture and composition
NEOPROTEROZOIC MACAÚBAS GROUP 529

(milky quartz, sandstone, granitoid, carbonate and mafic rock). the southwestern Araçuaı́ and Jequitinhonha valleys (Fig. 49.2).
Faceted and striated flat-iron-shaped cobbles and pebbles can Detailed sections carried out along creeks of the medium valley
also be found in the Serra do Catuni diamictite. Lenses of of the Araçuaı́ River suggest a local minimum thickness of
massive sandstone up to 2 m thick appear mainly in the upper 800 m for the Lower Chapada Acauã Fm. (Mourão & Pedrosa-
part of the section. This sandstone is poorly sorted and consists Soares 1992; Pedrosa-Soares & Grossi-Sad 1997). However, the
of quartz with minor K-feldspar, carbonate and iron oxides strong regional deformation hampers any accurate thickness
(Martins 2006). evaluation for the whole unit, which seems to be thicker than
The rock-assemblage included in the Nova Aurora Fm. 2 km (e.g. Grossi-Sad et al. 1997a; Uhlein et al. 1998, 1999).
was mapped and/or studied in detail by Viveiros et al. (1979), The following description is based on detailed studies carried
Pedrosa-Soares et al. (1984, 1992, 2000, 2007, 2008), Mourão & out in the middle valley of the Araçuaı́ River, the type-area
Grossi-Sad (1997), Oliveira et al. (1997a), Pedrosa-Soares & of the Lower Chapada Acauã Fm. (Mourão & Pedrosa-Soares
Oliveira (1997), Roque et al. (1997), Uhlein et al. (1998, 1999), 1992; Pedrosa-Soares et al. 1992; Pedrosa-Soares & Grossi-Sad
Drumond (2000) and Knauer et al. (2006). The Nova Aurora 1997), and in the Macaúbas River valley (Martins 2006). The
Fm. is well exposed in road cuts to the west of Salinas town Lower Chapada Acauã Fm. has many vertical and lateral variations
(Fig. 49.3), and part of its stratigraphic pile is also known in drill of lithofacies. It is essentially composed of fining-up cycles of
holes for iron ore prospecting (Viveiros et al. 1979; Uhlein et al. stratified diamictite, graded sandstone and pelitic rhythmite
1999; Vilela 2010). Viveiros et al. (1979) estimated the (Fig. 49.5). The diamictite layers show great variations in clast/
minimum thickness for the Nova Aurora Fm. to be c. 1200 m matrix ratios and the foliated matrix consists mainly of quartz,
based on detailed sections and data from drill holes. Again, the feldspar, muscovite, chlorite, biotite and garnet (according to the
strong regional deformation hampers any accurate thickness evalu- metamorphic zone). The clasts vary from granules to boulders
ation for the whole unit, which has independently been estimated and are composed of milky and blue quartz, metasandstone, car-
to be thicker than 2 km (e.g. Grossi-Sad et al. 1997a; Uhlein bonate, metapelite, granitoid, gneiss and meta-mafic rock. The dia-
et al. 1998, 1999). mictite layers envelope lenses of cross-laminated sandstone,
The following description is based mainly on detailed studies ranging in length from decimetres to a few metres and up to 1 m
carried out by Viveiros et al. (1979), Pedrosa-Soares et al. thick. At the top of the cycles the diamictite gives way to sandstone
(1984, 1992), Pedrosa-Soares & Oliveira (1997) and Vilela and pelite. Erosional contacts between cycles are common. The
(2010) in the region west of Salinas (white rectangle in sandstone layers and lenses show fining-up graded bedding
Fig. 49.3). The Nova Aurora Fm. is essentially a package of dia- and vary in composition from wacke-rich in carbonate + feldspar
mictites with minor intercalations of graded sandstone and + lithoclasts + micas to quartz sandstone. Conglomerate +
scarce pelite, with the most striking feature being the great conglomeratic sandstone can be present at the base of the sand-
amount of Fe-rich diamictite. The lower Nova Aurora Fm. is stone layers and lenses. The pelitic rhythmite layers locally show
dominated by Fe-rich diamictite followed by non-ferruginous isolated outsized clasts, mainly cobbles and boulders. These
diamictite (Fig. 49.5). Despite the Fe content, the diamictites layers have mineralogical compositions similar to the diamictite
vary considerably in grain size, matrix composition and clast/ matrix, with variable contents of metamorphic minerals, such as
matrix ratio. They are metamorphosed to greenschist facies and quartz, micas and garnet, in alternating laminae and bands.
their clasts are mainly of milky to blue quartz, metasandstone,
carbonate, granitoid, gneiss and metapelite. The Fe-rich
diamictite package consists of ferruginous diamictite (hematite þ
magnetite .5% in volume) and diamictitic Fe formation (Fe Boundary relations with overlying and underlying
.15 wt%), with minor intercalations of sulphide-bearing non-glacial units
diamictite, ferruginous and non-ferruginous sandstone, hema-
tite – magnetite schist and pelite with scattered outsized clasts. The Serra do Catuni Fn. was thrust over the Neoproterozoic pelite-
The foliated matrix of the Fe-rich diamictite consists of quartz, carbonate cratonic cover (Fig. 49.2), outlining a clear limit
hematite, magnetite, muscovite, biotite and chlorite. The amount between the Araçuaı́ orogen and São Francisco craton (Almeida
of magnetite and the total Fe content (in excess of 40% in 1977; Uhlein et al. 1998; Alkmim et al. 2006; Pedrosa-Soares
places) increases in shear zones. A thin carbonate intercalation, et al. 2008).
recently found in a drill hole, occurs in the upper section of the A regional disconformity occurs between the diamictite-
Fe-rich diamictite. Layers of sulphide-bearing diamictite occur at free Duas Barras Fm. and the overlying Serra do Catuni Fm.,
the base and top of the Fe-rich diamictite package. The sulphides the oldest diamictitic unit of the Macaúbas Group (Karfunkel
are mainly pyrite and pyrrhotite. The upper package consists of & Hoppe 1988; Grossi-Sad et al. 1997a; Uhlein et al. 1999,
stratified diamictite with minor layers and lenses of fining-up 2007; Martins-Neto et al. 2001; Martins 2006). Locally, the
graded sandstone and thin intercalations of pelite. This diamictite Serra do Catuni diamictite shows erosional contact with
has a foliated matrix composed mainly of quartz, biotite, musco- the Duas Barras Fm., but normal and tectonic (sheared) contacts
vite, garnet and carbonate. The arenaceous intercalations vary in are also observed (Grossi-Sad et al. 1997b; Noce 1997a;
composition from immature (micaceous, feldspathic, ferruginous) Martins 2006).
to quartz sandstone, and commonly show load structures. Ero- The Serra do Catuni Fm. gradually passes upwards and east-
sional contacts can be observed between diamictite and sandstone wards to the interlayered succession of diamictite, sandstone,
intercalations. pelite and mafic volcanic rocks that characterizes the Lower
The rock-assemblage included in the Lower Chapada Acauã Fm. Chapada Acauã Fm. (Fig. 49.4). Intraformational erosive contacts
was mapped and/or studied in detail by Hettich (1975, 1977), between diamictite and sand-pelite turbidites of both the Lower
Karfunkel & Karfunkel (1976, 1977), Schrank et al. (1978), Chapada Acauã and Nova Aurora formations can be locally
Uhlein & Chaves (1989), Mourão & Pedrosa-Soares (1992), observed (Pedrosa-Soares et al. 1992, 2008; Pedrosa-Soares &
Pedrosa-Soares et al. (1992, 1993, 2007, 2008), Chula et al. Oliveira 1997; Pedrosa-Soares & Grossi-Sad 1997; Uhlein et al.
(1996), Grossi-Sad et al. (1997b), Mourão & Grossi-Sad (1997), 1998, 1999).
Noce (1997b), Oliveira et al. (1997a), Pedrosa-Soares & Grossi- In the Araçuaı́ River valley and eastern escarpments of the
Sad (1997), Tupinambá & Grossi-Sad (1997), Uhlein et al. Acauã Plateau (Figs 49.2 & 49.3), the Lower Chapada Acauã
(1998, 1999, 2007), Gradim et al. (2005) and Martins (2006). Fm. gradually passes upwards to the diamictite-free package of
The Lower Chapada Acauã Fm. is well exposed along the the Upper Chapada Acauã Fm. To the east, this sandstone-
escarpments of the Acauã Plateau (Fig 49.3) and waterways in pelite unit transitions to the fine-grained turbidite package of
530 A. C. PEDROSA-SOARES ET AL.

the proximal Ribeirão da Folha Fm. (Pedrosa-Soares et al. 1992, Geochronological constraints
2008; Pedrosa-Soares & Grossi-Sad 1997; Uhlein et al.
1999, 2007). Although there are abundant data from detrital zircon grains
In the northern sector of the Macaúbas Group, the diamictite- extracted from rocks of the Matão, Duas Barras, Serra do Catuni
free Upper Chapada Acauã Formation conformably covers the and Chapada Acauã formations, and from intrusions related to dis-
Nova Aurora Formation (Viveiros et al. 1979; Roque et al. tinct extensional stages of the precursor basin of the Araçuaı́
1997; Mourão & Grossi-Sad 1997; Oliveira et al. 1997b; Pedrosa- orogen, these data yield few useful constraints on the timing of
Soares & Oliveira 1997). glacial sedimentation in the Macaúbas Group. The best maximum
age constraint on the onset of glacial sedimentation in the Macaú-
bas Group comes from a detrital study of the Duas Barras Fm., the
Chemostratigraphy lowermost stratigraphic unit, where the youngest zircon has an age
of 900 + 21 Ma (Babinski et al. 2011). The youngest concordant
No chemostratigraphic study is available for the Macaúbas Group, zircon grain in the Upper Chapada Acauã Fm. provides a
but there are isotopic data for older carbonate units located in the maximum depositional age of 864 + 30 Ma (Pedrosa-Soares
Araçuaı́ orogen and for carbonate clasts extracted from the Serra et al. 2000).
do Catuni diamictite. Santos et al. (2004) presented C-isotopic The Pedro Lessa mafic dykes and Salto da Divisa granite
data for metadolomite and metamarlstone samples of the Domin- (Fig. 49.2) are considered to be related to the continental rift
gas Fm., formerly interpreted as the oldest unit of the Macaúbas stage of the precursor basin of the Araçuaı́ orogen (Pedrosa-Soares
Group (Grossi-Sad et al. 1997b; Noce 1997a). In fact, the et al. 2008; Silva et al. 2008). The meta-dolerite dyke swarm of
shallow marine dolomite-pelite package of the Domingas Fm. is Pedro Lessa was dated at 906 + 2 Ma (zircon and baddeleyite,
unconformably covered by a fluvial sandstone-conglomerate U –Pb TIMS; Machado et al. 1989). The fluorite-bearing anoro-
succession of the Duas Barras Fm. (age , 900 + 21 Ma, see genic Salto da Divisa granite yielded a zircon U –Pb SHRIMP
‘Geochronological constraints’) and contains the stromatolite age of 875 + 9 Ma (Silva et al. 2008).
Conophyton metula Kirichenko that suggests a Mesoproterozoic The maximum age of deep sea sedimentation for the distal
age (Schöll 1976). The stromatolitic dolomite samples of the Dom- Ribeirão da Folha Fm. is constrained by the age of 660 + 29 Ma
ingas Formation yielded d13CPDB values from þ0.4 to þ0.7‰, (U –Pb LA-ICPMS), obtained for the magmatic crystallization of
similar to other Mesoproterozoic dolostones in the region (the euhedral zircon crystals extracted from meta-plagiogranite
Rio Pardo Grande Fm.; Santos et al. 2004). Other metadolomite patchs found in ophiolitic meta-mafic rocks (Queiroga et al.
samples of the Domingas Fm. yielded d13CPDB values from þ0.2 2007). The entire Macaúbas must be older than the regional meta-
to –1.9‰ and one sample of metamarlstone is as low as –4.0‰ morphism, which is dated at c. 580 Ma (Pedrosa-Soares et al. 2001,
(Santos et al. 2004). 2008; Silva et al. 2005).
Martins (2006) reported d13CPDB values ranging – 7.2 to 0‰
obtained on 14 carbonate pebbles extracted from the Serra do
Catuni diamictite. These data are similar to those obtained from Discussion
carbonate clasts of the Jequitaı́ diamictite (d13CPDB from – 0.6 to
–3.1‰) and from some samples of the Domingas Fm. by Santos The Macaúbas Group includes three widespread diamictitic units,
et al. (2004). the Serra do Catuni, Nova Aurora and Lower Chapada Acauã for-
mations (Figs 49.3, 49.4 & 49.5). The Serra do Catuni Fm. has been
interpreted as a glaciogenic unit deposited mostly in a proximal
Other characteristics marine environment based on multiple observations. First, it com-
prises massive diamictite occurring for over 400 km along the
Manganese deposits have been mined in the southern and central proximal region of the Macaúbas basin (Fig. 49.3). Second, it con-
sectors of the Macaúbas Group. They consist of manganese tains zones rich in chaotic boulders, faceted and striated cobbles
oxides formed by Cenozoic lateritic alteration and supergene con- and pebbles, and only scarce finer-grained sediments. Third, it is
centration on pelitic horizons of the Chapada Acauã Fm. (Dossin & correlated with the glacio-terrestrial deposits of the Jequitaı́ Fm.
Dardenne 1984b). located on the little deformed São Francisco cratonic cover
The thick carbonate lens of the Lower Chapada Acauã Fm. sequence to the west.
(Figs 49.4 & 49.5) has been sporadically mined (Pedrosa-Soares The Nova Aurora and Lower Chapada Acauã formations also
& Grossi-Sad 1997). occur in extensive areas, mainly to the east of the Serra do
The lower part of the Nova Aurora Fm. is rich in ferruginous Catuni Fm., along a more distal region of the Macaúbas basin
diamictite and diamictitic Fe formation that form deposits up to (Fig. 49.3). They essentially consist of stratified diamictite, sand-
200 m thick, with Fe content up to 40 wt% (Viveiros et al. 1979; stone and pelite, and show sedimentary features such as cyclic
Vilela 2010). These deposits are widely capped by thick lateritic deposition from coarse- to fine-grained sediments, graded bedding,
soils on the plateaus that characterize the northern sector of the erosional contacts between cycles and load structures that are evi-
Macaúbas Group. However, new aeromagnetic data suggest dence of debris flow and turbiditic sedimentation in a distal glacio-
significant Fe-rich diamictite deposits that are the target of marine environment. The clear relationship between the Serra do
current exploration. Catuni and Lower Chapada Acauã formations also points to a tran-
Martins (2006) suggested that the Serra do Catuni diamictite is sition from proximal to distal glaciomarine conditions, respect-
the source for diamonds found in alluvial deposits along ively (Figs 49.4 & 49.5). The Nova Aurora Fm., which overlies
watercourses of the Macaúbas River valley (Fig. 49.2). the pre-glacial Rio Peixe Bravo Fm., can be interpreted as a
distal correlative of the Serra do Catuni Fm. and, to some extent,
a lateral equivalent of the Lower Chapada Acauã Fm.
Palaeolatitude and palaeogeography Based on the occurrence of glaciogenic diamictites, the Macaú-
bas Group is subdivided into a pre-glacial diamictite-free succes-
No palaeomagnetic data are available for the Macaúbas Group. sion (Matão, Duas Barras and Rio Peixe Bravo formations), a
However, palaeomagnetic data and palaeogeographic settings for glaciogenic succession (Serra do Catuni, Nova Aurora and Lower
the São Francisco–Congo craton in late Mesoproterozoic and Chapada Acauã formations) and a post-glacial succession
Neoproterozoic times have been discussed in several papers (Upper Chapada Acauã and Ribeirão da Folha formations).
(e.g. Cordani et al. 2003; D’Agrella-Filho et al. 2004). These successions are related to different evolutionary stages of
NEOPROTEROZOIC MACAÚBAS GROUP 531

the Macaúbas basin (Fig. 49.3). The pre-glacial and glaciogenic In conclusion, it seems that the Macaúbas Group records at least
successions have been ascribed to the continental rift to transi- the early Cryogenian glacial event. The Serra do Catuni Fm. and its
tional stages of the Macaúbas basin based on multiple lines of probable correlative, the Jequitaı́ Fm., were likely deposited during
evidence: an early phase of that glacial event. The Lower Chapada Acauã
Fm. and its probable equivalents, the Carrancas diamictite and
† the predominance of immature sandstone-conglomeratic sedi- the Lower Mixtite Fm. then represents a late phase of the early
mentation with bimodal (NW – SE and SE –NW) palaeocurrent Cryogenian glacial event. Indeed, the very extensive and thick dia-
sets in the pre-glacial succession; mictitic package of the Macaúbas Group might contain equivalents
† syn-sedimentary normal faulting; of all the Neoproterozoic glaciogenic units preserved on the São
† clastic dykes related to the rudite-sandstone sedimentation of Francisco craton. The challenging question is how many discrete
the pre-glacial succession and Serra do Catuni Fm.; glacial epochs do they represent?
† the huge ferruginous diamictite deposits of the Nova Aurora
Fm.; The authors acknowledge financial support provided by CNPq (Conselho Nacio-
nal de Desenvolvimento Cientı́fico e Tecnológico), FINEP/CT-Infra (Financia-
† the transitional geochemical and isotopic signatures of the
dora de Estudos e Projetos), FAPEMIG (Fundação de Amparo à Pesquisa de
Chapada Acauã submarine basalts that also carry inherited
Minas Gerais) and FAPESP (Fundação de Amparo à Pesquisa do Estado de
zircon grains from continental sources. São Paulo). We also thank W. Altermann, M. A. Martins-Neto and an anonymous
reviewer, as well as Emmanuelle Arnaud, for corrections and suggestions. We
The post-glacial succession represents the passive margin stage of dedicate this work to our beloved friend and coauthor – the late Carlos Maurı́cio
the Macaúbas basin, when sea-floor spreading continued at least Noce – who contributed very significantly to understanding the Macaúbas basin.
until c. 660 Ma. This represents a contribution of the IUGS- and UNESCO-funded IGCP (Inter-
The onset of a rifting stage in the region is constrained by national Geoscience Programme) Project #512.
the U –Pb ages of the Pedro Lessa mafic dykes (c. 906 Ma;
Fig. 49.2) and Salto da Divisa anorogenic granites (c. 875 Ma;
see ‘Geochronological constraints’). Indeed, these anorogenic References
magmatic episodes provide solid evidence that an active continen-
tal rift developed during the late Tonian (900 –850 Ma) in this Alkmim, F. F., Marshak, S., Pedrosa-Soares, A. C., Peres, G. G.,
region (Fig. 49.2), and it is certain that at least the pre-glacial Cruz, S. C. & Whittington, A. 2006. Kinematic evolution of the
strata were deposited during this rift stage. The Serra do Catuni Araçuaı́-West Congo orogen in Brazil and Africa: nutcracker tec-
Fm. and the cratonic Jequitaı́ Fm. have also been assigned to the tonics during the Neoproterozoic assembly of Gondwana. Precam-
continental rift stage (Pedrosa-Soares et al. 1992, 2008; Uhlein brian Research, 149, 43 –63.
et al. 1998, 1999; Martins-Neto et al. 2001; Martins-Neto & Alkmim, F. F., Pedrosa-Soares, A. C., Noce, C. M. & Cruz, S. C. P.
Hercos 2002). However, if the regional disconformity between 2007. Sobre a evolução tectônica do Orógeno Araçuaı́-Congo
the pre-glacial and glaciogenic successions represents a significant Ocidental. Geonomos, 15, 35 – 43.
hiatus (Fig. 49.3; see ‘Boundary relations’), the Serra do Catuni Almeida, F. F. M. 1977. O Cráton do São Francisco. Revista Brasileira de
Fm. could be early Cryogenian in age (Uhlein et al. 2007). Geociências, 7, 349–364.
With regards to the age of glaciation, the cratonic Jequitaı́ and Babinski, M., Gradim, R. J., Pedrosa-Soares, A. C., Alkmim, F. F.,
Noce, C. M. & Liu, D. 2005. Geocronologia U– Pb (SHRIMP) e
Carrancas diamictites located in the southern sector of the
Sm–Nd de xistos verdes basálticos do Orógeno Araçuaı́: implicações
nearby São Francisco craton (Fig. 49.1a) should be considered.
para a idade do Grupo Macaúbas. Revista Brasileira de Geociências,
Although the Macaúbas, Jequitaı́ and Carrancas diamictites show 35 (4-supplement), 77– 81.
no direct field relation, they have been correlated with each other Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating of the
since the 1930s, owing to their shared glacial origin. The Sete Lagoas cap carbonate (Bambuı́ Group, Brazil) and implications
maximum sedimentation ages for the Jequitaı́ Fm. (c. 880 Ma; for the Neoproterozoic glacial events. Terra Nova, 19, 401 –406.
Rodrigues 2008) and the oldest Macaúbas diamictites (c. 900 Ma; Babinski, M., Pedrosa-Soares, A. C., Trindade, R. I. F., Martins, M.,
see ‘Geochronological constraints’) are similar, but they do not Noce, C. M. & Liu, D. 2011. Neoproterozoic glacial deposits from
provide any precise information on the glaciation age. On the the Araçuaı́ orogen, Brazil: age, provenance and correlations with
other hand, the Carrancas diamictite can be correlated to an early the São Francisco craton and West Congo belt. Gondwana Research,
Cryogenian glacial event because it is covered by the c. 740 Ma doi: 10.1016/j.gr.2011.04.008.
cap carbonate of the cratonic cover (Babinski et al. 2007, 2011). Chula, A. M., Knauer, L. G. & Almeida-Abreu, P. A. 1996. Estratigra-
In order to fully understand the timing and nature of glaciation fia do Supergrupo Espinhaço na região de Planalto de Minas, Diaman-
on the São Francisco craton, the entire Araçuaı́ –West Congo oro- tina, MG. Geonomos, 3, 69 –81.
genic system and its precursor basin should also be considered Cordani, U. G., Brito-Neves, B. B., D’agrella-Filho, M. S. & Trin-
(Fig. 49.1b). Because these now rifted margins were juxtaposed dade, R. I. F. 2003. Tearing-up Rodinia: the Neoproterozoic paleo-
prior to full opening of the Macaúbas basin, the Brazilian and geography of South American cratonic fragments. Terra Nova, 15,
African counterparts of the precursor basin must have lain in 343– 349.
similar palaeoclimatic zones (e.g. Cordani et al. 2003; Pisarevsky D’Agrella-Filho, M. S., Pacca, I. I. G., Trindade, R. I. F., Teixeira,
et al. 2003). In the West Congo belt, the Lower Mixtite Fm. is the W., Raposo, M. I. B. & Onstott, T. C. 2004. Paleomagnetism and
40
Ar39Ar ages of mafic dykes from Salvador, Brazil: new constraints
diamictite unit that has been related to the early Cryogenian glacial
on the São Francisco craton APW path between 1080 and 1010 Ma.
event (Trompette 1994; Tack et al. 2001; Frimmel et al. 2006).
Precambrian Research, 132, 55– 77.
Pedrosa-Soares et al. (2008) correlated the Lower Mixtite and Dossin, I. A. & Dardenne, M. A. 1984a. Geologia da borda ocidental da
Lower Chapada Acauã formations because they include glacio- Serra do Cipó, Minas Gerais. In: Congresso Brasileiro de Geologia
marine diamictites with intercalations of transitional submarine 33, Rio de Janeiro, Anais, 7. Sociedade Brasileira de Geologia,
basalts, representing a late rift stage. These authors also correlated 3104–3117.
the Lower Chapada Acauã and Lower Mixtite formations with the Dossin, I. A. & Dardenne, M. A. 1984b. Os depósitos supergênicos de
Carrancas diamictite, and ascribed them to the early Cryogenian manganês da borda ocidental da Serra do Cipó, MG (Quadrı́cula
glacial event. The Upper Mixtite Fm. in the West Congo Belt, Inhame). In: Congresso Brasileiro de Geologia 33, Rio de Janeiro,
which is ascribed to the late Cryogenian glacial event (Frimmel Anais, 3. Sociedade Brasileira de Geologia, 1129–1143.
et al. 2006), does not appear to have a counterpart in the Macaúbas Drumond, J. B. V. 2000. Folhas Cordeiros, Belo Campo e Curral de
Group (Pedrosa-Soares et al. 2008). Dentro. CODEMIG, Projeto Leste, Belo Horizonte.
532 A. C. PEDROSA-SOARES ET AL.

Fogaça, A. C. C. 1997. Geologia da Folha Diamantina. In: Grossi-Sad, Knauer, L. G. & Grossi-Sad, J. H. 1997. Geologia da Folha Presidente
J. H., Lobato, L. M., Pedrosa-Soares, A. C. & Soares-Filho, Kubtischek. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa-Soares,
B. S. (eds) Projeto Espinhaço. CODEMIG, Belo Horizonte, A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. CODEMIG,
1575– 1666. Belo Horizonte, 1902– 2056.
Frimmel, H. E., Tack, L., Basei, M., Nutman, A. & Boven, A. 2006. Knauer, L. G., Silva, L. L., Souza, F. B. B., Silva, L. R. & Carmo, R. C.
Provenance and chemostratigraphy of the Neoproterozoic West 2006. Folha Monte Azul 1:100.000. CPRM, Programa Geologia do
Congolian Group in the Democratic Republic of Congo. Journal Brasil, Brası́lia.
of African Earth Sciences, 46, 221–239. Lima, S. A. A., Martins-Neto, M. A., Pedrosa-Soares, A. C., Cordani,
Gradim, R. J., Alkmim, F. F., Pedrosa-Soares, A. C., Babinski, M. & U. G. & Nutman, A. 2002. A Formação Salinas na área-tipo, NE de
Noce, C. M. 2005. Xistos verdes do Alto Araçuaı́, Minas Gerais: Minas Gerais: uma proposta de revisão da estratigrafia da Faixa
vulcanismo básico do rifte neoproterozóico Macaúbas. Revista Araçuaı́ com base em evidências sedimentares, metamórficas e
Brasileira de Geociências, 35 (4-suplemento), 59 –69. idades U– Pb SHRIMP. Revista Brasileira de Geociências, 32,
Gravenor, C. P. & Monteiro, R. 1983. Ice-thrust features in the Proter- 491– 500.
ozoic Macaúbas Group, Jequitaı́ area, Minas Gerais, Brazil. Journal Machado, N., Schrank, A., Abreu, F. R., Knauer, L. G. &
of Geology, 91, 113– 116. Almeida-Abreu, P. A. 1989. Resultados preliminares da geocrono-
Grossi-Sad, J. H., Lobato, L. M., Pedrosa-Soares, A. C. & logia U– Pb na Serra do Espinhaço Meridional. Boletim do Núcleo
Soares-Filho, B. S. 1997a. Projeto Espinhaço. CODEMIG, Belo Minas Gerais-Sociedade Brasileira de Geologia, 10, 171–174.
Horizonte. Marshak, S., Alkmim, F. F., Whittington, A. & Pedrosa-Soares,
Grossi-Sad, J. H., Roque, N. C., Knauer, L. G., Noce, C. M. & A. C. 2006. Extensional collapse in the Neoproterozoic Araçuaı́
Fonseca, E. 1997b. Geologia da Folha Carbonita. In: Grossi-Sad, orogen, eastern Brazil: a setting for reactivation of asymmetric crenu-
J. H., Lobato, L. M., Pedrosa-Soares, A. C. & Soares-Filho, lation cleavage. Journal of Structural Geology, 28, 129–147.
B. S. (eds) Projeto Espinhaço. CODEMIG, Belo Horizonte, Martins, M. S. 2006. Geologia dos diamantes e carbonados aluvionares
1251–1372. da bacia do Rio Macaúbas, MG. PhD thesis, Instituto de Geociências,
Guimarães, M. L. V. 1997. Geologia da Folha Botumirim. In: Grossi- Universidade Federal de Minas Gerais, Belo Horizonte.
Sad, J. H., Lobato, L. M., Pedrosa-Soares, A. C. & Soares-Filho, Martins, M. S., Karfunkel, J., Noce, C. M., Babinski, M., Pedrosa-
B. S. (eds) Projeto Espinhaço. CODEMIG, Belo Horizonte, Soares, A. C., Sial, A. N. & Lyu, D. 2008. A seqüência pré-glacial
543– 610. do Grupo Macaúbas na área-tipo e o registro da abertura do rifte
Guimarães, M. L., Crocco-Rodrigues, F. A., Abreu, F. R., Oliveira, Araçuaı́. Revista Brasileira de Geociências, 38, 768– 779.
O. A. B. & Greco, F. M. 1993. Geologia do bloco Itacambira-Monte Martins-Neto, M. A. & Hercos, C. M. 2002. Sedimentation and tectonic
Azul entre Barrocão e Porteirinha, MG. Boletim do Núcleo Minas setting of Early Neoproterozoic glacial deposits in southeastern
Gerais-Sociedade Brasileira de Geologia, 12, 74– 78. Brazil. In: Altermann, W. & Corcoran, P. L. (eds) Precambrian
Guimarães, M. L. V., Grossi-Sad, J. H. & Fonseca, E. 1997. Geologia da Sedimentary Environments: A Modern Approach to Ancient Deposi-
Folha Francisco Sá. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa- tional Systems. International Association of Sedimentologists,
Soares, A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. Special Publications, 33, 383– 403.
CODEMIG, Belo Horizonte, 223– 313. Martins-Neto, M. A., Gomes, N. S., Hercos, C. M. & Reis, L. A. 1999.
Hettich, M. 1975. Zur stratigraphie und Genese des Macaúbas nordlich Fácies glácio-continentais (outwash plain) na Megasseqüencia
der Serra Negra, Espinhaço-Zone, Minas Gerais, Brasilien. Geolo- Macaúbas (Serra da Água Fria, MG) e seu contexto geotectônico.
gisches Jahrbuch, 14, 47 –85. Revista Brasileira de Geociências, 29, 179–188.
Hettich, M. 1977. A glaciação proterozóica no centro-norte de Minas Martins-Neto, M. A., Pedrosa-Soares, A. C. & Lima, S. A. A. 2001.
Gerais. Revista Brasileira de Geociências, 7, 87– 101. Tectono-sedimentary evolution of sedimentary basins from Late
Hettich, M. & Karfunkel, J. 1978. Um esker, um varvito e seixos Paleoproterozoic to Late Neoproterozoic in the São Francisco
estriados no Grupo Macaúbas, norte de Minas Gerais. Revista craton and Araçuaı́ fold belt, eastern Brazil. Sedimentary Geology,
Escola de Minas de Ouro Preto, 34, 5– 8. 141/142, 343–370.
Isotta, C. A. L., Rocha-Campos, A. C. & Yoshida, R. 1969. Striated Moraes, L. J. 1929. Geologia da região diamantina de Minas Gerais. In:
pavement of the Upper Precambrian glaciation in Brazil. Nature, Relatório Anual do Diretor 1928, Rio de Janeiro. Serviço Geológico
222, 466– 468. e Mineralógico, 29 – 34.
Karfunkel, J. & Karfunkel, B. 1976. Estudos petro-faciológicos do Moraes, L. J. 1932. Área ocupada pela Formação Macaúbas no
Grupo Macaúbas na porção mediana da Serra do Espinhaço, Minas norte de Minas Gerais. Anais Academia Brasileira de Ciências, 4,
Gerais. In: Congresso Brasileiro de Geologia 29, Ouro Preto, 111– 114.
Anais, 2. Sociedade Brasileira de Geologia, 179 – 188. Moraes, L. J. & Guimarães, D. 1930. Geologia da região diamantı́fera do
Karfunkel, J. & Karfunkel, B. 1977. Fazielle Entwicklung der mittle- norte de Minas Gerais. Anais Academia Brasileira de Ciências, 2,
ren Espinhaço-Zone mit besonderer Berücksichtigung des Tillit- 153– 186.
Problems (Minas Gerais, Brasilien). Geologisches Jahrbuch, 24, Moraes, L. J. & Guimarães, D. 1931. The diamond-bearing
3 –91. region of Northern Minas Gerais, Brazil. Economic Geology, 26,
Karfunkel, J. & Hoppe, A. 1988. Late Precambrian glaciation in central- 502– 530.
eastern Brazil: synthesis and model. Palaeogeography, Palaeoclima- Mourão, M. A. A. & Pedrosa-Soares, A. C. 1992. Evidências de sedi-
tology, Palaeoecology, 65, 1– 21. mentação turbidı́tica no Grupo Macaúbas, Faixa Araçuaı́. Revista
Karfunkel, J., Moreira, P. C., Ribeiro, M. C. & Franco, A. L. 1984. Escola de Minas de Ouro Preto, 45, 94– 96.
Aspectos genéticos e deposicionais do Grupo Macaúbas na região Mourão, M. A. A. & Grossi-Sad, J. H. 1997. Geologia da Folha Padre
da barragem do Paraúna e sua importância na contribuição para um Carvalho. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa-Soares,
modelo paleogeográfico e geotectônico. In: Congresso Brasileiro A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. CODEMIG,
de Geologia 33, Rio de Janeiro, Anais, 7. Sociedade Brasileira de Belo Horizonte, 316– 418.
Geologia, 3091– 3103. Mourão, M. A. A., Grossi-Sad, J. H. & Fonseca, E. 1997. Geologia da
Karfunkel, J., Pedrosa-Soares, A. C. & Dossin, I. A. 1985. O Grupo Folha Janaúba. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa-
Macaúbas em Minas Gerais: revisão dos conhecimentos. Boletim Soares, A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço.
do Núcleo Minas Gerais-Sociedade Brasileira de Geologia, 5, CODEMIG, Belo Horizonte, 11 – 123.
45 –59. Noce, C. M. 1997a. Geologia da Folha Curimataı́. In: Grossi-Sad, J. H.,
Karfunkel, J., Hoppe, A. & Noce, C. M. 2002. Serra da Água Fria e Lobato, L. M., Pedrosa-Soares, A. C. & Soares-Filho, B. S. (eds)
vizinhancas, MG: vestı́gios de glaciação neoproterozóica. In: Schob- Projeto Espinhaço. CODEMIG, Belo Horizonte, 1207– 1250.
benhaus, C., Campos, D. A., Queiroz, E. T., Winge, M. & Noce, C. M. 1997b. Geologia da Folha Itacambira. In: Grossi-Sad, J. H.,
Berbert-Born, M. (eds) Sı́tios Geológicos e Paleontológicos do Lobato, L. M., Pedrosa-Soares, A. C. & Soares-Filho, B. S. (eds)
Brasil, 1. DNPM, Brası́lia, Brazil, 165– 173. Projeto Espinhaço. CODEMIG, Belo Horizonte, 853–926.
NEOPROTEROZOIC MACAÚBAS GROUP 533

Noce, C. M., Pedrosa-Soares, A. C. et al. 1997. Nova subdivisão estra- Queiroga, G. N., Pedrosa-Soares, A. C., Quéméneur, J. & Casta-
tigráfica regional do Grupo Macaúbas na Faixa Araçuaı́: o registro de ñeda, C. 2006. A unidade metassedimentar do ofiolito de Ribeirão
uma bacia neoproterozóica. Boletim do Núcleo Minas Gerais- da Folha, Orógeno Araçuaı́, Minas Gerais: petrografia, geotermobar-
Sociedade Brasileira de Geologia, 14, 29 –31. ometria e calcografia. Geonomos, 14, 9 –12.
Oliveira, M. J. R., Grossi-Sad, J. H., Romano, A. W. & Lobato, L. M. Queiroga, G. N., Pedrosa-Soares, A. C. et al. 2007. Age of the
1997a. Geologia da Folha Grão Mogol. In: Grossi-Sad, J. H., Ribeirão da Folha ophiolite, Araçuaı́ Orogen: the U– Pb zircon
Lobato, L. M., Pedrosa-Soares, A. C. & Soares-Filho, B. S. dating of a plagiogranite. Geonomos, 15, 61 – 65.
(eds) Projeto Espinhaço. CODEMIG, Belo Horizonte, 611–642. Rocha-Campos, A. C. & Hasui, Y. 1981. Tillites of the Macaúbas
Oliveira, M. J. R., Fogaça, A. C. C. & Fonseca, E. 1997b. Geologia da Group (Proterozoic) in central Minas Gerais and southern Bahia,
Folha Baldim. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa- Brazil. In: Hambrey, M. J. & Harland, W. B. (eds) Earths’s
Soares, A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. Pre-Pleistocene Glacial Record. Cambridge University Press,
CODEMIG, Belo Horizonte, 2437– 2531. 933– 939.
Pedrosa-Soares, A. C. & Grossi-Sad, J. H. 1997. Geologia da Folha Rodrigues, J. B. 2008. Proveniência de sedimentos dos grupos Canastra,
Minas Novas. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa- Ibiá, Vazante e Bambuı́: um estudo de zircões detrı́ticos e idades-
Soares, A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. modelo Sm –Nd. PhD thesis, Instituto de Geociências, Universidade
CODEMIG, Belo Horizonte, 925– 1053. de Brası́lia, Brazil.
Pedrosa-Soares, A. C. & Oliveira, M. J. R. 1997. Geologia da Folha Rogers, J. W. & Santosh, M. 2004. Continents and Supercontinents.
Salinas. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa-Soares, Oxford University Press.
A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. CODEMIG, Roque, N. C., Grossi-Sad, J. H., Noce, C. M. & Fonseca, E. 1997. Geo-
Belo Horizonte, 419– 541. logia da Folha Rio Pardo de Minas. In: Grossi-Sad, J. H., Lobato,
Pedrosa-Soares, A. C., Leonardos, O. H. & Correia-Neves, J. M. L. M., Pedrosa-Soares, A. C. & Soares-Filho, B. S. (eds) Projeto
1984. Aspectos metamórficos de seqüências supracrustais da Faixa Espinhaço. CODEMIG, Belo Horizonte, 126–221.
Araçuaı́ em Minas Gerais. In: Congresso Brasileiro de Geologia Santos, R. V., Alvarenga, C. et al. 2004. Carbon isotopes of Mesopro-
33, Rio de Janeiro, Anais, 6. Sociedade Brasileira de Geologia, terozoic –Neoproterozoic sequences from southern São Francisco
3056– 3065. craton and Araçuaı́ Belt: paleogeographic implications. Journal of
Pedrosa-Soares, A. C., Noce, C. M., Vidal, P., Monteiro, R. & Leo- South American Earth Sciences, 18, 27 –39.
nardos, O. H. 1992. Toward a new tectonic model for the Late Pro- Schöll, W. U. 1972. Der südwestliche Randbereich der Espinhaço
terzoic Araçuaı́ (SE Brazil) — West Congolian (SW Africa) Belt. Zone, Minas Gerais, Brasilien. Geologische Rundschau, 61,
Journal of South American Earth Sciences, 6, 33 –47. 201– 216.
Pedrosa-Soares, A. C., Baars, F. J., Lobato, L. M., Magni, M. C. V. & Schöll, W. U. 1976. Estromatólitos Conophyton em dolomitos do Grupo
Faria, L. F. 1993. Arquitetura tectono-metamórfica do setor central Macaúbas. In: Congresso Brasileiro de Geologia 29, Ouro Preto,
da Faixa Araçuai. Boletim do Núcleo Minas Gerais-Sociedade Brasi- Anais, 2. Sociedade Brasileira de Geologia, 67– 73.
leira de Geologia, 12, 176– 182. Schrank, A., Dourado, B. V. & Biondi, J. C. 1978. Estudo preliminar
Pedrosa-Soares, A. C., Leonardos, O. H., Ferreira, J. C. & Reis, L. B. dos metavulcanitos do Grupo Macaúbas na região do Alto Jequitin-
1996. Duplo regime metamórfico na Faixa Araçuaı́: re- honha, MG. In: Congresso Brasileiro de Geologia 30, Recife,
interpretação à luz de novos dados. In: Congresso Brasileiro Anais, 3. Sociedade Brasileira de Geologia, 1323–1335.
de Geologia 39, Salvador, Anais, 6. Sociedade Brasileira de Geolo- Silva, L. C., Mcnaughton, N. J., Armstrong, R., Hartmann, L. &
gia, 5– 8. Fletcher, I. 2005. The Neoproterozoic Mantiqueira Province and
Pedrosa-Soares, A. C., Vidal, P., Leonardos, O. H. & Brito-Neves, its African connections. Precambrian Research, 136, 203– 240.
B. B. 1998. Neoproterozoic oceanic remnants in eastern Brazil: Silva, L. C., Pedrosa-Soares, A. C. & Teixeira, L. R. 2008. Tonian
further evidence and refutation of an exclusively ensialic evolution rift-related, A-type continental plutonism in the Araçuaı́ orogen,
for the Araçuaı́-West Congo orogen. Geology, 26, 519– 522. Eastern Brazil: new evidences for the breakup stage of the São Fran-
Pedrosa-Soares, A. C., Cordani, U. & Nutman, A. 2000. Constraining cisco– Congo Paleocontinent. Gondwana Research, 13, 527– 537.
the age of Neoproterozoic glaciation in eastern Brazil: first U –Pb Tack, L., Wingate, M. T .D., Liégeois, J. P., Fernandez-Alonso, M.
(SHRIMP) data from detrital zircons. Revista Brasileira de Geociên- & Deblond, A. 2001. Early Neoproterozoic magmatism
cias, 30, 58 – 61. (1000–910 Ma) of the Zadinian and Mayumbian Groups
Pedrosa-Soares, A. C., Noce, C. M., Wiedemann, C. M. & Pinto, C. P. (Bas-Congo): onset of Rodinian rifting at the western edge of
2001. The Araçuaı́-West Congo orogen in Brazil: an overview of a the Congo craton. Precambrian Research, 110, 277–306.
confined orogen formed during Gondwanland assembly. Precam- Tait, J., Delpomdor, F., Préat, A., Straathof, G., Kanda, V. & Tack,
brian Research, 110, 307–323. L. 2011. Neoproterozoic Sequences of the West Congo and Lindi/
Pedrosa-Soares, A. C., Noce, C. M., Alkmim, F. F., Silva, L. C., Ubangi Supergroups in the Congo Craton, central Africa. In:
Babinski, M., Cordani, U. & Castañeda, C. 2007. Orógeno Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
Araçuaı́: sı́ntese do conhecimento 30 anos após Almeida 1977. Geo- logical Record of Neoproterozoic Glaciations. Geological Society,
nomos, 15, 1– 16. London, Memoirs, 36, 185–194.
Pedrosa-Soares, A. C., Alkmim, F. F., Tack, L., Noce, C. M., Babinski, Trompette, R. 1994. Geology of Western Gondwana (2000 –500 Ma).
M., Silva, L. C. & Martins-Neto, M. A. 2008. Similarities and Pan-African–Brasiliano aggregation of South America and Africa.
differences between the Brazilian and African counterparts of the A. A. Balkema, Rotterdam.
Neoproterozoic Araçuaı́-West Congo orogen. In: Pankhurst, R. J., Tupinambá, M. & Grossi-Sad, J. H. 1997. Geologia da Folha Rio Ver-
Trouw, R. A. J., Brito Neves, B. B. & De Wit, M. J. (eds) West melho. In: Grossi-Sad, J. H., Lobato, L. M., Pedrosa-Soares,
Gondwana: Pre-Cenozoic Correlations Across the South Atlantic A. C. & Soares-Filho, B. S. (eds) Projeto Espinhaço. CODEMIG,
Region. Geological Society, London, Special Publications, 294, Belo Horizonte, 1667–1806.
153– 172. Uhlein, A. & Chaves, M. S. C. 1989. Geologia da borda nordeste da Serra
Pflug, R. & Schöll, W. U. 1975. Proterozoic glaciation in Eastern do Espinhaço (região de Mendanha a São Gonçalo do Rio Preto, MG).
Brazil: a review. Geologische Rundschau, 64, 287– 299. Boletim do Núcleo Minas Gerais-Sociedade Brasileira de Geologia,
Pisarevsky, S. A., Wingate, M. T. D., Powell, C. Mca., Johnson, S. & 10, 175– 179.
Evans, D. A. D. 2003. Models of Rodinia assembly and fragmentation. Uhlein, A., Trompette, R. & Egydio-Silva, M. 1998. Proterozoic
In: Yoshida, M., Windley, B. & Dasgupta, S. (eds) Proterozoic East rifting and closure, SE border of the São Francisco craton, Brazil.
Gondwana: Supercontinent Assembly and Breakup. Geological Journal of South American Earth Sciences, 11, 191– 203.
Society, London, Special Publications, 206, 35–55. Uhlein, A., Trompette, R. & Alvarenga, C. 1999. Neoproterozoic
Porada, H. 1989. Pan-African rifting and orogenesis in southern to glacial and gravitational sedimentation on a continental rifted
equatorial Africa and Eastern Brazil. Precambrian Research, 44, margin: the Jequitaı́-Macaúbas sequence (Minas Gerais, Brazil).
103–136. Journal of South American Earth Sciences, 12, 435– 451.
534 A. C. PEDROSA-SOARES ET AL.

Uhlein, A., Trompette, R., Egydio-Silva, M. & Vauchez, A. 2007. A oeste de Salinas, ME. MSc thesis, Instituto de Geosciências, Univer-
glaciação Sturtiana (750 Ma), a estrutura do rifte Macaúbas-Santo sidade Federal de Minas Gerais, Belo Horizonte.
Onofre e a estratigrafia do Grupo Macaúbas, Faixa Araçuaı́. Geono- Viveiros, J. F. M. & Walde, D. 1976. Geologia da Serra do Cabral, Minas
mos, 15, 45 – 60. Gerais, Brasil. Münsterische Forschungshefte Geologie und Palaeon-
Uhlein, A., de Alvarenga, C. J. S., Dardenne, M. A. & Tromp- tologie, 38/39, 15 –27.
ette, R. R. 2011. The Glaciogenic Jequitaı́ Formation, Viveiros, J. F. M., Sá, E. L., Vilela, O. V., Santos, O. M., Moreira,
Southeastern Brazil. In: Arnaud, E., Halverson, G. P. & J. M. P., Holder-Neto, F. & Vieira, V. S. 1979. Geologia dos vales
Shields-Zhou, G. (eds) The Geological Record of Neoprotero- dos rios Peixe Bravo e Alto Vacaria, norte de Minas Gerais. Boletim
zoic Glaciations. Geological Society, London, Memoirs, 36, do Núcleo Minas Gerais-Sociedade Brasileira de Geologia, 1, 75– 87.
541– 546. Walde, D. 1976. Neue Hinweise für eine proterozoische Vereisung in
Vilela, F. T. 2010. Caracterização de metadiamictitos ferruginosos da Ostbrasilien. Münsterische Forschungshefte Geologie und Palaeon-
Formação Nova Aurora (Grupo Macaúbas, Orógeno Araçuail) a tologie, 38/39, 47 –59.
Chapter 50

Moema laminites: a newly recognized Neoproterozoic (?) glaciogenic


unit, São Francisco Basin, Brazil

ANTONIO C. ROCHA-CAMPOS1*, BENJAMIN B. DE BRITO NEVES1, MARLY BABINSKI1, PAULO R. DOS


SANTOS1, SÔNIA M.B. DE OLIVEIRA1 & A. ROMANO2
1
Instituto de Geociências, Universidade São Paulo, Rua do Lago, 562, Cidade Universitária, São Paulo, SP, Brazil
2
Instituto de Geociências, Universidade Federal de Minas Gerais, Belo Horizonte, MG, Brazil
*Corresponding author (e-mail: acrcampo@usp.br)

Abstract: A recently identified diamictite and silt-clay laminite, which discordantly overlie the Archaean basement and underlie the
Neoproterozoic Bambui Group, have been informally named as the Moema Laminites. They are preserved at the southwestern
margin of the São Francisco Basin in southeastern Brazil, and are widely distributed in the central-western Minas Gerais state. They
crop out discontinuously over at least 140 km along a north–south direction. The nomenclature of and stratigraphic relationships
between the Moema Laminites and other isolated Neoproterozic occurrences of similar rocks are in a state of flux. Two exceptionally
good exposures of the Moema Laminites show good evidence for deposition under glacial conditions. At the Formiga locality, a
single glacial advance is registered by a deformation tillite, while overlying laminite records deposition in a post-glacial, probably
marine basin following deglaciation. At the SAFFRAN quarry, striations on a bedding plane may have been caused by floating sea-
ice that just touched the bottom of the basin. Much additional work is needed to establish relationships between the Moema Laminites
and other similar occurrences. If these and Moema Laminites are shown to be Cryogenian glacial deposits, the area covered by the
Cryogenian glaciations in the São Francisco basin is much larger than formerly believed.

The name Moema Laminites is being used informally to designate Stratigraphy


a newly identified, possibly Neoproterozoic unit, made up of dia-
mictite and silty-clayey laminites. The strata, so far known at The sparcity of outcrops (mostly road cuts), moderate topography
reconnaissance level, crop out extensively in the central-western and deep weathering make it difficult to log a complete represen-
State of Minas Gerais (Fig. 50.1), below carbonate assigned to tative stratigraphic section. A regional field reconnaissance was
the Sete Lagoas Formation (Fm.) of the Bambuı́ Group (Neoproter- performed along several roads in the area between the towns of
ozoic) (Fig. 50.2). Cristais and Bom Despacho, in order to evaluate the distribution
The Moema Laminites crop out discontinuously over a large and stratigraphic relations of the studied unit (Fig. 50.1). This
portion of the southern São Francisco Basin, from the latitude of was complemented by more detailed stratigraphic and sedimento-
the town of Onça do Pitangui (198430 S, 448W), southward for c. logical study of critical outcrops.
140 km, at least as far as the town of Cristais (208520 S, 45830 W) Laterally, the thickness of the diamictite varies from 30 cm on
(Fig. 50.1). the Formiga outcrop to, more commonly, c. 5 m, and may be
Two exceptionally good exposures of the strata are found at a missing in places. In other places, the unit is represented only by
road cut of BR-354, near the town of Formiga, 1 km north of the remains of the basal diamictite, the upper laminites having been
intersection with road MG-050, in the southern part of the study eroded away. The thickness of the overlying laminites reaches at
area, and at the SAFFRAN quarry (named after the owner least 60 m in the area of the town of Onça do Pitangui (Fig. 50.1).
company), some 100 km north of the first locality, in the vicinity Strata below and above the lower nonconformity are deeply
of Onça do Pitangui, 10 km north of road BR-262. weathered. Kaolinitized regolith often reaches considerable thick-
The thickest exposure of the unit is in the vicinity of the town of ness on top of Archaean basement in the study area and may record
Moema (Fig. 50.1), but because the name is provisional, no formal a significant and intense phase of pre-Neoproterozoic weathering,
type-section has yet been designated. superposed by widespread recent alteration.
The stratigraphic relations of the Moema Laminites are similar
Structural framework to that of other isolated outcrops of conglomerate/diamictite and
associated rocks that are found below carbonate rocks of the Neo-
The Moema Laminites are part of the sedimentary cover of the proterozoic Sete Lagoas Fm. (Bambuı́ Group). These are known
intracratonic São Francisco Basin, in the central-western State of from several places, fringing the western margin of the São Fran-
Minas Gerais. They occur immediately below the Neoproterozoic cisco Basin (e.g. Member/Facies Carrancas, Costa et al. 1961;
Bambui Group along the southwestern margin of the basin. Facies Samburá, Miranda 1953, and so on) and are traditionally
In this area, the horizontal to gently folded platformal cover of correlated with the Jequitaı́ Fm. The latter unit, of well established
the basin fills a depression in the cratonic foreland, located west glacial origin (Isotta et al. 1969), crops out some 400 km north of
of the Sete Lagoas High (Fig. 50.1). Deposition of the Moema sedi- the study area (Fig. 50.1; see Dardenne et al. (1978) and Rocha-
ments was contemporaneous with the sag phase of tectonic evol- Campos & Hasui (1981) for a fuller account of these occurrences).
ution of the basin (Alkmin 2004). At present, there are inconsistencies in the interpretation of the
Outcrops of the Moema Laminites are, in general, preserved in nomenclature, stratigraphic relations and depositional setting of
topographic lows on the irregular surface of the cratonic basement. the strata lying below the Sete Lagoas Fm. (Bambuı́ Group),
Strata are essentially horizontal but have been slightly affected by including the Moema Laminites.
normal faults. Near Onça do Pitangui they are part of an east –west Pierre Muzzi Magalhães, for example, in his Masters disser-
trending syncline with plunges up to 308E. tation of 1989, recognizes a Conglomeratic Facies (polymitic)
Rocks of the unit are essentially non-metamorphic. and a Pellitic Facies (argillite and siltstone), below carbonate

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 535– 540. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.50
536 A. C. ROCHA-CAMPOS ET AL.

Fig. 50.1. Geological map of the São Francisco


Basin, Southeastern Brazil, showing location of
the best exposures of the Moema Laminites.
Squares: 1, Formiga outcrop; 2, SAFFRAN
quarry. SFC, São Francisco Craton.

rocks of the Sete Lagoas Fm., that are probably equivalent to the Formiga area
Moema Laminites, as a basal unit of the Bambuı̀ Group
(Fig. 50.2). A glacial influence during deposition of these rocks A cut on the northeastern side of the road BR-364 exposes a lower,
was, however, not recognized. discontinuous, thin diamictite of variable thickness (30 –40 cm),
On recent geological maps from the southern part of the São resting non-conformably on topographic highs of Archaean ortho-
Francisco Basin (SEME/COMIG/ CPRM, 2003; CPRM, 2007), gneiss or on its regolith, followed with a sharp contact by 15 m of
rocks considered by us to be part of the Moema Lamites have laminites. The smooth surface of the gneiss bears fine, poorly pre-
been called the Serra de Santa Helena Fm. and Samburá Fm. served striae (Fig. 50.3).
The former is a pelitic unit lying above the Sete Lagoas Fm., Isolated or semi-detached round boulders of up to 1.5 m lie on
and the latter corresponds to one of the infra-Bambuı́ Group dia- top of the orthogneiss and regolith and seem to correspond to core-
mictite occurrences. As pointed out by Dardenne & Castro stones (Fig. 50.3). Deep weathering of local rocks has altered their
(2000), the Samburá diamictite is a fault-controlled, fan-delta colours, thus enhancing internal features.
deposit that occurs at different levels of the Bambui Group. The diamictite corresponds to a layer of tectonically deformed
Under these circumstances, the use of a provisional name seems regolith of the local basement, containing many centimetre-
more advisable. to-decimetre clasts, predominantly of the local gneiss, but also
of two different lithologies (fine and coarse quartz schist), in a
sandy-silty-clayey matrix. Deformation features recognized in
Glaciogenic deposits and associated strata the diamictite include boudins, stretched clasts, attenuated and
transposed folds and clast alignment. Small thrust faults cut the
The sedimentary characteristics of the Moema Laminites are based top of the gneissic basement (Fig. 50.3).
on the best exposures in the region: the Formiga area and the The contact between the deformed bed and overlying laminites
SAFFRAN quarry. seems to be sharp. Two subunits are recognized, a lower one of
MOEMA LAMINITES 537

Thin laminations are present in the lower subunit. Normal


Sandstones graded bedding was observed in centimetre-thick laminae of the
Serra de Santa Helena Fm.

upper subunit, as was evidence of local disharmonic, soft sediment


Mudstones and Siltstones
folding.
Stromatolitic structures Unequivocal evidence of dropstones was not found in the
laminites.
Dolomitic calcarenite

Calcarenite with parallel


SAFFRAN quarry
and cross bedding
Bambuí Group

Calcilutite with parallel At least 60 m of thin (cm) silty-clayey laminites of kaolinitic com-
and cross bedding position and interbedded, irregular, laterally discontinuous, ferru-
Sete Lagoas Fm.

Siltstone ginous (hematite) bands are exposed on the extraction faces of the
quarry (Fig. 50.4). A diamictite bed occurs under the laminites
nearby, but its contact is not exposed. The upper boundary with
the laminites was not seen in this area.
200 m The section is made up of a lower interval where ferruginous
beds are more common and an upper thicker, whiteish one,
where they are rare (Fig. 50.4).
The two intervals differ also in their sedimentary structures.
Whereas the lower interval is characterized by intercalation of
Basement 0
thin silt-clay lenses (starved ripples?) and normal graded and non-
graded laminae, the whiteish clay bands of the upper interval are
Fig. 50.2. Stratigraphy of the Bambuı́ Group in Southern São Francisco Basin.
mostly thinly laminated.
Conglomeratic (not shown) and pellitic facies (siltstone) at the base of the
Bambuı́ Group is thought to correspond to the Moema Laminites (Muzzi
A bedding plane at the bottom of the quarry contains two areas
Magalhães, pers. comm. 1989).
bearing sets of parallel, few-metres-long, fine striae and furrows
(1 –3 cm wide). The features are rectilinear or curved and their
orientation varies from 2898N to 3398N. On the same surface,
dark laminites, 2 m thick, overlain with a sharp contact by a section there are a few examples of striated areas surrounded on both
of light laminites, 12 m thick (Fig. 50.3). sides by more elevated non-striated ones on the same bedding
The lower interval seems to fill a wide and shallow depression plane. Rare features, such as masses of sediment laterally covering
on top of the deformed bed and orthogneiss. Laminae are deformed furrows (Rocha-Campos et al. 1994), suggest that the bed was not
around corestones of weathered gneiss. consolidated when the striae were produced.

Fig. 50.3. Partial diagrams (a and b) of the


Formiga outcrop. Deformation, thought to be
glaciotectonic in origin includes stretched
clasts, attenuated folds, boudins and
comminution (A –C). 1, Orthogneiss
(Archaean); 2, regolith; 3, deformed bed
(regolith and lower dark laminites); 4, upper
light laminites. Note lower dark laminites (unit
3) in (a) are undeformed on the left side of the
outcrop.
538 A. C. ROCHA-CAMPOS ET AL.

Top

//
Explanation
2 Laminites

Ferruginous layer
1
Interbedded laminites
0 and ferruginous layers
.m
Fractures/faults
Fig. 50.4. Stratigraphic log of laminites at the
Base Striae SAFFRAN quarry. showing distribution of
ferruginous layer (in black).

Boundary relations with overlying and underlying contribute to the understanding of their origin. This involved deter-
non-glacial units mination of major (FRX) and trace elements (FRX and ICP-MS),
REE (ICP-MS), content of Fe2þ and identification of mineralogi-
Rocks of the Moema Laminites rest non-conformably on an cal phases by X-ray diffraction. REE data were normalized to
erosive surface formed on top of predominantly granitic terranes NASC (North American Shale Composite).
of the Archaean basement of the São Francisco Basin (Fig. Clastic components of the laminites (silt-clay and kaolinite)
50.2). The upper contact with the overlying Sete Lagoas Fm. present a homogeneous composition and chemical signature of
carbonates was not directely observed in the studied area. more abundant weathering-resistant elements (Cr, Ga, Hf, Rb Sc,
Ta, Th and Zr), in comparison with the ferruginous layers (up to
48% hematite). Both clastic and ferruginous layers depict a nega-
Chemostratigraphy tive anomaly for Eu and a positive anomaly for Ce.

There are no known carbonate rocks associated directly with the


Moema Laminites and, as such, no chemostratigraphic studies of Other characteristics
C, O or Sr have been undertaken to date.
Eight samples of laminites from the two areas studied were geo- Laminites from the SAFFRAN Quarry are being exploited as a
chemically analysed to better characterize the two rocks and to source of kaolinite for the ceramic industry.
MOEMA LAMINITES 539

Palaeolatitude and palaeogeography of basement highs. Deformation seems to have been associated
with lodgement of clasts.
There are no palaeomagnetic data from the Moema Laminites. The glacier may have moved towards the east over the weath-
Palaeomagnetic studies carried out previously on carbonate units ered, deformable bed, in a glacial terrestrial setting that was sub-
of the São Francisco Basin revealed that a continental-scale fluid sequently inundated following glacier retreat and deglaciation. A
event could have been responsible for the resetting of the palaeo- second possibility is that a grounded glacier advanced into a pre-
magnetic, as well as the isotopic system, at c. 520 Ma (D́Agrella existing water-filled basin, where laminites had accumulated.
Filho et al. 2000; Trindade et al. 2004). Recently, however, The fact that the laminites were not involved in the deformation
Babinski et al. (2007) referred to good-quality palaeomagnetic and that the glacier caused deformation of the regolith favours
poles obtained from rocks of the Congo –São Francisco craton, the first hypothesis (Fig. 50.3).
penecontemporaneous with the Sete Lagoas Fm., that would Deposition of the laminites was probably through settling of silt
place the southern São Francisco craton at low-intermediate and clay, from suspended plumes fed by the glacier, as well as by
palaeolatitudes (20– 308). turbidity or other types of density currents during the deglaciation
phase. Instability of the unconsolidated deposit led to frequent
soft-sediment deformation. Lack (or rarity ?) of dropstones in
Geochronological constraints
the laminites may indicate that the glacial terminus was far
removed from the basin margin or the intervention of mechanisms
No direct radiometric data are yet available to constrain the deposi-
that either prevent the occurrence of icebergs or result in their fast
tional age of the Moema Laminites. However, the regional strati-
evacuation, as described from present glacial estuarine environ-
graphic position of other units with similar lithologies below
ments (Dowdeswell et al. 2000).
carbonate beds of the Sete Lagoas Fm. along all sections exam-
Sedimentation processes identified in the two localities studied
ined, and evidence of a glacial influence in the Moema Laminites,
seem to have been grossly similar, as shown by textures and sedi-
are suggestive of correlation between the Moema Laminites and
mentary structures present.
the Jequitaı́ Fm., which has a cap carbonate dated at 740 + 22 Ma
No ice-contact features were found at the SAFFRAN quarry.
(Babinski et al. 2007).
Interpretation of glacial influence in the area relies only on the pre-
Although the Moema Laminites were deposited on top of the
sumed origin of the striated laminites by erosion of floating ice
Archaean basement of the São Francisco Basin at both localities
masses touching the bottom of the sedimentary basin.
discussed (Teixeira et al. 2000), Sm –Nd TDM ages yielded by
Striae on bedding plane of laminites at the SAFFRAN quarry
rocks from the Formiga area range from 1.8 to 1.98 Ga. The
and characteristics such as bending or change of direction are com-
results indicate that besides the basement, other younger, probably
patible with their origin through scratching of unconsolidated
Palaeoproterozoic terranes, may also have contributed as a source
bottom sediments by free-floating ice masses (Woodworth-Lynas
of sediments.
& Dowdeswell 1994; Rocha-Campos et al. 1994), either free
A similar picture was obtained for samples of laminites and ferru-
sea-lake ice or icebergs.
ginous bands in the SAFFRAN Quarry. The 147Sm– 144Nd of these
The nature of the water body in which the silty-clayey laminites
rocks are usually very fractionated, with values higher than 0.14,
accumulated (either a lake or sea) is difficult to ascertain. The
especially for the ferruginous bands, where they can reach values
extensive distribution of the pellitic strata in the São Francisco
of 0.18. Samples with 147Sm– 144Nd close to 0.12 (normal crustal
Basin seems to favour a marine setting.
value) yielded Sm–Nd TDM ages of 1.6–1.9 Ga, again suggesting a
Geochemical data indicate that Fe from ferruginous layers of
much younger source area for the deposits, besides the Archaean
the laminites originated from weathering of basic rocks, under
basement. On the basis of these data, the age of deposition of the
more reducing conditions than the present ones and not from the
Moema Laminites cannot be older than Palaeoproterozoic.
Late Proterozoic and Archaean banded iron formations (BIFs)
that crop out in Central Minas Gerais State. Source for kaolinite
Discussion was probably deeply weathered feldspar-rich Archaean gneissic
terrane.
Several erosional and depositional features at the Formiga and Geochemical data also suggest a strong fluviatile contribution in
SAFFRAN quarry outcrops support the existence of glacial influ- the transportation of sedimentary particles to the depositional site
ence in this part of the São Francisco Basin. (Graf et al. 1994; Kato et al. 1998). These features denote differ-
At Formiga they include (i) a smooth surface of round topo- ences in source areas for the particles and in palaeoenvironmental
graphic highs of orthogneiss, bearing striae and (ii) the presence setting between the Formiga outcrop and SAFFRAN Quarry areas.
in the diamictite and gneissic basement of deformation features A more terrestrially influenced (less saline?) environment could
(Fig. 50.3). then be envisaged for laminites at the quarry.
These structures are interpreted as having resulted from ductile Data from the literature and from this study concur in showing
deformation of the basement rocks (regolith) under a simple shear that the Moema Laminites differ in stratigraphic position and litho-
regimen. Small thrusts affecting the orthogneiss point to brittle be- logical composition from the pelitic Serra de Santa Helena Fm.,
haviour of the rock. Vergence of the features indicates a direction situated above the basal carbonates (Sete Lagoas Fm.) of the
of strain oriented from west to east. Bambuı́ Group, in which they have been included in recent geo-
Deformation seems to have varied in intensity along the upper logical maps of the area.
surface of the basement, as demonstrated by changes in thickness Additional fieldwork is needed to clarify the relations of the
of the deformed bed, size and shapes of clasts and distribution of Moema Laminites with other isolated occurrences of infra-Bambuı́
deformation features. As a rule, regolith was deformed on topo- Group diamictite outcrops, including the Jequitaı́ Fm., known in
graphic highs of the orthogneissic substrate, but not in depressions the São Francisco Basin (Dardenne et al. 1978; Rocha-Campos
between them. & Hasui 1981), and to establish their mode of origin.
Such features are commonly found associated with sub-glacial, Possible correlation among them would significantly enlarge the
glacial –tectonic deformation of Recent and Pleistocene glacial area known to have been affected by the Early Cryogenian glacia-
deposits (e.g. Bennett & Glasser 1997; Hart 1998), which results tion in Southeastern Brazil.
in the generation of deformation tills. The same explanation
is applicable to the diamictite at the Formiga outcrop. Glacial We thank F. M. Canile for her help with the preparation of this text and figures and
displacement caused fracturing and small-scale thrusting of the I. McReath for kindly revising the English language. The original text was much
substrate and abrasion, resulting in smooth round surfaces on top improved by reviews from G. Young and E. Arnaud. This represents a
540 A. C. ROCHA-CAMPOS ET AL.

contribution of the IUGS- and UNESCO-funded IGCP (International Geoscience and SE Bolivia. Journal of South American Earth Sciences, 7,
Programme) Project #512. 115– 133.
Hart, J. 1998. The deforming bed/debris rich basal ice continuum
and its implications for the formation of glacial landforms (flutes)
References and sediments (melt-out till). Quaternary Science Reviews, 17,
737– 754.
Alkmin, F. F. 2004. O que faz de um cráton um cráton? O Cráton do São Isotta, C. A. L., Rocha-Campos, A. C. & Yoshida, R. 1969. Striated
Francisco e as revelações almedianas ao delimitá-lo. In: Mantesso pavement of the Upper Pre-Cambrian glaciation in Brazil. Nature,
Neto, V., Bartorelli, A., Carneiro, C., dal Ré, & Brito 222, 466– 468.
Neves, B. B. de (organizers) Geologia do Continente Sul-Americano. Kato, Y., Ohta, I., Tsunematsu, T., Watanabe, Y., Isozaki, Y., Mar-
Beca, 19 –35. uyama, S. & Imai, N. 1998. Rare earth element variation in mid-
Alkmin, F. F. & Marshak, S. 1998. Transamazonian Orogeny in the Archean banded iron formations: implications for the chemistry of
Southern São Francisco Craton Region, Brazil: evidence for Paleo- ocean and continent and plate tectonics. Geochimica et Cosmochi-
proterozoic collision and collapse in the Quadrilátero Ferrı́fero. mica Acta, 62, 3475– 3497.
Precambrian Research, 99, 29 –58. Miranda, J. 1953. Folha Bambuı́-Seção de Geologia. Departamento
Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating of Nacional da Produção Mineral, Divisão de Geologia e Mineralogia,
the Sete Lagoas cap carbonate (Bambuı́ Group, Brazil) and impli- Relatório Anual do Diretor, 1925, 25 – 27.
cations for the Neoproterozoic glacial events. Terra Nova, 19, Rocha-Campos, A. C. & Hasui, Y. 1981. Proterozoic diamictites of
401– 406. western Minas Gerais and east Goiás, central Brazil. In: Hambrey,
Bennet, M. R. & Glasser, N. F. 1997. Glacial Geology, Ice sheets and M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial
Landforms. Wiley. Record. Cambridge University Press, Cambridge, 920–923.
Costa, M. T. da & Branco, J. J. R. 1961 (eds). Introdução, Roteiro para a Rocha-Campos, A. C., Santos, P. R dos & Canuto, J. R. 1994. Ice
Excursão Belo Horizonte-Brası́lia. Contribuição ao XIV Congresso scouring structures in Late Paleozoic rhythmites, Paraná Basin,
Brasileiro de Geologia, Belo Horizonte, Instituto de Pesquisas Radio- Brazil. In: Deynoux, M., Miller, J. M. G., Domack, E., Eyles,
ativas, 15, 9– 25. N., Fairchild, I. J. & Young, G. M. (eds) Earth’s Glacial Record.
CPRM – Serviço Geológico do Brasil, 2007. Mapa Geológico, Folha Cambridge University Press, Cambridge, 234– 240.
Guapé, 1:100.000. SEME (Secretaria de Estado de Minas e Energia)/COMIG (Companhia
D’Agrella Filho, M. S., Babinski, M., Trindade, R. I. F., Van Mineradora de Minas Gerais)/CPRM (Serviço Geológico do
Schmus, W. R. & Ernesto, M. 2000. Simultaneous remagnetization Brasil), 2003. Projeto São Francisco, Geologia e Hidrologia, Text
and U– Pb isotope resetting in Neoproterozoic carbonates of the São and Maps, 1:250.000 (PDF and CD-ROM).
Francisco craton, Brazil. Precambrian Research, 99, 179– 196. Teixeira, W., Sabaté, P., Barbosa, J., Noce, C. M. & Carneiro, M. A.
Dardenne, M. A. & Castro, P. T. A. 2000. The sedimentology, stratigra- 2000. Archean and Paleoproterozoic tectonic evolution of the
phy and tectonic context of the São Franscisco Group at the southwest São Francisco Craton. In: Cordani, U. G., Milani, E. J.,
boundary of the São Francisco craton, Brazil. Revista Brasileira de Thomaz Filho, A. & Campos, D. A. (eds) Tectonic Evolution of
Geociências, 30, 439–441. South America. Sociedade Brasileira de Geologia, São Paulo,
Dardenne, M. A., Faria, A., Magalhães, L. F. & Soares, L. A. 1978. O 101– 137.
tilito da base do Grupo Bambuı́ na borda ocidental do cráton do São Trindade, R. I. F., D’Agrella-Filho, M. S., Babinski, M., Font, E. &
Francisco. Sociedade Brasileira de Geologia, Núcleo Centro-Oeste, Brito Neves, B. B. 2004. Paleomagnetism and geochronology of the
Boletim Informativo, 7/8, 85 –97. Bebedouro cap carbonate: evidence for continental-scale Cambrian
Dowdeswell, J. A., Whittington, R. J., Jennings, A. E., Andrews, remagnetization in the São Francisco craton, Brazil. Precambrian
J. T., Mackensen, A. & Marienfeld, F. 2000. An origin for lami- Research, 128, 83– 103.
nated glacimarine sediments through sea-ice build-up and suppressed Woodworth-Lynas, C. M. T. & Dowdeswell, J. A. 1994. Soft-
iceberg rafting. Sedimentology, 47, 557– 576. sediment striated surfaces and massive diamicton facies produced
Graff, J. L., Jr., O’Connor, E. A. & Van Leewen, P. 1994. Rare earth by floating ice. In: Deynoux, M., Miller, J. M. G., Domack, E.,
element evidence of origin and depositional environment of Late Pro- Eyles, N., Fairchild, I. J. & Young, G. M. (eds) Earth’s Glacial
terozoic ironstone beds and manganese-oxide deposits, SW Brazil Record. Cambridge University Press, Cambridge, 241–259.
Chapter 51

The glaciogenic Jequitaı́ Formation, southeastern Brazil

A. UHLEIN1*, C. J. S. ALVARENGA2, M. A. DARDENNE2 & R. R. TROMPETTE3


1
Universidade Federal de Minas Gerais – UFMG, Dept. de Geologia, Av. Antônio Carlos, 6627, Campus Pampulha,
Belo Horizonte, Minas Gerais, 31270-901, Brazil
2
Universidade de Brası́lia – UnB, Inst. de Geociências, Campus, Asa Norte, 70910-900, Brası́lia (DF), Brazil
3
35 rue Pascal, 75013, Paris, France
*Corresponding author (e-mail: uhlein@netuno.lcc.ufmg.br)

Abstract: Glaciogenic deposits of the Jequitaı́ Formation (Fm.) are well exposed along the margins of the Serra do Cabral on the São
Francisco Craton, southeastern Brazil. The Jequitaı́ Formation is thin (0–150 m thick), lenticular and overlies the Espinhaço Supergroup
on a discrete unconformity. Sandstones show subglacial erosional structures such as grooved and striated pavements oriented ENE–
WSW. The Jequitaı́ Fm. consists of massive and stratified diamictites with granules, pebbles and boulders of gneiss, granite, quartzite
and carbonate. At the base, the diamictites are massive, whereas the upper part contains many alternating beds of clast-rich and -poor
diamictites. They also contain discontinuous, fine-grained sandstones and a few laminated siltstone–mudstone intercalations. This dia-
mictite association indicates glaciomarine sedimentation. The Jequitaı́ Fm. covers the São Francisco cratonic domain and its equivalent
extends eastward over the Araçuaı́ fold belt where it is part of the metasedimentary Macaúbas Group, a thick Neoproterozoic unit with
metadiamictites, quartzites and schists. The diamictite–turbidite association of the Macaúbas Group was deposited on the border of the
Pan-African– Brasiliano rift as gravity flows.

Three distinct glaciogenic units are present in the Neoproterozoic schists (Karfunkel & Hoppe 1988; Uhlein et al. 1999; Pedrosa
of Brazil (Alvarenga & Trompette 1992; Nogueira et al. 2003; Soares et al. 2001, 2011). The Jequitaı́ Fm. crops out in the Serra
Alvarenga et al. 2007). The older and more widespread unit is do Cabral Anticlinorium, extending eastward under the Bambuı́
probably of middle Cryogenian age (665 –710? Ma). It is well Group and reappearing much thicker due to thrust duplication in
known on the São Francisco Craton where it is represented by the Araçuaı́ fold belt (Figs 51.1 & 51.2).
the Jequitaı́ Fm. at the base of the Bambuı́ Group in the São Fran- Tectonism and metamorphism related to the Brasiliano orogeny
cisco basin (Rocha-Campos & Hasui 1981; Karfunkel & Hoppe increases in intensity eastward, reaching epizonal metamorphism.
1988; Trompette 1994) and as the Bebedoura Fm. in the Una The Araçuaı́ fold belt (Macaúbas Group) shows west-vergent
Group, which occurs in a series of segmented sub-basins in the thrusts and folds, and greenschist- to amphibolites-facies meta-
NE of the craton (Guimarães et al. 2011). morphism (Fig. 51.2). Over the cratonic domain, rocks of the
The Jequitaı́ Fm. occurs in the Serra do Cabral region Mesoproterozoic Espinhaço Supergroup are overlain by glacio-
(Fig. 51.1), at the transition between the São Francisco craton genic rocks of the Jequitaı́ Fm. in the Serra do Cabral region
and the external zone of the Araçuaı́ fold belt, in southeastern (Walde 1978; Uhlein 1991), where they are exposed in large
Brazil (Minas Gerais State). Along the Brası́lia fold belt at the brachy-anticline structures. To the north, near the town of Jequitaı́,
western margin of the São Francisco Craton (Goiás State), these the Serra da Água Fria region displays an anticline structure with
glaciogenic rocks are also exposed in the Cristalina area. The sandstones overlain by Jequitaı́ Fm. diamictites and siltstones
Jequitaı́ Fm. constitutes the base of the São Francisco Supergroup, and carbonates of the Bambuı́ Group (Figs 51.1–51.3).
a cover sequence of the São Francisco craton. It crops out at
the lower part of the Bambuı́ Group along the margins of the
Serra do Cabral region, overlying the Palaeo- to Mesoproterozoic Stratigraphy
Espinhaço Supergroup. The Jequitaı́ Fm. was recognized early on
as being of glacial origin, first by Branner (1919) and Moraes & The São Francisco Basin includes the Espinhaço Supergroup and
Guimarães (1930). Subsequently, Isotta et al. (1969) discovered the unconformably overlying São Francisco Supergroup, with
a striated pavement and confirmed the glacial origin. the Jequitaı́ Fm. at the base overlain by marine siliciclastic and
carbonate sediments of the Bambuı́ Group. The Espinhaço Super-
group consists of metaquartzites and metasiltstones probably
Structural framework deposited c.1500 Ma (Uhlein et al. 1998; Martins Neto 2000).
They are cut by Neoproterozoic tholeiitic mafic dykes that do
Brasiliano – Pan-African orogenic belts surround the southern São not cut the overlying São Francisco Group.
Francisco craton: the Brası́lia fold belt to the west and the Araçuaı́ The Jequitaı́ Fm. is thin (0 –150 m thick) and lenticular and
fold belt to the east (Brito Neves et al. 1999; Alkmim et al. 2001). overlies the Espinhaço Supergroup, on a discrete unconformity,
The Jequitaı́ Fm. occurs on the São Francisco craton, with the type along the margins of the Serra do Cabral. To the north, at the
section in the Serra do Cabral region (Fig. 51.1). Meso- to Serra da Água Fria, between the towns of Jequitaı́ and Francisco
Neoproterozoic rocks form the sedimentary cover of the São Fran- Dumond, the basal sandstones show striated pavements (Isotta et al.
cisco Craton, the São Francisco basin, a sub-horizontal sequence 1969; Karfunkel & Karfunkel 1976; Rocha-Campos & Hasui
that was deformed at 560 –550 Ma near the cratonic margins 1981; Karfunkel & Hoppe 1988). Grooves and numerous striae
(Uhlein et al. 1998; Alkmim et al. 2006). The Jequitaı́ Fm. covers are oriented ENE – WSW (1008).
the cratonic domain and extends eastward into the Araçuaı́ fold The overlying Bambuı́ Group can be subdivided into two major
belt and the metasedimentary Macaúbas Group (Fig. 51.2), a units in the southeastern part of the São Francisco basin (Dardenne
thick Neoproterozoic unit with metadiamictites, quartzites and 1978; Martins 1999). The lower unit is characterized by lenticular

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 541– 546. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.51
542 A. UHLEIN ET AL.

Fig. 51.1. Simplified geological map


showing location of the Jequitaı́ Fm. on the
São Francisco Craton. Modified from
Uhlein et al. (1999) and Pedrosa Soares
et al. (2001).

carbonates alternating with fine clastics sediments. The limestones eastward to rhythmic facies, with metamorphosed sandstone, silt-
and dolomites are commonly oolitic, oncolitic and stromatolitic, stone and shale intercalations (Fig. 51.2).
and show abundant cross-bedding. The upper unit, the Três Marias
Fm., consists of green or brown fine-grained arkoses interlayered
with greenish argillaceous siltstones (Uhlein et al. 2004). Glaciogenic deposits and associated strata
Three lithostratigraphic units are recognized in the Araçuaı́ fold
belt: a polycyclic basement, the Palaeo – Mesoproterozoic Espin- The Jequitaı́ Fm., near its type location, is 0 –150 m thick and
haço Supergroup, and the Neoproterozoic Macaúbas Group. The comprises a diamictite-sandstone-pelite facies association (Figs
latter consists of metadiamictites with pebbles and boulders, some 51.3 & 51.4). It consists of massive diamictites, as well as very
of which are striated and polished (Hettich 1977; Karfunkel & rare and narrow centimetre- to metre-scale alternations of sand-
Hoppe 1988; Uhlein et al. 1998, 1999; Pedrosa Soares et al. stones and pelites (Uhlein et al. 1999; Martins-Neto & Hercos
2001). The lateral continuity between the Jequitaı́ Fm. into the 2002). The diamictites contain granules to large boulders, which
Macaúbas Group (Fig. 51.2) was originally suggested by Karfun- are angular to sub-rounded and composed of sandstones, carbon-
kel & Karfunkel (1976), Karfunkel & Hoppe (1988) and Uhlein ates, pelites, granites and gneisses, vein quartz and volcanic rocks.
et al. (1998, 1999) based on the geographical distribution of Carbonate clasts are more common towards the base, whereas
facies, the variation of thicknesses, and the rare indication pro- crystalline basement clasts dominate towards the top of the
vided by glacial abrasion features on the craton. The Macaúbas Jequitaı́ Fm. (Karfunkel & Hoppe 1988). Coarse material was
Group contains metadiamictites grading vertically and laterally probably derived from the underlying basement and the Espinhaço

Fig. 51.2. Schematic geological


cross-section of the (a) São Francisco Basin
and (b) Araçuaı́ Fold Belt. Modified from
Uhlein et al. (1999).
THE GLACIOGENIC JEQUITAÍ FORMATION 543

Fig. 51.4. Measured vertical sections in Jequitaı́ Fm., Serra da Água Fria,
around Jequitaı́ town, Minas Gerais State. See Figure 51.3 for locations of
the sections.

subaqueous mass-flow deposits and turbidity currents. Cross-


stratified sandstones (facies Sc in Fig. 51.4) are fine- to coarse-
grained. The diamictites also contain spherical or ovoidal,
discontinous, 10–12 m-thick, fine- to medium-grained, massive
sandstones intercalations variably interpreted as palaeo-eskers by
Karfunkel & Karfunkel (1976) and as large rafts of the Espinhaço
substratum (Martins-Neto & Hercos 2002). The pelites are com-
posed of millimetre-scale siltstone and claystone alternations
(rhythmites) or structureless mudstones, intercalated with the dia-
mictites (Fig. 51.4), and related to low-density turbidity currents or
vertical accretion.
In the Serra da Água Fria, 10 km SE of Jequitaı́, diamictites of
the Jequitaı́ Fm. rest directly on scratched, grooved and polished
pavements developed on sub-horizontal sandstones, interpreted
as basement rocks belonging to the Espinhaço Supergroup (Isotta
et al. 1969; Karfunkel & Hoppe 1988; Uhlein et al. 1999). The
erosional features vary from very fine scratches to grooves up
20 cm wide and 5 cm deep (Isotta et al. 1969). They show V-
Fig. 51.3. Geological map of the Jequitaı́ region, Minas Gerais State, with and U-shaped grooves, and crescent fractures have been observed.
locations of the studied sections. Modified from Cukrov et al. ( 2005). These features, which have an azimuth of 1008 indicate roughly
See Figure 51.1 for location. east – west ice palaeoflow. These glacial abrasion scars are inter-
preted as land-based and the diamictites are interpreted as tillites
resting directly on striated pavements (Isotta et al. 1969; Rocha-
Supergroup, which contains a few limestone and dolostone beds. Campos & Hasui 1981; Karfunkel & Hoppe 1988). Striated and
The diamictites are massive, sheet-like to lenticular bodies, with faceted clasts, varvites, inter-tillite striated pavements and ice-
a high variation in clast/matrix ratio. In the lower portions, the thrust features have been identified (Hettich & Karfunkel 1978;
diamictites are massive and show very little variation in clast Gravenor & Monteiro 1983). These observations indicate that
size or clast/matrix ratio. The upper portions contain many alter- the glaciers that deposited the tillites were continental ice sheets
nating beds of clast-rich and clast-poor diamictite (Fig. 51.4). (Isotta et al. 1969; Karfunkel & Karfunkel 1976; Rocha-Campos
Clasts float in a texturally and compositionally immature sand- & Hasui 1981; Karfunkel & Hoppe 1988).
stone –siltstone matrix with clay minerals, calcite and sericite. Rocha-Campos et al. (1996) suggested a soft-sediment origin
Some diamictite clasts are scratched, polished and faceted (Isotta for glacial striae and grooves found on top of the sandstones of
et al. 1969; Karfunkel & Hoppe 1988). the Serra da Água Fria, near Jequitaı́. The main lines of evidence
The stratified diamictites also contains discontinous sand- are (i) striae inside grooves laterally covered by slumped plough
stones and a few intercalated laminated pelites. The sandstones ridges; (ii) clasts inside furrows partially embedded in the sand-
are massive (facies Sm in Fig. 51.4), graded or horizontally strati- stone; (iii) sinuosity of furrows; (iv) crescent-like skip marks and
fied, fine- to coarse-grained, sheet-like to lenticular bodies, poorly ridges transverse to furrows; (v) occurrence of striated surfaces
sorted, and overall upward fining. They are interpreted as on two bedding planes that are separated by a 25-cm-thick bed
544 A. UHLEIN ET AL.

of quartzite; (vi) contiguous striated and ripple-marked areas on Fm. are overlain by siltstones and shales of the Bambuı́ Group,
the same bedding plane; ripple marks a few centimetres below which represent deep-marine platform deposits. Thus it is not poss-
the striated surface are, however, not deformed. This ible to recognize a capping carbonate overlying the glacial rocks of
re-interpretation indicates that the striated sandstones may be the Jequitaı́ Fm. in the Serra do Cabral and Serra da Água
part of the Jequitaı́ Fm. rather than the Espinhaço Supergroup. Fria regions.
At the same time, the terrestrial setting of the Jequitaı́ Fm. has C- and O-isotopic studies have been carried out in Sete Lagoas
been questioned. Uhlein et al. (1999) pointed to the scarcity of Fm., basal unit of the Bambuı́ Group in southern São Francisco
outwash deposits around Jequitaı́ to support reinterpretation of Craton (Santos et al. 2000, 2004; Vieira et al. 2005, 2007),
the monotonous, massive, clast-poor diamictites as representing around the cities of Sete Lagoas and Belo Horizonte. The Sete
glaciomarine sedimentation. Rocha-Campos et al. (1996) re- Lagoas Fm. shows sea-floor precipitates with aragonitic pseudo-
interpreted the glacial abrasion marks as having been formed in morphs (Peryt et al. 1990) and negative d13C values (c. – 4‰),
the fluctuating grounding zone of a marine ice sheet or tongue which permit the characterization of a basal carbonate unit of the
just grazing the sediment surface. Finally, Martins-Neto & Bambuı́ Group as a capping carbonate (Vieira et al. 2007). The
Hercos (2002) describe wide grooves with inner steps that yield Sete Lagoas Fm. displays two upward-shallowing megacycles,
the palaeo-direction of glacier movement (2858 azimuth) on the with very high d13C (10‰) in the second megacycle (Vieira
western border of the Serra da Água Fria, suggesting a more et al. 2007). Sr-isotopic compositions of least altered carbonates
complex dispersion pattern. in the Bambguı́ Group range from 0.7074 to 07.7076 (reviewed
The metadiamictites of the Macaúbas Group in the Araçuai in Misi et al. 2007).
fold belt contain clasts from granule to boulder size in a
dominantly muddy and/or sandy matrix. The clasts are mainly
quartzites, granitoids and gneisses, limestones, quartz and Mineralization and other characteristics
schists. Clasts are commonly tectonically stretched. Clast-rich
and clast-poor beds of diamictites alternate and commonly The Macaúbas Group contains economic ferruginous and manga-
show different contents of sandy and muddy matrix. The diamic- niferous deposits in the Araçuaı́ Fold Belt. The Fe deposits
tite facies are interpreted as subaqueous debris flows deposited form the Riacho dos Poções Member of the Nova Aurora Fm.,
along basin margins during a period of tectonic activity. Two between Peixe Bravo and Vacaria Rivers (Viveiros et al.
main sequences of metadiamictites, 300–2000 m thick, are sep- 1978). Manganiferous levels originate from the lateritic weather-
arated by an interval of metarhythmites. The latter consist of ing of rhythmites and schists. Ferruginous diamictites, hematite
parallel-bedded, centimetre- to metre-scale schists, metasilt- quartzites and banded iron formations (BIFs), rhythmites, and
stones, calc-silicate rocks and quartzites, which in places show hematite schists form the ferruginous facies of the Macaúbas
normal grading. Hence, the Macaúbas Group is interpreted as Group. They were deposited in the transition zone between a
resedimented glacial material deposited by subaqueous debris glaciated continental domain and a relatively deep marine
flow and turbidity currents (Uhlein et al. 1999). From west to environment with turbidity currents. Locally, mafic lavas with
east, a depositional model can be proposed for the Jequitaı́ – pillow structures are intercalated in the Macaúbas Group,
Macaúbas units: a cratonic diamictite sequence interpreted as suggesting that hydrothermal activity in the basin may have
glaciomarine (Jequitaı́ Fm.) grading to a diamictite –turbidite been important.
sequence deposited on the border of a Neoproterozoic rift and
subsequently passive margin basin, the precursor to the
Araçuaı́ fold belt (Uhlein et al. 1998, 1999; Pedrosa Soares
et al. 2001, 2011). Palaeolatitude and palaeogeography

Palaeomagnetic studies on basic dykes in the eastern São Francisco


Craton (D’Agrella Filho et al. 1990) place the southern part of the
Boundary relations with overlying and underlying craton between 408 and 608 latitude at c. 1.0 Ga, but there are
non-glacial units no available data to constrain palaeolatitudes during deposition
of the Jequitaı́ Fm. or Bambuı́ Group. The Jequitaı́ glaciation
In the Serra do Cabral region, the lower contact is a very slight occurred on the elevated Rodinia supercontinent, where larger
angular unconformity and an erosional disconformity separating areas may have been covered by a continental ice sheet or
the Jequitaı́ Fm. from the underlying Espinhaço Supergroup several ice caps (Trompette 1994).
(Walde 1978; Uhlein 1991). Palaeotopographical relief of 300 to
400 m is developed on this contact. The lenticular character of
the Jequitaı́ Fm. along the margins of the Serra do Cabral can be Geochronological constraints
related to infilling of palaeotopography. Block tilting, indicated
by angular relationships across the Serra do Cabral, reflects Neo- Radiometric age determinations on the São Francisco craton
proterozoic extensional tectonics related to Rodinia break-up provide few concrete age constraints on glaciogenic diamictites
(Uhlein et al. 1999; Martins-Neto & Hercos 2002). The Jequitaı́ of the Jequitaı́ Fm. and equivalent units. Rb –Sr dating of clay
Fm. is conformably overlain by metasiltstone of the Bambuı́ minerals yields ages between 932 + 30 Ma and 570 + 7 Ma
Group in the Serra do Cabral region. A hiatus and a discontinuity (Macedo & Bonhomme 1984) and their geological interpretation
between the Jequitaı́ Fm. and the Bambuı́ Group can be recognized remains difficult. Badelleyite from mafic rocks intruded into the
(Dardenne 1978; Martins 1999). Espinhaço Supergroup yield an age of 906 + 2 Ma (Machado
et al. 1989), while the youngest concordant detrital zircon in the
Macaúbas Group is 864 + 30 Ma (Pedrosa Soares et al. 2000).
Chemostratigraphy Rocks inferred to represent related oceanic crust in the Macaúbas
Group have produced a Sm– Nd age of 816 + 72 Ma (Pedrosa
The lack of biostratigraphic controls and suitable minerals for Soares et al. 2001) and a 660 + 29 Ma U –Pb age by LA-
radiometric age measurements make chemostratigraphy a poten- ICPMS of a plagiogranite (Pedrosa Soares et al. 2011). The Sete
tial alternative for the establishment of a reliable isotope strati- Lagoas Fm., which overlies the Jequitaı́ Fm. on the São Francisco
graphic profile for Neoproterozoic sequences. Meanwhile, in the craton, has been dated at 740 + 22 Ma by the Pb/Pb carbonate
Serra do Cabral region, the glaciomarine diamictites of the Jequitaı́ method (Babinski et al. 2007).
THE GLACIOGENIC JEQUITAÍ FORMATION 545

Discussion Babinski, M., Vieira, L. C. & Trindade, R. I. F. 2007. Direct dating


of the Sete Lagoas cap carbonate (Bambuı́ Group, Brazil) and
Based on the different characteristics of the diamictites and associ- implications for the Neoproterozoic glacial events. Terra Nova, 19,
ated sediments, two distinct depositional environments can be 1– 6.
recognized: (i) a glaciomarine environment on the São Francisco Branner, J. C. 1919. Outlines of the geology of Brazil to accompany the
palaeocontinent, with striated bedrock pavements and a 0 –150- Geological Map of Brazil. Geological Society of American Bulletin,
m-thick diamictite-sandstone-pelite association (i.e. the Jequitaı́ 30, 189– 338.
Brito Neves, B. B., Campos Neto, M. C. & Fuck, R. A. 1999. From
Fm.) and (ii) a deepwater environment dominanted by debris
Rodinia to Western Gondwana: an approach to the Brasiliano-
flows (metadiamictites) and turbidites (quartzites, metarhythmites)
Panafrican Cycle and orogenic collage. Episodes, 22, 155– 166.
represented by the Macaúbas Group in the Araçuaı́ fold belt, a Cukrov, N., Alvarenga, C. J. S. de & Uhlein, A. 2005. Litofácies da
Neoproterozoic orogen adjacent to the São Francisco craton. glaciação neoproterozóica na porção sul do Cráton do São Francisco:
The Jequitaı́ Fm. was initially interpreted as glaciocontinental exemplos de Jequitaı́, MG e Cristalina, GO. Revista Brasileira de
and the diamictites regarded as tillites resting directly on striated Geociências, 35, 69 –76.
pavements (Isotta et al. 1969; Karfunkel & Karfunkel 1976; D’Agrella Filho, M. S., Pacca, I. G., Teixeira, W., Onstott, T. C. &
Rocha-Campos & Hasui 1981; Karfunkel & Hoppe 1988). Renne, P. R. 1990. Paleomagnetic evidence for the evolution
However, the massive clast-poor diamictites and the stratified of Meso to Neoproterozoic glaciogenic rocks in central-eastern
diamictites of the Jequitaı́ Fm. have been reinterpreted to represent Brazil. Palaeogeography, Palaeoclimatology, Palaeoecology, 80,
glaciomarine deposits because of its significant thickness 255– 265.
(0 –150 m), fine-grained sediment intercalations, and absence of Dardenne, M. A. 1978. Sı́ntese sobre a estratigrafia do Grupo Bambuı́
outwash facies. no Brasil Central. Anais XXX Congresso Brasileiro de Geologia,
In the Jequitaı́ area, clast-poor diamictites overlying the striated SBG, Recife, 2, 597–610.
pavement suggests two phases for the record of the glacial event. Guimarães, J. T., Misi, A., Pedreira, A. J. & Dominguez, J. M. L. 2011.
The first phase, probably on the continent, produced ice erosion The Bebedouro Formation, Una Group, Bahia (Brazil). In: Arnaud,
(striated pavement); during the second phase, clast-poor diamictite E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
was deposited by gravity flows during ice retreat and sea-level rise Record of Neoproterozoic Glaciations. Geological Society, London,
(Cukrov et al. 2005) at a time of active extension (Uhlein et al. Memoirs, 36, 503–508.
1998, 1999; Martins-Neto & Hercos 2002). Gravenor, C. P. & Monteiro, R. L. B. P. 1983. Ice-thrust features and a
possible intertillite pavement in the Proterozoic Macaúbas Group,
The lateral transition eastward to the Macaúbas Group is based
Jequitaı́ area, Minas Gerais, Brazil. Journal of Geology, 91, 113– 116.
on the geographical distribution of facies, thickness changes and
Hettich, M. 1977. A glaciação proterozóica no centro-norte de Minas
the presence of glacial abrasion features (Karfunkel & Hoppe Gerais. Revista Brasileira Geociências, 7, 87– 101.
1988; Uhlein et al. 1998, 1999). The Macaúbas Group in the Hettich, M. & Karfunkel, J. 1978. Um esker, um varvito e seixos
Araçuaı́ fold belt is composed of metadiamictites, quartzites, estriados no Grupo Macaúbas—norte de Minas Gerais. Revista da
rhythmites, ferruginous diamictites, BIF and schists. It consists Escola de Minas de Ouro Preto, REM, 34, 5– 8.
of resedimented glacial material deposited by subaqueous debris Isotta, C. A. L., Rocha-Campos, A. C. & Yoshida, R. 1969. Striated
flows and turbidity currents. The Macaúbas Group represents syn- pavement of the Upper-Precambrian glaciation in Brazil. Nature,
glacial, rift-to-drift sedimentation related to the break-up of the 222, 466–468.
Rodinia supercontinent. A rift system opened and was filled by Karfunkel, B. & Karfunkel, J. 1976. Estudos petro-faciológicos do
thick gravitational sedimentation, (i.e. debris flows and turbidites; Grupo Macaúbas na porção mediana da Serra do Espinhaço, MG.
Uhlein et al. 1999; Pedrosa Soares et al. 2001). The elevated rift XIX Congresso Brasileiro de Geologia, 2, 179–188.
shoulders with ice caps provided glacial material for subaqueous Karfunkel, J. & Hoppe, A. 1988. Late proterozoic glaciation in central-
resedimentation. Tectonic activity on the edge of the rift zone eastern Brazil: synthesis and model. Palaeogeography, Palaeoclima-
may have triggered slumps that developed into debris flow and tology, Palaeoecology, 65, 1 –21.
turbidity currents. Macedo, M. H. F. & Bonhomme, M. G. 1984. Contribuição à cronoestra-
tigrafia das Formações Caboclo, Bebedouro e Salitre na Chapada
Diamantina (BA) pelos métodos Rb– Sr e K– Ar. Revista Brasileira
We acknowledge financial support from the CNPq-Brazilian National Research
Council (grant 301732/2005-1 PQ) and FAPEMIG-Research Support Foundation
de Geociências, 14, 153– 163.
Machado, N., Schrank, A., Abreu, F. R. de, Knauer, L. G. & Abreu,
of Minas Gerais State (grants CRA-1321/98 and CRA-80772/05), Brazil.
F. J. Baars and L. M. Lobato helped improve the English of the original manu-
P. A. A. 1989. Resultados preliminares da geocronologia U/Pb
script. This represents a contribution of the IUGS- and UNESCO-funded IGCP
na Serra do Espinhaço Meridional. An 5 Simp. Geologia Núcleo
(International Geoscience Programme) Project #512.
Minas Gerais – 1 Simp. Núcleo Brası́lia, Belo Horizonte, 171– 174.
Martins, M. 1999. Análise estratigráfica das sequências mesoproterozói-
cas (borda oeste) e Neoproterozóicas da bacia do São Francisco.
References MSc thesis, UFRGS.
Martins-Neto, M. A. 2000. Tectonics and sedimentation in a Paleo-
Alkmim, F. F., Marshak, S. & Fonseca, M. A. 2001. Assembling West Mesoproterozoic rift-sag basin (Espinhaço Basin, southeastern
Gondwana in the Neoproterozoic: clues from the São Francisco Brazil). Precambrian Research, 103, 147–173.
craton region, Brazil. Geology, 29, 319– 322. Martins-Neto, M. A. & Hercos, C. M. 2002. Sedimentation and tectonic
Alkmim, F. F., Marshak, S., Pedrosa Soares, A. C., Peres, G. G., Cruz, setting of Early Neoproterozoic glacial deposits in south-eastern
S. & Whittington, A. 2006. Kinematic evolution of the Araçuaı́— Brazil. International Association of Sedimentology, Special Publi-
West Congo orogen in Brazil and África: Nutcracker tectonics during cations, 33, 383–403.
the Neoproterozoic assembly of Gondwana. Precambrian Research, Misi, A., Kaufman, A. J. et al. 2007. Chemostratigraphic correlation
149, 43– 64. of Neoproterozoic successions in South America. Chemical Geology,
Alvarenga, C. J. S. & Trompette, R. 1992. Glacially influenced sedi- 237, 143–167.
mentation in the Later Proterozoic of the Paraguay belt (Mato Moraes, L. J. & Guimarães, D. 1930. Geologia da região norte de Minas
Grosso, Brazil). Palaeogeography, Palaeoclimatology, Palaeoecol- Gerais. Anais Academia Brasileira de Ciências, 2, 153–186.
ogy, 92, 85– 105. Nogueira, A. C. R., Riccomini, C., Sial, A. N., Moura, C. A. V. &
Alvarenga, C. J. S. de, Figueiredo, M. F., Babinski, M. & Pinho, Fairchild, T. R. 2003. Soft-sediment deformation at the base of
F. E. C. 2007. Glacial diamictites of Serra Azul Formation (Edia- the Neoproterozoic Puga cap carbonate (southwestern Amazon
caran, Paraguai belt): evidence of the Gaskiers glacial event in craton, Brazil): confirmation of rapid icehouse to greenhouse tran-
Brazil. Journal of South American Earth Sciences, 23, 236– 241. sition in snowball Earth. Geology, 31, 613– 616.
546 A. UHLEIN ET AL.

Pedrosa Soares, A. C., Cordani, U. G. & Nutman, A. 2000. Constrain- Francisco craton and Araçuaı́ belt, Brazil: paleogeographic impli-
ing the age of Neoproterozoic glaciation in Eastern Brazil: first U –Pb cations. Journal of South American Earth Sciences, 18, 27– 39.
(SHRIMP) data for detrital zircons. Revista Brasileira de Geociên- Trompette, R. 1994. Geology of Western Gondwana (2000– 500 Ma).
cias, 30, 58 –61. Pan-African –Brasiliano Aggregation of South America and Africa.
Pedrosa Soares, A. C, Noce, C. M., Vidal, P., Wiedemann, C. M. & Balkema, Rotterdam.
Piva-Pinto, C. 2001. The Araçuaı́—West Congo Orogen in Brazil: Uhlein, A. 1991. Transição cráton – faixa dobrada: exemplo do cráton
an overview of a confined orogen formed during Gondwanaland do São Francisco e da Faixa Araçuaı́ (Ciclo Brasiliano) no Estado
assembly. Precambrian Research, 110, 307– 323. de Minas Gerais. Aspectos estratigráficos e estruturais. PhD thesis,
Pedrosa-Soares, A. C., Babinski, M., Noce, C., Martins, M., Queir- USP, São Paulo.
oga, G. & Vilela, F. 2011. The Neoproterozoic Macaúbas Group Uhlein, A., Trompette, R. & Egydio-Silva, M. 1998. Proterozoic
(Araçuaı́ orogen, SE Brazil) with emphasis on the diamictite for- rifting and closure, SE border of the São Francisco craton, Brazil.
mations. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, Journal of South American Earth Sciences, 11, 191– 203.
G. (eds) The Geological Record of Neoproterozoic Glaciations. Geo- Uhlein, A., Trompette, R. & Alvarenga, C. J. S. de 1999. Neoproter-
logical Society, London, Memoirs, 36, 523– 534. ozoic glacial and gravitational sedimentation on a continental rifted
Peryt, T. M., Hoppe, A., Bechstadt, T., Koster, J., Pierre, C. J. & margin: the Jequitaı́ –Macaúbas sequence (Minas Gerais, Brazil).
Richter, D. K. 1990. Late Proterozoic aragonitic cement crusts, Journal of South American Earth Sciences, 12, 435– 451.
Bambuı́ Group, Minas Gerais, Brazil. Sedimentology, 37, 279– 286. Uhlein, A., Lima, O. N. B., Fantinel, L. M. & Baptista, M. C. 2004.
Pimentel, M. M., Alvarenga, C. J. S. de & Armstrong, R. 2002. Estratigrafia e evolução geológica do Grupo Bambuı́, Minas Gerais.
Proveniência da Formação Jequitaı́, Brasil Central, com base em Roteiro de Excursão. Excursão 2. Congresso Brasileiro de Geologia.
dados U/Pb Shrimp em zircões detrı́ticos. Anais XLI Congresso CD-Rom.
Brasileiro de Geologia, SBG, 503, João Pessoa (PB). Vieira, L. C., Trindade, R. I. F. & Nogueira, A. C. R. 2005. Quimio-
Rocha-Campos, A. C. & Hasui, Y. 1981. Tillites of the Macaúbas Group estratigrafia da Formação Sete Lagoas, Grupo Bambuı́, Minas Gerais.
(Proterozoic) in central Minas Gerais and Southern Bahia, Brazil. In: III Simposio Cráton do São Francisco, Salvador (BA), 299–302.
Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Vieira, L. C., Trindade, R. I. F., Nogueira, A. C. R. & Ader, M. 2007.
Glacial Record. Cambridge University Press, Cambridge, 924– 927. Identification of a Sturtian cap carbonate in the Neoproterozoic Sete
Rocha-Campos, A. C., Young, G. M. & Santos, P. R. 1996. Re-examin- Lagoas carbonate platform, Bambuı́ Group, Brazil. Comptes Rendus
ation of a striated pavement near Jequitaı́, MG: implications for pro- Geoscience, 339, 240–258.
terozoin stratigraphy and glacial geology. Anais Academia Brasileira Viveiros, J. F. M., Sá, E. L., Vilela, O. V., Santos, O. M. & Moreira,
de Ciências, 68, 593. J. M. P. 1978. Geologia dos Vales do rios Peixe Bravo e Vacaria,
Santos, R. V., Alvarenga, C. J. S. de, Dardenne, M. A., Sial, A. N. & Norte de Minas Gerais. XXX Congresso Brasileiro de Geologia,
Ferreira, V. P. 2000. Carbon and oxigen isotope profiles across Recife, 1, 243–254.
Meso-Neoproterozoic limestones from central Brazil: Bambuı́ and Walde, D. H. G. 1978. Desenvolvimento faciológico do Precambriano
Paranoá Groups. Precambrian Research, 104, 107– 122. entre a Serra Mineira e a Serra do Cabral (região sudoeste da Serra
Santos, R. V., Alvarenga, C. J. S. de et al. 2004. Carbon isotopes do Espinhaço, Minas Gerais). Anais XXX Congresso Brasileiro de
of Mesoproterozoic – Neoproterozoic sequences from Southern São Geologia, Recife, 2, 711–725.
Chapter 52

The Playa Hermosa Formation, Playa Verde Basin, Uruguay

P. J. PAZOS1,2*, A. E. RAPALINI1,2, L. SÁNCHEZ BETTUCCI3 & O. R. TÓFALO1


1
Departamento de Ciencias Geológicas, FCEN, Universidad de Buenos Aires, Ciudad Universitaria,
Buenos Aires 1428, Argentina
2
Consejo Nacional de Investigaciones Cientı́ficas y Técnicas (CONICET)
3
Facultad de Ciencias, Universidad de la República, Iguá 4225, 11400 Montevideo, Uruguay
*Corresponding author (e-mail: pazos@gl.fcen.uba.ar)

Abstract: The Playa Hermosa Formation (Fm.) (Playa Verde Basin) is a volcano-sedimentary unit that crops out in the extreme south of
the Dom Feliciano Belt in Uruguay. This formation has traditionally been interpreted as non-glacial in origin, but recently it has been
suggested that the lower part at least may be glacially influenced. The stratigraphic position of the Playa Hermosa Fm. and its correlation
with other Neoproterozoic units of Uruguay, Argentina and southern Brazil remains in dispute. An age of c. 580 Ma is indicated by a
hornblende 39Ar/40Ar age on magmatism regarded as coeval with syn-sedimentary volcanism. Preliminary palaeomagnetic data
suggest a primary remanence and a mean geomagnetic pole consistent with the proposed apparent polar wander path for the Rı́o de
La Plata craton. In combination, these data suggest mid-Ediacaran glacial sedimentation in low to intermediate latitudes on the Rı́o
de la Plata craton. However, this conclusion needs to be reinforced by more thorough studies of the age and origin of the Playa
Hermosa Fm.

The Ediacaran Playa Hermosa Formation (Masquelin & Sánchez terrane (c. 2.1 Ga) of low- to medium-grade metamorphic orogenic
Bettucci 1993), regarded as part of the Las Ventanas Formation belts, virtually unaffected by younger orogenies, (ii) the Nico
by Blanco & Gaucher (2005) and Blanco et al. (2009), is com- Pérez terrane represented by Palaeoproterozoic medium- to high-
monly omitted in regional correlations and palaeoclimatic grade metamorphic orogenic belts affected by the Neoproterozoic
interpretations (e.g. Eyles & Janusczack 2004). This formation is Brasiliano – Pan-African orogenic cycle, and (iii) the Neoprotero-
exposed in southern Uruguay, close to the city of Piriápolis zoic to Early Palaeozoic Dom Feliciano belt that developed on
(between 348490 4100 S, 0558190 0000 W and 348500 0500 S, the eastern border of the Nico Pérez terrane. The fourth and east-
0558180 4500 W; see Figs 52.1 & 52.2) and the best outcrops are ernmost basement unit in Uruguay is the Punta del Este tectono-
found along the coast. It was originally interpreted by Masquelin stratigraphic terrane (also known as Cuchilla Dionisio terrane),
& Sánchez Bettucci (1993) as a shallow turbiditic sequence associ- which has been interpreted as a fragment of the Kalahari craton
ated with coarse-grained beds deposited in distal fans and canyons accreted to the Rı́o de La Plata craton c. 540 Ma and left behind
sourced from a delta system. Subsequently, Sánchez Bettucci & after the opening of the South Atlantic (Sánchez Bettucci et al.
Pazos (1996) ascribed the Playa Hermosa Fm. to the base of the 2010). Evidence of the suture is seen in the Sierra Ballena shear
latest Proterozoic –Ordovician Playa Verde basin. The upper part zone (Figs 52.1 & 52.2), a continental-scale structure (Fernandes
of the lower member of the Playa Hermosa Fm. was deposited con- & Koester 1999; Basei et al. 2005).
temporaneously with volcanic flows and intrusions of the middle The Playa Hermosa Fm. is located in the southwestern extreme
Ediacaran Sierra de Las Animas Complex (Sánchez Bettucci of the Dom Feliciano belt. Sánchez Bettucci & Ramos (1999)
1998; Sánchez Bettucci et al. 2009), which provide the most accu- suggested that the Playa Verde basin originated during post-
rate time constraints available for its deposition. orogenic collapse. In contrast, Gaucher et al. (2003) interpreted
A glacial influence in the deposition of the Playa Hermosa was this succession as the remnant of a passive margin.
originally proposed by Pazos et al. (1998) and later reinforced by The Playa Hermosa Fm. is associated with bimodal (basalt-
Pazos et al. (2003), although this interpretation was initially not rhyolite) volcanism rich in sub-alkaline subvolcanic (syenite)
accepted (e.g. Gaucher et al. 2003, 2005). Pazos et al. (2008) pub- intrusions. This magmatism has been related to the post-orogenic
lished an exhaustive analysis of the available literature on the phase of the Rio Doce Orogeny (Brasiliano Cycle) that developed
Playa Hermosa Fm. and concluded that a glaciogenic origin for between c. 590 and 510 Ma (Campos Netto & Figueiredo 1995).
the lower member of the Playa Hermosa Fm. is unequivocal, The Playa Hermosa Fm. records the opening of the Playa Verde
although these sediments appear to be dominantly subaqueously basin in a transtensional tectonic setting (Pazos et al. 2003;
reworked glacial debris. Only rainout processes, evidenced by Pecoits et al. 2008).
dropstones and coarse-grained rhythmite intervals, are regarded The Playa Hermosa Fm. is exposed in a homocline. Its base
as directly related to ice transport. Correlation between this (348510 S, 0558180 W) is underlain by tonalitic gneisses of the
formation and other glaciogenic and suspected glaciogenic units 1735 + 32 Ma Palaeoproterozoic Campanero Unit (Sánchez
in the Rı́o de La Plata craton was critically analysed by Pazos Bettucci et al. 2004, 2009; Mallmann et al. 2007). Its upper
et al. (2008). contact, however, is covered variably by modern beach sands
and Quaternary sediments (Pazos et al. 2003). Tectonic defor-
mation is characterized by extension-related stretching and fractur-
Structural framework ing. The succession lacks regional metamorphism, despite lying
adjacent to the conspicuous tectonic boundary between the Dom
The Uruguayan Precambrian basement is separated into four main Feliciano belt and the Punta del Este terrane (i.e. the Sierra
tectonic units (Fig. 52.1). From west to east, three units are part of Ballena Shear Zone, Oyhantçabal et al. 2007). This major shear
the Rı́o de La Plata craton: (i) the Palaeoproterozoic Piedra Alta zone was apparently active after deposition of the Playa Hermosa

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 547– 553. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.52
548 P. J. PAZOS ET AL.

Fig. 52.1. Location of the studied area and


morphotectonic divisions of the eastern part
of the Rı́o de la Plata Craton. Modified from
Pazos et al. (2008). SYSZ, Sarandı́ del Yı́
shear zone; MAFMSZ, Marı́a Albina-Fraile
Muerto shear zone; SBSZ, Sierra Ballena
shear zone; CSZ, Cordillera shear zone.

sediments, recording docking of the Punta del Este terrane in the The San Carlos and the Las Ventanas formations were originally
Early Cambrian. Gaucher et al. (2007) proposed that the Nico assigned to the Ordovician by Sánchez Bettucci (1998) based on
Pérez terrane (Fig. 52.1) was displaced some 500 km northward the occurrence of basalt and syenite clasts in the Las Ventanas
during the Cambrian, along the Sarandı́ del Yı́ major shear zone Fm. that were thought to be sourced from the Ediacaran to Cam-
that separates this terrane from the Piedra Alta terrane to the brian Sierra de Las Ánimas Complex (Sánchez Bettucci 1998;
west. However, the geotectonic setting and the detailed geological Sánchez Bettucci & Rapalini 2002). Later, Pecoits et al. (2008)
evolution of these rocks is still far from resolved. suggested that conglomerate units and rhythmite intervals of the
Las Ventanas Fm. are Neoproterozoic with evidence of glacial
influence. This succession was later folded during the Early
Stratigraphy Palaeozoic. Both Pazos et al. (2008) and Pecoits et al. (2008) cor-
related the lower member of the Playa Hermosa Fm. with the fine-
The Playa Verde basin (Sánchez Bettucci & Pazos 1996) infill con- grained deposits of the Las Ventanas Fm. based on the similarity
stitutes the Maldonado Group (Pecoits et al. 2008), which is sub- of rainout deposits (Fig. 52.3).
divided into the Playa Hermosa, the Las Ventanas and the San
Carlos formations (Fig. 52.3). The Las Ventanas and San Carlos
formations are discussed in a separate chapter (Pecoits et al. Glaciogenic deposits and associated strata
2011). The glacially influenced Playa Hermosa Fm. comprises
two members. The lower member contains conglomerate, fine- The sedimentology of the Playa Hermosa Fm. has been studied by
grained sandstone, pelite and rhythmically laminated fine-grained Sánchez Bettucci & Pazos (1996), Pazos et al. (1998, 2003) and
sediments resembling varves. Coarse-grained lithologies are Fambrini et al. (2003). The Playa Hermosa is divided into two
locally deformed by loading and other syn-sedimentary defor- members, with the lower one containing deposits thought to be gla-
mation (Pazos et al. 2003). The upper member is composed of ciogenic. In the logged section of the lower member analysed by
conglomerate, rare sandstone, diamictite beds and volcanic rocks Pazos et al. (2003), two facies associations (FA I and II) were
that intruded and brecciated the sediments and produced hyalo- defined (Fig. 52.4). The lowermost, FA I, is composed of rhyth-
clastic features indicative of syn-sedimentary volcanism (Sánchez mite, contorted conglomerate and rare normally graded sandstone
Bettucci et al. 2009). The Las Ventanas Fm. is characterized by and pelite beds. The conglomerates have variable matrix content
conglomerate, sandstone and rhythmites containing out-sized and include sub-rounded and angular pebbles of felsic magmatic
clasts (Pecoits et al. 2008). The San Carlos Fm. is a succession rocks as well as intrabasinal rhythmite clasts that are disaggregated
of fluvial metaconglomeratic-sandstone, quartzite, feldspar sand- and deformed. Massive diamictites contain abundant rip-up
stone and subordinate metapelite (Sanchez Bettucci & Pazos rhythmite clasts, as well as some angular feldspar clasts. Evidence
1996) and rhyolithic flows (Pecoits et al. 2008). This unit is for syn-sedimentary deformation is seen in folded stringers of
found to the east of the Sierra Ballena shear zone (see Fig. 52.2). gravels. Conglomerate beds have irregular, sometimes loaded,
THE PLAYA HERMOSA FORMATION 549

Fig. 52.2. Geology of the southern area of


the Dom Feliciano Belt. Modified from
Pazos et al. (2008).

basal contact, are lenticular at outcrop scale, and have clasts fine-grained rhythmite units rarely occur in FA I and are composed
forming aggregates or clusters with clast-supported fabric. of dark and light green silt and clay couplets.
The most distinctive lithology of FA I is the fine- to coarse- FA II (Fig. 52.4) consists of fine-grained rhythmites, some with
grained rhythmite. The coarse rhythmite consists of couplets of brecciated tops, normally graded and rippled sandstones, some
gravel and sand/mud layers. These rhythmite units contain large, containing lenticular, linsen and moderate- to high-angle ripple
out-sized, bullet-shaped boulders of quartzite. The intervals of drift cross-lamination. Diamictite beds are usually sharp based,
550 P. J. PAZOS ET AL.

Fig. 52.3. Comparative stratigraphic chart of the Playa Verde basin, Arroyo del
Soldado Group and Tandilia System. Taken and modified from Gaucher et al.
(2003), Pecoits et al. (2008) and Pazos & Rapalini (2011). Triangles indicate
glaciogenic intervals. The brick symbol corresponds to carbonates. Grey boxes
are siliciclastic units.

although some units are interbedded with normal graded sandstone


and dark claystone beds. All palaeocurrents measured from ripples
indicate palaeoflow towards the NE. Horizons of completely disag-
gregated mixtures of gravel, sand and mud with relict rhythmite
clasts are also present (Bm(i) in Fig. 52.4). Rare out-sized clasts
of felsic magmatic origin occur. Close to the top, metre-thick con-
volute bedding occurs in fine-grained rhythmitic deposits below a
second interval of distorted conglomerate. This distorted conglom-
erate deposit is overlain by extremely thin, tectonically deformed
rhythmite beds.

Boundary relations with overlying and underlying


non-glacial units

The basal contact of the Playa Hermosa Fm. is a disconformity


overlying 1.7 Ga tonalitic gneiss (Sánchez Bettucci et al. 2009).
The contact between the lower glaciogenic member and the
upper member and the contact at the top of the Playa Hermosa
Fm. are covered by beach sands.

Chemostratigraphy
Fig. 52.4. Logged section of the lower member of the Playa Hermosa Fm.
Modified from Pazos et al. (2003). Cl, claystone; Fs, fine-grained sandstone;
No chemostratigraphic studies have been carried out in the Playa
Ms, medium-grained sandstone; Cs, coarse-grained sandstone; Fcg,
Hermosa Fm. or other units within the Playa Verde Basin. fine-grained conglomerate; Mcg, medium-grained conglomerate.

Palaeolatitude and palaeogeography Isothermal remanent acquisition curves confirm that magnetite is
the likely carrier of the remanence in this formation. Lack of
Preliminary palaeomagnetic data on fine-grained sandstone units resemblance of the characteristic directions (both in situ and
of the Playa Hermosa Fm. were reported by Sánchez Bettucci & after untilting) with post-Ordovician expected directions for
Rapalini (2002). This study included just two sites in the lower South America, coupled with the tectonic stability of the region
part of this succession near Piriápolis (Fig. 52.1). The mean geo- after the Ordovician suggests a possible primary origin for the
magnetic pole suggests a low palaeolatitude (12.7 þ 9.5/ –8.18S) characteristic remanence of the Playa Hermosa Fm. However,
for deposition of this succession, although these data must be con- the very small number of sites (two) and samples (six) make this
firmed with further analysis before any palaeogeographical ‘palaeomagnetic pole’ a simple mean geomagnetic pole with no
interpretation is verified. The entire succession is exposed as a certainty that non-dipole components of the Earth’s magnetic
homoclinal sequence tilted some 408 towards WNW, which does field have been fully compensated. Under such conditions, ‘true’
not permit the performance of any fold or tilt test to date the palaeolatitudes from palaeomagnetic data are not certain.
remanence. Detailed alternating field (AF) and thermal cleaning Despite this uncertainty, the geomagnetic pole for the Playa
and principal component analysis (Kirschvink 1980) assure a Hermosa Fm. is consistent with the distribution of other Neoproter-
reliable isolation and definition of the magnetic components. ozoic palaeomagnetic poles for the Rı́o de La Plata craton
THE PLAYA HERMOSA FORMATION 551

(Sánchez Bettucci & Rapalini 2002; Rapalini 2006; Pazos & Rapa- sedimentary succession. Oyhantçabal et al. (2007) obtained an
lini 2011). Comparison of this pole with the apparent polar wander age of 579 + 2 Ma (39Ar/40Ar in hornblende) from a syenitic
path suggests an age of c. 590 Ma. This apparent polar wander path body (Pan de Azúcar pluton) that is part of this magmatic cycle
also indicates that the Rı́o de La Plata craton was located at inter- and is exposed 12 km from the Playa Hermosa outcrops. This
mediate southern latitudes during the Ediacaran and Cambrian age and the geological relationships suggest a roughly similar
(Fig. 52.5). age for the deposition of the Playa Hermosa sediments. The pres-
Palaeogeographic reconstruction of the basin is highly uncertain ence of boulders of conglomerate and sandstone lithologies of the
because the Playa Hermosa Fm. only crops out as a narrow belt on lower member of the Playa Hermosa Fm., enclosed in quartz –
the coast, with very little obtainable information on the geometry syenite breccia of the Sierra de Las Animas complex (Sánchez
of the original basin. The systematic palaeocurrent pattern from Bettucci et al. 2009), also supports this interpretation.
the SW and quartzite composition of the outsized clasts (drop-
stones) suggests that the Tandilia System and its eastern pro-
longation (Rapela et al. 2007; Pazos & Rapalini 2011) may have Discussion
been the source area. Mylonitized magmatic clasts within the
Playa Hermosa Fm. also indicate provenance from an intensely The lower member of the Playa Hermosa Fm. is attributable to a
tectonized region. glacially influenced environment that records high rates of sedi-
mentation in a tectonically active depositional system. It is separ-
ated into two facies associations. FA I is thought to record a
Geochronological constraints pro-glacial depositional setting that grades upward into the more
distal FA II (Pazos et al. 2003). Identification of glacial influence
The age of the Playa Hermosa Fm. is not constrained by any direct is based on a combination of sedimentary facies and features,
radiometric dates. However, the interaction of magmatism with including ice-rafted debris and rhythmites forming couplets of
unconsolidated sediments in the upper part of the lower member angular grains (coarse layer) immersed in a fine-grained matrix
of the Playa Hermosa Fm. (Pazos et al. 2008) strongly suggests and dark clay. These couplets are common in Carboniferous degla-
that deposition was contemporaneous with this magmatic event. cial to post-glacial transitions in Argentina (see Pazos et al. 2008
The characteristics of this interaction were described by Sánchez for references) and in the Quaternary. A large out-sized, elongate
Bettucci et al. (2009). The bimodal magmatism corresponds to clast of quartzite that disrupted laminations was interpreted by
the first cycle of the Sierra de Las Animas complex and has been Pazos et al. (2003) as the most robust evidence for ice-rafting pro-
loosely constrained between 615 and 550 Ma based on K –Ar cesses. A debris flow origin is ruled out due to the lateral disconti-
and Rb – Sr ages (Sánchez Bettucci & Linares 1996; Sánchez nuity of disruption to bedding around the dropstone and the
Betucci & Rapalini 2002) from samples taken from different well-laminated nature of the couplets hosting the clast. Very fine-
units of this complex, including flows intercalated in the grained rhythmite units present in the FA II are similar to varves

Fig. 52.5. Palaeogeographic reconstruction


for the Rı́o de la Plata Craton and other
blocks during the Ediacaran–Cambrian
periods. (a) 600 Ma, (b) 575 Ma,
(c) 550 Ma. AM, Amazonia; AR, Arabia;
Bal, Baltica; C, Congo; K, Kalahari;
Lau, Laurentia; PA, Pampia; RP, Rı́o de la
Plata; SF, San Francisco; WA, Western
Africa; WN, Western Nile (or Sahara
metacraton). Modified from Sánchez
Bettucci & Rapalini (2002).
552 P. J. PAZOS ET AL.

and suggest a cyclic control on deposition. No tillite deposits have Gaucher, C., Boggiani, P. C., Sprechmann, P., Sial, A. N. & Fair-
been reported and for this reason, the succession was considered as child, T. 2003. Integrated correlation of the Vendian to Cambrian
only glacially influenced by Pazos et al. (2008), consistent with Arroyo del Soldado and Corumbã Groups (Uruguay and Brazil)
earlier interpretations of a pro-glacial depositional setting (Pazos palaeogeographic, palaeoclimatic and palaeobiologic implications.
et al. 2003; Pecoits et al. 2008) for the lower member of the Precambrian Research, 120, 241– 278.
Playa Hermosa Fm. Gaucher, C., Poiré, D., Peral, L. & Chiglino, L. 2005. Litoestratigra-
Gravity-driven deposits, such as normally graded sandstone fı́a, Bioestratigrafı́a y correlaciones de las sucesiones sedimentarias
beds with ripple drift cross-lamination and giant convolutions, as del Neoproterozoico-Cámbrico del Cratón del Rı́o de La Plata
(Uruguay y Argentina). Latin American Journal of Sedimentology
well as irregular conglomerate beds, suggest that deposition was
and Basin Analysis, 12, 145– 160.
abrupt on an unstable slope related to a pro-glacial environment. Gaucher, C., Poiré, D., Finney, S. C., Valéncia, V., Blanco, G.,
Top-brecciated rhythmite beds and syn-depositional folding and Pamoukaghlian, K. & Gómez Peral, L. 2007. Zircones detrı́ticos
slumping are interpreted to record syn-sedimentary seismicity. de secuencias neoproterozoicas de Uruguay y Argentina: Inferencias
It is difficult to discern between a marine and non-marine deposi- sobre la evolución paleogeográfica del craton del Rı́o de La Plata. V
tional setting for the Playa Hermosa Fm. Fambrini et al. (2003) Congreso Uruguayo de Geologı́a, Montevideo, Uruguay. CD-Rom
argued for a marine environment based on micro-hummocky struc- (extended paper, 25).
tures. However, these structures seem to represent ripple-drift cross- Kirschvink, J. L. 1980. The least-squares and plane and the analysis of
lamination rather than storm-wave processes, and hence do not dis- palaeomagnetic data. Geophysical Journal of the Royal Astronomical
tinguish between marine and lacustrine environments (Pazos et al. Society, 62, 699– 718.
2003). Sánchez Bettucci et al. (2009) described orange palagonite Mallmann, G., Chemale, F., Jr., Avila, J. N., Kawashita, K. & Arm-
glass in the volcanism–sediment interaction that is a common altera- strong, R. A. 2007. Isotope geochemistry and geochronology of the
tion product where basalts extrude into the marine environment. Nico Pérez Terrane, Rı́o de La Plata Craton, Uruguay. Gondwana
The relationship between the Playa Hermosa Fm. and other units Research, 12, 489– 508.
within the Maldonado Group (Playa Verde basin; Fig. 52.3), in Masquelin, H. & Sánchez-Bettucci, L. 1993. Propuesta de evolución
particular the purportedly glaciogenic Las Ventanas Fm. (Pecoits tectono-sedimentaria para la fosa tardi-Brasliana en la región
et al. 2011), remains subject to debate. Several other Ediacaran- de Piriápolis, Uruguay. Revista Brasileira de Geociencias, 23,
aged successions occur on the Rı́o de la Plata craton (Fig. 52.3), 313– 322.
including the Cerro do Buggio Fm. (Camaquã basin) and the Oyhantçabal, P., Siegesmund, S., Wemmer, K., Frei, R. & Layer, P.
Arroyo del Soldado Group in the Ribeira belt and the Sierras 2007. Post-collisional transition from calc-alkaline to alkaline mag-
matism during transcurrent deformation in the southernmost Dom
Bayas Group and equivalent units in the Tandilia System (Pazos
Feliciano Belt (Braziliano –Pan-African, Uruguay). Lithos, 98,
& Rapalini 2011). Possible correlations of the Playa Hermosa 141– 159.
Fm. with other Ediacaran units cropping out in Brazil have been Pazos, P. J. & Rapalini, A. 2011. The controversial stratigraphy of the
discussed in detail by Pazos et al. (2003, 2008). In summary, no glacial deposits in the Tandilia System, Argentina. In: Arnaud, E.,
conclusive correlations can yet be made due to highly fragmentary Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
exposures and insufficient radiometric control. Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 565–569.
We are grateful to M. A. S. Basei for a constructive review and especially to Pazos, P. J., Tofalo, R. & Sánchez Bettucci, L. 1998. Procesos
E. Arnaud for her thorough review and editorial improvements. This is a contri- sedimentarios e indicadores paleoclimáticos en la sección inferior
bution R-39 to the Instituto Don Pablo Groeber. This represents a contribution of de la Formación Playa Hermosa, Cuenca Playa Verde, Piriápolis,
the IUGS- and UNESCO-funded IGCP (International Geoscience Programme) Uruguay. II Congreso Uruguayo de Geologı́a, Punta del Este,
Project #512. Uruguay, 64 –69.
Pazos, P. J., Sánchez Bettucci, L. & Tófalo, R. O. 2003. The record of
the Varanger glaciation at the Rı́o de La Plata craton, Vendian-
References Cambrian of Uruguay. Gondwana Research, 6, 65 –78.
Pazos, P. J., Sánchez Bettucci, L. & Loureiro, J. 2008. The Neopro-
Basei, M. A. S., Frimmel, H. E., Nutman, A. P., Preciozzi, F. & Jacob, terozoic glacial record in the Rı́o de La Plata Craton: a critical reap-
J. 2005. A connection between the Neoproterozoic Dom Feliciano praisal. In: Pankhurst, R. J., Trouw, R. A. J., de Brito Neves,
(Brazil/Uruguay) and Gariep (Namibia/South Africa) orogenic B. B. & de Witt, M. J. (eds) West Gondwana Pre-Cenozoic Corre-
belts – evidence from a reconnaissance provenance study. Precam- lations Across the South Atlantic Region. Geological Society,
brian Research, 139, 195–221. London, Special Publications, 294, 343–364.
Blanco, G. & Gaucher, C. 2005. Estratigrafı́a, paleontologı́a y edad de Pecoits, E., Gingras, M., Aubet, N. & Konhauser, K. 2008. Ediacaran
la Formación Las Ventanas (Neoproterozoico, Uruguay). Latin Amer- in Uruguay: palaeoclimatic and palaeobiological implications. Sedi-
ican Journal of Sedimentology and Basin Analysis, 12, 115–131. mentology, 55, 689– 721.
Blanco, G., Rajesh, H. M., Gaucher, C., Germs, G. J. B. & Chemale, Pecoits, E., Gingras, M. K. & Konhauser, K. O. 2011. Las Ventanas
F., Jr. 2009. Provenance of the Arroyo del Soldado Group (Ediacaran and San Carlos formations, Maldonado Group, Uruguay. In:
to Cambrian, Uruguay): implications for the paleogeographic evolu- Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
tion of southwestern Gondwana. Precambrian Research, 171, 57 – 73. logical Record of Neoproterozoic Glaciations. Geological Society,
Campos Neto, M. C. & Figueiredo, M. C. 1995. The Rı́o Doce Orogeny, London, Memoirs, 36, 555–564.
Southeastern Brazil. Journal of South American Earth Sciences, 8, Rapalini, A. E. 2006. New Late Proterozoic paleomagnetic pole for the
143– 162. Rı́o de La Plata craton: implications for Gondwana. Precambrian
Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for Research, 147, 223– 233.
Neoproterozoic glaciations during the break up of Rodinia after Rapela, C. W., Pankhurst, R. J. et al. 2007. The Rı́o de La Plata craton
750 Ma. Earth-Science Reviews, 65, 1– 73. and the assembly of SW Gondwana. Earth Science Reviews, 83,
Fambrini, G. L., Paes-de-Almeida, R., Riccomini, C. & Fragoso- 49 – 82.
Cesar, A. R. S. 2003. Tempestitos com Influência Glacial da Forma- Sánchez Bettucci, L. 1998. Evolución tectónica del Cinturón Dom
ção Playa Hermosa (Neoproterozóico), Piriápolis, Uruguai. Revista Feliciano en la región Minas – Piriápolis, Uruguay. PhD thesis,
Brasileira de Geociências, 33, 1 –12. Universidad de Buenos Aires, Argentina.
Fernandes, L. A. D. & Koester, E. 1999. An overview of the Neoproter- Sánchez Bettucci, L. & Linares, E. 1996. Primeras Edades en Basaltos
ozoic Dorsal de Canguç u strike-slip shear zone and its role in the del Complejo Sierra de Ánimas, Uruguay. XIII Congreso Geológico
tectonic evolution of the continental crust in southern Brazil. Journal Argentino and III Congreso de Exploración de Hidrocarburos I,
of African Earth Sciences, 29, 3– 24. Buenos Aires, Argentina, I, 399–404.
THE PLAYA HERMOSA FORMATION 553

Sánchez Bettucci, L. & Pazos, P. 1996. Análisis Paleoambiental y Sánchez Bettucci, L., Oyhantçabal, P., Preciozzi, F., Loureiro, J.,
Marco tectónico en la Cuenca Playa Verde, Piriápolis, Uruguay. Ramos, V. A. & Basei, M. A. S. 2004. Mineralizations of the Laval-
XIII Congreso Geológico Argentino, Buenos Aires, Argentina, I, leja Group (Uruguay), a Neoproterozoic volcano– sedimentary
405– 412. sequence. Gondwana Research, 7, 745–751.
Sánchez Bettucci, L. & Ramos, V. A. 1999. Aspectos Geológicos de las Sánchez Bettucci, L., Koukharsky, M., Pazos, P. J. & Stareczek, S.
rocas metavolcánicas y metasedimentarias del Grupo Lavalleja, 2009. Neoproterozoic subaqueous extrusive –intrusive rocks in the
Sudeste de Uruguay. Revista Brasilera de Geociencias, 29, 557– 570. Playa Hermosa Formation in Uruguay: regional and stratigraphic
Sánchez Bettucci, L. & Rapalini, A. E. 2002. Palaeomagnetism of the significance. Gondwana Research, 16, 134– 144.
Sierra de Las Animas Complex, Southern Uruguay: Its Implications Sánchez Bettucci, L., Peel, E. & Masquelin, H. 2010. Neoproterozoic
in the Assembly of Western Gondwana. Precambrian Research, tectonic synthesis of Uruguay. International Geology Review, 52,
118, 243– 265. 51– 78.
Chapter 53

Las Ventanas and San Carlos formations, Maldonado Group, Uruguay

ERNESTO PECOITS*, MURRAY K. GINGRAS & KURT O. KONHAUSER


Department of Earth and Atmospheric Sciences, University of Alberta, 1-26 Earth Sciences Building, Edmonton, AB, T6G 2E3, Canada
*Corresponding author (e-mail: epecoits@ualberta.ca)

Abstract: Together with the Playa Hermosa Formation (Fm.), the Las Ventanas and San Carlos formations constitute the Maldonado
Group, which is better developed in the southeastern part of Uruguay and covers an area of c. 200 km2. The total thickness of both units
(i.e. Las Ventanas and San Carlos formations) reaches c. 1500 m, and comprises mafic and acidic volcanic rocks, pyroclastic rocks, dia-
mictite, sandstone, conglomerate and pelite. Structurally, the Maldonado Group is extensively deformed, although variably, throughout
the region. Strike–slip faults, westward-verging detachment faults, and folds with axis sub-parallel to the strike–slip planes are common
features. The presence of pumpellyite, prehnite, chlorite and epidote in mafic rocks indicates very low- to low-grade metamorphic con-
ditions. The Las Ventanas Fm. is characterized by basal conglomerate, diamictite, sandstone and siltstone that pass upwards into fine-
grained rhythmites (pelite), and is the thickest unit of the Maldonado Group (c. 1250 m). The San Carlos Formation (c. 250 m) comprises
fine-grained conglomerate, sandstone and mudstone towards the top. Both units lie on an angular unconformity above Palaeo- and Neo-
proterozoic basement and are overlain unconformably by late Ediacaran– lowermost Cambrian units. Reliable palaeomagnetic data indi-
cate that the Maldonado Group accumulated at high palaeolatitudes; however, the palaeogeographical evolution of the Rı́o de la Plata
Craton during the Neoproterozoic remains conjectural. Radiometric data from intrusive bodies and cross-cutting strike– slip faults
place the minimum age of the group at c. 565 Ma, whereas basement volcanic rocks dated at 590 + 2 Ma interbedded with meta-sand-
stones hosting detrital zircons c. 600 million years old provide the best constraint on the maximum age of deposition. Given the absence
of carbonate rocks, no chemostratigraphic studies (e.g. C, O, Sr) are available.
The Las Ventanas and San Carlos formations are largely interpreted as units within a thick glacially influenced fan-delta sedimentary
system formed during the early Ediacaran in a strike–slip basin. Based on stratigraphic and sedimentological characteristics it has been
suggested that this succession, containing glacially influenced diamictite and dropstones, records a glacial period that occurred sometime
between c. 570 and 590 million years ago. Ongoing research is focused on establishing the precise age of deposition of the Maldonado
Group and on reconstructing the tectonic evolution of the basin. Further palaeomagnetic studies will be especially useful for determining
the palaeogeography of the Rı́o de la Plata Craton during the Ediacaran and establishing its relationships with neighbouring strata hosting
similar successions.

The Maldonado Group was formally erected by Pecoits et al. fluvio-lacustrine. The stratotype of the formation is located 6 km
(2005) to include the Playa Hermosa (Pazos et al. 2011) and Las south of San Carlos town (Fig. 53.2), where 220 m of San Carlos
Ventanas formations, which are better exposed near the towns of strata are exposed with the base and top of formation not visible
Piriápolis and Pan de Azúcar (Figs 53.1 & 53.2). The San Carlos (Pecoits et al. 2008). The sedimentary facies and volcanic associ-
Fm. was informally included. Subsequent work showed that the ation of the San Carlos Fm. are similar to those of the middle Las
succession continues to the SW and NE of Minas and Melo Ventanas Fm. Likewise, palynological macerations carried out in
cities (Fig. 53.1). The group reaches a maximum thickness of the pelites of both units reveal the occurrence of similar microbiota
c. 1500 m and covers an area of c. 200 km2. It comprises acidic (Pecoits et al. 2005). These observations led Pecoits et al. (2005) to
and basic volcanic rocks, pyroclastic and sedimentary strata gener- propose a correlation between the San Carlos and Las Ventanas
ated in a tectonically active and glacially influenced basin (for a formations. Whether both units were deposited in the same
recent review see Pecoits et al. 2008). basin, and subsequently dismantled by the displacement of the
Midot (1984) originally suggested the Las Ventanas Fm. Sierra Ballena Shear Zone (Fig. 53.2), or were developed within
included conglomerate, sandstone and pelite outcropping at De different depocentres remains uncertain.
Las Ventanas Hill and in the surrounding areas (Fig. 53.3). This The first evidence of glacial influence in the Las Ventanas Fm.
unit was considered by Midot (1984) and various other authors was recorded by Pecoits (2003a), comprising faceted, outsized
to be Ordovician in age (e.g. Bossi & Navarro 1991; Masquelin clasts in finely laminated rhythmites that were interpreted as drop-
& Sánchez 1993; Pazos et al. 2003). This assumption was stones. Recently, Gaucher et al. (2008) reported glacial diamictite
mainly founded on the inferred development of alluvial fans with associated dropstones occurring in laminated siltstone to the
sourced from the Cambrian Sierra de las Ánimas Complex, south of Minas (Fig. 53.2). Additionally, glacial diamictite and
located westward (Fig. 53.2). However, detailed mapping and stra- fine-grained rhythmites containing striated dropstones are well
tigraphic analysis of the Las Ventanas Fm. led to its redefinition exposed c. 15 km NW (El Perdido area) of this locality (Pecoits
as a Neoproterozoic volcanic/sedimentary succession (Pecoits et al. 2008). No glacial evidence has yet been recorded in the
2003a). The sections exposed in the northern and southern parts San Carlos Fm.
of the type area were designated as the stratotype and parastrato-
type of the unit, respectively (Pecoits 2003a). The former is
located near Paso del Molino, where 1200 m of Las Ventanas Structural framework
strata are continuously exposed. The parastratotype is situated
near the Burgueño Quarry and Apolonia Mine, where the uncon- In Uruguay, a significant extensional and synkinematic magmatic
formable contact between the Las Ventanas Fm. and the basement event corresponding to the final stages of the SW-Gondwana
(Lavalleja Group) is exposed (Fig. 53.3). assembly occurred during the Neoproterozoic– lowermost Cam-
The San Carlos Fm. was erected by Masquelin (1990), who brian (Bossi & Campal 1992; Pecoits 2003b; Oyhantçabal 2005).
documented that the unit consists of conglomerate, sandstone From a structural perspective, the Sierra Ballena Shear Zone
and pelite. The depositional environment was likely lacustrine or (Figs 53.1 & 53.2) constitutes the largest remnant of the

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 555– 564. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.53
556 E. PECOITS ET AL.

50° W
WA
SOUTH 3
AMERICA
A SF

C
Florianópolis
Z

ne
RP

Zo
K

r
ea
Sh
(a) (b) Serra Azul Fm

o
in
rc
N

Ge
54° W

r
ajo
BRAZIL Porto Alegre

M
30° S

Rivera 2
ARGENTINA

Paysandú Melo
e
on
r Z
a
She

an
ce
Ballena

O
URUGUAY
ic
ra

nt

Fig. 53.1. (a) Distribution of cratonic


Sier

Minas 34° S
tla

blocks of west Gondwana: A, Amazonian;


A

1 C, Congo; K, Kaoko; SF, São Francisco;


RP, Rı́o de la Plata; WA, West African; Z,
Buenos Montevideo Zaire. (b) Regional map showing the
Aires Fig. 53.2 location of the units discussed in the text
(see (c) for more detail). (c) Distribution of
volcano-sedimentary early Ediacaran units
of Uruguay and southeastern Brazil: 1, Las
(c) 58° W 54° W 0 100 200 Km Ventanas-San Carlos formations; 2,
Camaquã Basin; 3, Itajaı́ Basin.

Brasilian –Pan African Orogeny (c. 700– 500 Ma). This high- (overlying Arroyo del Soldado Group), where strike– slip shear
strain transcurrent structure, which was operative primarily zones of crustal scale played a major role in the evolution of
between c. 600 and 580 Ma, contributed significantly to the basin- the orogen.
fill architecture of the early Ediacaran units (Oyhantçabal 2005). In At the outcrop scale, the Las Ventanas and San Carlos for-
this regard, the Las Ventanas and San Carlos formations were mations show evidence of both brittle and ductile deformation.
deposited in a strike –slip basin, as indicated by (i) the diverse Small- and large-scale strike –slip faults, westward-verging
depositional facies and their abrupt lateral changes; (ii) apparent detachment faults, and folds with axis sub-parallel to the strike –
migration of the primary depocentre towards the south; (iii) the slip planes are common features (Fig. 53.3). Axial plane slaty
subparallel trend of the basin with respect to its strike –slip and sporadic millimetre-spaced fracture cleavages are present in
margins; and (iv) synchronous timing with regional shearing fine-grained facies (pelite). The basic volcanic and pyroclastic
(Sierra Ballena Shear Zone). Following the development of the rocks show abundant chlorite and epidote as well as pumpellyite
widespread transcurrent system, a gravitational orogenic collapse and prehnite, demonstrating very low- to low-grade metamorphic
characterized by high-angle normal faulting and accompanied by conditions (Pecoits 2003a).
marine transgression occurred during the late Ediacaran –Early
Cambrian (Pecoits et al. 2008).
A similar geotectonic evolution is observed in the associated Stratigraphy
magmatism, which was initiated with highly fractionated
calc-alkaline granite (c. 584 Ma), followed by mildly alkaline The Las Ventanas and San Carlos formations lie on an angular
granite and shoshonitic volcanics (c. 575 Ma), and concluded unconformity above a crystalline basement of undetermined
with peralkaline intrusions and volcanics (c. 540 –520 Ma) age and the Lavalleja Group (Figs 53.4 & 53.5). Relatively well-
(Oyhantçabal et al. 2007). Therefore, the Ediacaran –Early Cam- dated basement granitoids are represented by the Campanero
brian in Uruguay is characterized by a transition from a back-arc Complex and the Cerro Olivo Complex, with ages of c. 1750 Ma
basin (underlying Lavalleja Group), followed by a strike – and 1006 + 37 Ma, respectively (Table 53.1; Oyhantçabal
slip-related basin (Maldonado Group), to a foreland basin 2005 and references therein). The lithostratigraphy of the Lavalleja
LAS VENTANAS AND SAN CARLOS FORMATIONS 557

Fig. 53.2. Simplified geological map showing the distribution of the Las Ventanas and San Carlos formations (Maldonado Group) and selected age determinations of
southeastern Uruguay (see Table 53.1 and text for explanation). Sources of geological information: Bossi & Navarro (1991), Pecoits et al. (2005), Oyhantçabal (2005) and
references therein. SYPSZ, Sarandı́ del Yı́-Piriápolis Shear Zone; SBSZ, Sierra Ballena Shear Zone.

Group, although poorly known, is different from that of the Las The latter, however, shows geochemical signatures, radiometric
Ventanas and San Carlos formations. According to Midot ages and structural features indicating anorogenic magmatism,
(1984), the Lavalleja Group is a volcanosedimentary succession which was extruded after the main deformational phase that
dominated by immature fine-grained siliciclastics, marl, basalt affected the Maldonado Group (Oyhantçabal 2005). In fact, the
and limestone towards the top. This limestone hosts columnar stro- Sierra de las Ánimas Complex systematically displays radiometric
matolites assignable to Conophyton (Poiré et al. 2005), and ages younger than those of the volcanics assigned to the Maldo-
although their occurrence extended from the early Proterozoic to nado Group (see ‘Geochronological constraints’) and it displays
the Ediacaran, preliminary radiometric studies suggest an early neither ductile deformation nor metamorphism. Field relationships
Ediacaran age for this unit. SHRIMP U – Pb detrital zircon analyses show that the Sierra de las Ánimas Complex intrudes the Las
from the Lavalleja Group display ages between 3.4 and 0.6 Ga Ventanas Fm., providing definite evidence of the older age of the
(Basei et al. 2008). Likewise, interbedded basalt shows a crystal- Las Ventanas.
lization age of 590 + 2 Ma (U–Pb SHRIMP; Mallmann et al. Conglomerate-dominated lithofacies (proximal facies associ-
2007), indicating that the deposition must have occurred c. 590 ation of Pecoits 2003a) dominate the basal part of the Las Ventanas
million years ago. Fm., including clast-supported conglomerate and breccia, diamic-
The whole succession (i.e. the Las Ventanas and San Carlos for- tite, massive sandstone and conglomerate –sandstone couplets.
mations) can be divided (from base to top) into three informal Upwards, pelites are abundant with occasional conglomerate
intervals: (i) volcanic and pyroclastic deposits, (ii) conglomerate- beds (pelites-dominated lithofacies or distal facies association
dominated lithofacies and (iii) pelite-dominated lithofacies. Vol- of Pecoits 2003a). This lithofacies includes laminated siltstone
canic and pyroclastic rocks including basalt, mafic hyaloclastic and sandstone-pelite rhythmites, and massive sandstone and
breccias and subaqueous tuff, as well as rhyolite, acidic volcano- conglomerate. Likewise, the San Carlos Fm. consists of basal
clastic and pyroclastic rocks, have been recognized as part of the conglomerate- and upper pelite-dominated lithofacies, but the
Maldonado Group (Fig. 53.4a). This bimodal volcanism has long clast size in the lower conglomerate never reaches that of the
been thought to represent part of the Sierra de las Ánimas Las Ventanas Fm. In this sense, two possible explanations can
Complex (e.g. Sánchez & Rapalini 2002 and references therein). be drawn. First, the San Carlos Fm. represents a lateral equivalent
558 E. PECOITS ET AL.

Fig. 53.3. Geological map of the type area of the Las Ventanas Fm. The inset shows the location of the stratotype (1) and parastratotype (2) (modified from Pecoits
et al. 2008).
LAS VENTANAS AND SAN CARLOS FORMATIONS 559

Fig. 53.4. (a) Stratigraphic column of the Las Ventanas Fm. at its stratotype (point 1, Fig. 53.3). (b) Stratigraphic section for the lower Las Ventanas Fm. at its
parastratotype (point 2, Fig. 53.3) (modified from Pecoits et al. 2008). K-Pg, potassium and plagioclase.

to the middle and uppermost part of the Las Ventanas Fm., The Las Ventanas is thought to be overlain by the Arroyo del
where coarse-grained conglomerate is rare (see ‘Glaciogenic Soldado Group, a thick (3000 m) mixed siliciclastic –carbonate
deposits and associated strata’). Second, both units, although succession, mainly represented by intercalating conglomerate,
potentially contemporaneous (see below), were deposited in sandstone, siltstone, thick carbonate, Fe-formation, black- and
different basins. iron-rich shale and chert. It contains a rich fossil assemblage
560 E. PECOITS ET AL.

proposed for the latter, a glacially influenced system in a strike –


slip setting was suggested for the Las Ventanas and San Carlos
formations (Pecoits et al. 2005, 2008). The presence of abundant
organic-walled microfossils and shelly fauna in the Arroyo del
Soldado Group points to high biological productivity and an
elevated nutrient supply, possibly related to increased weathering
during warmer conditions (Gaucher et al. 2004). Based on avail-
able geochronology, chemostratigraphy and biostratigraphy,
Pecoits et al. (2008) proposed a maximum depositional age of
c. 560 Ma for the group, which is younger than earlier suggestions
(Gaucher et al. 2004).

Glaciogenic deposits and associated strata

Palaeoenvironmental interpretations indicate that the Las Venta-


nas –San Carlos system records sheet flood-dominated fan-delta
deposits in a glacially influenced setting. Direct evidence of
ancient glacial activity comes from the basal and uppermost
facies of the Las Ventanas Fm., where some deposits have been
described and interpreted as ice-rafted diamictite with striated
and faceted clasts and rhythmite-hosting dropstones (Pecoits
2003a; Pecoits et al. 2008).
Diamictites are mainly coarse-grained and matrix-supported
lithofacies with a massive structure displaying normal and
occasionally reversed grading. Clasts are rounded to angular and
range from granule to boulder size. Compositionally, the diamic-
tite dominantly contains extrabasinal clasts (rhyolite, basalt,
granitoid, gabbro, quartzite). Two types of diamictite can be dis-
tinguished: sub-rounded pebbles and cobbles in a massive
muddy matrix, and angular to sub-rounded cobbles to boulders
in a massive to horizontally laminated silty/clayey matrix.
These deposits have a polymodal texture and form relatively
thick (2 –10 m) tabular beds with faceted and occasionally
striated clasts.
Although common in the uppermost part of the succession, fine-
grained rhythmites are also found interbedded with diamictite at
the base of the Las Ventanas Fm. The ,3-m-thick beds are charac-
terized by a millimetre- to centimetre-intercalation of silty and
sandy material with clay and also by the presence of out-sized
clasts deforming the layering. Based on the diverse composition
of these large clasts (granite, basalt, gabbro, etc.), and on the pres-
ence of pre-depositional foliation, an interpretation of the clasts as
volcanic bombs or other ballistic/pyroclastic material is discarded.
These lithofacies (i.e. rhythmites with out-sized clasts and diamic-
tite) were first described at the parastratotype section of the unit
(Burgueño Quarry and Apolonia Mine; Fig. 53.3), but later discov-
ery in other localities (e.g. NE Minas and Melo; Fig. 53.1)
suggesting that they are more extensive than originally thought.
The type section of the Las Ventanas Fm. is largely dominated
by conglomerate, sandstone and finely laminated siltstone. It begins
with a 690-m-thick fining- and thinning-upward cycle (Fig. 53.4a).
Conglomerate, sandstone and laminated siltstone dominate the
lowermost, medial and uppermost sub-cycles, respectively. The
conglomerate is typically granitic and clast-supported with
arkosic sandstone present at the top of each sub-cycle. The follow-
Fig. 53.5. Simplified stratigraphic column of the San Carlos Fm. at its ing changes occur up-section within the lower major cycle: (i) bed
stratotype, 6 km SE of San Carlos town (see Fig. 53.2) (modified from Pecoits thickness progressively decreases, from metre-scale to a few
et al. 2008). millimetres (laminae); (ii) average grain size decreases from
pebbles to silt; (iii) the proportion of granitic clasts becomes
smaller; and (iv) planar parallel stratification and lamination
composed of organic-walled microfossils and small shelly fauna, become a common feature in the siltstone but are absent in the
including the index fossil Cloudina riemkeae (Gaucher et al. lower and middle part of the cycle. This finely laminated siltstone
2004). The distinct lithological differences between Las Ventanas shows similar features to those described in the basal part of the
and San Carlos formations with the Arroyo del Soldado Group unit. Here, the lithofacies is considerably thicker, but the out-sized
have been explained by different prevailing climatic conditions clasts (dropstones) identified in it are smaller and rarely reach more
and tectonic settings. Whereas an evolution towards tropical con- than 10 cm.
ditions (Gaucher et al. 2004) and a marine transgression in a fore- The formation passes up-section into a second major cycle that
land setting (Pecoits et al. 2005; Basei et al. 2008) have been is nearly 560 m thick, and is composed of minor subcycles of
LAS VENTANAS AND SAN CARLOS FORMATIONS 561

Table 53.1. Summary of geochronological data available from southeastern Uruguay (see Fig. 53.2)

Stratigraphic unit Rock type Method Age (Ma)

Sierra de las Ánimas Complex Porphyres Rb–Sr 520 + 5


Syenite Ar–Ar 579 + 1.5
A8 del Soldado Group* Depositional age: c. 560– 530 Ma
El Renegado Granite Rb–Sr 559 + 28
Puntas del Pan de Azúcar Lineament ‘Mylonite’ K–ArMusc 572 + 7
Aiguá Batholith Granite Rb–Sr 587 + 16
Solı́s de Mataojo Tonalite U–Pb 584 + 13
Aguas Blancas Mylonite K–ArMusc 594 + 13
Las Ventanas and San Carlos formations* Depositional age: c. 590– 570 Ma
Las Flores Trachybasalt K–Ar 615 + 30
Quartz-sericite schist U–Pb(SHRIMP) 600–3400
Metabasalt U–Pb(SHRIMP) 590 + 2
Lavalleja Group*
Metarhyolite U–Pb 624 + 14
Metarhyolite U–Pb 667 + 4
Cerro Olivo Complex Orthogneiss U–Pb 1006 + 37
Zanja del Tigre Fm.* Meta-sandstone U–Pb(SHRIMP) 1800–3400
Campanero Complex Orthogneiss U–Pb 1735 + 2
Orthogneiss U–Pb 1754 + 7

Source: Oyhantçabal (2005) and references therein, Mallmann et al. (2007), Basei et al. (2008), Pecoits et al. (2008). *Volcanic-sedimentary units.

sandstone and fine-grained conglomerate. The sandstone has a Chemostratigraphy


tabular geometry, is massive in appearance, and occasionally has
non-erosive basal contacts. The conglomerate is clast-supported, Owing to the lack of carbonate rocks directly associated with the
polymictic and has a modal grain size of 3–10 cm. The clast com- Las Ventanas or San Carlos formations, no chemostratigraphic
position is variable, and includes rhyolite (32%), granite (2%), studies (e.g. C, O, Sr) have been performed.
quartz (12%), basic volcanic rocks (11%), alkaline feldspar (10%),
plagioclase feldspar (8%) and schist (5%). Clasts, either in con-
glomerate or sandstone, are fresh and show no signs of chemical Other characteristics
weathering.
No evidence of mineral deposits was found in any of the units
described here.
Boundary relations with overlying and underlying
non-glacial units
Palaeolatitude and palaeogeography
The Las Ventanas Fm. rests unconformably on the Lavalleja Group.
Palaeoproterozoic orthogneiss (Campanero Complex) and Monzo- The location and kinematic history of the blocks involved during
granite (La Nativa) are of unknown age (Pecoits et al. 2008). A non- the assembly of West-Gondwana in the late Neoproterozoic is
conformity separates the San Carlos Fm. from Palaeoproterozoic poorly known (Rapela et al. 2007). In particular, the palaeogeogra-
ortho- and paragneiss (Cerro Olivo Complex; Fig. 53.2). Particularly phical position of the Rı́o de la Plata Craton (Fig. 53.1 inset) during
important is the relationship with the Lavalleja Group. This unit is the Ediacaran is highly disputed (e.g. Cordani et al. 2000). In this
largely dominated by basalt and immature fine-grained siliciclastic regard, no palaeomagnetic studies have been performed either in
rocks (Midot 1984). An evolution towards warm climate and the sedimentary rocks of Las Ventanas Fm. or in the San Carlos
stable tectono-magmatic conditions is evidenced by thick stromato- Fm. Preliminary mean geomagnetic poles were only obtained
litic limestones developed in the uppermost part of the unit (Poiré from sedimentary rocks of the Playa Hermosa Fm., volcanics
et al. 2005), upon which the Las Ventanas Fm. uncomformably from the Las Ventanas Fm., and volcanics and intrusives from
lies. However, as discussed below, the hiatus between both units the Sierra de las Ánimas Complex (Sánchez & Rapalini 2002).
is poorly constrained. Despite this, the transition from the Lavalleja According to the same authors, the new data support the Apparent
Group to the Las Ventanas Fm. is not only indicated by an angular Polar Wander Path (APWP) previously suggested for the entire
unconformity but also by a strong change in climatic and Gondwana since c. 550 Ma, indicating that the Rı́o de la Plata
tectono-magmatic activity. Craton was indeed at that time part of the supercontinent. Further-
The relationship and nature of the contact between Las Ventanas more, it was suggested that a mean geomagnetic pole obtained
and San Carlos formations with the overlying late Ediacaran Arroyo from the Playa Hermosa Fm. (12.7 þ 9.5/–8.18) meant that
del Soldado Group is not firmly established. The best locality to another example of Neoproterozoic low-latitude glaciation is
study such transition is to the north of Minas (Fig. 53.2) where evident in Uruguay.
both units are closely exposed. Detailed geological mapping of the The most reliable palaeomagnetic pole for early Ediacaran units
area indicates stratal juxtaposition due to tectonic shortening in a of the Rı́o de la Plata Craton is derived from the Campo Alegre
zone of major thrusting. In other words, both units would be separ- lavas (Sánchez & Rapalini 2002). The Campo Alegre Fm., dated
ated by major structural discontinuities (i.e. thrusts oriented SW– by the U –Pb method at 595 + 5 Ma (Citroni et al. 1999), is
NE), and no conformable contacts have been recorded. The located in the Itajaı́ Basin, SE Brazil (Fig. 53.1). Palaeomagnetic
contact between the units – although not well exposed – suggests reconstructions indicate a moderate palaeolatitude of 33.3 +
the presence of an angular unconformity separating the uppermost 9.58S (D’Agrella & Pacca 1988), in marked contrast to the low
fine-grained rhythmites of the Las Ventanas Fm. and the sandstone palaeolatitudes (12.7 þ 9.5/ –8.18) proposed for the Playa Hermosa
of the basal Arroyo del Soldado Group (Yerbal Fm.). Fm. (Sánchez & Rapalini 2002).
562 E. PECOITS ET AL.

In contrast to the scarce database for the early Ediacaran units, couplets and diamictite, while the distal facies includes massive
the APWP for Gondwana since 550 Ma is better known. Since and normally graded sandstone, pebbly sandstone, laminated
550 Ma, poles for the Rı́o de la Plata Craton and other Gondwanan siltstone, and fine-grained massive and graded conglomerate.
continents have tended to form a single APWP ranging from The proximal facies association was interpreted by Pecoits
c. 308S in the late Ediacaran towards lower palaeolatitudes during (2003a) as a subaerial alluvial fan in which debris-flow deposits
the Lower Cambrian (Meert & Van der Voo 1996). (diamictite) and sheet flood deposits (stratified conglomerate and
sandstone) constitute the dominant facies. The subaerial alluvial
fan succession is characterized by upward-coarsening and
Geochronological constraints upward-thickening trends resulting from fan progradation. The
restricted occurrence of debris flow beds and the comparatively
Since the definition of both units, the age of the Las Ventanas and high roundness of the clastic fraction indicate that the preserved
San Carlos formations was considered Ordovician (e.g. Midot succession represents middle and outer regions of the alluvial-fan
1984; Masquelin & Sánchez 1993; Pazos et al. 2003). This complex. The distal facies association is thought to represent a sub-
assumption was challenged by Pecoits et al. (2008) in reporting marine delta subenvironment with sediment gravity-flow deposits
cross-cutting relationships with several intrusive Ediacaran bodies occasionally interbedded with turbidites (massive and graded
and major faults that indicated a minimum depositional age of conglomerate and sandstone) and suspension fallout deposits
c. 570 Ma. The age of deposition of the Las Ventanas and San (laminated siltstone). Although some conglomeratic levels are
Carlos formations is now relatively well constrained to c. 590– interpreted to represent shoreline deposits along the distal fan,
575 Ma by radiometric data based on K – Ar, Rb –Sr and U –Pb no evidence of wave reworking has been observed (Pecoits 2003a).
dating techniques on basement rocks, interbedded basalt, The proximal facies association offers evidence of sedimen-
intrusive syenite, granitic and trachytic dykes, and cross-cutting tation under arid climatic conditions, as shown by exceptionally
faults (Table 53.1). fresh well-rounded clasts (e.g. basalt) in conglomerate and sand-
This inference is supported by the 590 + 2 Ma age (SHRIMP stone. Glacial sedimentary evidence comes from the distal facies
U –Pb) obtained for a metabasalt (Mallmann et al. 2007), the det- association. Therein, outsized clasts within finely laminated silt-
rital zircons from the Lavalleja Group in the basement of the Las stone have been recorded and interpreted as dropstones. This litho-
Ventanas Fm., which show U– Pb (SHRIMP) ages between 3.4 facies has been described at the base and top of the Las Ventanas
and 0.6 Ga (Basei et al. 2008), and an intrusive syenite that Fm. In the first case, the laminated strata are ‘sandwiched’ between
yields Ar/Ar ages of 579 +1.5 Ma (Oyhantçabal et al. 2007). Fur- massive and bedded diamictite, mostly containing extrabasinal
thermore, the ages are corroborated by (i) basic volcanics inter- clasts. Faceted, striated and bullet-shaped clasts are consistent
bedded with sedimentary rocks of the Las Ventanas Fm., which with glacial transport and suggest a glacial influence during the
display ages between 615 + 30 and 565 + 30 Ma (K –Ar deposition of laminated siltstone and diamictite facies. In contrast,
method; Sánchez & Linares 1996); (ii) Rb –Sr ages from intrusive the laminated siltstone described at the top of the unit overlies
granite with an age of 559 + 28 Ma (Preciozzi et al. 1993) and tra- laminated siltstone and fine-grained sandstone, which are inter-
chyte of the Sierra de Las Ánimas Complex, which intrude and preted as turbiditic deposits (Pecoits et al. 2008). Here, the gla-
overly the Las Ventanas Fm., dated between 520–530 Ma (Bossi cially influenced laminated siltstone is differentiated from the
et al. 1993; Linares & Sánchez 1997); (iii) basic dykes cross- turbidites by the lack of turbidity current structures, finer grain
cutting the Las Ventanas Fm. to the south of Minas yielding a size and numerous dropstones with impact-induced deformation
K –Ar age of 485 +12.5 Ma (Poiré et al. 2005); (iv) the last reac- of underlying laminae.
tivation of the Puntas del Pan de Azúcar Lineament, which cross- Despite the lack of evidence for glacially influenced sedimen-
cuts the Las Ventanas Formation (Fig. 53.3), and occurs at tation in the San Carlos Fm., the structural and geochronological
572 + 7 Ma (K –Ar in syn-kinematic muscovites) (Bossi & framework, stratigraphy and fossil content support the premise
Campal 1992); and (v) the San Carlos Fm., which is intensively that the San Carlos Fm. is correlative with the middle– upper
deformed by the Sierra Ballena Shear Zone, in which the third part of the Las Ventanas Fm. (Figs 53.4a & 53.5). This would
and last deformation phase occurred at c. 550 –500 Ma (Oyhantça- explain the absence of the basal and uppermost glacially influ-
bal 2005). Finally, the transcurrent tectonics that occurred during enced facies described for the Las Ventanas Fm.
the early Ediacaran, which is closely related to the generation Only one systematic palaeomagnetic study has been performed
of strike –slip basins recorded for example by the Las Ventanas in glacial Ediacaran units and associated rocks of Uruguay
and San Carlos formations, is also associated with a voluminous (Sánchez & Rapalini 2002); however, none of the sampled sites
syn-kinematic magmatism (Pecoits 2003a; Pecoits et al. 2005). corresponds to the Las Ventanas and San Carlos formations.
Radiometric studies performed in all these bodies yield ages sys- Unfortunately, all the samples from this work were collected
tematically between 570 and 590 Ma (Rb–Sr and U–Pb methods) near the border of the Dom Feliciano belt, in an area affected by
(Bossi et al. 1993; Hartmann et al. 2002; Oyhantçabal 2005). intense Neoproterozoic –Cambrian tectono-magmatic activity,
According to Gaucher et al. (2008), the acritarch assemblage and thus probably affected by widespread remagnetization (e.g.
recovered from the Las Ventanas Fm. indicates a depositional Rapalini & Sánchez 2008). The two palaeopoles obtained from
age between 635 and 582 Ma, supporting previous data (Pecoits volcanic and hypabyssal rocks of the Sierra de las Ánimas
2003a, b). This assemblage, however, comprises and is dominated Complex are, from geochronological and structural points of
by individuals with no stratigraphic value, such as Leiosphaeridia view, poorly constrained. Much of the radiochronology is based
and others of doubtful origin (e.g. Soldadophycus) (Butterfield, on the K – Ar method, which usually provides a minimum age,
pers. comm. 2008). and recent dating on the same lithologies using more precise
methods has yielded ages as much as 30 Ma older (Oyhantçabal
et al. 2007). Although the early Ediacaran units of Uruguay are
Discussion extensively deformed, the interbedded basalt samples for palaeo-
magnetism were not corrected with respect to the palaeohorizontal.
The facies associations point to the development of sheet flood- Field relationships have extensively shown that even the youngest
dominated alluvial fans (Blair & McPherson 1994) intercalated rocks of the Sierra de las Ánimas Complex (i.e. c. 525 million years
with minor lake deposits in a glacially influenced, transtensional old), although not folded, are tilted. Therefore the integrity and
tectonic setting. The proximal facies association comprises usability of these palaeomagnetic data are problematic.
massive and horizontally stratified clast-supported conglomerate Recent palaeogeographical reconstructions locate the Rı́o de
and rare breccia, massive sandstone, conglomerate – sandstone la Plata Craton at high palaeolatitudes c. 580 million years ago
LAS VENTANAS AND SAN CARLOS FORMATIONS 563

(Trinidade & Macouin 2007). Inclination data from deposits Rı́o de la Plata Craton was probably already amalgamated into a
slightly older than the Gaskiers equivalent in the Avalon Terrane single crustal block by the Ediacaran (e.g. Cordani et al. 2000).
(Newfoundland) similarly indicate a palaeolatitude of 358S during Unlike the Las Ventanas –San Carlos formations and their Brazi-
the early Ediacaran (c. 608 Ma, U –Pb zircon age; Myrow & lian correlatives Cerro do Bugio –Bom Jardim allogroups, the
Kaufman 1999). However, palaeogeographical models between Serra Azul Fm. was deposited on a passive margin (Alvarenga
590 and 560 Ma (i.e. when Gaskiers deposits and their possible et al. 2007), showing no comparable tectono-magmatic activity
Ediacaran correlatives, Squantum, Loch na Cille, and Moelv, but probably similar palaeolatitude (see above). The correlation
were formed) are controversial due to the ambiguous results pre- between Ediacaran (c. 590–570 Ma) glacially influenced succes-
sented by the Laurentian palaeopoles (Trinidade & Macouin sions of Brazil and Uruguay strengthen the notion of a post Cryo-
2007 and references therein). For instance, both low and high lati- genian glaciation and suggests that these deposits are distributed
tudes for Laurentia at c. 580 Ma have been proposed. If the latter more extensively than previously recognized in South America.
configuration is confirmed, the glacial strata observed in Laurentia, Future efforts focused on the sedimentological constraints of
Baltica, Cadomia, Avalonia and Rı́o de la Plata cratons are compa- these and other successions (e.g. Itajaı́ Basin in Brazil) are required
tible with a palaeoclimatic scenario similar to the Phanerozoic gla- to determine the glacial influence on the Brazilian units.
ciations rather than ‘snowball’ conditions.
Although early proposals promoted a ‘Marinoan’ age for Las
The authors wish to thank Natalie Aubet who actively participated in the
Ventanas and San Carlos formations (Pecoits et al. 2005), an Edia- fieldwork and in the preparation of the figures. Constructive comments were pro-
caran event seems to be a more reasonable alternative (Pecoits vided by the reviewers Ruben Rieu, Pablo Pazos and editor Emmanuelle Arnaud.
et al. 2008) based on the radiometric constraints. Indeed, this The research was partially funded by the Natural Sciences and Engineering
would explain the absence of thick ‘cap carbonate’ facies immedi- Research Council of Canada Discovery Grants to KOK and MKG. This is a con-
ately overlying these deposits as is distinctive of Ediacaran glacial tribution to IGCP 512 ‘Neoproterozoic Ice Ages”.
deposits. Alternatively, the absence of cap carbonates might be due
to deposition in a highly active tectonic setting, characterized by
high rates of subsidence and high accumulation rates of siliciclas- References
tic sediments, lack of preservation because they were eroded; or,
they may simply not have been found yet. However, the typical Alvarenga, C. J. S. de., Figueiredo, M. F., Babinski, M. & Pinho,
facies of the Las Ventanas and San Carlos formations indicates F. E. C. 2007. Glacial diamictites of Serra Azul formation (Ediacaran,
an active participation of the hydrological cycle that is incompati- Paraguay belt): evidence of the Gaskier event in Brazil. Journal of
ble with the ‘Snowball Earth’ model for Cryogenian glaciations South American Earth Sciences, 23, 236– 241.
(e.g. Hoffman & Schrag 2002). Alvarenga, C. J. S., Boggiani, P. C. et al. 2011. Glacially-influenced
Recently, a similar glacial succession (Tacuarı́ Fm.) was sedimentation of the Puga Formation, Cuiabá Group and Jacadigo
described in NE Uruguay (Veroslavsky et al. 2006). These depos- Group, and associated carbonates of the Araras and Corumbá
its were long-considered a classic example of the Carboniferous – groups, Paraguay Belt, Brazil. In: Arnaud, E., Halverson, G. P.
Permian glaciation in Gondwana (e.g. Bossi & Navarro 1991). & Shields-Zhou, G. (eds) The Geological Record of Neopro-
According to Veroslavsky et al. (2006) and mainly based on a terozoic Glaciations. Geological Society, London, Memoirs, 36,
very similar fossil content to that described for the Las Ventanas 487– 497.
Fm., this unit was tentatively assigned to the Neoproterozoic. Basei, M. A. S., Frimmel, H. E., Nutman, A. P. & Preciozzi, F. 2008.
West Gondwana amalgamation based on detrital zircon ages from
Ongoing research using radiometric dating on cross-cutting grani-
Neoproterozoic Ribeira and Dom Feliciano Belts of South America
tic dykes (U–Pb TIMS) and detrital zircons (LA-ICP-MS) has
and comparison with coeval sequences from SW Africa. In:
confirmed the Ediacaran age of the succession. The obtained Pankhurst, R. J., Trouw, R. A. J., de Brito Neves, B. B. & de
ages constrain the deposition of the unit between 590 and Wit, M. J. (eds) West Gondwana: Pre-Cenozoic Correlations
570 Ma. This is in agreement with the age proposed for the Las Across the South Atlantic Region. Geological Society, London,
Ventanas Fm. (570 –590 Ma) and would suggest regional glacial Special Publications, 294, 239–256.
conditions. Blair, T. C. & McPherson, J. G. 1994. Alluvial fans processes and
Sedimentological, tectonic and magmatic evidence, supported forms. In: Abrahams, A. D. & Parsons, A. J. (eds) Geomorphology
by radiometric ages, suggests some similarities between the Mal- of Desert Environments. Chapman and Hall, London, 354–402.
donado Group and other successions in Brazil. In this regard, the Bossi, J. & Navarro, R. 1991. Geologı́a del Uruguay. Departamento de
Las Ventanas and San Carlos formations have been correlated publicaciones de la Universidad de la República, Montevideo.
with the Bom Jardim (c. 592–573 Ma) and Cerro do Bugio Bossi, J. & Campal, N. 1992. Magmatismo y tectónica transcurrente
(573 –559 Ma) allogroups of the Camaquã Basin located in Rio durante el Paleozoico Inferior en Uruguay. In: Gutiérrez-Marco,
Grande do Sul, southern Brazil (Pecoits 2003b; Fig. 53.1). The J. G., Saavedra, J. & Rabano, I. (eds) Paleozoico Inferior de Iber-
Bom Jardim Allogroup is composed of basic to intermediate oaméric. Mérida, Universidad de Extremadura, 343– 356.
volcanic rocks, alluvial conglomerate and turbidites (Paim et al. Bossi, J., Cingolani, C., Lambias, E., Varela, R. & Campal, N. 1993.
2000). SHRIMP age dating of the volcanics yielded an age of Caracterı́sticas del magmatismo post-orogénico finibrasiliano en el
c. 580 Ma (Paim et al. 2000). The Cerro do Bugio Allogroup con- Uruguay: formaciones Sierra de Rı́os y Sierra de Ánimas. Revista
sists of acidic and basic rocks, alluvial conglomerate, rhythmites Brasileira de Geociências, 23, 282– 288.
(sandstone-pelite) and pelite. Geochronological studies on acidic Citroni, S. B., Basei, M. A. S., Sato, K. & Siga, O., Jr. 1999. Petrogen-
esis of the Campo Alegre Basin magmatism, based on geochemical
rocks yielded a U – Pb age of 573 + 8 Ma (Paim et al. 2000).
and isotopic data. In: 2nd South-American symposium on isotope
Both the Bom Jardim and Cerro do Bugio allogroups were
geology. Extended abstracts, Córdoba, Argentina, 1, 174– 177.
deformed by sinistral transcurrent displacement dated to c. 570 Ma Cordani, U. G., Milani, E. J., Thomaz Filho, A. & Campos, D. A. (eds)
and were subsequently intruded by granitic bodies dated at 2000. Proceedings of the 31st International Geological Congress on
559 + 7 Ma and 565 + 14 to 561 + 6 Ma (Paim et al. 2000 and Tectonic Evolution of South America. Rio de Janeiro, Brazil.
references therein). Evidence of a seasonal glacial influence D’Agrella, M. S. F. & Pacca, I. G. 1988. Paleomagnetism of the Itajaı́,
has been suggested by Eerola (2001, 2006) for the Bom Jardim Castro and Bom Jardim Groups from Southern Brazil. Geophysical
Allogroup. Journal, 93, 365–376.
Recently, Alvarenga et al. (2007) reported glacial deposits in the Eerloa, T. 2001. Climate change at the Neoproterozoic – Cambrian tran-
Ediacaran Serra Azul Fm. in the Paraguay belt, Brazil (Alvarenga sition. In: Zhuravlev, A. Y. & Riding, R. (eds) The Ecology of the
et al. 2011). The Paraguay belt is located on the south-eastern edge Cambrian Radiation. Perspectives in Paleobiology and Earth
of the Amazon Craton (Fig. 53.1), which in conjunction with the History. Columbia University Press, 90– 106.
564 E. PECOITS ET AL.

Eerola, T. 2006. Myöhäisproterotsooiset ilmastonmuutokset Tutkimuk- Pazos, P., Tófalo, R. & Sánchez Bettucci, L. 2003. The record of the
sia Etelä-Brasiliassa. Geologi, 58, 164– 174. Varanger Glaciation at the Rio de la Plata Craton, Vendian-Cambian
Gaucher, C., Sial, A. N., Blanco, G. & Sprechmann, P. 2004. Che- of Uruguay. Gondwana Research, 6, 65 – 77.
mostratigraphy of the lower Arroyo del Soldado Group (Vendian, Pazos, P. J., Rapalini, A. E., Sánchez-Bettucci, L. & Tófalo, O. R.
Uruguay) and palaeoclimatic implications. Gondwana Research, 7, 2011. The Playa Hermosa Formation, Playa Verde Basin, Uruguay.
715– 730. In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The
Gaucher, C., Blanco, G., Chiglino, L., Poire, D. & Germs, G. J. B. Geological Record of Neoproterozoic Glaciations. Geological
2008. Acritarchs of the Las Ventanas Formation (Ediacaran, Society, London, Memoirs, 36, 547–553.
Uruguay): implications for the timing of coeval rifting and Pecoits, E. 2003a. Sedimentologı́a y consideraciones estratigráficas de la
glacial events in western Gondwana. Gondwana Research, 13, Formación Las Ventanas en su área tipo, departamento de Maldo-
488– 501. nado, Uruguay. Revista de la Sociedad Uruguaya de Geologı́a,
Hartmann, L., Santos, O., Bossi, J., Campal, N., Schipilov, A. & Special Issue, 1, 124–140.
McNaughton, N. J. 2002. Zircon and Titanite U– Pb SHRIMP geo- Pecoits, E. 2003b. Age and preliminary correlation of the Las
chronology of Neoproterozoic felsic magmatism on the eastern Ventanas Formation and Bom Jardim-Cerro do Bugio allogroups
border of the Rı́o de la Plata Craton, Uruguay. Journal of South Amer- (Vendian, Uruguay and Brazil). In: Frimmel, H. E. (ed.) 3rd Inter-
ican Earth Sciences, 15, 229– 236. national Colloquium Vendian-Cambrian of W-Gondwana. University
Hoffman, P. F. & Schrag, D. P. 2002. The Snowball Earth hypo- of Cape Town, Cape Town, South Africa, 32 – 34.
thesis: testing the limits of global change. Terra Nova, 14, Pecoits, E., Aubet, N., Oyhantçabal, P. & Sánchez Bettucci, L.
129– 155. 2005. Estratigrafı́a de sucesiones sedimentarias y volcanosedimentar-
Linares, E. & Sánchez Bettucci, L. 1997. Edades Rb/Sr y K/Ar del ias Neoproterozoicas del Uruguay. Revista de la Sociedad Uruguaya
cerro Pan de Azúcar, Piriápolis, Uruguay. South American Sym- de Geologı́a, 11, 18 –27.
posium on Isotope Geology, San Pablo, 1, 176– 180. Pecoits, E., Gingras, M., Aubet, N. & Konhauser, K. 2008. Ediacaran
Mallmann, G., Chemale, F., Jr., A ´ vila, J. N., Kawashita, K. & in Uruguay: palaeoclimatic and palaeobiologic implications. Sedi-
Armstrong, R. A. 2007. Isotope geochemistry and geochronology mentology, 55, 689– 719.
of the Nico Pérez Terrane, Rı́o de la Plata Craton, Uruguay. Gond- Poiré, D. G., González, P. D., Canalicchio, J. M., Repetto, F. G. &
wana Research, 12, 489– 508. Canessa, N. D. 2005. Estratigrafı́a del Grupo Mina Verdún, Proter-
Masquelin, H. 1990. Análisis estructural de las zonas de cizalla en las ozoico de Minas, Uruguay. Latin American Journal of Sedimentology
migmatitas de Punta del Este – Uruguay. Acta Geologica Leopolden- and Basin Analysis, 12, 125– 143.
sia, 30, 139 –158. Preciozzi, F., Masquelin, H. & Sánchez Bettucci, L. (eds) 1993.
Masquelin, H. & Sánchez, L. 1993. Propuesta de evolución tectono- Geologı́a de la Porción Sur del Cinturón Cuchilla de Dionisio. In:
sedimentaria para la fosa tardi-brasiliana en la región de Piriápolis, 1st Simp. Int. Neoproterozoico – Cámbrico de La Cuenca del
Uruguay. Revista Brasileira de Geociências, 23, 313– 322. Plata. Guı́a de Excursiones, La Paloma, Uruguay.
Meert, J. G. & Van der Voo, R. 1996. Paleomagnetic and 40Ar– 39Ar Rapalini, A. E. & Sánchez, L. B. 2008. Widespread remagnetization of
study of the Sinyai dolerite, Kenya: implications for Gondwana late Proterozoic sedimentary units of Uruguay and the apparent polar
assembly. Journal of Geology, 104, 131– 142. wander path for the Rio de La Plata craton. Geophysical Journal
Midot, D. 1984. Etude geologique et diagnostic metallogénique pour International, 174, 55– 74.
l’exploration du secteur de Minas (Uruguay). Diplome de Docteur Rapela, C. W., Pankhurst, R. J. et al. 2007. The Rı́o de la Plata craton
de 3è Cycle, Universitè Pierre et Marie Curie. and the assembly of SW Gondwana. Earth-Science Reviews, 83,
Myrow, P. M. & Kaufman, A. J. 1999. A newly discovered cap carbonate 49 – 82.
above Varanger-age glacial deposits in Newfoundland, Canada. Sánchez, L. & Linares, E. 1996. Primeras edades en basaltos del
Journal of Sedimentary Research, 6, 784– 793. Complejo Sierra de las Ánimas, Uruguay. In: 13th Congreso Geoló-
Oyhantçabal, P. 2005. The Sierra Ballena Shear Zone: kinematics, gico Argentino y 3rd Congreso de Exploración de Hidrocarburos,
timing and its significance for the geotectonic evolution of southeast Actas, La Plata, Argentina, I, 399– 404.
Uruguay. PhD thesis, Georg-August-Universität zu Göttingen. Sánchez, L. & Rapalini, A. E. 2002. Paleomagnetism of the Sierra de Las
Oyhantçabal, P., Siegesmund, S., Wemmer, K., Frei, R. & Layer, P. Animas Complex, southern Uruguay: its implications in the assembly
2007. Post-collisional transition from calc-alkaline to alkaline of western Gondwana. Precambrian Research, 118, 243–265.
magmatism during transcurrent deformation in the southernmost Trindade, R. I. F. & Macouin, M. 2007. Palaeolatitude of glacial depos-
Dom Feliciano Belt (Braziliano –Pan-African, Uruguay). Lithos, 98, its and palaeogeography of Neoproterozoic ice ages. Comptes Rendus
141– 159. Geoscience, 339, 200–211.
Paim, P. S. G., Chemale, F., Jr. & Lopes, R. da 2000. A Bacia do Veroslavsky, G., de Santa Ana, H. & Daners, G. 2006. Tacuarı́
Camaquã. In: Holz, M. & De Ros, L. F. (eds) A geologı́a do Rio Formation (Nov. Nom.): lithostratigraphy, facies, environment, age
Grande do Sul. CIGO-Universidade Federal Rio Grande do Sul, and geological significance (Cerro Largo-Uruguay). Revista de la
Rio Grande do Sul, 231– 274. Sociedad Uruguaya Geologı́a, 13, 23 –35.
Chapter 54

The controversial stratigraphy of the glacial deposits in the Tandilia System, Argentina

P. J. PAZOS1,2* & A. RAPALINI1,2


1
Departamento de Ciencias Geológicas, FCEN, Ciudad Universitaria, Buenos Aires 1428, Argentina
2
Consejo Nacional de Investigaciones Cientı́ficas y Técnicas (CONICET)
*Corresponding author (e-mail: pazos@gl.fcen.uba.ar)

Abstract: In the Ediacaran–Cambrian Tandilia System of central Argentina, the glacial origin of the thin (10 m) Sierra del Volcán For-
mation (Fm.) has been recognized for many years (Spalletti & del Valle 1984), being the first undisputed glacial deposits recorded in the
Rı́o de la Plata craton. It consists of three units: (i) a basal polymictic diamictite with a kaolin-rich matrix, (ii) a middle pelite with
heterolithic levels, undulatory stratification, symmetric ripples, and outsized clasts previously interpreted as dropstones, and (iii) an
upper polymictic diamictite with subtle normal grading. Here, this poorly age-constrained formation is reviewed from a stratigraphic
and regional context and compared with other putative glaciogenic intervals at the base of the Cerro Largo Fm. and in the Punta
Mogotes borehole.

The Tandilia System consists of an Ediacaran – Ordovician 2008). The last confirmed evidence of magmatism in the Tandilia
package of sedimentary rocks that crops out in a 300-km-long, System is a series of diabase sills (Los Barrientos) dated by
NW –SE-trending belt south of Buenos Aires, Argentina Rapela et al. (1974) as lower Ordovician and probably associated
(Fig. 54.1). The Tandilia System comprises the Sierras Bayas with the Ordovician Famatinian magmatic cycle in the western
Group and the unconformably overlying Cerro Negro and Balcarce Proto-Andean continental margin of Gondwana (Pankhurst &
formations. Whereas the stratigraphic and tectonic evolution of the Rapela 1998).
Tandilia System has received some recent attention (e.g. Gómez Teruggi & Kilmurray (1975) proposed that the present landscape
Peral et al. 2007), its glaciogenic record remains poorly described of the Tandilia system is due to Miocene Andean reactivation of
in the international literature. The glaciogenic origin of the Sierra basement faults. However, from a detailed geomorphological
del Volcán Fm., which occurs in the southern part of the Tandilia study, Demoulin et al. (2005) recognized two regional planation
belt, has long been accepted. However, its age and stratigraphic surfaces, which suggests that most of the present landscape is sig-
position, reviewed recently by Pazos et al. (2008), have remained nificantly older (Palaeogene or even Cretaceous), and that the
ambiguous, having been variably assigned to the Precambrian, region has experienced slow denudation rates.
Cambrian and Ordovician. A putative glacial diamictite occurs at
the base of the Cerro Largo Fm. in the lower Sierras Bayas
Group and crops out near the town of Olavarrı́a in the northwestern Stratigraphy
part of the Tandilia belt (Fig. 54.1). The relationship between this
diamictite and the Sierra del Volcán Fm., and more generally the The main geological units within the Tandilia belt are a Palaeopro-
stratigraphic framework for the Tandilia System, is poorly under- terozoic igneous and metamorphic basement (The Buenos Aires
stood, but a relatively complete section preserved in the Punta Complex) and a Neoproterozoic to Early Palaeozoic cover com-
Mogotes borehole in the SE of the belt (Marchese & di Paola prised of Ediacaran (?) calcareous and clastic sediments of the
1975) sheds light on the Tandilia stratigraphy. Sierras Bayas Group, the clastic (Cambrian) Cerro Negro Fm.,
and quartzite of the Ordovician Balcarce Fm., which gently over-
laps the underlying units and marks an extensive transgression
Structural framework across the basin.
The Buenos Aires Complex includes various igneous and meta-
The Tandilia system (Dalla Salda et al. 1988; Cingolani & Dalla morphic suites. Sm–Nd model ages averaging 2620 + 80 Ma on
Salda 2000) crops out in a 300-km-long chain of low hills extend- several magmatic units suggest a Late Archaean history (Pankhurst
ing in a NW –SE direction from the town of Olavarrı́a in the centre et al. 2003). Hartmann et al. (2002) recognized orogenic peaks
of the Buenos Aires province of Mar del Plata on the Atlantic coast around 2.16 Ga and 2.08 Ga in Uruguay (eastern part of the craton).
(Fig. 54.1). It rests on a Palaeoproterozoic basement assigned to The Sierras Bayas Group corresponds approximately to what was
the Buenos Aires Complex (Marchese & di Paola 1975). It is originally described as the La Tinta Fm. (Iñiguez Rodriguez et al.
unmetamorphosed (Gómez Peral et al. 2007) and only lightly 1989; Marchese & di Paola 1975) and the Sierras Bayas Fm.
folded locally, such as near Olavarrı́a in the NW part of the belt (Borrelo 1962). This group, according to Poiré & Spalletti (2005),
(Massabie & Nestiero 2005). A low-angle unconformity occurs presents a simple stratigraphy and covers the basement rocks of
between the Cambrian Cerro Negro Fm. and the overlying Ordovi- the Buenos Aires Complex (Fig. 54.2). The basal Villa Mónica
cian Balcarce Fm. Glaciogenic deposits of the Sierra del Volcán Fm. (52–70 m thick) is composed of quartz-arenite, pelite, con-
Fm. are completely undisturbed, cropping out with subhorizontal glomerate and cherty dolomite beds with stromatolites (Poiré &
bedding and no metamorphic overprint. Spalletti 2005). It is overlain by the Cerro Largo Fm., which is c.
The Sierras Bayas Group records sedimentation in a shallow 40 m thick and includes a diamictite containing mainly chert
marine basin with strong tidal influence and reworking (see clasts and slumped beds at the base. Poiré & Spalletti (2005) inter-
Poiré & Spalletti 2005). The Balcarce Fm. was deposited during pret this diamictite as glaciogenic in origin. The diamictite passes
a subsequent sedimentary cycle in a more extensive basin. It over- upward into multicolour pelite, glauconitic sandstone units and
lies palaeohighs of basement or the Sierras Bayas Group rocks, whitish quartzite beds containing hummocky, planar and trough
with a very low angle of unconformity in places (Pazos et al. cross-stratification. The succession ends with a shallowing-upward

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 565– 569. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.54
566 P. J. PAZOS & A. RAPALINI

Fig. 54.1. The Tandilia System and


outcrops of the Sierra del Volcán Fm. at del
Volcán Hill. The Punta Mogotes Borehole
is located just south of Mar del Plata.

claystone cycle (Gómez Peral et al. 2007). The Olavarrı́a Fm., Fm. is laterally equivalent to the Las Águilas, the Punta Mogotes
which follows upwards, is composed of red, fine-grained rippled and the Sierra del Volcán formations. We propose a distinct strati-
sandstone, pelite and heterolithic sediments deposited in a graphic framework (Fig. 54.2) based on a description by Marchese
shallow, marginal marine environment judging from the occurrence & di Paola (1975) of the Tandilia stratigraphy in the Punta Magotes
of bimodal palaeocurrents (Poiré et al. 2005). The youngest unit of borehole in the southestern part of the belt. These authors recog-
the Sierras Bayas Group is the Loma Negra Fm., which is c. 40 m nized eight units within the well. The lowermost, unit 8 (89 m
thick and composed almost exclusively of reddish and black lime- thick), is composed of metapelite beds that they redefined as the
stone deposited in an open marine ramp to lagoonal environment. It Punta Mogotes Fm. and correlated with the regional basement.
contains Cloudina, which indicates a late Ediacaran age and tropical Units 7– 5 comprise quartz arentite and heterolithic sediments.
depositional environment (Poiré & Spalletti 2005; Gaucher et al. Unit 4 is a diamictite (6 m thick), which we suspect to be glacio-
2003, 2005; Gómez Peral et al. 2007, Pecoits et al. 2008). genic (see below). It is followed by more siliciclastic units 2–3,
The overlying Cerro Negro Fm. is disconformably separated which may correlate with either the Cerro Negro or Balcarce for-
from the upper Sierras Bayas Group by a palaeokarst surface mations. Unit 1 is assigned to the Balcarce Fm.
(Barrio et al. 1991). This .150-m-thick unit is composed of
reddish fine-grained sandstone to claystone deposited in a
shallow marine to tidal environment. Poiré & Spalletti (2005) cor- Glaciogenic deposits and associated strata
related the Cerro Negro Fm. with the Sierra del Volcán and Punta
Mogotes formations, which occur in the southeastern part of the The glaciogenic deposits of the Sierra del Volcán Fm. are extre-
belt, and suggested a Cambrian age for all three units. Gómez mely thin (10 m thick), and the outcrops are almost entirely
Peral et al. (2007, fig. 2) instead argued that the Cerro Negro covered by vegetation and debris from the overlying Balcarce
Fm. Exposures occur on private property with restricted access
(Estancia Cerro Del Volcán). The sedimentological description
is based on recent fieldwork and descriptions given in Spalletti
ERAS LITOSTRATIGRAPHIC UNITS SE REGION & del Valle (1984). The unit was divided into three intervals: (i)
PERIODS URUGUAY
NW REGION CENTRAL REGION Borehole a lower diamictite, (ii) a middle heterolithic and pelitic unit, and
upper
ORDOVICIAN
Balcarce Formation
Balcarce Formation
Balcarce Formation
Las Ventanas (iii) an upper diamictite. The lower diamictite is polymictic with
Formation
clasts of the Buenos Aires complex, but no carbonate clasts. The
matrix contains abundant kaolinite, reminiscent of the saprolitized
CAMBRIAN
Cerro Negro Cerro Negro CerroNegro basement and the matrix of the Balcarce Fm. The middle interval is
Formation Formation Arroyo
Formation (units 2-3)
del Soldado
rich in outsized clasts disrupting the stratification, which exhibits
Group undulatory and faintly developed ripples. The overlying diamictite
Loma Negra Loma Negra
Formation
Palaeohigh ? is normally graded (Spalletti & del Valle 1984). Evidence of a
Sierras Bayas Group

Formation
LATE Del Volcán Playa Hermosa glacial influence in deposition of the Sierra del Volcán Fm.
Olavarría
EDIACARAN
Formation Las Águilas
Formation UNIT 4 Formation includes glacial pavements, rainout deposits, faceted clasts and
and Sierras Bayas matrix-supported diamictite beds with immature composition
Cerro Largo Cerro Largo Group
Formation Formation (units 5-7) (Pazos et al. 2008).
Another diamictite occurs at the contact between the Villa
Monica and Cerro Largo formations (Gómez Peral et al. 2007,
Fig. 54.3) in the northwestern sector of the belt. The slumped dia-
Villa Villa mictite contains mainly chert clasts and was interpreted by Poiré &
EDIACARAN
Mónica
Formation
Mónica
Formation
Spalletti (2005) as glacial in origin, although no direct evidence of
glacial influence has been reported. Another diamictite occurs at a
depth of 264 –270 m (unit 4) in the Punta Mogotes borehole
PALAEO-MESO
BUENOS AIRES COMPLEX Punta Mogotes
Formation
ARCHAEAN
COMPLEX
(Marchese & di Paola 1975). This 6-m-thick diamictite contains
PROTEROZOIC (unit 8)
fragments of feldspar and of metamorphosed pelite believed to
be derived from the Punta Mogotes Fm. and is compositionally dis-
Fig. 54.2. Integrated stratigraphy (after Poiré & Spalletti 2005), including the tinct from bracketing sediments. No unequivocal evidence for gla-
Punta Mogotes borehole stratigraphy suggested in this chapter, and the ciation has been observed, but based on the stratigraphic
suggested stratigraphic position for the Sierra del Volcán Fm. Diamictite arguments described below, we tentatively correlate unit 4 with
intervals marked with triangles. the Sierra del Volcán Fm.
THE TANDILIA SYSTEM, ARGENTINA 567

low-angle unconformity (Spalletti & del Valle 1984). The nature


of the contacts of the diamictite comprising unit 4 in the Punta
Mogotes borehole is unkown.

Chemostratigraphy

Gómez Peral et al. (2007) published reconnaissance chemostrati-


graphic data on the Villa Mónica and Loma Negra formations of
the Sierras Bayas Group. To summarize, the Villa Mónica carbon-
ates are heavily diagenetically altered, with very low Sr concen-
trations and enrichments in Fe and Mn. d13C values define a
narrow range from –0.5 to 2.2‰. The Loma Negra Fm. is moder-
ately altered, with Sr contents of 300 –400 ppm. d13C carbonate
values cluster between 3 and 4‰, whereas organic matter values
lie between – 27 and –28‰.

Palaeolatitude and palaeogeography

The available Neoproterozoic palaeomagnetic data from the Rı́o


de la Plata (RP) craton were reviewed in Tohver et al. (2006)
and Rapalini (2005, 2006). A preliminary apparent polar wander
track for the interval 600– 500 Ma has been proposed recently
for this craton (Rapalini 2006; Fig. 54.4). The available palaeo-
magnetic database to constrain the palaeogeographical evolution
of the RP craton consists of four palaeomagnetic poles and one
mean geomagnetic pole. The latter means that there is no guarantee
of full compensation of non-dipole components of the Earth
magnetic field.
Fig. 54.3. Detailed logged section of the Sierra del Volcán Fm. taken from
The oldest available pole was obtained from the Campo Alegre
Pazos et al. (2008). (a) Location of the interval studied. (b) Logged section
lavas (CA, D́Agrella Filho & Pacca 1988) in Brazil. These rocks
originally from Spalletti & del Valle (1984).
were accurately dated as 595 + 5 Ma (U – Pb, Citroni et al.
1999). The similar position of the mean geomagnetic pole for
the Playa Hermosa sediments (PH) to CA (Fig. 54.4) suggests a
Boundary relations with overlying and underlying broadly similar age for both (Pazos et al. 2011).
non-glacial units The other poles in the APWP are nearly identical to those of the
older units of the Sierra de Animas Complex (SA2, Sanchez Bet-
The Sierras Bayas Group is bounded by stratigraphic unconformi- tucci & Rapalini 2002) and Los Barrientos claystone (LB, Rapalini
ties (e.g. Poiré & Spalletti 2005; Gómez Peral et al. 2007). The 2006) in the Tandilia System (Table 54.1). Originally, an age of c.
non-conformity that separates the basement (Buenos Aires 550 Ma was assigned to SA2 based on old K –Ar and Rb –Sr ages.
Complex) from the Sierras Bayas Group represents hundreds of Oyhantçabal et al. (2007) published an Ar/Ar age of 579 + 2 Ma
millions of years during which the basement was intensely chemi- on hornblende for the Pan de Azucar syenite, interpreted as a
cally weathered. An unconformity also separates the Villa Mónica cooling age of that body. LB is a high-quality pole with no firm
Fm. from the overlying Cerro Largo Fm., but the nature and mag- age constraints, although its position being similar to SA2 suggests
nitude of the gap is subject to debate (Pazos et al. 2008). an age between 580 and 570 Ma.
The Sierra del Volcán Fm. sits non-conformably on the Taking into consideration the apparent polar wander path in
basement and is separated from the overlying Balcarce Fm. by a Figure 54.4 and Ediacaran to Cambrian poles from other

Table 54.1. Available Neoproterozoic palaeomagnetic poles for the Rio de la Plata

Geographical Formation/lithology Palaeomagnetic pole (PP) Palaeomagnetic results Palaeopoles age Q Ref.
location (lat.) (long.) A95 (dp/dm)

Uruguay Sierra de las Animas 5.98 338.18 (19.68/26.78) AF/Th demag, N ¼ 7 (n ¼ 33), IRM Cambrian (520– 3 1
Complex (SA1) 500 Ma)
Uruguay Sierra de las Animas –16.98 250.98 (15.98/21.58) AF/Th demag, N ¼ 6 (n ¼ 27), IRM, both Ediacaran 4 1
Complex (SA2) polarities (579 Ma)
Tandilla System Sierra de los –15.18 252.68 (10.98/14.28) AF/Th demag, N ¼ 12 (n ¼ 56), both polarities, Ediacaran? 6 2
Barrientos reversal of EMF recorded. Correction for
claystones (LB) remanence anisotropy
Uruguay Playa Hermosa Fm. –43.08 198.48 (8.68/16.08) AF/Th demag, N ¼ 2 (n ¼ 6), IRM, small number Early Ediacaran? 3 1
(PH) of samples
Brazil Campo Alegre lavas –57.08 223.08 9.08 AF/Th demag, N ¼ 6 (n ¼ 46), both polarities. Early Ediacaran 5 3
(CA) (595 Ma)

Q, quality factor of the pole (maximum, 7) according to criteria set by Van der Voo (1990). 1, Sánchez Bettucci & Rapalini (2002); 2, Rapalini (2006); 3, D́Agrella Filho &
Pacca (1988). EMF, Earth magnetic field; IRM, isothermal remanent magnetization; AF, alternating field; Th, thermal.
568 P. J. PAZOS & A. RAPALINI

observations and interpretations of the Sierra del Volcán Fm. can


be summarized as follows:
† The diamictite at the base and top of the unit (Fig. 54.3) appears
500
to have been deposited by debris flows, as is particularly
525 SA1 evident in the normal grading of the upper diamictite.
† The clasts in the diamictite beds are derived mainly from the
metamorphic rocks of the Buenos Aires Complex, as well as
unmetamorphosed quartzite clasts and blocks, probably
510
belonging to the quartzite beds of the Sierras Bayas Group
535 present in the Villa Mónica (lower quartzite) and Cerro Largo
(upper quartzite) formations.
† The micromorphology of quartz grains is consistent with a
glacial origin (van Staden et al. 2005).
† The middle unit contains dropstones that clearly disrupt
550 bedding (Spalletti & del Valle 1984) and symmetric ripples
indicating a combination of open water and floating ice.
LB
SA2 (579) The diamictite documented at the base of the Cerro Largo Fm. is a
cherty breccia that lacks evidence of glacially influenced sedimen-
tation. The diamictite interval documented in the Punta Mogotes
CA (595 ) borehole (unit 4) described in detail by Marchese & di Paola
(1975) is very thin (6 m thick) and differs compositionally with
PH
both underlying and overlying sediments. It contains fresh feldspar
and angular basement clasts floating in a clayish matrix. These fea-
tures indicate a low degree of weathering of clast material, but not
Fig. 54.4. Palaeomagnetic poles and their respective circles (or ovals) of sufficient to assign an unambiguous glacial origin (e.g. Eyles &
confidence for the Rio de la Plata (RP) craton (black squares with white ovals), Januszczak 2004). The absence of carbonate clasts, despite the
and other Gondwana blocks (white circles with grey ovals). Poles from low weathering index, imply that the unit 4 diamictite is older
remagnetized units (Rapalini & Sanchez Bettucci 2008) have been omitted. The than the Loma Negra Fm., and thus indisputably Neoproterozoic
hypothetical apparent polar wander track for this craton is shown in dashed in age. The Sierra del Volcán Fm. similarly lacks carbonate
lines. Numbers indicate approximate ages of poles in Ma. Data for RP poles are clasts and we suggest that it correlates with the unit 4 diamictite
presented in Table 54.1. Gondwana poles selected from Trindade et al. (2006). (Fig. 54.2). This correlation requires an explanation for the
SA1, Sierra de Animas 1; SA2, Sierra de Animas 2; PH, Playa Hermosa; LB, absence of the Loma Negra carbonates in the Punta Mogotes bore-
Los Barrientos; CA, Campo Alegre poles.
hole. We interpret this to be the result of a topographic high in the
Punta Mogotes region, which was likely also a source area for
Gondwanan cratonic blocks (Trindade et al. 2006), the RP craton clasts in the Sierra del Volcán Fm.
appears to have lain between 308S and 608S through the Ediacaran
and Early Cambrian. We thank Dr R. Somoza and Professor V. Ramos for providing bibliography, dis-
cussions and stimulation, although we stress that we accept full responsibility for
the ideas presented here. I want to express my gratitude (P.P.) to Cheryl Turking-
Geochronology ton for her unconditional support. This is the contribution R-38 of the Instituto de
Estudios Andinos Don Pablo Groeber. This represents a contribution of the IUGS
In the Villa Mónica Fm., Rb –Sr data on diagenetic clay minerals and UNESCO-funded IGCP (International Geoscience Programme) project #512.
yielded an age of 795 + 28 Ma on a green shale. A Rb – Sr age of
769 + 12 Ma was obtained on shale deposits of the overlying
Cerro Largo Fm. (Bonhomme & Cingolani 1980), while Cingolani References
& Bonhomme (1988) and Kawashita et al. (1999) suggested a
depositional age of c. 730 Ma for the same formation based on Barrio, C. A., Poiré, D. G. & Iñiguez Rodrı́guez, A. M. 1991. El contacto
Rb –Sr whole-rock data. However, these ages are all regarded as entre la Formación Loma Negra (Grupo Sierras Bayas) y la Formación
too old and, based on the palaeomagnetic considerations described Cerro Negro, un ejemplo de paleokarst, Olavarrı́a, provincia de Buenos
above, an Ediacaran age for the Sierra Bayas Group is more likely. Aires. Revista de la Asociación Geológica Argentina, 46, 69–76.
Detrital zircon geochronology (Rapela et al. 2007; Gaucher Bonhomme, M. G. & Cingolani, C. 1980. Mineralogı́a y geocronologı́a
Rb –Sr y K–Ar de fracciones finas de la Formacion ‘La Tinta’,
et al. 2007) has provided no meaningful age constraints on the
provincia de Buenos Aires. Revista de la Asociación Geológica
Sierras Bayas Group, but has established a maximum Early Ordo-
Argentina, 35, 519–538.
vician age for the Balcarce Fm. (Rapela et al. 2007). Borello, A. 1962. Formación Punta Mogotes (Eopaleozoico Provincia de
The body fossil Cloudina riemkeae occurs in the Loma Negra Buenos Aires). Comisión Nacional de Investigaciones Cientı́ficas.
Fm. (Gaucher et al. 2005; Pecoits et al. 2008) and confirms an Notas 1, 1. La Plata.
Ediacaran age, at least for the upper Sierras Bayas Group. Cingolani, C. & Bonhomme, M. G. 1988. Resultados geocronológicos
Chauria circularis has been reported from Villa Mónica and en niveles pelı́ticos intercalados en las dolomı́as de las Sierras
Cerro Negro formations and leiosphaeridia-dominated acritarch Bayas (Grupo La Tinta), provincia de Buenos Aires. Actas Segundas
assemblages are found from the upper Villa Mónica Fm. to the Jornadas Geológicas, Actas I, 2283– 2389.
middle Cerro Negro Fm. (Gaucher et al. 2005). Cingolani, C. & Dalla Salda, L. 2000. Buenos Aires Cratonic region.
In: Cordani, U., Milani, E., Thomaz Filho, A. & Campos, D. (eds)
Tectonic Evolution of South America. 31st International Geological
Discussion Congress, Brazil, Rio Janeiro, 139–147.
Citroni, S. B., Basei, M. A. S., Sato, K. & Siga, O. Jr. 1999. Petrogen-
The diamictite beds of the Sierra del Volcán Fm. are identified as esis of the Campo Alegre basin magmatism, based on geochemical
glacial in origin based on a range of features and interpretations and isotopic data. II South American Symposium on Isotope
described in detail in Pazos et al. (2008). The important Geology, Cordoba, Argentina, Actas, 174– 177.
THE TANDILIA SYSTEM, ARGENTINA 569

D’Agrella Filho, M. S. & Pacca, I. I. G. 1988. Paleomagnetism of the B. B. & de Witt, M. J. (eds) West Gondwana Pre-Cenozoic Corre-
Itajaı́, Castro and Bom Jardim Groups from Southern Brazil. Geophy- lations Across the South Atlantic Region. Geological Society,
sical Journal International, 93, 365–376. London, Special Publications, 294, 343–364.
Dalla Salda, L., Bossi, J. & Cingolani, C. 1988. The Rio de la Plata cra- Pazos, P. J., Rapalini, A., Sánchez Bettucci, L. & Tófalo, R. O. 2011.
tonic región of southwestern Gondwana. Episodes, 11, 263–269. The Playa Hermosa Formation, Playa Verde Basin, Uruguay.
Demoulin, A., Zárate, M. & Rabassa, J. 2005. Long-term landscape In: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The
development: a perspective from the southern Buenos Aires ranges Geological Record of Neoproterozoic Glaciations. Geological
of east central Argentina. Journal of South American Earth Sciences, Society, London, Memoirs, 36, 547– 553.
19, 193– 204. Pecoits, E., Gingras, M., Aubet, N. & Konhauser, K. 2008. Ediacaran
Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model in Uruguay: palaeoclimatic and palaeobiological implications.
for Neoproterozoic glaciations during the break up of Rodinia after Sedimentology, 55, 689– 721.
750 Ma. Earth-Science Reviews, 65, 1 – 73. Poiré, D. G. & Spalletti, L. A. 2005. La cubierta sedimentaria precám-
Gaucher, C., Boggiani, P. C., Sprechmann, P., Sial, A. N. & Fiar- brica/paleozoica inferior del Sistema de Tandilia. In: Barrio, R. E.,
child, T. 2003. Integrated correlation of the Vendian to Cambrian Etcheverry, R. O., Caballé, M. F. & Llambı́as, E. J. (eds) Geolo-
Arroyo del Soldado and Corumbã Groups (Uruguay and Brazil) – gı́a y Recursos Minerales de la provincia de Buenos Aires. Relatorio
palaeogeographic, palaeoclimatic and palaeobiologic implications. del XVI Congreso Geológico Argentino, 51 –68.
Precambrian Research, 120, 241– 278. Poiré, D. G., Canalicchio, J. & Alonso, G. 2005. Las calizas del
Gaucher, C., Poiré, D. G., Peral, L. & Chiglino, L. 2005. Litoestrati- Sistema de Tandilia y su utilización en la industria cementera. In:
grafı́a, Bioestratigrafı́a y correlaciones de las sucesiones sedimentar- Barrio, R. E., Etcheverry, R. O., Caballé, M. F. & Llambı́as, E.
ias del Neoproterozoico –Cambrico del Cratón del Rı́o de la Plata J. (eds) Geologı́a y Recursos Minerales de la provincia de Buenos
(Uruguay y Argentina). Latin American Journal of Sedimentology Aires. Relatorio del XVI Congreso Geológico Argentino, 387– 396.
and Basin Analysis, 12, 145–160. Rapalini, A. E. 2005. The accretionary history of Southern South
Gaucher, C., Poiré, D., Finney, S. C., Valencia, V., Blanco, G., America from the latest Proterozoic to the Late Paleozoic: some
Pamoukaghlian, K. & Gómez Peral, L. 2007. Zircones detrı́ticos paleomagnetic constraints. In: Vaughan, A., Leat, P. & Pan-
de secuencias neoproterozoicas de Uruguay y Argentina: inferencias khurst, R. J. (eds) Terrane Processes at the Margins of Gondwana.
sobre la evolución paleogeográfica del cartón del Rı́o de la Plata. Geological Society, London, Special Publication, 246, 305– 328.
V Congreso Uruguayo de Geologı́a, Montevideo, Uruguay. Rapalini, A. E. 2006. New Late Proterozoic paleomagnetic pole for the
(CD-ROM, extended paper). Rio de la Plata craton: implications for Gondwana. Precambrian
Gómez Peral, L., Poiré, D. G., Strauss, H. & Zimmermann, U. 2007. Research, 147, 223–233.
Chemostratigraphy and diagenetic constraints on Neoproterozoic Rapalini, A. E. & Sanchez Bettucci, L. 2008. Widespread remagneti-
carbonate successions from the Sierras Bayas Group, Tandilia zation of Late Proterozoic sedimentary units of Uruguay and the
System, Argentina. Chemical Geology, 237, 109–128. apparent polar wander path for the Rio de la Plata craton. Geophysical
Hartmann, L. A., Santos, J. O. S., Cingolani, C. A. & MacNaug- Journal International, 174, 55 – 74.
thon, N. J. 2002. Two Paleoproterozoic orogenies in the evolution Rapela, C., Dalla Salda, L. & Cingolani, C. 1974. Un intrusivo básico
of the Tandilia Belt, Buenos Aires, as evidenced by zircon U –Pb Ordovı́cico en la Formación La Tinta (Sierra de los Barrientos),
SHRIMP geochronology. Int. Geology Rev., 44, 528– 543. Provincia de Buenos Aires. Revista de la Asociación Geológica
Iñiguez Rodrı́guez, A. M., Del Valle, A., Poiré, D., Spalletti, L. A. Argentina, 29, 319–331.
& Zalba, P. 1989. Cuenca Precámbrica/Palaeozoica inferior de Rapela, C. W., Pankhurst, R. J. et al. 2007. The Rio de la Plata craton
Tandilia, Provincia de Buenos Aires. In: Chebli, W. & Spalletti, and the assembly of SW Gondwana. Earth Science Reviews, 83,
L. A. (eds) Cuencas Sedimentarias Argentinas. Serie Correlación 49– 82.
Geológica, San Miguel del Tucumán, 6, 245– 263. Sánchez Bettucci, L. & Rapalini, A. E. 2002. Paleomagnetism of the
Kawashita, K., Varela, R. et al. 1999. Geocronology and chemostrati- Sierra de Las Animas Complex, Southern Uruguay: Its Implications
graphy of ‘La Tinta’ Neoproterozoic sedimentary rocks, Buenos in the Assembly of Western Gondwana. Precambrian Research,
Aires Province, Argentina. Actas II. South American Symposium on 118, 243–265.
Isotope, 403–407. Spalletti, L. & Del Valle, A. 1984. Las diamictitas del sector oriental
Marchese, H. G. & di Paola, E. C. 1975. Miogeosinclinal Tandil. de Tandilia: caracteres sedimentológicos y origen. Revista de la
Revista de la Asociación Geológica Argentina, 30, 161–179. Asociación Geológica Argentina, 39, 188– 206.
Massabie, A. C. & Nestiero, O. E. 2005. La estructura del Grupo Sierras Teruggi, M. E. & Kilmurray, J. 1975. Tandilia. In: Geologı́a de la
Bayas en el sector norte de las Sierras Septentrionales de Buenos Provincia de Buenos Aires. Relatorio del VI Congreso Geológico
Aires. Revista de la Asociación Geológica Argentina, 60, 135–146 Argentino, Buenos Aires, 55 –77.
Oyhantçabal, P., Siegesmund, A., Wemmer, K., Frei, R. & Layere- Tohver, E., D’Agrella Filho, M. S. & Trindade, R. I. F. 2006. Paleo-
tal, P. 2007. Post-collisional transition from calc-alkaline to alkaline magnetic record of Africa and South America for the 1200–500 Ma
magmatism during transcurrent deformation in the southernmost interval, and evaluation of Rodinia and Gondwana assemblies.
Dom Feliciano Belt (Braziliano –Pan-African, Uruguay). Lithos, 98, Precambrian Research, 147, 193– 222.
141– 159. Trindade, R. I. F., D’Agrella-Filho, M. S., Epof, I. & Brito Neves,
Pankhurst, R. J. & Rapela, C. W. (eds) 1998. The Proto-Andean B. B. 2006. Paleomagnetism of Early Cambrian Itabaiana mafic
Margin of Gondwana. Geological Society of London, Special dikes (NE Brazil) and the final assembly of Gondwana. Earth
Publication, 142. Planetary Science Letters, 244, 361– 377.
Pankhurst, R. J., Ramos, A. & Linares, E. 2003. Antiquity of the Rı́o de Van der Voo, R. 1990. The reliability of paleomagnetic data. Tectonophy-
la Plata craton in Tandilia, southern Buenos Aires Province, Argen- sics, 184, 1– 9.
tina. Journal of South American Earth Sciences, 16, 5 –13. van Staden, A., Zimmermann, U. & Germs, G. J. 2005. Provenance and
Pazos, P. J., Sánchez Bettucci, L. & Loureiro, J. 2008. The Neopro- depositional study on tillites from the Volcan Hill, Tandilia System in
terozoic glacial record in the Rı́o de la Plata Craton: a critical reap- east Argentina: preliminary results. XVI Congreso Geológico
praisal. In: Pankhurst, R. J., Trouw, R. A. J., de Brito Neves, Argentino, Actas, I, 239–246.
Chapter 55

Glacial sediments and associated strata of the Polarisbreen Group,


northeastern Svalbard

GALEN P. HALVERSON1,2
1
School of Earth and Environmental Sciences, The University of Adelaide North Terrace, Adelaide, SA 5005, Australia
2
Present address: Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montreal,
QC, H3A 2A7, Canada
(e-mail: galen.halverson@mcgill.ca)

Abstract: Northeastern Svalbard hosts exceptionally well-preserved Neoproterozoic sediments. The glaciogenic Petrovbreen Member
and Wilsonbreen Formation (Fm.) of the Polarisbreen Group crop out in a narrow, Caledonian-aged fold-and-thrust belt spanning from
Olav V Land on Spitsbergen in the south to western Nordaustlandet in the north. The older Petrovbreen Member is thin (0– 52 m) and
patchily preserved, comprising mainly poorly stratified, dolomite-matrix diamictite likely deposited in a marine setting. The basal contact
of the Petrovbreen Member erosionally truncates the upper Russøya Member, which preserves a large (12‰) negative C-isotope
anomaly. The Petrovbreen Member is overlain by 200 m of dark, monotonous shales of the MacDonaldryggen Member, followed by
cherty dolomite grainstone and microbiolamintes of the Slangen Member. The upper Slangen Member is an exposure surface in the
southern part of the belt, but in northern Spitsbergen and on Nordaustlandet is transitional into sands of the northward-thickening
Bråvika Member. The Wilsonbreen Fm. is typically 100– 150 m thick and consists mostly of massively to poorly stratified diamictite,
with subordinate sand beds, conglomerate lenses, and carbonates deposited in a terrestrial environment. It is overlain by the colourful
Dracoisen Fm., which records, at its base, a typical post-glacial negative d13C anomaly. There are no direct radiometric age constraints
or reliable palaeomagnetic data from the Polarisbreen Group, but it is widely accepted that northeastern Svalbard was contiguous with
East Greenland during the Neoproterozoic.

Late Neoproterozoic glaciogenic rocks occur in both western and sediments have been carried out (e.g. Chumakov 1968; Edwards
northeastern Svalbard (Harland et al. 1993). The much better 1976; Hambrey 1982; Fairchild 1983; Dowdeswell et al. 1985),
studied and preserved glaciogenic sediments and bracketing with the results culminating in a monograph by Harland et al.
strata of the upper Hecla Hoek succession in northeastern Svalbard (1993). Carbonates within and bounding the Polarisbreen diamic-
are the subject of this chapter. These rocks occur in Olav V Land tites have also been the focus of coupled sedimentological –
and Ny Friesland in northeastern Spitsbergen and in Gustav V geochemical investigations (Fairchild & Hambrey 1984; Fairchild
Land in northwestern Nordaustlandet (Fig. 55.1). These rocks & Spiro 1987; Halverson et al. 2004). More broadly, Neoprotero-
are best exposed on nunataks in Olav V Land and southern Ny zoic sediments in Svalbard and East Greenland were the subject
Friesland, but occur also over a large area in Nordaustlandet of the seminal chemostratigraphic study by Knoll et al. (1986)
along the eastern side of Hinlopenstretet, and to a lesser extent, that first firmly established the intimate association between nega-
in coastal exposures in northern Ny Friesland. In Spitsbergen, tive C-isotope anomalies and Precambrian glaciation. The upper
two distinct glacial units, named the Petrovbreen Member Hecla Hoek succession also hosts an unusually rich microfossil
(older) and Wilsonbreen Fm. (younger) occur within the middle record, which has been studied in detail by A. H. Knoll, N. J. Butter-
of the mixed clastic-carbonate Polarisbreen Group (Fig. 55.2). field and colleagues (e.g. Knoll 1982, 1984; Butterfield et al. 1988,
Both the Petrovbreen Member and the Wilsonbreen Fm., together 1994). The excellent preservation of these sediments and their role
with bracketing strata, are readily identified in Nordaustlandet in the ongoing debate over the cause and severity of Neoproterozoic
(Halverson et al. 2004), where a separate nomenclature has histori- glaciations overcompensate for the logistical difficulty involved in
cally been used (Kulling 1934). Because correlation across Hinlo- studying them. This contribution is intended to update the review
penstretet (Fig. 55.1) is unambiguous (Kulling 1934; Harland et al. by Hambrey et al. (1981) in the context of more recent research
1993; Fairchild & Hambrey 1995), only the better known nomen- on the Polarisbreen Group and significant advances in our under-
clature from Spitsbergen (Harland 1997) is used here (Fig. 55.2). standing of the Neoproterozoic Earth system since that time.
The history of geological investigation of Precambrian glacial
deposits was recently summarized in Harland (2007). In short,
the Neoproterozoic sedimentary succession in northeastern Sval- Structural framework
bard was first studied by Nordenskiöld (1863). A glacial origin
for the Polarisbreen diamictites (specifically the Wilsonbreen The Svalbard archipelago, situated on the northwestern corner of
Fm.) was originally proposed by Kulling (1934), who mapped the Barents Shelf, comprises three tectonic terranes that were
and named these rocks in Nordaustlandet and northern Ny Friesland amalgamated in the Silurian – Devonian Ny Friesland (Caledonian)
during the Swedish– Norwegian Arctic Expedition of 1931. orogeny (Harland & Gayer 1972; Harland et al. 1992; Gee & Page
Fleming & Edmonds (1941) further investigated coastal exposures 1994; Lyberis & Manby 1999). The thick Hecla Hoek succession
of the Hecla Hoek, including the Wilsonbreen Fm., but it was not occurs in the northern part of the Eastern Terrane, which is the
until W. B. Harland and C. B. Wilson worked systematically on only exposed part of the Barentsia microcontinent (Breivik et al.
the upper Hecla Hoek succession on both coastal and inland sec- 2002). The Polarisbreen Group is part of the relatively undeformed
tions on Spitsbergen (Harland & Wilson 1956; Wilson 1961; upper Hecla Hoek succession, which includes the Lomfjorden and
Wilson & Harland 1964) that specific attention was focused on Hinlopenstretet supergroups (Fig. 55.2). In Ny Friesland, the basal
the Neoproterozoic glacial rocks of Svalbard and their significance Lomfjorden Supergroup is juxtaposed against high-grade meta-
in Earth history (Harland 2007). Since this time, and with their sediments of the Stubendorfbreen Supergroup along the Eolus-
global importance established (Harland 1964), several detailed sletta Shear Zone (Fig. 55.1; Lyberis & Manby 1999). The
sedimentological studies on the Polarisbreen glaciogenic nature of this contact and the relationship between the lower and

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 571– 579. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.55
572 G. P. HALVERSON

Fig. 55.1. Geological sketch map of the


region in northeastern Svalbard where the
Polarisbreen Group and equivalent rocks
crop out. The mapping is from Dallmann
et al. (2002), with modifications to the
Neoproterozoic strata based on mapping by
the author. Letters a to e refer to principal
regions in which the Polarisbreen Group has
been studied: a, Akademikerbreen area;
b, Ditlovtoppen; c, Dracoisen/Oslobreen;
d, Sveanor (south coast of Murchisonfjord),
e, Aldousbreen (north coast of
Wahlenbergfjord). ESZ, Eolusletta Shear
Zone (cf. Lyberis & Manby 1999); LFZ,
Lomfjorden Fault Zone.

upper Hecla Hoek has been the subject of a long-standing debate, Witt-Nilsson et al. 1998), an argument that is supported by
with one side arguing that the two packages form a single, con- mapping and geochronology in Nordaustlandet where the Lomf-
formable stratigraphic sequence (e.g. Harland 1959; Harland jord Supergroup rests unconformably on early Neoproterozoic vol-
et al. 1992). Most recent workers, however, agree that this canics and subvolcanic intrusives (Fig. 55.2), which themselves
contact represents a significant hiatus associated with Grenvillian intrude and overlie high-grade, late Mesoproterozoic metasedi-
orogenesis (e.g. Gee et al. 1995; Manby & Lyberis 1995; ments (Gee et al. 1995; Johansson et al. 2005). The broad,
THE POLARISBREEN GROUP, NORTHEASTERN SVALBARD 573

is sharp (Halverson et al. 2004). The top of the Polarisbreen Group


is bound by the Cambrian Tokammane Fm. (base of the Oslobreen
Group; Fig. 55.2). Although no erosional break at this contact is
evident from mapping or outcrop exposures, a large depositional
hiatus is implied by the biostratigraphy (Knoll & Swett 1987).
The Polarisbreen Group (Figs 55.2 & 55.3) is subdivided into
three formations (Wilson & Harland 1964): Elbobreen, Wilsonb-
reen and Dracoisen. The Elbobreen Fm. is further subdivided
into four regionally mappable members, each of which could be
accorded formation status in its own right: (from bottom to top)
the E1 or Russøya Member, the E2 or Petrovbreen Member, the
E3 or MacDonaldryggen Member, and the E4 or Slangen
Member (Fig. 55.2). A fifth unit that only occurs in the north of
Fig. 55.2. Stratigraphic nomenclature and correlation of the Neoproterozoic the outcrop belt, the Bråvika member, was tentatively assigned
successions in northeastern Spitsbergen and northwestern Nordaustlandet. Note to the Wilsonbreen Fm. (Halverson et al. 2004), but is clearly tran-
that only the nomenclature from Spitsbergen is used in this chapter. Hinlopen. sitional below with Slangen Member, and thus may more appropri-
Spgp., Hinlopenstretet Supergroup; Lom. S., Lomfjorden Supergroup; St. S., ately belong to the Elbobreen Fm. The Petrovbreen Member and
Stubendorfbreen Supergroup. The ‘upper Hecla Hoek succession’, referred to in Wilsonbreen Fm. comprise the two glaciogenic units, both of
the text and many publications, is the Lomfjorden plus Hinlopenstretet which are distinct from one another sedimentologically and
supergroups.
within the context of their bounding strata (Figs 55.3 & 55.4).

conformable and aggradational nature of the upper Hecla Hoek


and its continuity with the equivalent units in East Greenland Glaciogenic deposits and associated strata
(Fairchild & Hambrey 1995) imply that the glacial sediments of
the Polarisbreen Group were deposited in a thermally subsiding Russøya Member (E1)
basin (Maloof et al. 2006).
The upper Hecla Hoek crops out in a 120-km-long, north – south The Russøya Member varies from 75 to 170 m in thickness,
belt (Fig. 55.1) that was deformed during the Caledonian orogeny. thinning from south to north (Halverson et al. 2004). It is a litho-
The Polarisbreen Group sits in the core of the north-plunging Hin- logically variable unit, being predominantly carbonate in Nordaus-
lopenstrettet Synclinorium, which spans the length of the belt tlandet and containing an increasing fraction of sandstone and
(Fig. 55.1). The eastern limb of the synclinorium, exposed in siltstone to the south. Despite its heterogeneity, the stratigraphic
Nordaustlandet, is a fold-and-thrust belt, dissected by a network geometry of the Russøya Member is consistent, comprising a rela-
of conjugate NW –SE and SW –NE strike –slip faults (Sokolov tively thick (40– 160 m), basal, shoaling-upward sequence, over-
et al. 1968). The western limb of the synclinorium is less lain by up to four complete parasequences, 5–20 m thick
deformed, but is similarly cut by Caledonian-aged reverse and (Fig. 55.3). The lowest parasequence consists of a basal dark shale
strike –slip faults. (Lyberis & Manby 1999). In the central part that grades upward into increasingly carbonate-rich sediments,
of the belt, the upper Hecla Hoek is intruded by Silurian granitoids including limestone ribbonites and grainstones. Molar tooth struc-
(Teben’kov et al. 1996). The upper Hecla Hoek was eroded and tures are prominent in this sequence, but occur exclusively within
draped by Carboniferous sediments during post-orogenic, west– the ribbonite facies (Fig. 55.3).
east extension (Harland 1959), but normal faulting was largely The upper parasequences typically consist of shale at the base
concentrated along the prominent Lomfjorden Fault Zone (LFZ; and sandstone or dolomitic grainstone, stromatolites or microbial
Fig. 55.1). Dolerites related to the Late Jurassic – Early Cretaceous laminites at the top. The uppermost parasequence is capped by a
High Arctic Large Igneous Province (Maher 2001) intrude the 2– 6-m-thick biostrome of a spiralling, columnar stromatolite
Hecla Hoek within the northern part of the belt in the Hinlopenstre- (Kussiella) and brecciated and variably truncated on an outcrop
tet region (Fig. 55.1). The Late Cretaceous –Early Tertiary West scale (Halverson et al. 2004). Although erosional truncation
Spitsbergen Orogen, which heavily deformed the Western below the Petrovbreen Members is ,2 m on the outcrop scale,
Terrane (Harland 1959), only lightly affected the Eastern regional correlations, assisted by chemostratigraphy (see below),
Terrane, but did result in reactivation of major lineaments, such suggest as much as 50 m of erosion beneath the Petrovbreen
as the LFZ (Larsen 1987; Lyberis & Manby 1993). Member (Halverson et al. 2004).

Stratigraphy Petrovbreen Member (E2)

The Polarisbreen Group (formerly Series) was originally defined The Petrovbreen Member is on average only about 10 m thick
by Wilson & Harland (1964) in the Dracoisen area in southern (Fig. 55.4), but is highly variable (,52 m), its thickness broadly
Ny Friesland, near the confluence of the Polarisbreen and Chyde- reciprocating the depth of erosion on the sub-glacial sequence
niusbreen glaciers (Fig. 55.2). It conformably overlies the thick boundary (Halverson et al. 2004). In Nordaustlandet, the Petrovb-
stack (2 km) of Akademikerbreen Group carbonates and broadly reen Member is typically very thin and had been recognized only
marks a transition from predominantly carbonate to mixed in northernmost sections (Flood et al. 1969; Hambrey 1982;
carbonate-clastic sedimentation. The contact between these two Harland et al. 1993), but recent mapping has shown that it can be
groups is easily identified throughout the outcrop belt, with the identified in most sections (Halverson et al. 2004). The yellow- to
lithologically distinctive, pale yellow ‘Dartboard Dolomite’ orange-weathering Petrovbreen Member is predominantly com-
(Wilson 1961; Wilson & Harland 1964) of the uppermost Akade- posed of poorly stratified and massive diamictite and wackestone
mikerbreen Group contrasting with the laminated black shale and with a yellow- to orange-weathering dolomite matrix. Other
bluish-grey to black limestone of the lower Polarisbreen Group common facies include dolomitic rhythmites, tabular clast conglom-
(Halverson et al. 2004). The nature of this contact is variable. In erates and sandy dolostone (Fairchild & Hambrey 1984). Climbing
some places, it is unambiguously transitional. Elsewhere, the ripples occur near the top of the member in some sections. The
upper Dartboard Dolomite is brecciated and silicified, and contains clasts in the diamictite range up to 90 cm in diameter, are sub-
tepee-like structures, and the contact with the overlying limestone angular to angular, and consist mainly of light grey dolomite
574 G. P. HALVERSON

Fig. 55.3. Composite stratigraphic column


through the Polarisbreen Group (after
Halverson et al. 2004) and associated
chemostratigraphic and biostratigraphic
data. Sr-isotope data are from Halverson
et al. (2007), C-isotope data (v. VPDB) on
organic matter (d13Corg) from Knoll et al.
(1986) and Kaufman et al. (1997) (as
compiled in the latter), and C-isotope data
on carbonates (d13Ccarb) from Fairchild &
Spiro (1987), Halverson et al. (2004) and
Kaufman et al. (1997). The two principal
acritarch assemblages in the Polarisbreen
Group (right side of figure) are summarized
from Knoll (1982) and Knoll & Swett
(1987).

(grainstone and stromatolite) and black chert. Minor sandstone, silt- rosette-shaped and have canted and square prismatic habits (Halver-
stone and rare volcanic clasts are also found (Hambrey 1982; son et al. 2004).
Harland et al. 1993). The upper MacDonaldryggen shales and siltstones are every-
Evidence for direct glacial influence in the Petrovbreen Member where transitional into flaggy dolomites that mark the bottom of
comes from rare striated clasts and the abundance of dropstones in the Slangen Member. In southern sections, this transition includes
finely laminated sediments (Hambrey 1982; Harland et al. 1993), an increase in carbonate, including limestone ribbons. Mud-cracks
as well as evidence of glacial rock flour in the form of sub- have been reported from this level (Fairchild & Hambrey 1984),
micrometre-sized dolomite in the matrix of the diamictite but appear to be rare.
(Fairchild 1983). Fe and 18O-enrichment of this dolomitic matrix
relative to clasts suggests early marine alteration (Fairchild et al.
1989). The absence of features indicative of a terrestrial environ- Slangen Member (E4)
ment, coupled with rapid facies changes and the occurrence of
dropstones, suggest glaciomarine (or glaciolacustrine) deposition The Slangen Member is typically a 20 –30 m regressive parase-
near an ice-grounding line (Hambrey 1982; Fairchild & quence of cherty dolostone. The dominant facies in the Slangen
Hambrey 1984; Harland et al. 1993). Member is a vuggy, trough and tabular cross-bedded grainstone;
however, fenaestral microbial laminates also occur as thin inter-
beds within the grainstone and as distinct beds up to 5 m thick in
MacDonaldryggen Member (E3) the upper Slangen Member. Ribbons occur mainly in the transition
with the underlying MacDonaldryggen Member, but are also
The contact between the upper Petrovbreen Member and the base of important in the uppermost Slangen Member in Nordauslandet.
the MacDonaldryggen Member is everywhere sharp but conform- The occurrence of length-slow chalcedony, high Na concen-
able. In Olav V Land, where the MacDonaldryggen Member was trations, and rare anhydrite in the Slangen Member suggest depo-
named (Harland et al. 1993), the base consists of a thin (10–40 sition in a restricted environment such as a coastal sabkha
cm), finely laminated carbonate (variably limestone and dolomite) (Fairchild & Hambrey 1984).
with wispy organic-rich laminae (Halverson et al. 2004). Else- The upper contact of the Slangen Member is variable. In most
where, the upper Petrovbreen Member is directly overlain by the Spitsbergen sections, it is sharp, silicified and brecciated, with
same finely laminated, olive green to dark grey mudstone that com- clasts showing no evidence for reworking (Fairchild & Hambrey
prises the bulk of the c. 200-m-thick MacDonaldryggen Member. 1984). At Ditlovtoppen (Fig. 55.1) the grainstone passes transition-
The fraction of silt increases up-section, with thin, fine sandstone ally upward into 2.5 m of mud-cracked siltstone beneath the basal
beds present in some sections. In Nordaustlandet, authigenic diamictites of the Wilsonbreen Fm. In Nordaustlandet, the grain-
calcite nodules up to 2 cm in diameter, and in places densely clus- stones and microbial laminites are commonly overlain by silty
tered within distinct layers, speckle the upper MacDonaldryggen dolomite ribbons, which grade upwards through ripple cross-
Member some 20–40 m beneath the contact with the overlying laminated dolomitic siltstone and fine sandstone and into dolomitic
Slangen Member (Fig. 55.3). These nodules are commonly sandstone of the basal Bråvika Member (Fig. 55.4).
THE POLARISBREEN GROUP, NORTHEASTERN SVALBARD 575

Fig. 55.4. Characteristic stratigraphic


sections of the Wilsonbreen Fm. and
Petrovbreen Member glacials from each of
principal regions in which the glacial
intervals have been studied (locations a to e
are linked to Fig. 55.1; the Petrovbreen
Member does not occur at location e).
Stratigraphic depth is given relative to a 0 m
datum at the top of each glacial unit. Note
the difference in scale between the
Wilsonbreen Fm. and Petrovbreen Member.
The measured sections are compiled from
published columns in Edwards (1976),
Harland et al. (1993), Halverson et al.
(2004) and unpublished data of the author.
The breakdown and naming of members
here follows that of Harland et al. (1993),
but is slightly different from that used in
previous papers, where the names of the
Ormen and Gropbreen Members are
inverted and a fourth Wilsonbreen member
is included (e.g. Hambrey 1982).

The Wilsonbreen Fm. abundant at Aldousbreen (Fig. 55.4e; Edwards 1976). Sandstone
wedges occur near or at the top of the Ormen Member in the Aka-
In most sections, the Wilsonbreen Fm. consists of c. 100–150 m of demikerbreen area (Fig. 55.4a) and at Ditlovtoppen (Fig. 55.4b)
green- to red-weathering, massively to poorly bedded diamictite; (Hambrey 1982; Fairchild & Hambrey 1984).
however, the diamictite thins dramatically northwards in Nordaus- The Middle Carbonate Member is thin (,25 m) and comprises
tlandet, being reciprocated by a northward-thickening wedge of mainly dolomitic, medium-grained quartz sandstone. It is,
quartz sands, informally named the Bråvika member (Halverson however, recognized and distinguished by the presence of thin
et al. 2004; Fig. 55.4). In the northernmost exposure of the Wil- (,1.5 m) beds, lenses and fragments of primary carbonate. The
sonbreen Fm. (at Langrunesset), the Bråvika member is 250 m nature of the carbonate is variable, being corrugated silty dolomitic
thick and diamictite is absent. microbial laminites at Ormen (in the Akademikerbreen area),
In Spitsbergen, the Wilsonbreen Fm. is subdivided into the mixed limestone and dolomite rhythmites and ribbons at Ditlov-
Ormen (W1), Middle Carbonate (W2) and Gropbreen (W3) toppen (Fig. 55.4b), and isolated limestone stromatolites within
members (Harland et al. 1993; Fig. 55.4). The Ormen Member sandstone at Dracoisen (Fig. 55.4c). Based on their geochemistry,
varies from 20 to 75 m in thickness and consists mainly of these carbonates are interpreted to have formed under the influence
poorly stratified, shaley diamictite, but also includes lenses and of evaporation in periglacial lakes (Fairchild et al. 1989).
interbeds of sandstone, conglomerate and breccia (Fig. 55.4; Fair- The Gropbreen Member is lithologically similar to the Ormen
child & Hambrey 1984). Clast lithology is variable, but dominated Member, with dominantly weakly stratified to massive maroon-
by carbonates derived from the underlying Elbobreen Fm. and weathering diamictite, and subordinate sandstone lenses and
Akademikerbreen Group. Chert, silt, sandstone and basement beds. Clast composition is also similar to the Ormen Member,
clasts, including exotic fragments of gneiss and undeformed pink but with a higher proportion of basement clasts. In many sections,
granite, are also found. Striated clasts occur throughout the unit a medium brown sandstone up to a few metres thick occurs at the
and dropstones, although not common in most sections, are very top of the Gropbreen Member and is continuous with well
576 G. P. HALVERSON

developed sandstone wedges that penetrate up to 2 m into the et al. (2004, 2005). Separate studies have focused on the isotopic
underlying diamictite. Whether the sand is present or not, the composition of dolomites and limestones within the Wilsonbreen
contact between the Wilsonbreen Fm. and the overlying Dracoisen Fm. as a means of reconstructing the palaeoenvironment of
Fm. is always sharp. syn-glacial carbonate precipitation (Fairchid & Spiro 1987,
In Nordaustlandet, the Wilsonbreen is subdivided into two 1990; Fairchild et al. 1989).
members. The Bråvika member overall is a coarsening-upwards A compilation of C-isotope data from these various publications
package of quartz sandstone, with irregular dolomite lenses and is plotted alongside a composite stratigraphic column in
intraclasts, and in the northernmost sections, minor limestone Figure 55.3. The Polarisbreen Group d13Ccarb and d13Corg records
rhythmites and ribbons. The northward-thickening sandstone con- broadly mirror one another (Fig. 55.3), with an average isotopic
tains both planar and trough cross-bedding throughout, indicating a difference of c. 30‰, which is typical for the Neoproterozoic
palaeoflow direction to the north (Halverson et al. 2004). The (Hayes et al. 1999). The most striking feature of the C-isotope
prevalence of low-angle cross-sets, coupled with pinstripe lami- record is a 12‰ negative anomaly just below the Petrovbreen
nations and a bimodal grain population, suggests an aeolian Member. This anomaly is reproduced in multiple sections and
origin, although the trough cross-bedding and dolomite precipi- essentially spans the upper two parasequences of the Russøya
tation indicate that the sand was in places reworked subaqueously. Member (Halverson et al. 2004). d13C values are still negative in
In the southern exposures in Nordaustlandet, the Bråvika sand- carbonates within the lowermost MacDonaldyrggen Fm., but
stones are transitional with overlying diamictites, but in the quickly rise to positive values, where they remain into the Wil-
Sveanor (Fig. 55.4d) are also transitional below with the Slangen sonbreen Fm. Intraglacial carbonates show a range in d18O and
Member (Halverson et al. 2004). d13C values between –11 and þ11‰ and þ1 and 5‰, respectively
The diamictite-dominated unit overlying the Bråvika member in (Fairchild & Spiro 1987; Halverson et al. 2004). Fairchild & Spiro
Nordaustlandet resembles the Ormen and Gropbreen members on (1987, 1990) argued that the Wilsonbreen carbonates were precipi-
Spitsbergen, only containing a higher percentage of undeformed tated by evaporation in a periglacial lake based, in part, on these
igneous clasts. Rhythmites with dropstones and fine interbedded unusually high d18O compositions. More recently, Bao et al.
turbidites are well preserved in the southernmost Nordaustlandet (2009) reported D17O (triple oxygen isotopes) compositions in
section at Aldousbreen (Fig. 55.4e), in Wahlenbergfjorden carbonate-associated sulphate (CAS) within these carbonates of
(Edwards 1976). Unstratified siltstones, interpreted as loessites as low as –1.6‰, which are the lowest values ever reported
(Edwards 1976), also occur in the same section (Fig. 55.4). from natural specimens. Bao et al. (2009) interpreted these anom-
alous values to record either extraordinarily high pCO2 or unu-
The Dracoisen Fm. sually sluggish O2 cycling during the glaciation.
A second negative C-isotope anomaly is recorded in the basal
The colourful lower part of the Draocoisen Fm. is a superb marker Dracoisen dolostone and reproduced in multiple sections, where
interval (Wilson & Harland 1964). A yellow-weathering dolomite d13C values gradually decline from –3 to – 5‰ (Halverson et al.
bed everywhere overlies the Wilsonbreen Fm., with no evidence 2004). d13C values then spike to .10‰ in evaporitic dolomites
for reworking or depositional hiatus. This dolostone varies in in the upper Dracoisen Fm. (Fig. 55.3).
thickness from as much as 18 m at Ditlovtoppen to less than 3 m Four Sr-isotope data from the Polarisbreen Group were first
in parts of Nordaustlandet (Halverson et al. 2004) and always com- published by Derry et al. (1989), and a few new analyses from
prises the transgressive component of a thick (c. 170 m) sequence the original sample set were added in Jacobsen & Kaufman
(Fig. 55.3). The dolostone is mostly finely laminated dololutite, (1999). Additional 87Sr/86Sr data from the Polarisbreen were pub-
with minor low-angle erosional truncations and cross-laminations. lished in Halverson et al. (2005, 2007). The highest-quality data
Inversely graded laminae and pockets of peloids are also common. from all these studies consistently show 87Sr/86Sr ¼ 0.7067–
Large-scale oscillation ripples with amplitudes of up to 40 cm and 0.7068 in the Russøya Member (Fig. 55.3). The absence of unal-
wavelengths up to 6 m (Allen & Hoffman 2005) occur in the upper tered marine limestone above the Russøya Member precludes
part of the dolostone in most northern sections. application of Sr-isotope chemostratigraphy to the remainder of
The top of the dolostone is transitional into marly red siltstone, the Polarisbreen Group.
which then passes upwards into green then black shales. The
change to black shale, at c. 30–40 m above the base of the Dracoi-
sen Fm., marks the maximum flooding surface of the sequence. Palaeolatitude and palaeogeography
Black shales with common carbonate concretions persist for over
100 m and are transitional into mud-cracked, variegated, finely Barentsia is widely assumed to have been contiguous with East
bedded siltstone and shale that constitutes much of the upper Greenland during the Neoproterozoic based on the remarkable
part of the formation. The interval of mud-cracked siltstone is stratigraphic similarity between the two regions (Harland &
interrupted by 9 m of microbial-laminated dolomite, internally Gayer 1972; Knoll et al. 1986; Fairchild & Hambrey 1995).
deformed by cauliflower chert. However, the precise location of Eastern Svalbard relative to the
East Greenland Caledonides (Johansson et al. 2005) and the geo-
metry and origin of the East Greenland –eastern Svalbard platform
Chemostratigraphy within the broader framework of the connection and eventual
rifting between Laurentia, Baltica and Amazonia are controversial.
In one of the earliest systematic chemostratigraphic studies of a In the conventional configuration, western Baltica is the conjugate
Neoproterozoic succession, Knoll et al. (1986) produced a margin to eastern Greenland (and Barentsia), virtually identical to
coupled d13Ccarb – d13Corg record through the Veteranen, Akademi- the Caledonian fit (e.g. Weil et al. 1998). In a different reconstruc-
kerbreen and Polarisbreen groups in both Spitsbergen and Nor- tion, Hartz & Torsvik (2002) placed the East Greenland –eastern
daustlandet (as well as East Greenland). Although by modern Svalbard platform adjacent to the southern peri-Urals of Baltica
standards the resolution of this record is low, with only a and proposed that it developed as a sinistral pull-apart basin
handful of data points from the Polarisbreen Group, this paper between Baltica and Amazonia. This model has been challenged
established the concurrence of glaciations and negative by more recent palaeomagnetic data from Baltica (Cawood &
C-isotope anomalies, as well as the prevalence of high d13C Pisarevsky 2006), and in a modification of earlier west-Baltica –
values during the much of the Neoproterozoic. Subsequent eastern Greenland fits, Pisarevsky et al. (2008) proposed that
C-isotope stratigraphy on the Polarisbreen Group was published the west Norwegian margin was conjugate to southeastern
by Fairchild & Spiro (1987), Kaufman et al. (1997) and Halverson Greenland. This model implies that the eastern Svalbard and the
THE POLARISBREEN GROUP, NORTHEASTERN SVALBARD 577

Neoproteroozic successions in northern Norway were part of a 1982; Fairchild & Hambrey 1984; Dowdeswell et al. 1985) effec-
long passive margin isolated from the other coeval sedimentary tively put to rest any doubt. Evidence for a glacial environment,
basins to the south. specifically in the Wilsonbreen Fm., is found in the abundance
Direct palaeomagnetic constraints for Svalbard’s Neoproterozoic of striated and faceted clasts, dropstones, the extrabasinal origin
palaeography are limited to a new suite of palaeomagnetic data of some stones, and sandstone wedges.
from the Akademikerbreen Group that show a stable, pre- The Petrovbreen Member diamictite and associated lithofacies
Caledonian remanent magnetization indicating that the EGES are typically interpreted to have been deposited in relatively
platform resided in tropical latitudes (I  158) during the mid- deep water, most likely in a marine environment (Fairchild &
Neoproterozoic (Maloof et al. 2006). No reliable palaeomagnetic Hambrey 1984; Harland et al. 1993). This interpretation is
results have been obtained from the Polarisbreen Group, and Sval- consistent with the lack of evidence for any obvious change in
bard’s complicated and uncertain tectonic history since this time relative sea level at the contact between the Petrovbreen and
precludes any extrapolation of palaeolatitude for the glacial depos- MacDonaldryggen Members (Halverson et al. 2004). In contrast,
its from elsewhere on the Laurentia craton (Evans 2000). the heterogeneous Wilsonbreen Fm., which includes diamictites
interpreted as lodgement tills (Dowdeswell et al. 1985), direct
evidence for subaerial exposure (sandstone wedges), and lacus-
Geochronological constraints trine carbonates is inferred to have been deposited in a pre-
dominantly terrestrial setting (Fairchild & Hambrey 1984;
No direct radiometric ages have been obtained on the upper Hecla Dowdeswell et al. 1985; Halverson et al. 2004). Dowdeswell
Hoek. U –Pb ages on detrital zircons in the Planetfjella Group in et al. (1985) argued that the Wilsonbreen diamictites were depos-
Spitsbergen and presumed primary zircons from volcanics and ited beneath ice sheets or ice caps (rather than valley glaciers)
subvolcanic intrusives beneath the Veteranen Group in Nordaus- based on the preponderance of exotic clasts and a lack of
tlandet indicate a maximum age for the base of the Lomfjorden angular supraglacial debris.
Supergroup of c. 946 Ma (Johansson et al. 2000, 2005; A great deal of effort has been expended on attempting to corre-
Fig. 55.2). Similar ages have been obtained from igneous clasts late the Polarisbreen diamictites with other Neoproterozoic glacial
in the Wilsonbreen Fm. in Nordaustlandet (Johansson et al. 2000). units in the North Atlantic (e.g. Harland & Gayer 1972; Hambrey
The top of the Polarisbreen Group, which is in disconformable 1983; Nystuen 1985; Fairchild & Hambrey 1995) and globally
contact with the overlying Cambrian Tokamanne Fm. (Knoll & (e.g. Kaufman et al. 1997; Kennedy et al. 1998; Halverson et al.
Swett 1987), is inferred to be ,542 Ma (the age of the 2005). Based on biostratigraphic and chemostratigraphic data, it
Precambrian –Cambrian boundary; Amthor et al. 2003). If the has generally been argued that the Polarisbreen Group is late
Dracoisen Fm. is correlative with the Maieberg Fm. in northern ‘Vendian’ in age, meaning essentially that it post-dates the earliest
Namibia and the Doushantuo Fm. in South China (Hoffmann Neoproterozoic glaciations. Unfortunately, the initial optimism that
et al. 2004; Condon et al. 2005), as implied by the chemostrati- the chronostratigraphy and correlations would be firmed up by geo-
graphy and biostratigraphy of the Polarisbreen Group, then the chronological and palaeomagnetic data (Hambrey 1983) has not
Wilsonbreen glaciation ended at 635 Ma (Halverson et al. 2005). been realized, and no undisputed correlation scheme exists for the
North Atlantic region, let alone the whole of the globe.
Biostratigraphy In a twist on the conventional Vendian interpretation of the age of
the Polarisbreen Group diamictites, Halverson et al. (2004) argued
In contrast to the rich and diverse fossil assemblage from the that both the Petrovbreen Member and Wilsonbreen Fm. belonged
underlying Akademikerbreen Group (Knoll 1982; Butterfield to a single glacial episode, ending at 635 Ma (the approximate age
et al. 1994, and references therein), the biostratigraphic record of of the end of the Ghaub glaciation in Namibia and Nantuo glaciation
the Polarisbreen Group is most notable for its depauperate micro- in South China; Hoffmann et al. 2004; Condon et al. 2005). This
fossil assemblage, despite the generally mild thermal history of the model was based on (i) the similar stratigraphic context and magni-
rocks (Knoll & Swett 1987). In the MacDonaldryggen Member tude of the upper Russøya negative d13C anomaly (Fig. 55.3) and
and in shales within the Wilsonbreen Fm., the acritarch Bavlinella the so-called Trezona anomaly, which precedes the Ghaub glacia-
faveolata dominates the microfossil assemblage (Knoll 1982; tion in Namibia (Halverson et al. 2002), coupled with lack of evi-
Knoll & Swett 1987). B. faveolata also occurs in the Dracoisen dence for an analogous anomaly preceding any of the older
Fm. and is locally abundant, but overall is subordinate to a low glaciations; (ii) the occurrence of glendonites in the upper MacDo-
diversity assemblage of small, smooth-walled leiosphere acri- naldryggen Member; and (iii) the virtually identical geochemistry,
tarchs (Fig. 55.3). The diverse assemblage of the large, acantho- sedimentology and stratigraphy of the Dracoisen and post-Ghaub
morphic acritarchs that characterize middle– upper Ediacaran Maieberg (and other) cap-carbonate sequences. In light of new
successions in Australia (Grey et al. 2003) and elsewhere (Knoll Sr-isotope data from the level of the negative anomaly in the
2000) is absent in the Polarisbreen Group (Knoll & Swett 1987). Russøya Member (87Sr/86Sr ¼ 0.7068 v. 0.7072 in the Trezona
Coupled with the lack of Ediacaran fauna, this lacuna strongly anomaly; Halverson et al. 2007) and evidence that the lowermost
suggests that the contact between the Polarisbreen Group and the of three distinct glacial units in the Dalradian Supergroup is
overlying Oslobreen Group represents a significant depositional preceded by a negative C-isotope anomaly of similar magnitude
hiatus (Knoll & Swett 1987) and that at least the last 35 myr of to that in the upper Russøya Member (McCay et al. 2006), it now
the Ediacaran Period are missing in Svalbard. seems more likely that the Petrovbreen Member diamictite must
represent a separate, older glaciation. Although it remains likely
that the Wilsonbreen Fm. is c. 635 Ma, the ongoing ambiguity in
Discussion Cryogenian–Ediacaran age constraints precludes any unequivocal
assignment of age to either of the Polarisbreen glaciations.
The Neoproterozoic succession in Svalbard has played a promi- The more conventional two-glaciation interpretation also raises
nent role in questions and controversies regarding Precambrian the question of the climatic significance of the glendonite nodules
glaciations and environmental change since Harland (1964) first in the upper MacDonaldryggen Member. Glendonites have also
proposed a pan-global infra-Cambrian glaciation. Although a been reported in possibly equivalent-aged strata in the Windermere
few researchers maintained scepticism over the glacial origin of Supergroup in the Mackenzie Mountains (James et al. 2005).
diamictites in the Polarisbreen Group (Krasil’shchov 1973; Scher- Insofar as the MacDonaldryggen pseudomorphs originated as
merhorn 1974), the detailed sedimentological work of Edwards, ikaite, which forms at near-freezing temperatures in alkaline-
Hambrey, Fairchild and others (e.g. Edwards 1976; Hambrey charged waters (Buchart et al. 1997), then their occurrence in
578 G. P. HALVERSON

continental shelf sediments with no other evidence of glaciation, Fairchild, I. J. & Hambrey, M. J. 1995. Vendian basin evolution in East
and just below sabkha deposits, poses yet another Neoproterozoic Greenland and NE Svalbard. Precambrian Research, 73, 217– 233.
climatic conundrum. Thus, it seems certain that the Polarisbreen Fairchild, I. J. & Spiro, B. 1987. Petrological and isotopic implications
Group will continue to play a central role in the ongoing effort of some contrasting Late Precambrian carbonates, NE Spitsbergen.
to understand Neoproterozoic climate. Sedimentology, 34, 973– 989.
Fairchild, I. J. & Spiro, B. 1990. Carbonate minerals in glacial sedi-
This chapter stems in part from the author’s PhD and subsequent research in ments: geochemical clues to palaeoenvironment. In: Dowdeswell,
Svalbard, funded by NSF and NASA. Original mapping, measured sections, J. A. & Scourse, J. D. (eds) Glacimarine Environments: Processes
and other contributions presented here are the results of collaborative fieldwork and Sediments. Geological Society, London, Special Publications,
with A. C. Maloof, P. F. Hoffman, E. W. Domack, M. T. Hurtgen, J. Eigenbrode, 53, 201– 216.
and many field assistants. M. Hambrey and I. Fairchild provided constructive Fairchild, I. J., Hambrey, M. J., Spiro, B. & Jefferson, T. H. 1989. Late
reviews of the manuscript. The contribution forms TRaX record #169. This rep- Proterozoic glacial carbonates in northeast Spitsbergen: new insights
resents a contribution of the IUGS and UNESCO-funded IGCP (International into the carbonate –tillite association. Geological Magazine, 126,
Geoscience Programme) project #512. 469– 490.
Fleming, W. L. S. & Edmonds, J. M. 1941. Hecla Hoek rocks of New
Friesland (Spitsbergen). Geological Magazine, 78, 405– 428.
Flood, B., Gee, D. G., Hjelle, A., Siggerud, T. & Winsnes, T. 1969.
References The geology of Nordaustlandet, northern and central parts. Norsk
Polarinstitutt Skrifter, 146, 1– 139 (þ1:250,000 map).
Allen, P. A. & Hoffman, P. F. 2005. Extreme winds and waves in the Gee, D. G. & Page, L. M. 1994. Caledonian terrane assembly on Svalbard:
aftermath of a Neoproterozoic glaciation. Nature, 433, 123– 127. new evidence from 40Ar/39Ar dating in Ny Friesland. American
Amthor, J. E., Grotzinger, J. P., Schröder, S., Bowring, S. A., Rame- Journal of Science, 294, 1166–1186.
zani, J., Martin, M. W. & Matter, A. 2003. Extinction of Cloudina Gee, D. G., Johansson, Å. et al. 1995. Grenvillian basement and a major
and Namacalathus at the Precambrian – Cambrian boundary in Oman. unconformity within the Caledonides of Nordaustlandet, Svalbard.
Geology, 31, 431–434. Precambrian Research, 70, 215– 234.
Bao, H., Fairchild, I. J., Wynn, P. M. & Spötl, C. 2009. Stretching the Grey, K., Walter, M. R. & Calver, C. R. 2003. Neoproterozoic biotic
envelope of past surface environments: Neoproterzoic glacial lakes diversification: Snowball Earth or aftermath of the Acraman
from Svalbard. Science, 323, 119–122. impact. Geology, 31, 459– 462.
Breivik, A. J., Mjelde, R., Grogan, P., Shimamura, H., Murai, Y., Halverson, G. P., Hoffman, P. F., Schrag, D. P. & Kaufman, A. J.
Nishimura, Y. & Kuwana, A. 2002. A possible Caledonide arm 2002. A major perturbation of the carbon cycle before the Ghaub
through the Barents Sea imaged by OBS data. Tectonophysics, 355, glaciation (Neoproterozoic) in Namibia: prelude to snowball
67 –97. Earth? Geochemistry, Geophysics, Geosystems, 3, doi: 10.1029/
Buchart, B., Seaman, P. et al. 1997. Submarine columns of ikaite tufa. 2001GC000244.
Nature, 390, 129–130. Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice,
Butterfield, N. J., Knoll, A. H. & Swett, K. 1988. Exceptional pres- A. H. 2005. Towards a Neoproterozoic composite carbon-isotope
ervation of fossils in an Upper Proterozoic shale. Nature, 334, record. Geological Society of America Bulletin, 117, 1181–1207.
424– 427. Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
Butterfield, N. J., Knoll, A. H. & Swett, K. 1994. Paleobiology of the glaciation (Neoproterozoic) in Svalbard. Basin Research, 16,
Neoproterozoic Svanbergfjellet Formation, Spitsbergen. Fossils and 297– 324.
Strata, 34, 1– 81. Halverson, G. P., Dudas, F. O., Maloof, A. C. & Bowring, S. A. 2007.
Cawood, P. A. & Pisarevsky, S. A. 2006. Was Baltica right-way-up or Evolution of the 87Sr/86Sr composition of Neoproterozoic
upside-down in the Neoproterozoic? Journal of the Geological seawater. Palaeogeography, Palaeoclimatology, Palaeoecology,
Society, London, 163, 753–759. 256, 103– 129.
Chumakov, N. 1968. On the character of the Late Precambrian glaciation Hambrey, M. J. 1982. Late Precambrian diamictites of northeastern Sval-
of Spitsbergen (translated title). Doklady An SSSR, Seriya Geologi- bard. Geological Magazine, 119, 527– 551.
cheskaya, 180, 1446– 1449. Hambrey, M. J. 1983. Correlation of late Proterozoic tillites in the North
Condon, D., Zhu, M., Bowring, S., Jin, Y., Wang, W. & Yang, A. 2005. Atlantic region and Europe. Geological Magazine, 120, 290– 320.
From the Marinoan glaciation to the oldest bilaterians: U– Pb ages Hambrey, M. J., Harland, W. B. & Waddams, P. 1981. Late Precam-
from the Doushantou Formation, China. Science, 308, 95– 98. brian tillites of Svalbard. In: Hambrey, M. J. & Harland, W. B.
Derry, L. A., Keto, L. S., Jacobsen, S. B., Knoll, A. H. & Swett, K. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge University
1989. Strontium isotopic variations in Upper Proterozoic carbonates Press, 592–600.
from Svalbard and East Greenland. Geochimica et Cosmochimica Harland, W. B. 1959. The Caledonian sequence in Ny Friesland, Spits-
Acta, 53, 2331– 2339. bergen. Quarterly Journal of the Geological Society, 114, 307– 342.
Dallmann, W. K., Ohta, Y., Elvevold, S. & Blomeier, D. 2002. Harland, W. B. 1964. Critical evidence for a great infra-Cambrian
Bedrock map of Svalbard and Jan Mayen (1:750,000 scale geological glaciation. Geologische Rundschau, 54, 45 –61.
map). Norsk Polarinstitutt Temekart, 33. Harland, W. B. 1997. The Geology of Svalbard. Geological Society of
Dowdeswell, J. A., Hambrey, M. J. & Wu, R. 1985. A Comparison of London Memoir, 17, London, 521.
clast fabric and shape in Late Precambrian and Modern glaciogenic Harland, W. B. 2007. The Ny Friesland Orogen, Spitsbergen. Geological
sediments. Journal of Sedimentary Petrology, 55, 691– 704. Magazine, 144, 633– 642.
Edwards, M. B. 1976. Sedimentology of Late Precambrian Sveanor Harland, W. B. & Gayer, R. A. 1972. The Arctic Caledonides and
and Kapp Sparre Formations at Aldousbreen, Wahlenbergfjorden, earlier oceans. Geological Magazine, 109, 289–314.
Nordaustlandet. Norsk Polarinstitutt Årbok, 1974, 51 –61. Harland, W. B. & Wilson, C. B. 1956. The Hecla Hoek succession in Ny
Edwards, M. B. 1978. Late Precambrian glacial loessites from north Friesland, Spitsbergen. Geological Magazine, 93, 265– 286.
Norway and Svalbard. Journal of Sedimentary Petrology, 49, 85 – 91. Harland, W. B., Scott, R. A., Auckland, K. A. & Snape, I. 1992.
Evans, D. A. D. 2000. Stratigraphic, geochronological and paleomagnetic The Ny Friesland Orogen, Spitsbergen. Geological Magazine, 129,
constraints upon the Neoproterozoic climatic paradoxes. American 679– 708.
Journal of Science, 300, 347– 443. Harland, W. B., Hambrey, M. J. & Waddams, P. 1993. Vendian
Fairchild, I. J. 1983. Effects of glacial transport and neomorphism on Geology of Svalbard. Norsk Polarinstitutt Skrifter (Oslo), 193, 150.
Precambrian dolomite crystal sizes. Nature, 304, 714– 716. Hartz, E. & Torsvik, T. 2002. Baltica upside down: a new plate tectonic
Fairchild, I. J. & Hambrey, M. J. 1984. The Vendian succession of model for Rodinia and the Iapetus Ocean. Geology, 30, 225– 258.
northeastern Spitsbergen: petrogenesis of a dolomite– tillite associ- Hayes, J. M., Strauss, H. & Kaufman, A. J. 1999. The abundance of C
ation. Precambrian Research, 26, 111–167. in marine organic carbon and isotopic fractionation in the global
THE POLARISBREEN GROUP, NORTHEASTERN SVALBARD 579

biogeochemical cycle of carbon during the past 800 Ma. Chemical Lyberis, N. & Manby, G. 1999. Continental collision and lateral escape
Geology, 161, 103– 125. deformation in the lower and upper crust: an example from Caledo-
Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L. nide Svalbard. Tectonics, 18, 40 – 63.
2004. A U– Pb zircon date from the Neoproterozoic Ghaub Maher, H. D. Jr. 2001. Manifestations of the Cretaceous High Arctic
Formation, Namibia: constraints on Marinoan glaciation. Geology, Large Igneous Province in Svalbard. Journal of Geology, 109,
32, 817–820. 91– 104.
Jacobsen, S. B. & Kaufman, A. J. 1999. The Sr, C and O isotopic evol- Maloof, A. C., Halverson, G. P., Kirschvink, J. L., Schrag, D. P.,
ution of Neoproterozoic seawater. Chemical Geology, 161, 37 –57. Weiss, B. P. & Hoffman, P. F. 2006. Combined paleomagnetic, iso-
James, N. P., Narbonne, G. M., Dalrymple, R. W. & Kyser, C. 2005. topic and stratigraphic evidence for true polar wander from the Neo-
Glendonites in Neoproterozoic low-latitude, interglacial sedimentary proterozoic Akademikerbreen Group, Svalbard, Norway. Geological
rocks, northwest Canada: insights on the Cryogenian ocean and Pre- Society of America Bulletin, 118, 1099– 2014.
cambrian cold-water carbonates. Geology, 33, 9 –12. Manby, G. & Lyberis, N. 1995. Discussion on the Ny Friesland Orogen,
Johansson, Å., Larianov, A. N., Tebenkov, A. M., Gee, D. G., White- Spitsbergen. Geological Magazine, 132, 351– 356.
house, M. J. & Vestin, J. 2000. Grenvillian magmatism of western McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial
and central Nordaustlandet, northeastern Svalbard. Transactions of trinity: Neoproterozoic Earth history within the British –Irish Caledo-
the Royal Society of Edinburgh, 90, 221– 234. nides. Geology, 34, 909– 912.
Johansson, A., Gee, D. G., Larionov, A. N., Ohta, Y. & Tebenkov, Nordenskiöld, A. E. 1863. Geografisk och geognostisk beskrifning over
A. M. 2005. Grenvillian and Caledonian evolution of eastern Sval- noröstra delarna af Spetbergen och Hinlopen Strait [Geographic and
bard – a tale of two orogenies. Terra Nova, 17, 317– 325. geognostic descriptions of the northeast part of Spitsbergen and
Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes, ice Hinlopen Straight]. Kungliga Svenska Vetenskapsakademiens Han-
ages and terminal Proterozoic earth history. Proceedings of the dlingar, 4. Stockholm.
National Academy of Science (USA), 94, 6600–6605. Nystuen, J. P. 1985. Facies and preservation of glaciogenic sequences
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K.-H. & from the Varanger ice age in Scandinavia and other parts of the
Arthur, M. 1998. Two or four Neoproterozoic glaciations? North Atlantic region. Palaeogeography, Palaeoclimatology,
Geology, 26, 1059 – 1063. Palaeoecology, 51, 209–229.
Knoll, A. H. 1982. Microfossil based biostratigraphy of the Precambrian Pisarevsky, S. A., Murphy, J. B., Cawood, P. A. & Collins, A. S. 2008.
Hecla Hoek sequence of Nordaustlandet, Svalbard. Geological Late Neoproterozoic and Early Cambrian palaeogeography: models
Magazine, 199, 269– 279. and problems. In: Pankhurst, R. J., Trouw, R. A. J., de Brito
Knoll, A. H. 1984. Microbiotas of the late Precambrian Hunnberg Neves, B. B. & de Wit, M. (eds) West Gondwana: Pre-Cenozoic
Formation Nordaustlandet, Svalbard. Journal of Palaeontology, Correlations Across the South Atlantic Region. Geological Society,
58, 131– 162. London, Special Publications, 294, 9 –31.
Knoll, A. H. 2000. Learning to tell Neoproterozoic time. Precambrian Schermerhorn, L. J. G. 1974. Late Precambrian mixtites: glacial and/or
Research, 100, 3– 20. non-glacial. American Journal of Science, 274, 673–824.
Knoll, A. H. & Swett, K. 1987. Micropaleontology across the Precam- Sokolov, V. N., Krasil’chov, A. A. & Livshits, YU. YA. 1968. The main
brian – Cambrian boundary in Spitsbergen. Journal of Palaeontology, features of the tectonic structure of Spitsbergen. Geological Maga-
61, 898– 926. zine, 105, 95– 115.
Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K. & Lambert, I. B. Teben’kov, A. M., Ohta, Y., Balashov, J. A. & Sirotkin, A. N. 1996.
1986. Secular variation in carbon isotope ratios from Upper Protero- Newtontoppen granitoid rocks; their geology, chemistry and Rb –Sr
zoic successions of Svalbard and east Greenland. Nature, 321, age. Polar Research, 15, 67– 80.
832– 837. Weil, A. B., Van der Voo, R., Mac Niocaill, C. & Meert, J. G. 1998.
Krasil’shchikov, A. A. 1973. The stratigraphy and paleotectonics of the The Proterozoic supercontinent Rodinia: paleomagnetically-derived
Precambrian – Early Paleozoic of Spitsbergen (translated title). Trudy reconstructions for 1100– 800 Ma. Earth and Planetary Science
Arkticheskogo Nauchno-Issledovatel’skogo Instituta, 172, 1 –120. Letters, 154, 13– 24.
Kulling, O. 1934. The Hecla Hoek Formation round Hinlopenstredet. Wilson, C. B. 1961. The upper Middle Hecla Hoek rocks of Ny Friesland,
Geografiska Annaler, 14, 161–253. Spitsbergen. Geological Magazine, 98, 89 – 116.
Larsen, V. B. 1987. A synthesis of tectonically related stratigraphy in the Wilson, C. B. & Harland, W. B. 1964. The Polarisbreen Series and other
North Atlantic – Arctic region from Aalenian to Cenomanian time. evidences of late Pre-Cambrian ice ages in Spitsbergen. Geological
Norsk Geologisk Tidsskrift, 67, 323– 338. Magazine, 101, 198– 219.
Lyberis, N. & Manby, G. 1993. The origin of the West Spitsbergen Fold Witt-Nilsson, P., Gee, D. G. & Hellman, F. J. 1998. Tectonostrati-
Belt from geological constraints and plate kinematics: implications graphy of the Caledonian Atomfjella Antiform of the northern Ny
for the Arctic. Tectonophysics, 224, 371–391. Friesland, Svalbard. Norsk Geologiske Tidsskrift, 78, 67 –80.
Chapter 56

Neoproterozoic (Cryogenian –Ediacaran) deposits in East and North-East Greenland

SVEND STOUGE1*, JØRGEN LØYE CHRISTIANSEN2, DAVID A. T. HARPER1, MICHAEL HOUMARK-NIELSEN3,


KASPER KRISTIANSEN4, CONALL MACNIOCAILL5 & BJØRN BUCHARDT-WESTERGÅRD6
1
Geological Museum, University of Copenhagen, DK-1350 Copenhagen K, Denmark
2
University College of Zealand, DK-4300 Holbaek, Denmark
3
GeoGenetics Centre, Natural History Museum, University of Copenhagen, DK-1350 Copenhagen K, Denmark,
4
Maersk Oil and Gas, Denmark
5
Department of Earth Sciences, Oxford University, UK
6
Department of Geography and Geology, University of Copenhagen, DK-1350 Copenhagen K, Denmark
*Corresponding author (e-mail: svends@snm.ku.dk)

Abstract: The Neoproterozoic succession of East and North-East Greenland (over 14 000 m thick) includes the Eleonore Bay Super-
group (?Tonian– Cryogenian) and the Tillite Group (Cryogenian–Ediacaran). The upper units of the Eleonore Bay Supergroup
consist of shallow to deeper-water carbonates, succeeded by siliciclastic fine-grained sediments (Bedgroup 19) that characterize the
top unit of the supergroup.
The Tillite Group includes two diamictite-bearing units (Ulvesø and Storeelv formations) of glaciogenic origin and two upper,
upwards-shallowing strata (Canyon and Spiral Creek formations) that were deposited during semiarid conditions and concluded the
Neoproterozoic depositional cycle. Diamictite is preserved on the craton and compares with the Storeelv Formation (Fm.) of the
Tillite Group.
Detailed investigations of the diamictite-bearing units (i.e. Ulvesø and Storeelv formations) demonstrate that the lower of the two for-
mations is mainly of marine origin, whereas the upper one has both marine and terrestrial origins. Chemostratigraphic data include ana-
lyses on total carbon (TC), total organic carbon (TOC), total sulphur (TS) and d13C. The data set for d13C shows a substantial and abrupt
shift towards negative values of 10%, from below Bedgroup 19. Low-diversity acritarch assemblages (Cryogenian) are recorded from
the Andrée Land and Tillite groups; a thin cherty dolostone unit present above the Storeelv Fm. suggests that the diamictite units are of
late Cryogenian age and the upper part of the Tillite Group is Ediacaran.
Bedgroup 19 disconformably overlies older carbonates and the unit is a prelude to the succeeding (upper Cryogenian– lower
Ediacaran) diamictite sediments of the Tillite Group. A disconformity separates the Tillite Group from the overlying Lower
Palaeozoic sediments. Both disconformities are, according to palaeomagnetic data, related to rift–drift episodes that occurred during
the late Neoproterozoic. Alternatively, the isotope data suggest that the diamictites were deposited during the late Cryogenian glaciation
and the older disconformity may be interpreted as a significant gap developed by the lowering of sea level during an early
Cryogenian glaciation.

The Neoproterozoic sediments in East and North-East Greenland Ulvesø on Ella Ø (728520 N, 258050 W) and with a coastal section
are well preserved and, despite their remote location, the area is beginning in Tømmerbugt and extending into Bastion Bugt,
well known and has served as key area for the study of Neoproter- serving as the reference section (Figs 56.3 & 56.4). It formed
ozoic successions in the Arctic (Figs 56.1 & 56.2). It should be part of the Cape Oswald Fm. of Poulsen (1930), and Schaub
noted that this region was previously referred to as East Greenland (1950) placed the unit in his Tømmerbugt Group. The Storeelv
in the literature but is now referred to as East and North-East Fm. takes its name from Storeelv on Ella Ø (728510 N, 258070 W)
Greenland according to new administrative boundaries. (Fig. 56.1).
Two diamictite bearing units (Ulvesø and Storeelv formations, Systematic research of the upper Neoproterozoic sediments
formerly referred to as lower and upper tillite respectively) are began in the mid-1920s and continued in the 1950s, when several
thought to record glacial conditions in the region during this expeditions visited the area under the leadership of Lauge Koch
time period. Other diamictite units occur in para-autochthonous (Koch 1929, 1930; Koch & Haller 1971; for a summary see
windows and are broadly correlated with the Storeelv Fm. The Haller 1971 and Henriksen & Higgins 2008). At that time, ship
Ulvesø and Storeelv formations are part of the Tillite Group and aircraft support had become widely available, which made it
(Haller 1971; Henriksen & Higgins 1976; Hambrey & Spencer possible to visit these remote parts of East and North-East
1987), which Kulling (1929) initially referred to as the Tillite Greenland. Stratigraphical, sedimentological and facies studies
Series. Even though the name does not follow modern guidelines from the mid-1970s onwards were undertaken by the Geological
for stratigraphic terminology, it has been used continuously in Survey of Greenland as part of its general mapping programme
the literature since 1929, and it is retained here following the (Henriksen & Higgins 1976; Bengård 1991; Henriksen 1999,
recommendation of Hambrey & Spencer (1987, p. 6). The Tillite 2003; Henriksen & Higgins 2008). Geologists from Harvard and
Group is exposed in a north – south-trending belt extending Cambridge University participated in expeditions to the region
through the central fjord region from Canning Land in the south in the 1970s and 1980s, which resulted in several new stratigraphi-
(728250 N) to Payer Land in the north (748200 N). The exposure is cal and sedimentological findings (Hambrey 1983, 1989; Hambrey
excellent in many places, especially along the coastlines of & Spencer 1987; Moncrieff & Hambrey 1988, 1990; Hambrey et al.
the fjords. 1989; Herrington & Fairchild 1989; Manby & Hambrey 1989;
The Ulvesø Fm. was formally defined by Hambrey & Spencer Swett & Knoll 1989) and descriptions of microfossils assemblages
(1987) with its type locality in Kløftelv, c. 200 m north of (Knoll et al. 1986; Green et al. 1987, 1988, 1989). The present

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 581– 592. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.56
582 S. STOUGE ET AL.

Fig. 56.1. Location map showing positions


of exposures, and the structure and geology
of East and North-East Greenland.
(Modified from Frederiksen 2000 and
Higgins et al. 2004). Used with permission
from the National Geological Survey of
Denmark and Greenland (GEUS) #.

lithostratigraphic scheme for the upper Neoproterozoic deposits in Structural framework


North-East Greenland (Hambrey & Spencer 1987) appeared at
around the same time, when Knoll et al. (1986) published their Most of the Neoproterozoic deposits of East and North-East
geochemistry-driven data for the broadly correlative successions Greenland accumulated in the Eleonore Bay basin that existed
on Spitsbergen. More recent research on stratigraphy, sedimentol- for c. 350 Ma. This sedimentary basin was nearly 500 km long,
ogy, micropalaeontology, palaeomagnetism and geochemistry NE –SW trending and 200– 400 km wide, and a pile of c. 14–
started in the late 1990s (Frederiksen et al. 1999; Frederiksen 15 km of sediments accumulated (Sønderholm et al. 2008).
2000) and continues now into the 21st century (Stouge et al. Eleonore Bay Basin sediments, preserved today in the Franz
2001; Sønderholm et al. 2008). Joseph Allochton of the Caledonian orogen were deposited
NEOPROTEROZOIC NE GREENLAND 583

Era Period Supergroup Group Formation Sequence Ma Events

Palaeozoic Cambrian (not named) Kong Oscar Fjord Kløftelv Sauk Iapetus passive margin
542
Hiatus Uplift
Ediacarian

580
Spiral Creek

Canyon III
635
Tillite
Storeelv
Neoproterozoic

Eleonore Bay

Group Maximal separation Fig. 56.2. Stratigraphical nomenclature


SQB used in this chapter and summary of the
Arena
series of events that occurred in the
II
Eleonore Bay basin during the Cryogenian
Ulvesø
Cryogenian

Drift and Ediacaran. The succession comprises


SQB
three stratigraphic sequences, and the hiatus
Bedgroup 20 Final carbonate deposition
between the Tillite Group and the Lower
I
Foundering of Palaeozoic Kløftelv Fm. is extensive. No
Bedgroup 19 Rifting
carbonate platform radiometric ages are available in this
Andrée Land SQB
succession; ages are inferred based on
Bedgroup 18 Carbonate platform interpretation of the stratigraphic and
biostratigraphic record. See text
for discussion.

c. 200– 400 km away, farther to the east (Higgins & Leslie 2000, The upper part of the Andrée Land Group is of special interest in
2008) than the present day. The Neoproterozoic succession in this context, and together with the Tillite Group crop out over an
the basin is thought to record a rift –drift transition associated area of c. 200 km from north to south and c. 50 km in an east –
with the fragmentation of Rodinia (Herrington & Fairchild 1989; west direction (Fig. 56.1). The outcrop area is bounded to the
Stouge et al. 2001; Eyles & Januszczak 2004; Nystuen et al. west by the older sediments of the Eleonore Bay Supergroup and
2008; Sønderholm et al. 2008), although volcanic deposits typi- to the east by the Lower Palaeozoic outcrops or by major north –
cally associated with rifted margins have not been reported from south-trending faults. The sediments are gently folded into syn-
the succession (Soper 1994; Higgins & Leslie 2008; Sønderholm forms and antiforms within the East and North-East Greenland
et al. 2008). Caledonides and are only mildly affected by Caledonian meta-
This Neoproterozoic sedimentary pile is incorporated in the morphism; the succession is locally jointed and faulted by
Caledonian mountain belts of eastern Greenland, extending from younger late Palaeozoic (Carboniferous) orogenic events.
708 to 828N and flanking the eastern edge of the Greenland ice sheet Additional occurrences of upper Neoproterozoic sediments are
(Henriksen & Higgins 1976, 2008; Henriksen 1999; Higgins et al. preserved in the Charcot Land, Gåseland and Målebjerg windows
2004; Gee et al. 2008; Higgins & Leslie 2008). The outcrop of Cale- in the lower allochthon and are para-autochthonous deposits
donian bedrock in eastern Greenland (Fig. 56.1) is c. 300 km wide (Fig. 56.1; Wenk 1961; Phillips et al. 1973; Montcrieff 1989;
in the south, at 708N, and narrows northwards to c. 100 km, striking Smith & Robertson 1999; Higgins et al. 2001, 2004; Sønderholm
obliquely offshore into the northern Greenland continental shelf. It et al. 2008). The deposits are overlain by the Hagar Bjerg and the
forms the western part of the Caledonian Orogen dominated by Niggli Spids thrust sheets (Higgins et al. 2004, 2008; Fig. 56.1).
thrust sheets, which were emplaced west-northwestwards onto These sediments accumulated at least 200 km to the west of the
the Laurentian Craton with its Lower Palaeozoic cover. Eleonore Bay basin but are thought to be broadly correlative
The allochthons are composed of two major thrust sheets, the with the Tillite Group (Higgins & Leslie 2000, 2008).
Niggli Spids thrust sheet and the Hagar Bjerg thrust sheet (Higgins
et al. 2001, 2004; Higgins & Leslie 2008) (Fig. 56.1). The lower
Niggli Spids thrust sheet comprises Archaean and Palaeoproterozoic Stratigraphy
basement gneisses overlain by upper Mesoproterozoic to lower
Neoproterozoic metasediments or the Krummedal supracrustal The allochthonous upper Neoproterozoic sediments in East and
sequence, which is several kilometres thick. The Hagar Bjerg North-East Greenland are collectively referred to as the Eleonore
thrust sheet overlies the Niggle Spids thrust sheet; it includes the Bay Supergroup (Andrée Land Group) and the overlying Tillite
Krummedal supracrustal sequence intruded by c. 950–920 Ma gran- Group (Haller 1971; Sønderholm & Tirsgaard 1993; Sønderholm
ites (Kalsbeek et al. 2001, 2008). These are overlain by the et al. 2008) (Fig. 56.2). Associated para-autochtonous deposits
c. 18.5-km-thick Neoproterozoic siliciclastic and carbonate succes- are broadly correlated with the Storeelv Fm. of the Tillite Group
sions (Eleonore Bay Supergroup), siliciclastic sediments with dia- in the Eleonore Bay Basin (Montcrieff 1989).
mictite (Tillite Group), and Cambrian–Middle Ordovician
siliciclastic and platform carbonate formations (Kong Oscar Fjord
Group). This sedimentary package comprises the upper part of the Andrée Land Group
Hagar Bjerg thrust sheet and is recognized as the Franz Joseph
allochthon (Higgins et al. 2004; Higgins & Leslie 2008). The top unit of the Eleonore Bay succession is the Andrée Land
All the transported rocks in eastern Greenland were derived Group (Teichert 1933; Tirsgaard 1993, 1996; Sønderholm & Tirs-
from the Laurentian margin, including both Archaean and Palaeo- gaard 1993; Tirsgaard & Sønderholm 1997) (Fig. 56.2), which
proterozoic crystalline basement and the younger, shallow-marine consists primarily of shallow to relatively deeper-water marine
sedimentary cover. WNW-directed thrust transport of the alloch- carbonates succeeded by marine deep-water siltstone-shale,
thons has been estimated to be at least 200 km (Higgins & Leslie cherty rhythmites and carbonate. It totals c. 1500 m in thickness
2000, 2008), along with sinistral strike displacements; hence, and was recognized previously as the ‘Die Kalk-Dolomit Serie’
prior to the Caledonian Orogeny, the platform margin of Laurentia (i.e. ‘Limestone-Dolomite Series’; Teichert 1933), which was sub-
was nearly twice as wide as it is today. divided into a number of ‘bed groups’ (Teichert 1933; Schaub
584 S. STOUGE ET AL.

Fig. 56.3. Stratigraphy of the uppermost


Andrée Land Group (Bedgroups 19 and 20)
and Tillite Group. The sedimentary logs
from Andrée Land and Ella Ø represent the
northern and southern development of the
succession respectively. AS, sandstone
(aeolian) unit forming at the top of
Bedgroup 20; CD, cap dolomite following
above the diamictite of the Storeelv Fm.

1950; Katz 1952; Eha 1953; Fränkl 1953; Haller 1953, 1971; shallow-water carbonates that accumulated on a ramp situated
Cowie & Adams 1957; Sommer 1957). These informal bed along a former low-latitude passive continental margin (Frederik-
groups have since been widely used as standard reference nomen- sen et al. 1999; Frederiksen 2000; Stouge et al. 2001; Mac Niocaill
clature in the literature for the Eleonore Bay Supergroup (Haller et al. 2004, 2008; Sønderholm et al. 2008). The uppermost strata of
1971; Hambrey & Spencer 1987; Herrington & Fairchild 1989; the Andrée Land Group are referred to as Bedgroup 19 and Bed-
Swett & Knoll 1989; Fairchild & Hambrey 1995). group 20 (¼ AL6 and AL7 of Sønderholm & Tirsgaard 1993).
The Andrée Land Group was formally established by Sønder- These are up to c. 400 m thick and consist of fine-grained clastic
holm & Tirsgaard (1993), who also introduced informal litho-units sediments and carbonate.
(AL1 –7), which should serve as replacements for the bedgroups.
However, the older stratigraphic terminology summarized by
Haller (1971) is still applicable and is used here as stratigraphical Tillite Group
framework (Fig. 56.2).
The upper Andrée Land Group, named Bedgroup 18 (¼ AL5 The upper part of the Neoproterozoic succession comprises the
of Sønderholm & Tirsgaard 1993), is composed of marine, siliciclastic deposits and minor carbonates of the Tillite Group
NEOPROTEROZOIC NE GREENLAND 585

Fig. 56.4. Detailed stratigraphic logs of the


Ulvesø and Storeelv formations. The Ulvesø
Formation is from the type section at Ulvesø
in Kløftelv on Ella Ø and the reference
section from Tømmerbugt til Bastion Bugt,
Ella Ø. Stratigraphic log of the Storeelv
Formation is from the coast section in
Bastion Bugt, Ella Ø.
586 S. STOUGE ET AL.

(c. 900 m), representing terrestrial and marine depositional settings In detail, Unit A begins with quartzitic, fine-grained sandstone
in five different formations (Figs 56.2 & 56.3). The marine depos- overlying the black limestone of Bedgroup 18. The quartzitic sand-
its accumulated at some distance from the continental margin in a stone is succeeded by bedded chert and rhythmically bedded grey
deep-water, oceanic setting. The diamictite facies of the Ulvesø chert. The chert beds are 1–5 cm thick and have millimetre-scale
and Storeelv formations are separated by mudstone and sandstone partings of grey-green shale. This unit is succeeded by rhythmi-
facies of the Arena Fm. and succeeded, first by deep-water sedi- cally bedded chert with partings consisting of 0.1 –0.5 mm thin
ments and then by shallowing-upwards successions, where the green-grey shale as well as interbeds of grey-red-brown shale over-
uppermost unit concludes with evaporitic environments (Canyon lain by finely laminated chert and grey shale. The chert is charac-
and Spiral Creek formations; Fig. 56.2). terized by light yellow to brown weathering surfaces. Besides
A regional synthesis of the depositional environments and occasional ripples, sedimentary structures are rare to absent in
stratigraphical nomenclature was reviewed by Haller (1971) and the unit.
revisions were published by Cowie & Adams (1957), Hambrey The basal quartzitic unit of Unit A is sparsely exposed, so it has
& Spencer (1987), Fairchild & Hambrey (1995), Sønderholm & not been reported previously. It is well exposed in the region from
Tirsgaard (1993) and Sønderholm et al. (2008). The switch from Andrée Land and east of Eleonore Bay and towards the north
deposition of carbonate to deep-water, mostly fine-clastic, turbi- (Figs 56.1 & 56.3). On Ella Ø, Schaub (1950) and Hambrey &
dites on the slope in the upper part of the Andrée Land Group her- Spencer (1987) described Unit A in detail and divided it into
alded the deposition of the Cryogenian –Ediacaran diamictite units several beds, but Unit A has also an extensive lateral distribution.
in the region. It is recorded across the whole basin and with a nearly uniform
The Eleonore Bay Supergroup disconformably overlies the development. Many of the individual beds can be traced over
Krummedal supracrustal succession (Higgins et al. 2004, 2008; long distances; one prominent chert/carbonate horizon with
Nystuen et al. 2008). Granites, which are between 950 and shale partings, 7 to 10 m thick, can easily be traced laterally
920 Ma old (Kalsbeek et al. 2001, 2008), intrude the Krummedal from Lyell Land to Ella Ø to Ole Rømer Land and farther to the
supracrustal sequence, and the c. 900 Ma date for the granites north (i.e. for more than 200 km); it can thus be used as a
may represent a maximum age for the initial start of deposition marker bed across the whole basin.
in the Eleonore Bay Basin. The late Neoproterozoic succession Unit B is composed mainly of grey and black relatively
is unconformably overlain by the quartzitic sandstone of the Kløf- organic-rich shale that is interbedded with subordinate parted to
telv Fm. (Lower Cambrian) of the Kong Oscar Fjord Group in the ribbon limestone and limestone breccia. The ribbon and parted
central fjord zone (Fig. 56.2). limestone forms the base of the unit and is very fine-grained,
dark grey to black, finely laminated and with limestone beds
varying in thickness from 4 to 40 cm. Several horizons of the
Associated para-autochthonous strata ribbon and parted limestone are slumped and contorted. Above
the interval with parted to ribbon limestone and breccia beds
Diamictite beds have been recorded from Charcot Land (Mon- follows a unit up to 100 m thick, dark grey to black shale.
tcrieff 1989), Gåseland (Wenk 1961; Phillips et al. 1973) and Mål- Unit B is recorded only in the southern extension of the basin
ebjerg windows (Smith & Robertson 1999) (Fig. 56.1). The (Sønderholm et al. 2008).
diamictite deposits are preserved in pockets on the surface of The limestone breccias of Unit B have a chaotic fabric with a
the basement and below the higher thrusts. The Støvfanget Fm. matrix composed of grey to black, dirty green to dull brown,
(Montcrieff 1989) is named from the succession in the Gåseland purple-to-yellow weathering calcareous mud and lime mudstone.
window. The strata from the Charcot Land and Målebjerg Purple-to-yellow weathering carbonate and elongated plates of
windows were named the Tillite Nunatak Fm. (Montcrieff 1989; low lithological variety are the most common clasts. They are com-
Smith & Robertson 1999). The sediments are considered to be posed of grey, blue-grey, to dark grey limestone and dolomite
lateral equivalents, so the name Støvfanget Fm. is used here for plates. The lime breccias are mostly thin but well exposed. On
these deposits. Ella Ø, one breccia horizon up to 10 m thick serves as a local
lithological marker band in this region (Hambrey & Spencer
1987, fig. 6; Herrington & Fairchild 1989).
Glaciogenic deposits and associated strata

Bedgroup 19, Andrée Land Group Bedgroup 20, Andrée Land Group

Bedgroup 19 is characterized by deep-water black, green, grey and Bedgroup 20 (AL7 as per Sønderholm & Tirsgård 1993) is mainly
red shales, interbedded with chert, dolostone, minor limestone, a carbonate unit (Katz 1952; Eha 1953; Fränkl 1953; Haller
interbedded limestone and shale and carbonate breccias and con- 1953; Hambrey & Spencer 1987; Herrington & Fairchild 1989;
glomerates (Fig. 56.2). Bedgroup 19 is overlain either by Bedgroup Sønderström et al. 2008), which conformably overlies Unit A of
20 or the Ulvesø Fm. of the Tillite Group (Sønderholm et al. 2008). Bedgroup 19. Bedgroup 20 is typically composed of (i) silty,
The lateral distribution of facies is heterogeneous and the succes- planar bedded ribbon limestone at the base, (ii) lime mudstone
sion shows a northern and a southern development (Fig. 56.3). Bed and bedded limestone in the main part of the unit (Fig. 56.3) and
group 19 is here subdivided into two informal units: Unit A and (iii) planar lamellites associated with small domal stromatolites
Unit B. Unit A is well developed throughout the basin, whereas and minor nodular cherts. Oolitic-pisolitic limestone, up to 40 m
Unit B is only found in the southern portion of the basin. thick, prevails in the higher part of the unit on Ole Rømers Land
Unit A is the stratigraphically oldest strata of Bedgroup 19 of the (Swett & Knoll 1989; Frederiksen 2000).
upper Andrée Land Group (Sønderholm et al. 2008). It is 200 m or Bedgroup 20 varies in thickness from 240 m in Ole Rømer Land
more thick, and is characterized by mixed chert and fine-grained and decreases towards the south where 2 m are recorded on the
to very fine-grained siliclastic sediments associated with minor southern part of Ymer Ø. The unit extends northwards as far as
carbonate. Lithologies include argillite (shale); siltstone; bedded Hudson Land. Farther to the south and towards the southern exten-
light grey, dense thin-bedded chert, developed as rhythmites; sion of the Eleonore Bay Basin, the carbonates of Bedgroup
and siliceous argillite interbedded with variable amounts of thin- 20 are absent.
bedded and very fine-grained grey, green to violet chert. Yellow The uppermost part of Bedgroup 20 is composed of yellow,
to yellow-brown weathering is characteristic of the finely lami- medium-grained sandstone with well-developed and large cross-
nated grey to green to violet chert. beds (Fig. 56.3). This sandstone subunit (c. 40 m) is shown in
NEOPROTEROZOIC NE GREENLAND 587

Fig. 56.3 as ‘aeolian sandstone’ (AS). This subunit displays the alternate with breccias of limestone and sandstone. More com-
same geographical distribution as Bedgroup 20 (Sønderholm monly, they interfinger with cross-bedded sandstone and laminated
et al. 2008). shale with outsized clasts (Figs 56.3 & 56.4). Conglomerate-filled
mega-channels truncate the diamictite, mainly in the middle
part of the formation. Towards the top, diamictite units once
Ulvesø Fm., Tillite Group more dominate the succession. However, here it interfingers
with minor beds of laminated shale with dropstones and
At Ella Ø, the Ulvesø Fm. comprises a generally coarsening bedded sandstone.
upwards succession composed of diamictite, associated with
minor laminite and shale, massive sandstone and conglomerate
(Figs 56.3 & 56.4). The crudely bedded diamictite is composed Canyon Fm., Tillite Group
predominantly of carbonate clasts derived mainly from the older
Andrée Land Group and sediments of Bedgroup 19 in the southern The Canyon Fm. was first named the Canyon Series by Schaub
part of the basin, and from carbonate units of the older Andrée (1955) and Fränkl (1953). Later, Haller (1971) modified it to the
Land Group and Bedgroup 20 to the north of Ymer Ø. Beds are Canyon Fm. The name is derived from Tillite Canyon Fm.
normally 30–50 cm thick, and clasts are occasionally up to 1 m (Poulsen 1930), but the name Tillite was dropped by subsequent
in size. The matrix is grey-green and composed of shale, silt and authors.
sand, which weather characteristically as dull yellow to brown. In the northern part of the basin, this formation includes a cherty
The formation varies laterally in thickness from nearly 10 m dolostone horizon at the base. The cherty carbonate unit, up to
in the north to a maximum thickness of c. 320 m on northern 10 m thick, is situated at the base of the formation and extends
Scoresby Land. laterally from Andrée Land (Tillite Kløft and west of Kap
The erosive-based diamictite is rich in striated outsized clasts Weber) and northwards to Albert Heim Bjerge (Cowie & Adams
up to a metre in size and slump folds and mega breccias 1957). The unit is composed of light grey, dense and finely lami-
occur frequently throughout the succession. Towards the top, dia- nated chert or siliceous dolostone. Above follows a rhythmically
mictite interfingers with erosive-based sandstone and clast- bedded shale and chert. The shale is commonly very red, and the
supported conglomerate, beds of shale and brecciated limestone chert beds are pale grey to nearly white, so the shale/chert unit
derived from the breakdown of the underlying carbonate-shale is easy to recognize from a distance. This cherty carbonate
platform. Together with the lowermost part of the sandstones of unit is not present on Ella Ø, and red shale follows above the
the overlying Arena Fm., the diamictite of the Ulvesø Fm. Storeelv diamictite. The characteristic, rhythmically bedded, red
appears strongly deformed, occurring as metre-sized ball and shale and grey finely laminated chert appear c. 10 m above the
pillow structures. Storeelv Fm.
The subsequent and main part of the succession of the Canyon
Fm. is composed of red, grey green and black shale. Towards
Arena Fm., Tillite Group the top and on Ella Ø, the Canyon Fm. becomes increasingly
sandy and the sandstone and shale display well-developed,
The Arena Fm. is composed of grey, dark-grey to green shale, upwards coarsening sequences.
siltstone and sandstone (Figs 56.2 & 56.3). The formation was On Ella Ø, the top of the formation is developed as c. 50 m
formally defined by Hambrey & Spencer (1987), with its type yellow weathering dolostone with small stromatolitic reefs and
locality at Arenaen on Gunnar Andersson Land, Ymer Ø algal laminations. This facies of the Canyon Fm. is only seen on
(738200 N, 248500 W). The formation was previously included in Ella Ø. Elsewhere, black shale with prominent yellow-weathering
the Tillite Series (Kulling 1929; Schaub 1955; Katz 1954, 1961) surfaces marks the top of the formation.
and the Cape Oswald Fm. of Poulsen (1930). Later it was named
the Inter Tillite Member (Schaub 1950) or Inter-Tillite (Haller
1971; Henriksen & Higgins 1976; Higgins 1981). Spiral Creek Fm., Tillite Group
The formation varies from c. 100 m in thickness in Strindberg
Land and Ole Rømer Land, through 220– 360 m in thickness in Spiral Creek Fm. is the topmost unit of the Neoproterozoic
Albert Heim Bjerge (Hudson Land). The formation begins with succession in East and North-East Greenland (Fig. 56.2). The for-
shale and silt beds followed by thick-bedded green to grey sand- mation is named after Spiral Creek, a locality on the northern shore
stone. The shale and siltstone are the dominant lithologies, with of Andrée Land (Poulsen 1930, p. 307), where the unit is well dis-
a minor increase in the amount of sandstone occurring towards played. It is also known from Lyell Land, Ella Ø and Ole Rømer
the top of the Arena Fm. The recessive Arena Fm. separates the Land. Spiral Creek Fm. is up to 45 m thick and is composed of silt-
diamictite units and is not easy to access in inland outcrops, and stone, sandstone and stromatolitic dolostone. The colours of the
the beds of the unit are commonly covered. sediments vary from yellow, grey to green and maroon. The for-
On Ella Ø, the lowest part of the unit is composed of yellow mation is divided into several beds or horizons (Poulsen 1930;
sandstone, which is similar to the sandstone on Scoresby Schaub 1950; Cowie & Adams 1957). Halite pseudomorphs,
Land. The latter includes thin diamictite horizons (Hambrey & mudcracks, ripples and intraformational breccia are frequent in
Spencer 1987). Sedimentary structures in the Arena Fm. comprise the beds. The sediments probably deposited in a playa-like
cross-lamination, wave ripples, graded bedding, load-casts and setting and under semiarid conditions (Hambrey & Spencer
slump structures. 1987; Fairchild & Hambrey 1995).

Storeelv Fm., Tillite Group Støvfanget Fm., para-autochthonous deposits

The Storeelv Fm. ranges in thickness from 220 m on Ella Ø The sediments of the Støvfanget Fm. include diamictite, sandstone
(Fig. 56.3) and Arenaen to 60 m in Strindberg Land in the north and laminated mudstone with dropstone (Smith et al. 2004;
and Canning Land in the south. On Ella Ø, the formation conform- Sønderholm et al. 2008). The composition of the clasts is hetero-
ably rests on a striated glacial boulder pavement of outsized geneous and derived from local sources; granitic blocks are
far travelled erratics. Crudely bedded reddish diamictite with common. The blocks vary in size and are up to 2 m, rarely 6 m,
clasts of local and extra-basinal origin, including igneous and in size. The succession varies from a couple of tens of metres to
metamorphic rocks of the Archaean basement, occasionally a maximum of c. 200 m in thickness.
588 S. STOUGE ET AL.

Boundary relations with overlying and underlying Canyon and Spiral Creek formations (Tillite Group) revealed nega-
non-glacial units tive d13Corg values between –20 and – 35‰PDB. The d13Ccarb
values were predominantly positive throughout the Eleonore Bay
Boundary relations within the Neoproterozoic succession Supergroup with a sharp decrease to negative values ( –4 to
–9‰PBD) in the uppermost Bedgroup 19/20 and negative
The lower boundary of Bedgroup 19 is not well exposed in the values (c. 0 to –5‰PDB) in the Canyon and Spiral Creek for-
southern part of its extent of outcrop. The boundary relationships, mations. Both geochemical data sets followed very similar profiles
however, are well displayed in Andrée Land and Strindberg Land, suggesting a primary signature and an unusual secular variation in
where the boundary is a disconformity. The carbonate top surface the Neoproterozoic carbon cycle (Knoll et al. 1986).
of Bedgroup 18 is dissolved and has karst-like features (Frederik- Total carbon (TC), total organic carbon (TOC) and total sulphur
sen 2000; Sønderholm et al. 2008). This carbonate surface is (TS) have been recorded through an interval ranging from the
overlain by a thin, quartzitic and medium-grained sandstone. upper part of the Andrée Land Group and up to and including
The quartzitic bedded sandstone is succeeded by bedded fine- the top of Tillite Group; the results will be presented elsewhere.
grained to grey chert deposits with thin shale partings. The lower Stable C-isotopes from the upper part of the Andrée Land Group
boundary of Bedgroup 20 is conformable, with a gradual transition have also been analysed and will be presented elsewhere (see
from dark grey to black shale, through interbedded silty laminated Halverson et al. 2005; Sønderholm et al. 2008 for further details).
limestone to bedded limestone.
The base of the overlying Tillite Group is usually sharp in the
deep part of the basin, where black shale (Unit B) of Bedgroup Other characteristics
19 is overlain directly by the Ulvesø Formation diamictite
(Hambrey & Spencer 1987). The boundary towards the north is Organic constructed organisms (acritarchs) in the succession
different and the top carbonate strata of Bedgroup 20 are separated represent the plankton of the Andrée Land –Tillite Group basin.
from the diamictite of the Ulvesø Fm. by a yellow (aeolian), quart- The microflora assemblage is of low diversity and consists of
zitic sandstone. This sandstone is conformably overlain by the both small and large smooth leosperoid types associated with
Ulvesø Fm. of the Tillite Group. On Ymer Ø, Bedgroup 20 is in only few acanthomorphs (Vidal 1976, 1979, 1985). The biostrati-
faulted contact with the overlying Ulvesø Fm. and boundary graphic value of the acritarch assemblage is at present limited, but
relationships are obscure (Hambrey & Spencer 1987; Moncrieff in general the microflora from Andrée Land Group is assigned to
& Hambrey 1988). the Cryogenian (Vidal 1976, 1979; Green et al. 1987, 1988,
The upper boundary of the Ulvesø Fm. is in most places 1989). Other microfossils reported from the succession are vase-
gradual and is marked by a transition from dark grey diamictite, shaped protists (Green et al. 1988), similar to those reported in
conglomerate or sandstone to the lower sandstone of the Arena other Neoproterozoic successions (Porter & Knoll 2000; Porter
Fm. On Ella Ø the base of the overlying Arena Fm. is sharp and et al. 2003).
strongly deformed with the underlying Ulvesø into metre-sized
load-casts.
The upper Neoproterozoic para-authochthonous sediments rest Palaeolatitude and palaeogeography
directly on basement rocks and are tectonically constrained
upwards by thrust sheets; however, the diamictite in the Målebjerg Palaeogeographical reconstructions of East and North-East Green-
window is overlain disconformably by Lower Cambrian quartzitic land have long been tied to the palaeogeographical evolution of the
sandstone referred to as the Slottet Fm. (Smith et al. 2004). Neoproterozoic succession in northeastern Svalbard based on
detailed lithostratigraphic similarity between these two basins
(Harland & Gayer 1972; Fairchild & Hambrey 1995; Halverson
Boundary relations with the Lower Palaeozoic succession et al. 2004; Sønderholm et al. 2008; Halverson 2011).
The Hekla Hoek succession in northeastern Spitsbergen is
The Lower Cambrian Kløftelv Fm. of the Kong Oscar Fjord Group similar to the Eleonore Bay Supergroup of northeastern Greenland
(Lower Cambrian to Middle Ordovician) unconformably overlies and shows development from siliciclastic to carbonate deposition
the Tillite Group across the whole Eleonore Bay basin and the followed by glaciogenic deposits. This similarity is also strength-
boundary is a low-angle angular unconformity (Henriksen & ened by comparable and substantial thicknesses of the lithological
Higgins 1976; Hambrey 1989; Hambrey et al. 1989; Stouge units in the two regions (Harland & Geyer 1972; Hambrey 1989;
et al. 2001, 2002; Sønderholm et al. 2008). Thus the boundary Fairchild & Hambrey 1995; Harland 1997).
relationships from the north to south are variably developed. The chemostratigraphy shows similarities (Knoll et al. 1986)
In Andrée Land and on Ella Ø, the upper Neoproterozoic eva- suggesting close affinities between the two areas. Halverson
poritic sediments of Spiral Creek Fm. are overlain unconfomably et al. (2005) identified the distinctive pre-Marinoan negative
by the Lower Cambrian Kløftelv Fm., and the Lower Palaeozoic d13C anomaly or the ‘Trezona anomaly’ beneath the lowest dia-
succession begins with a basal conglomerate (Stouge et al. 2001, mictite in northeastern Svalbard. The same anomaly has been
2002). Away from Andrée Land and Ella Ø, the Spiral Creek recognized from within the upper unit of the Eleonore Bay Super-
Fm. and the upper stromatolitic limestone facies of the Canyon group in East and North-East Greenland (Sønderholm 2008; Kris-
Fm. are absent and black shales of Canyon Fm. are directly over- tiansen unpublished data), which also provides a strong argument
lain by a basal conglomerate followed by the arenitic sandstone for the similarity of the two regions. Palynology studies in both
of the Lower Cambrian Klöftelv Fm. The hiatus separating the East and North-East Greenland and Svalbard (Vidal 1985; Green
Neoproterozoic from the Palaeozoic strata is proposed to have et al. 1989; Swett & Knoll 1989) also indicate close relationships
lasted for about 35 Ma (see Sønderholm et al. 2008 for further between the two areas and as a whole the two regions are con-
discussion). sidered to be deposited within the same basin or series of ensialic
basins, with East and North-East Greenland located near eastern
Svalbard (Sønderholm et al. 2008).
Chemostratigraphy Recent and ongoing palaeomagnetic studies have shown that
East and North-East Greenland was situated at equatorial latitudes
Early studies of carbonate and organic carbon geochemistry were at about 38S during the Cryogenian, with palaeomagnetic data
carried out by Knoll et al. (1986). Samples from Bedgroup 7 from 23 sites (105 specimens) towards the top of the Eleonore
through 20 (Eleonore Bay Supergroup) as well as from the Bay Supergroup yielding a mean palaeomagnetic inclination
NEOPROTEROZOIC NE GREENLAND 589

of 58 (k ¼ 12; a95 ¼ 98) (Mac Niocaill et al. 2004, 2008; Kilner Bedgroup 20) was due to a marked regression and perhaps also
2005). These data are constrained by a positive fold test (Caledo- due to a change in palaeo-climate.
nian folding) and a positive conglomerate test on clasts in the The diamictites of the Tillite Group have long been considered
overlying lower Tillite. At that time the carbonate sedimentary glaciogenic deposits (Kulling 1929; Poulsen 1930 and subsequent
succession of the Andrée Land Group was deposited in a platform authors). Hambrey & Spencer (1987) and Moncrieff & Hambrey
setting on a ramp. The region formed the eastern part of the north- (1988, 1990) provided detailed models for glacial depositional
eastern margin of the Rodinia palaeocontinent. A change in the environments. Halverson et al. (2004, 2005) placed the Tillite
latitudinal position began at the base of Bedgroup 19 as the conti- Group deposits within the scope of ‘Snowball Earth’.
nental margin region started to move southwards. During depo- The Ulvesø and Storeelv formations are indeed dominated by
sition of the Storeelv Fm., the area was situated at c. 668 diamictite, with megabreccias and sculptured clasts alternating
southern latitude, with sparse palaeomagnetic data from six sites with minor quantities of shale, sandstone and conglomerate. A ter-
(nine specimens) in the Tillite Group yielding a mean palaeomag- restrial origin can most probably be applied to the lowermost and
netic inclination of 788 (k ¼ 10; a95 ¼ 228) (Mac Niocaill et al. middle parts of the Storeelv Fm.; the facies successions here indi-
2004, 2008; Kilner 2005), although this awaits confirmation with cate a highly diverse depositional system. Glaciers grounded on
further analyses in progress. The movement towards higher, the shelf or in fjords delivered till and glacio-fluvial deposits inter-
southern latitudes is associated with the shift from stable carbonate bedded with flow diamictites and lacustrine and aeolian facies.
accumulation of Andrée Land Group to the siliciclastic deposition Other parts of the Storeelv and most of the Ulvesø Fm. were
of Bedgroup 19 and the start of foundering of the platform. The most probably deposited below the grounding line in sub-aqueous
presence of Bedgroup 20 signifies that carbonate accumulation – environments. Consequently, the Tillite Group contains a limited
for a short while – returned to the region, but finally ceased in amount of tillite, and although most components in the succession
the late Cryogenian. originated from glacial debris, most diamictites finally settled in
marine settings, especially those of the Ulvesø Fm. A marine
origin is supported by the occurrence of algal life forms throughout
Geochronological constraints the Tillite Group, except where sediments show signs of oxidation
(Stouge & Piasecki unpublished data). The diamictite units were
Radiometric ages have not been obtained from the Andrée Land principally deposited from suspensions of glacial debris released
and Tillite groups. A Cryogenian age is confirmed based on the beneath floating ice shelves and by ice-rafting in a glaciomarine
biostratigraphic record (see previous section). environment. These deposits suffered slumping and reworking
by sediment gravity flows in a near shelf break environment.
Only parts of the Storeelv Fm. indicate that glaciers occasionally
Discussion reached grounding line or more rarely that terrestrial ice sheets
deposited tillite and glaciofluvial deposits. It is not clear how
Depositional setting much conglomerate and sandstone that originated in on-shore
fluvial and aeolian environments was actually reworked to even-
The stratigraphy of the upper Andrée Land Group has been con- tually settle in submarine canyons and in current- and wave-
sidered to conform to a normal superposition succession, where dominated shallow water.
Bedgroup 20 overlies Bedgroup 19. However, Eha (1953) pro-
posed that the carbonate sediments of Bedgroup 19 (¼ base of
Unit B of this chapter) represented the lateral equivalents of Bed- Sequence stratigraphy
group 20. In this review, the upper part of Bedgroup 19, that is, the
sediments of Unit B and Bedgroup 20 (¼ AL7), are considered The upper part of the Andrée Land Group and the Tillite Group
coeval (Fig. 56.3), as this interpretation is supported by the sedi- represent three depositional sequences that mostly accumulated
mentology, intragroup stratigraphy and the overall stratigraphical as deeper- to deep-water sediments in oceanic settings (Fig. 56.2;
succession developed during recent fieldwork. Unit B is con- Sønderholm et al. 2008). The oldest sequence is represented by
sidered the fine-grained and deep-water equivalent of Bedgroup Bedgroups 19 and 20. The lower sequence boundary is an erosional
20, dominated by parted to ribbon limestone, debris-flow deposits, or a dissolution surface. Siliciclastic silt, shale and chert were
lime-breccias containing intrabasinal clasts and distal turbiditic deposited in the transgressive system tract during the progradation
shale at the basin margin and organic-rich shale in the basin. of the coastline outside the study area. The maximum flooding
Bedgroup 19 (Eha 1953; Fränkl 1953; Haller 1971) of the upper- surface can be traced across the whole basin, and the following
most Andrée Land Group significantly provides the first record of highstand deposition created the regressive platform of Bedgroup
fine-grained to very fine-grained siliciclastic sediment and chert 20 and the distal deposits of Unit B. The regressive top of
rhythmites (i.e. Unit A), which followed disconformably above Bedgroup 20 is composed of shallow-water silt and sandstone
the shallow-water stromatolite-bearing limestone and slightly deposited in tidal flat environment and aeolian sandstones.
deeper-water carbonate rocks. Limestone (Bedgroup 20) and The second sequence comprises the Ulvesø and Arena for-
shale and ‘lime-breccias’ (i.e. Bedgroup 19, Unit B) underlie the mations (Fairchild & Hambrey 1995: Sønderholm et al. 2008),
first diamictite unit (i.e. Ulvesø Fm.) of the Tillite Group. This suc- which are developed exclusively in clastic facies. The third
cession is interpreted as recording deepening within the basin and sequence comprises the transgressive Storeelv Fm. – including
foundering of a carbonate platform that had developed on a stable the thin carbonate unit – and the Canyon Fm., which becomes
craton and had been in existence for a long period of time (i.e. regressive upwards, permitting shallow-water carbonates to
during deposition of the underlying Ymer Ø Group and most of accumulate. The overlying Spiral Creek Fm. represents the ulti-
the Andrée Land Group). In addition, the rapid facies change mate part of the highstand, and the regressive top of the Tillite
from carbonate accumulation to siliciclastic deposition shows Group accumulated in shoreface and semi-arid, evaporitic lagoo-
that rift-related subsidence was initiated in the Cryogenian, creat- nal environments (Fairchild & Herrington 1989).
ing north –south elongated basins with isolated and carbonate-
capped platforms and deep-water troughs. Bedgroup 20 carbonates
probably developed a low-relief bank profile with mobile pisolite Timing of glaciation
shoals at the margin, and shallow sub-tidal carbonate and peritidal
deposits in the interior. The final carbonate deposition in the region Possibly two widespread or even global glaciations took place
(i.e. before the occurrence of diamictite and represented by during the late Neoproterozoic (Hambrey & Harland 1995; Knoll
590 S. STOUGE ET AL.

2000; Hoffman & Schrag 2002). It is, however, uncertain if the two The authors thank the reviewers M. J. Hambrey and A. H. Knoll for their valuable
diamictite units in East and North-East Greenland represent one or comments and suggestions, from which the manuscript greatly benefited. Field-
two separate glaciations and in what respect they correlate with work in Greenland was funded both by the Carlsberg Foundation and the
other Neoproterozoic phases of glaciations (e.g. Kennedy et al. Danish Research Council. The Geological Survey of Denmark and Greenland
1988; Kaufman et al. 1997; Brasier & Shields 2000; Hoffmann (GEUS) gave permission to reproduce elements in Figure 56.1. This represents
& Schrag 2002; Robb et al. 2004). a contribution of the IUGS- and UNESCO-funded IGCP (International
Geoscience Programme) Project #512.
Halverson et al. (2004, 2005) suggested that the two diamictite
units on Svalbard represent one glaciation. This idea is based on
the d13C-isotope curve (Knoll et al. 1986; Halverson et al. 2004,
2005) and the presence of a distinctive d13C anomaly (¼‘Trezona References
anomaly’) found beneath the lowest tillite unit in Svalbard. In
addition, the presence of the thin carbonate unit, which should Bengaard, H.-J. 1991. Upper Proterozoic (Eleonore Bay Supergroup) to
be equal to a ‘cap dolomite’ (Fairchild & Hambrey 1995), is Devonian central fjord zone, East Greenland, geological map, 1:
typical of the late Cryogenian glaciation. The diamictite units in 250000. Copenhagen, Geological Survey of Greenland.
Svalbard were therefore related genetically to the late Cryogenian Brasier, M. D. & Shields, G. 2000. Neoproterozoic chemostratigraphy
glaciation (Marinoan) by Halverson et al. (2004, 2005). and correlation of the Port Askaig glaciation, Dalradian Supergroup
Based on their stratigraphic similarity (mentioned previously), of Scotland. Journal of the Geological Society, London, 157,
909– 914.
the two diamictite units in East and North-East Greenland may
Cowie, J. W. & Adams, P. J. 1957. The geology of the Cambro-
also contain tillite deposits of the Marinoan (late Cryogenian) gla-
Ordovician rocks of central Greenland. Part 1: Stratigraphy and struc-
ciation. The trend of the d13C-isotope curve (Knoll et al. 1986; ture. Meddelelser om Grønland, 153, 193.
Kristiansen et al. unpublished data) from the upper part of the Eha, S. 1953. The pre-Devonian sediments on Ymers Ø, Suess Land,
Andrée land Group carbonates, combined with the presence of a and Ella Ø (East Greenland) and their tectonics. Meddelelser om
possible cap dolomite above the Storeelv Formation in the northern Grønland, 111, 105.
part of the basin, suggests that the two diamictite units in East and Eyles, N. & Januszczak, N. 2004. ‘Zipper-rift’: a tectonic model for
North-East Greenland could be related to one glaciation. Neoproterozoic glaciations. Earth Science Reviews, 65, 1 –73.
A problem with this interpretation arises, however, because Fairchild, I. J. & Herrington, P. M. 1989. A tempestite-stromatolite-
any trace of an older Cryogenian glaciation often found in other evaporite association (Late Vendian, East Greenland): a shoreface-
Neoproterozoic successions is apparently missing in the Greenland lagoon model. Precambrian Research, 43, 101– 127.
succession. Although no diamictite-bearing units occur below the Fairchild, I. J. & Hambrey, M. J. 1995. Vendian basin evolution in
Ulvesø Fm., the extensive disconformity developed at the bound- East Greenland and NE Svalbard. Precambrian Research, 73,
ary between Bedgroup 18 and Bedgroup 19 may be the response to 217– 223.
a global or eustatic lowering of sea level causing subaerial Fränkl, E. 1953. Geologische Untersuchungen in Ost-Andrees Land
exposure and subsequent dissolution of the shallow marine sea (NE-Grønland). Meddelelser om Grønland, 113, 160.
bed in the region. The subsequent deepening (see ‘Sequence stra- Frederiksen, K. S. 2000. Evolution of a Late Proterozoic carbonate ramp
tigraphy’ section) could thus be a result of the rapid melting of the Ymer Ø and Andrée Land Groups, Eleonore Bay Supergroup, East
ice sheet producing fine- to very fine-grained sedimentation. Greenland: response to relative sea-level rise. Polarforschung, 68,
125– 130.
Frederiksen, K. S., Craig, L. E. & Skipper, C. B 1999. New observations
of the stratigraphy and sedimentology of the Upper Proterozoic
Palaeogeographical setting Andrée Land Group, East Greenland: supporting evidence for a
drowned carbonate ramp. Danmarks og Grønlands Geologiske
The transition from the upper Andrée Land Group to Bedgroup 19 Undersøgelse Rapport, 1999/19, 145– 158.
and further to the Tillite Group has also been interpreted as depo- Gee, D. G., Fossen, H., Henriksen, N. & Higgins, A. K. 2008.
sition in an early rift basin related to the splitting of Rodinia and From the Early Paleozoic platforms of Baltica and Laurentia
before the opening phase of the Iapetus Ocean (Fig. 56.2; Herring- to the Caldonide Orogen of Scandinavia and Greenland. Episodes,
ton & Fairchild 1989; Stouge et al. 2001; Eyles & Januszczak 31, 1 – 8.
2004; Nystuen et al. 2008; Sønderholm et al. 2008). According Green, J. W., Knoll, A. H., Golubic, S. & Swett, K. 1987. Paleobio-
to the interpretation of Eyles & Januszczak (2004), the upper logy of distinctive benthic microfossils from the Upper Proterozoic
part of the Andrée Land Group and the main part of the Tillite limestone-dolomite ‘Series’, central East Greenland. American
Group accumulated during active rifting, and deep-water sedi- Journal of Botany, 74, 928– 940.
ments accumulated in oceanic settings bordered by glaciated con- Green, J. W., Knoll, A. H. & Swett, K. 1988. Microfossils from oolites
tinents. Upwards, the uppermost part of the Andrée Land Group and pisolites of the Upper Proterozoic Eleonore Bay Group, central
and the top of the Tillite Group became shallower and the regres- East Greenland. Journal of Paleontology, 62, 835– 852.
sive top of the Andrée Land Group is marked by a prominent Green, J. W., Knoll, A. H. & Swett, K. 1989. Microfossils from silici-
dissolution surface, which is overlain by marine sandstone. fied stromatolitic carbonates of the Upper Proterozoic Limestone-
The palaeomagnetic signal supports rifting and the development Dolomite ‘Series’, central East Greenland. Geological Magazine,
of an extension-related basin and the rapid drift of the region from 126, 567– 585.
low to higher southern latitudes (Mac Niocaill et al. 2004, 2008). A Haller, J. 1953. Geologie und Petrographie von West-Andrées Land
und Ost-Frænkels Land (NE-Grönland). Meddelelser om Grønland,
Cryogenian age for the onset of rifting as potentially recorded in
113, 196.
the East and North-East Greenland succession is earlier than that
Haller, J. 1971. Geology of the East Greenland Caledonides. Inter-
initially proposed for the region by previous authors (i.e. during science Publishers, New York.
the Ediacaran; e.g. Soper 1994; Higgins & Leslie 2000). An Halverson, G. P. 2011. Glacial sediments and associated strata of the
additional and more significant problem with the rift-basin Polarisbreen Group, northeastern Svalbard. In: Arnaud, E.,
model is that no evidence or structures such as rift-related dolerite Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
dykes or volcanic or pyroclastic deposits have been observed in the Record of Neoproterozoic Glaciations. Geological Society, London,
region, which might suggest a rifting stage (Soper 1994; Higgins & Memoirs, 36, 571–580.
Leslie 2008; Sønderholm et al. 2008). One explanation could Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
be that the eastern Laurentian margin in East and North-East Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
Greenland was, in fact, located in a cratonward position relative isotope record. Geological Society of America, Bulletin, 117,
to the rift zone. 1181– 1207.
NEOPROTEROZOIC NE GREENLAND 591

Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan Kalsbeek, F., Higgins, A. K. & Jepsen, H. F. 2008. Granites and granites
glaciation (Neoproterozoic) in northeast Svalbard. Basin Research, in the East Greenland Caledonides. 227– 249. In: Higgins, A. K.,
16, 297– 324. Gilotti, J. A. & Smith, M. P. (eds) The Greenland Caledonides:
Hambrey, M. J. 1983. Correlation of late Proterozoic tillites in the Evolution of the Northeast Margin of Laurentia. Memoir, Geological
North Atlantic region and Europe. Geological Magazine, 120, Society of America, 202, 369.
290– 320. Katz, H. R. 1952. Zur Geologie von Strindbergs Land (NE-Grönland).
Hambrey, M. J. 1989. The Late Proterozoic sedimentary record of East Meddelelser om Grønland, 111, 150.
Greenland: its place in understanding the evolution of the Caledonide Katz, H. R. 1954. Einige Benmerkungen zur Lithologie und Stratigraphie
Orogen. In: Gayer, R. A. (ed.) The Caledonide Geology of Scandina- der Tillitprofile in Gebiet des Kejser Franz Josephs Fjord, Ostgrön-
via. Graham & Trotman, London, 352, 257– 262. land. Meddelelser om Grønland, 72, 63.
Hambrey, M. J. & Spencer, A. M. 1987. Late Precambrian glaciation Katz, H. R. 1961. Late Precambrian to Cambrian stratigraphy in East
of central East Greenland: Meddelelser om Grønland, Geoscience, Greenland. In: Raasch, G. O. (ed.) Geology of the Arctic. Vol. 1.
19, 50. University of Toronto Press, Toronto, 299– 328.
Hambrey, M. J. & Harland, W. B. 1995. The Late Proterozoic glacial Kaufman, A. J., Knoll, A. H. & Narbonne, G. M. 1997. Isotopes, ice
era. Palaeogeography, Palaeoclimatology, Palaeoecology, 51, ages, and terminal Proterozoic earth history. Proceedings of the
255– 272. National Academy of Science, 94, 6600– 6605.
Hambrey, M. J., Peel, J. S. & Smith, M. P. 1989. Upper Proterozoic and Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffman, K. H. &
Lower Palaeozoic in northern East Greenland. Rapport Grønlands Arthur, M. 1998. Two or four Neoproterozoic glaciations?
Geologiske Undersøgelse, 145, 103– 108. Geology, 26, 1059– 1063.
Harland, W. B. 1997. The Geology of Svalbard. Geological Society of Kilner, B. R. 2005. Low latitude Neoproterozoic glaciation: palaeomag-
London, Memoirs, 17, 521. netic investigations of the Precambrian sequences of Oman and East
Harland, W. B. & Gayer, R. A. 1972. The Arctic Caledonides and Greenland. Unpublished D.Phil thesis, University of Oxford.
earlier oceans. Geological Magazine, 109, 289–314. Knoll, A. H. 2000. Learning to tell Neoproterozoic time. Precambrian
Henriksen, N. 1999. Conclusion of the 1 : 500000 mapping project in the Research, 100, 3– 20.
Caledonian fold belt in North-East Greenland. Geology of Greenland Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K. & Lambert, I. B.
Survey Bulletin 183, 10– 22. 1986. Secular variation in carbon isotope ratios from Upper Proterozoic
Henriksen, N. 2003. Caledonian orogen East Greenland 708– 828N, litho- successions of Svalbard and East Greenland. Nature, 321, 832–838.
tectonic map 1: 1 000000. Geological Survey of Denmark and Green- Koch, L. 1929. The geology of East Greenland. Meddelelser om Grøn-
land, Copenhagen. land, 73, II(I), 204.
Henriksen, N. & Higgins, A. K. 1976. East Greenland Caledonian fold Koch, L. 1930. Die technische Entwicklung Grönlands. Geologische
belt. In: Escher, A. & Watt, W. S. (eds) Geology of Greenland. Rundschau, 21, 345– 347.
Geological Survey of Greenland, Copenhagen, 182–246. Koch, J. & Haller, J. 1971. Geological map of East Greenland 72 –76
Henriksen, N. & Higgins, A. K. 2008. Geological research and N. Lat. (1 : 250.000). Meddelelser om Grønland, 183, 26, 13 maps.
mapping in the Caledonian orogen in East Greenland, 708N–828N. Kulling, O. 1929. Stratigraphic studies of the geology of North-east
1 – 28. In: Higgins, A. K., Gilotti, J. & Smith, P. M. (eds) The Greenland. Meddelelser om Grønland, 74, 317–346.
Greenland Caledonides: Evolution of the Northeast Margin of Laur- Mac NioCaill, C., Stouge, S., Harper, D. A. T., Christiansen, J.,
entia. Memoir Geological Society of America, 202, 368. Kilner, B., Johnson, A. & Watts, C. 2004. Preliminary paleomag-
Herrington, P. M. & Fairchild, I. J. 1989. Carbonate shelf and slope netic results from the late Neoproterozoic of eastern Greenland: A
facies evolution prior to Vendian glaciation, central East Greenland. low-latitude Sturtian glaciation? EOS Transactions of the American
In: Gayer, R. A. (ed.) The Caledonide Geology of Scandinavia. Geophysical Union, 85, 165.
Graham & Trotman, London, 263–273. Mac Niocaill, C., Kilner, B., Stouge, S., Knudsen, M. F., Harper,
Higgins, A. K. 1981. The Late Precambrian Tillite Group of the Kong D. A. T. & Christiansen, J. L. 2008. The Neoproterozoic drift
Oscars Fjord and Kejser Franz Josefs Fjord region of East Greenland. history of Laurentia: a critical evaluation and new palaeomagnetic
In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene data from Northern and Eastern Greenland. EOS Transactions of
Glacial Record. Cambridge University Press, Cambridge, 778– 781. the American Geophysical Union, 89, S514– 05.
Higgins, A. K. & Leslie, A. G. 2000. Restoring thrusting in the East Manby, G. M. & Hambrey, M. J. 1989. The structural setting of the Late
Greenland Caledonides. Geology, 28, 1019– 1022. Proterozoic tillites of East Greenland. In: Gayer, R. A. (ed.) The
Higgins, A. K. & Leslie, A. G. 2008. Architecture and evolution of the Caledonide Geology of Scandinavia. Graham & Trotman, London,
East Greenland Caledonides – an introduction. In: Higgins, A. K., 352, 299–312.
Gilotti, J. A. & Smith, M. P. (eds) The Greenland Caledonides: Moncrieff, A. C. M. 1989. The Tillite Group and related rocks of East
Evolution of the Northeast Margin of Laurentia. Geological Society Greenland: implications for Late Proterozoic palaeogeography. In:
of America, Memoir, 202, 369, 29– 53. Gayer, R. A. (ed.) The Caledonian Geology of Scandinavia.
Higgins, A. K, Leslie, A. G. & Smith, M. P. 2001. Neoproterozoic – Graham and Trotman, London, 352, 285–297.
Lower Palaeozoic stratigraphical relationships in the marginal Moncrieff, A. C. M. & Hambrey, M. J. 1988. Late Precambrian
thin-skinned thrust belt of the East Greenland Caledonides: compari- glacially-related grooved and striated surfaces in the Tillite Group
sons with the foreland of Scotland. Geological Magazine, 138, of Central East Greenland. Palaeogeography, Palaeoclimatology,
143– 160. Palaeoecology, 65, 183–200.
Higgins, A. K., Elvevold, S. et al. 2004. The foreland-propagating Moncrieff, A. C. M. & Hambrey, M. J. 1990. Marginal-marine glacial
thrust sheet architecture of the East Greenland Caledonides 728 – sedimentation in the late Precambrian succession of East Greenland.
758N. Journal of the Geological Society (London), 161, 1009– 1026. In: Dowdeswell, J. A. & Scourse, J. D. (eds) Glacimarine Environ-
Higgins, A. K., Gilotti, J. A. & Smith, M. P. (eds) 2008. The Greenland ments: Processes and Sediments. Geological Society, London,
Caledonides: Evolution of the Northeast Margin of Laurentia. Geo- Special Publication, 53, 397– 410.
logical Society of America, Memoir, 202, 369. Nystuen, J. P., Andresen, A., Kumpulainen, R. A. & Siedlecka, A.
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: 2008. Neoproterozoic basin evolution in Fennoscandia, East Green-
testing the limits of global change. Terra Nova, 14, 129– 155. land, and Svalbard. Episodes, 31, 35 – 43.
Kalsbeek, F., Thrane, K., Nutman, A. P. & Jepsen, H. F. 2000. Late Phillips, W. E. A, Stillman, C. J., Friderichsen, J. D. & Jemelin, L.
Mesoproterozoic to early Neoproterozoic history of the East Green- 1973. Preliminary results of mapping in the western gneiss and
land Caledonides: evidence for Grenvillian orogenesis? Journal of schist zone around Vestfjord and inner Gåsefjord, south-west Scor-
the Geological Society (London), 157, 1215– 1225. esby Sund. Rapport Grønlands Geologiske Undersøgelse, 58, 17– 32.
Kalsbeek, F., Jepsen, H. F. & Jones, K. A. 2001. Geochemistry and Porter, S. M. & Knoll, A. H. 2000. Testate amoebae in the Neoproter-
petrogenesis of S-type granites in the East Greenland Caledonides. ozoic era: evidence from vase-shaped microfossils in the Chuar
Lithos, 57, 91 – 109, doi: 10.1016/S0024–4937(01)00038-X. Group, Grand Canyon. Paleobiology, 26, 360–385.
592 S. STOUGE ET AL.

Porter, S. M., Meisterfeld, R. & Knoll, A. H. 2003. Vase-shaped North-East Greenland. Bulletin Grønlands Geologiske Undersøgelse,
microfossils from the Neoproterozoic Chuar Group, Grand Canyon: 167, 38
a classification guided by modern testate amoebae. Journal of Sønderholm, M., Frederiksen, K. S., Smith, M. P. & Tirsgaard, H.
Paleontology, 77, 409–429. 2008. Neoproterozoic sedimentary basins with glacigenic deposits
Poulsen, C. 1930. Contributions to the stratigraphy of the Cambro – Ordo- of the East Greenland Caledonides. In: Higgins, A. K., Gilotti, J.
vician of East Greenland. Meddelelser om Grønland, 74, 297–316. A. & Smith, M. P. (eds) The Greenland Caledonides: Evolution of
Robb, L. J., Knoll, A. H., Plumb, K. A., Shields, G. A., Strauss, H. & the Northeast Margin of Laurentia. Memoir, Geological Society of
Veizer, J. 2004. The Precambrian: the Archaean and Proterozoic America, 202, 99– 136.
eons. In: Gradstein, F. M., Ogg, J. G. & Smith, A. G. (eds) A Geo- Swett, K. & Knoll, A. H. 1989. Marine pisolites from Upper Proterozoic
logic Time Scale 2004. Cambridge University Press, Cambridge, carbonates of East Greenland and Spitsbergen. Sedimentology, 36,
129– 140. 75 – 93.
Schaub, H. P. 1950. On the pre-Cambrian sedimentation in NE- Teichert, C. 1933. Untersuchungen zum Bau des kaledonischen Gebirges
Greenland. Meddelelser om Grønland, 114, 50. in Ostgrönland. Meddelelser om Grønland, 95, 121.
Schaub, H. P. 1955. Tectonics and morphology of Kap Oswald Tirsgaard, H. 1993. The architecture of Precambrian high energy tidal
(NE-Greenland). Meddelelser om Grønland, 103, 1– 33. channels: an example from the Lyell Land Group (Eleonore Bay
Smith, M. P. & Robertson, S. 1999. Vendian – Lower Palaeozoic strati- Supergroup), East Greenland. Sedimentary Geology, 88, 137–152.
graphy of the parautochthon in the Målebjerg and Eleonore Sø Tirsgaard, H. 1996. Cyclic sedimentation of carbonate and siliciclastic
windows, East Greenland Caledonides. Danmarks og Grønlands deposits on a late Precambrian ramp: the Elisabeth Bjerg Formation
Geologiske Undersøgelse Rapport, 1999/19, 169– 182. (Eleonore Bay Supergroup), East Greenland. Journal of Sedimentary
Smith, M. P., Rasmussen, J. A., Robertson, S., Higgins, A. K & Leslie, Research, 66B, 699– 712.
A. G. 2004. Lower Palaeozoic stratigraphy of the East Greenland Tirsgaard, H. & Sønderholm, M. 1997. Lithostratigraphy, sedimentary
Caledonides. Geological Survey of Denmark and Greenland Bulletin, evolution and sequence stratigraphy of the Upper Proterozoic Lyell
6, 5 –28. Land Group (Eleonore Bay Supergroup) of East and North-East
Sommer, M. 1957. Geologie von Lyells Land (NE-Grönland). Meddelel- Greenland. Geology of Greenland Survey Bulletin, 178, 60.
ser om Grønland, 155, 157. Vidal, G. 1976. Late Precambrian acritarchs from the Eleonore Bay
Soper, N. J. 1994. Neoproterozoic sedimentation on the NE margin of Group and Tillite Group in East Greenland. Grønlands Geologiske
Laurentia and the opening of Iapetus. Geological Magazine, 131, Undersøgelse Rapport, 78, 19.
291– 299. Vidal, G. 1979. Acritarchs from the Upper Proterozoic and Lower
Stouge, S., Boyce, W. D., Christiansen, J. L., Harper, D. A. T. & Cambrian of East Greenland. Grønlands Geologiske Undersøgelse
Knight, I. 2001. Vendian – Lower Ordovician stratigraphy of Ella Rapport, 134, 40.
Ø, North-East Greenland: new investigations. Geology of Greenland Vidal, G. 1985. Biostratigraphic correlation of the Upper Proterozoic and
Survey Bulletin, 189, 107–114. Lower Cambrian of the Fennoscandian Shield and the Caledonides of
Stouge, S., Boyce, W. D., Christiansen, J. L., Harper, D. A. T. & East Greenland and Svalbard. In: Gee, D. G. & Sturt, B. A. (eds)
Knight, I. 2002. Lower –Middle Ordovician stratigraphy of North- The Caledonide Orogen – Scandinavia and Related Areas. John
East Greenland. Geology of Greenland Survey Bulletin, 191, Wiley and Sons Ltd., London, 331– 338.
117– 125. Wenk, E. 1961. On the crystalline basement and the basal part of the pre-
Sønderholm, M. & Tirsgaard, H. 1993. Lithostratigraphic framework Cambrian Eleonore Bay Group in the southwestern part of Scoresby
of the Upper Proterozoic Eleonore Bay Supergroup of East and Sund. Meddelelser om Grønland, 168, 54.
Chapter 57

Glaciogenic rocks of the Neoproterozoic Smalfjord and Mortensnes formations,


Vestertana Group, E. Finnmark, Norway

A. HUGH N. RICE1 *, MARC B. EDWARDS2, TOR A. HANSEN3, EMMANUELLE ARNAUD4 & GALEN P. HALVERSON5,6
1
Vienna University, Structural Processes Group, Department of Geodynamics and Sedimentology, Althanstrasse 14,
1090 Vienna, Austria, Europe
2
5430 Dumfries Drive, Houston, Texas, USA
3
Talisman Energy Norge AS, Verven 4, PO Box 649, Stavanger, Norway
4
Department of Land Resource Science, University of Guelph, Guelph, Ontario, N1G 2W1, Canada
5
Department of Geology & Geophysics, School of Earth & Environmental Sciences, University of Adelaide,
North Terrace, Adelaide, SA 5005, Australia
6
Present address: Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montreal,
Quebec, H3A 2A7, Canada
*Corresponding author (e-mail: alexander.hugh.rice@univie.ac.at)

Abstract: The Vestertana Group in East Finnmark, North Norway, contains two Neoproterozoic glaciogenic sequences, the Smalfjord
and Mortensnes formations, preserved on the northern edge of Baltica. The former comprises up to 420 m of aeolian, fluvioglacial and
glaciomarine sediments and terrestrial diamictite. The latter consists of up to 50 m of predominantly diamictite. The Smalfjord Formation
(Fm.) is underlain by dolostones (Grasdalen Fm., Tanafjorden Group), only locally preserved due to the sub-Smalfjord Fm. unconfor-
mity, which cuts down-section through a c. 2.5 km dominantly clastic sequence to rest on Baltic Shield gneisses. The two glaciogenic
successions are separated by c. 350 m of mostly clastic sediments (Nyborg Fm.), with thin dolostones at the base and towards the top. The
latter are generally absent due to the sub-Mortensnes Fm. unconformity, which also cuts down southwards through the Nyborg and
Smalfjord formations to the Baltic Shield. No robust isotopic age constraints are available for the succession. d13C data, together
with cap dolostone characteristics, offer paradigmic correlations with other areas (Smalfjord ; Marinoan; Mortensnes ; Gaskiers).
A limited Ediacaran fauna, including Aspidella, give only broad age constraints. Palaeomagnetic data are ambiguous; some suggest
Baltica lay at equatorial (158S) to mid-latitudes (508S) for the period 750– 550 Ma, respectively, while other interpretations place it
at either 308N or S at c. 550 Ma.

This chapter is concerned with the glaciogenic sediments of the the Varanger Epoch (cf Harland et al. 1989). Use of the term
Smalfjord and Mortensnes formations (Lower and Upper Tillites, Varang(er)ian is here discouraged for reasons given below.
respectively, in older literature), lying at the base of the Vestertana
Group, within a c. 4.7-km-thick, predominantly clastic sequence of
Tonian/Cryogenian to Tremadocian age. The former glaciogenic Structural framework
unit only occurs in three areas, all in E Finnmark; Laksefjordvidda,
Tanafjord and Varangerfjord (Fig. 57.1). In contrast, the latter unit During the Neoproterozoic in northern Scandinavia, the NNE –
has tentatively been correlated with many other diamictite-bearing SSW-trending western Baltica continental shelf, now preserved
units in the Scandinavian Caledonides (Kumpulainen 2011; Kum- in the Scandinavian Caledonides, was joined to the WNW –ESE-
pulianen & Greiling 2011; Nystuen & Lamminen 2011; Stodt et al. trending Timan Basin in NE Baltica (Siedlecka 1985; Gayer &
2011). Owing to significant facies variations, no single outcrop/ Rice 1989; Siedlecka et al. 2004). Both of these margins are
area can be taken as a type locality/region. thought to have experienced extensional rift tectonics from
The rocks have been studied by many authors (see references in c. 1000 to c. 630 Ma, associated with the break-up of Rodinia, fol-
Føyn 1937; Bjørlykke 1967; Føyn & Siedlecki 1980; Laajoki lowed by relatively quiescent passive margin conditions during the
2002), but detailed sedimentological studies remain few deposition of the two glaciogenic units (Røe 2003; Siedlecka et al.
(Reading & Walker 1966; Edwards 1972, 1975, 1979, 1984; 2004). The Timan Basin has been divided into northerly basinal
Hansen 1992; Arnaud & Eyles 2002; Baarli et al. 2006; Levine and southerly shelf regions (Siedlecka et al. 1995), with the
et al. 2006; Arnaud 2008). An overview of d13C data from associ- WNW – ESE-trending dextral strike – slip Trollfjorden – Koma-
ated dolostones was given by Halverson et al. (2005). gelva Fault (cf Rice et al. 1989) forming the boundary between
The Smalfjord Fm. is significant in having provided some of the these two regions (Fig. 57.1). The Smalfjord Fm. was deposited
first evidence for a glacial event specifically attributed to a pre- (or has only been preserved in; but see Kumpulainen 2011) the
Pleistocene age, at Oaibaččanjar’ga (formerly Bigganjar’ga; eastern part of the Gaissa Basin, which palaeogeographically
Reusch 1891). Interpretation of this outcrop has remained a bone formed a linking area between the Timanian shelf region and the
of contention since its discovery; recent results indicate that western Baltica continental shelf. In contrast, the Mortensnes
intense high-temperature brittle deformation occurred during Fm. and its presumed correlatives were deposited and preserved
rapid glacial movement (Bestmann et al. 2006). over much of western Baltica.
The two glaciogenic sequences became an informal ‘type The rocks were shortened predominantly during Silurian-
sequence’ for late Precambrian glaciations in the North Atlantic Devonian Caledonian orogenesis (Scandian event). Deformation
region (e.g. Hambrey 1983), known as the Varang(er)ian was essentially in-sequence, with c. 50% shortening in the external
glaciation (Varangeristiden; Kulling 1951), later developed into imbricate zone (Gaissa Thrust Belt; imbricated Gaissa Basin

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 593– 602. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.57
594 A. H. N. RICE ET AL.

Fig. 57.1. Map of NE Scandinavian Caledonides showing the main localities and distribution of glaciogenic lithologies. A, Alduskai’di; An, Andabakoai’vi;
C, Čikkojåkka; D, Digermul; G, Gaessenjar’ga; Gr, Grasdal; Gu, Gulgofjord; H, Handelsneset; L, Lappaluokoai’vi; Lp, Leirpollen; Lv, Laksefjordvidda; M, Mortensnes;
Mi, Miel’kejåkka; Mn, Manjunnas; N, Nyborg; Ne, Nesseby; Nj, Njukčagai’sa; O, Oaibaččannjar’ga; R, Ruos’soai’vi; S, Selesnjar’ga; Sk, Skjåholm; Sm, Smalfjord;
St, Stappugied’di; U, Uccaskai’di; V, Vestertana; Vb, Vaððasbak’te; Vr, Vieranjar’ga.

sediments) west of Laksefjordvidda (Townsend et al. 1986). East & Siedlecki 1980) and in the intervening Tanafjord area. The
of Laksefjordvidda, shortening decreased to c. 15% (Chapman maximum preserved thickness occurs in the Krokvatn palaeoval-
et al. 1985) and dies out very gradually across Varangerhalvøya. ley (c. 420 m), but facies variations preclude formation-wide
In the Digermul –Tanafjord region, large-scale folds developed, member names. The Mortensnes Fm. (and equivalents; see Kum-
sometimes with a back-thrusting vergence (Reading 1965; pulainen 2011; Kumpulainen & Greiling 2011; Nystuen & Lammi-
Chapman et al. 1985; Siedlecka 1987). These likely formed by but- nen 2011; Stodt et al. 2011), which is more widespread across the
tressing, as the sole thrust attempted to footwall-shortcut through orogen, often lying unconformably on autochthonous or allochtho-
NNE –SSW-trending extensional structures. nous basement, is ,50 m thick. The two glaciogenic units are sep-
Although the strain decreases to the east, most pelitic rocks have a arated by c. 350 m of clastic and minor carbonate sediments of
cleavage, locally with crenulations of sedimentary laminae in the Nyborg Fm. (Edwards 1984). The uppermost sediments of
flexural-slip fold hinges, while diamictites may carry a spaced the Tanafjorden Group that underlie the Smalfjord Fm. are the
(approximately centimetre scale) anastamosing cleavage. Pressure- mixed carbonate and clastic sediments of the Grasdalen and Fad-
solution was common in carbonate rocks, including dolomite-rich nuvag’gi formations. The Mortensnes Fm. is overlain by mud-
diamictites. However, in sandier lithologies, finite strains are gener- stones and sandstones of the Stappogiedde Fm. (Lillevannet
ally low, with sedimentary structures well preserved. Member).
The Precambrian –Cambrian boundary lies near the base of the
overlying Breidvika Fm. (Føyn & Glaessner 1979; Farmer et al.
Stratigraphy 1992; Crimes & McIlroy 1999).
The glaciogenic sequences in E. Finnmark lie at the base of the
Vestertana Group, within the c. 4.7-km-thick, predominantly Glaciogenic deposits and associated strata
shallow marine and siliciclastic, Tonian (inferred) to Tremadocian
East Finnmark Supergroup (Fig. 57.2; Føyn 1937; Johnson et al. Grasdalen Fm., Tanafjord
1978; Siedlecka et al. 2004). The sub-Smalfjord Fm. unconformity
at the base of the Vestertana Group cuts down-section to the south The Grasdalen Fm. (230 m), which only crops out in the Grasdal-
until it lies on the Baltic Shield around Varangerfjord (Fig. 57.1). Gulgofjord area, very close to the top of the Tanafjorden Group,
Between the Tana River west of Varangerfjord and Andabakoaivi was divided into three major parts by Siedlecka & Siedlecki
(Fig. 57.1), the Vestertana Group forms the most southerly Caledo- (1971) and Johnson et al. (1978): a lower mixed sand/silt-marl-
nian Neoproterozoic rocks, although the basement –cover contact dolostone unit (111 m), a massive dolostone (58 m) and an inter-
is not exposed. bedded dolostone-clastic sequence (61 m). In detail, however,
The Smalfjord Fm. has been preserved in two major palaeoval- the stratigraphy is considerably more complex (Rice, unpublished
leys (Varanger and Krokvatn palaeovalleys; Bjørlykke 1967; Føyn data; Fig. 57.2).
NEOPROTEROZOIC SMALFJORD AND MORTENSNES FORMATIONS 595

Fig. 57.2. Regional stratigraphic profiles showing position of glaciogenic units (modified from Johnson et al. 1978; Føyn & Siedlecki 1980; Edwards 1984; Rice &
Townsend 1996; Halverson et al. 2005). The Lille Molvika Fm. and Ekkerøya group (formerly known as the Ekkerøya Fm., within the Vadsø Group) were informally
defined based on regional correlations by Rice & Townsend (1996). UE, LE, position of uppermost and lowermost Ediacaran fauna; Pl, position of Platysolenites
antiquissimus (Eichw.). Note different vertical scales for columns (a) and (c) compared to (b).

Fadnuvag’gi Fm., Tanafjord deposited within an essentially north –south-oriented palaeovalley


cut into the Tanafjorden Group on Laksefjordvidda (Fig. 57.1).
Although the Grasdalen Fm. is usually taken to continue up to the Figure 57.3a shows a semi-schematic profile across the northern
base of the Smalfjord Fm. (Siedlecka & Siedlecki 1971), the inter- part of the palaeovalley (Uccaskai’di-Vaððasbak’te); further
bedded dolostone-clastic sequence is overlain by c. 20 m of a dis- south (Čikkojokka; Fig 57.1), the sequence comprises only the
tinctly more fissile, darker, pyritiferous and finely interbanded middle diamictite and upper sandstone (Føyn & Siedlecki 1980).
shale silt/sandstone without carbonate interlayers, informally The Tanafjorden Group dips to the NW compared to the
termed the Fadnuvaggi Fm. (Halverson et al. 2005). 420-m-thick palaeovalley infill, with 600 m of tilted stratigraphy
eroded.
The Lower and Middle Krovatn Diamictites are both ,25 m
Smalfjord Fm., Krokvatn Palaeovalley, Laksefjordvidda thick, with a grey to reddish colour and a sandy matrix. The
former includes a 2-m-thick sandstone, whereas the latter has inter-
Føyn & Siedlecki (1980) documented three diamictite units with layered sandstones in the basal part and has faint banding and rare
intervening sandstones (Lower, Middle and Upper Krokvatn dropstones (Føyn & Siedlecki 1980). Clasts, sometimes faceted,
Diamictites, Lower and Upper Krokvatn Sandstones; Fig. 57.2) are predominantly from the Tanafjorden Group, with rare
596 A. H. N. RICE ET AL.

Fig. 57.3. (a) Schematic profile from Uccaskai’di to Vaððasbak’te through the glacial successions in the Krokvatn Palaeovalley (modified from Føyn & Siedlecki
1980). LKD, MKD, UKD, Lower, Middle and Upper Krokvatn Diamictites; (b) Schematic profile from Varangerfjord to Gulgofjord (modified from Edwards & Føyn
1981). Nyborg facies: N2, fan channel; N3, N4, fine and coarse submarine fans, respectively; N5, tidal distributary; N6, bay/lagoon; N7, offshore barrier. lsm, usm, lower
and upper sub-members of the Lillevannet Member.

kilometre-sized dolostone blocks at the base (Edwards et al. 1973) outcrops at Vieranjar’ga, Nesseby, Handelsneset, Skjåholm, Oai-
and some basement material. The upper diamictite (,100 m thick) baččanjar’ga and Selesnajar’ga (Edwards 1975, 1984; Arnaud &
is more variable, with dolomitic and sandy matrices and a buff or Eyles 2002; Baarli et al. 2006; Arnaud 2008).
grey colour. These are interbedded with sandstones and siltstones Two diamictite and four sandstone and conglomerate (S1– S4)
with a few lonestones. On Uccaskai’di, the Upper Krokvatn Dia- facies have been identified in this area. The first type of diamictite
mictite has abundant dolostone clasts, possibly reflecting the is a massive, matrix-supported, pink diamictite with a medium-
nearby presence of Tanafjorden Group dolostones (Fig. 57.1; to coarse-grained matrix with only a few granitic and sandstone
Føyn & Siedlecki 1980). In contrast, in the northeastern part of clasts. These are up to 20 cm across, sub-rounded to sub-angular,
the valley, the Nyborg Fm. rests unconformably on the Lower and without striae or faceting. This unit, in beds up to 1 m thick,
Krokvatn Diamictite and Lower Sandstone, the higher units can be traced over several hundred metres. The second diamictite,
having been removed either during deposition of the third diamic- which is also mostly matrix-supported and massive, with a dis-
tite elsewhere, or afterwards. tinctly grey colour in outcrop, consists of a muddy, medium
The massive, homogeneous very fine-grained sandstones and grained sandstone matrix and abundant granite and sandstone
very coarse-grained siltstones are essentially devoid of sedimen- clasts, ,50 cm across, some with faceting. Bjørlykke (1967)
tary structures throughout the succession; ripples occur at only also noted striated clasts in grey diamictite.
one locality. Beds are from 1 m to several metres thick, without The S1 facies comprises well-stratified pebble conglomerates,
sorting of the angular grains, although larger grains (.0.2 mm) occasionally imbricated, sandstones showing cross-bedding and
are rounded. Hand-sized basement lonestones occur at two parallel lamination, with fining-upwards trends between these
localities. Directly above the Middle Krokvatn Diamictite, the lithologies and rare mud-drapes. S2 is typified by large-scale
sediments are red (Føyn & Siedlecki 1980). inclined bedding, 5–10 m thick and continuous for up to several
kilometres. These include steeply dipping poorly sorted pebble
conglomerates to well-sorted gently dipping parallel-laminated
Smalfjord Fm., Varanger Palaeovalley, Varangerfjord sandstones. S3 comprises medium-bedded parallel-sided to slightly
lenticular sandstones, usually internally massive, but at times
The Smalfjord Fm. in the Varangerfjord area (Bjørlykke 1967; parallel-laminated or rippled. Mudstone partings are thin or absent
Edwards 1975, 1984; Arnaud & Eyles 2002; Arnaud 2008; but soft-sediment deformation, with convolute bedding, recum-
Laajoki 2001, 2002; Baarli et al. 2006) has been correlated bent folds and faulting occur. Steep-sided isolated channels filled
with the lower to middle part of the Krokvatn succession (Føyn with poorly sorted conglomerate are also present. S4 comprises
& Siedlecki 1980). The following descriptions are based on interbeddded sandstones and mudstones, the former in thin to
NEOPROTEROZOIC SMALFJORD AND MORTENSNES FORMATIONS 597

medium, laterally continuous beds. These display many turbiditic Edwards (1984), but ordered them differently, without any
features (erosive surfaces, sole marks, grading, convolute bedding). depositional cyclicity.
A 3-m-thick, crudely stratified breccia within S4 contains angular In the Grasdal area, Edwards (1979) described a 3 –6-m-thick
sandstone clasts ,40 cm across, sometimes graded, in a coarse- dark grey siltstone with an unusually high silt and low clay
grained sandstone matrix. content; this may also occur on the southern side of Leirpollen
The stratigraphic and lateral distribution of these facies vary (Fig. 57.1). Clasts consist of angular quartz and feldspar grains
from place to place. Laterally, lithofacies can often be traced for of very fine sand to coarse silt size, forming c. 77% of the rock,
several tens of metres, although abrupt facies changes are common with a statistically non-random preferred north –south to NE–
(cf Edwards 1984). Complex soft-sediment deformation in sand- SW orientation. This unit has sharp contacts with the over- and
stone and conglomerate facies is particularly notable at the base underlying units (the latter is a very dark diamictite with abundant
of the Smalfjord Fm. at Handelsneset (Arnaud 2008). Overall, the dolomitic clasts) and an internal stratification defined by isolated,
diamictite, breccia and S4 facies tend to occur near the base of typically dolomitic, clasts (,7.5 mm sized) in discrete layers
the sequence, whereas the S1 facies occurs near or at the top. c. 30 cm apart, together with very faint grey colour variations.

Smalfjord Fm., Tanafjord area Nyborg Fm., E. Finnmark (0– 350 m)

The Smalfjord Fm. in this region (Reading & Walker 1966; The Smalfjord and Mortensnes formations are separated by the
Edwards 1984; Hansen 1992; Arnaud & Eyles 2002) has been cor- Nyborg Fm. The lowest part of this unit (Member A) comprises
related with the Upper Krokvatn Diamictite (Laksefjordvidda lateral and vertical gradations between massive buff-weathering
region; Føyn & Siedlecki 1980; Edwards 1984). The best outcrops dolomicrites with sheet cracks and pseudo-tepees, through
in this area are those exposed at Gaessenjar’ga and Luovtat massive to thinly bedded dolomicrites to interbeddded red sand-
(Fig. 57.1). stones/shales and dolomicrites, locally reworked as edgewise
Based on outcrops in the Vestertana to Smalfjord area, with breccias, to dolomite-cemented orange-red sandstones, as well as
additional important outcrops between Njukčagai’sa and east of white-weathering green-grey sandstones and red and dark grey
the Tana River, Edwards (1984) recognized a fivefold (A –E) shales (Edwards 1984; Reading & Walker 1966; Rice &
repetitive cycle of erosion surface, diamictite and laminated mud- Hofmann 2001). Such sequences are usually only few tens of
stone, occasionally with intervening sandstones. Some units can be metres thick (Reading & Walker 1966; Edwards 1984) although
traced over hundreds of square kilometres. Diamictite units are from the succession is c. 160 m thick at Alduskaidi (Fig. 57.1; Føyn &
2–40 m thick, structureless or stratified, and have erosive bases, Siedlecki 1980). Generally, successions become more clastic
at both outcrop and regional scales. Diamictites are all upwards. On Ruos’soai’vi, west of Varangerfjord, 10-mm sulphate
matrix-supported, with clasts up to 1 m in diameter (most are crystal fans (now pseudomorphed by quartz) grew on the base-
,35 cm) in a siltstone to sandstone matrix. Material from underly- ment, preserved within a 12-mm-thick biotite extraclast recrystal-
ing beds was often incorporated into the overlying diamictite, lized dolostone overlain by white-weathering sandstones. Similar
forming inclusions of diamictite, conglomerate, sandstone or mud- white-weathering sandstones, likely cannibalized from the Smal-
stone, folds and faults, load-casts, boudinage, shear bands and fjord Fm., also directly overlie very irregular basement surfaces on
flame structures. Some diamictite units have gradational contacts Ruos’soai’vi. White-weathering green-grey sandstones (18.5 cm
(Hansen 1992; Arnaud & Eyles 2002). Stratified sandstone/con- thick) also unconformably overlie c. 14 m of massive sheet-
glomerate bodies are present within the diamictite; these deposits cracked dolomicrite in the Miel’kejåkka area, SW of Tanafjord
resemble the facies in the Varangerfjord area (see above). In some (Fig. 57.1). Members B –D of the Nyborg Fm. comprise, respect-
cases, the basal diamictite represents mixing of far-travelled and ively, interbedded reddish-purple shales/sandstones (c. 200 m),
locally derived material. The dominant clast and matrix source interbedded grey-green shales/sandstones (c. 150–200 m) and
areas determine the diamictite lithology; buff and brown colours purple sandstone/grey-green shales (c. 70 m; Edwards 1984).
indicate a dolostone source (cycles B, C and E). Cycle A is purple, Member E (c. 25 m thick), exposed only in the Gulgofjord-Grasdal
with a hematitic matrix, and was derived from ferruginous sand- area (Fig. 57.1), consists of white to grey sandstones with two thin
stones of the Dakkovarre Fm. (Fig. 57.2). Diamictites are also (c. 1 m thick) buff-weathering siliclastic-extraclast dolomicrites at
often distinguished by variable clast abundance. Some diamictite the base (Edwards 1984).
units exhibit normal grading with a sandier matrix at the base and
muddier matrix at the top (Hansen 1992; Arnaud & Eyles 2002).
The Smalfjord Fm. in the Tanafjord area has a much higher Mortensnes Fm., E. Finnmark (,50 m)
proportion of diamictite and is laterally much more consistent
compared to outcrops and lithofacies distributions in the Varan- Edwards (1984) recognized three members in this formation, the
gerfjord area; in the Gaessenjar’ ga area, over 85% of the sections lowest (,30 m) being a northwards-thinning wedge of predomi-
consist of diamictite units and these tend to be traceable over nantly grey-green to purple massive diamictite (depending on
several kilometres (Edwards 1984; Arnaud & Eyles 2002). One the substrate colour) of highly variable clast/matrix composition,
exception is that of a 1–5-m-thick purple diamictite that grades with basement clasts and minor intrabasinal clasts from the under-
into a laminated mudstone with outsized clasts within several lying substrate. This dies out around the southern end of Vester-
hundred metres of lateral exposure at Gaessenjar’ga (Arnaud & tana. Brecciation of the substrate has been found below some
Eyles 2002). diamictites, with substrate blocks up to 20 m long and 1 m thick
Sandstone beds are massive, graded (normal or reverse), rippled incorporated into the overlying diamictite. Lenses and bands of
and cross-bedded (trough and planar). Laminated mudstones, 0.3– extra-basinal dark diamictite occur within more mixed lithologies.
10 m thick, vary in colour, lamination prominence, whether this is Basic clasts may show facets and striations (Banks et al. 1971),
random or rhythmic, and size, abundance and composition of lone- while clast sizes decrease overall to the north (Edwards 1984).
stones. The latter are predominantly dolostones, ,1–30 cm in size The middle member has a gradational contact to the underlying
and are either dispersed throughout the lithology or lie in discrete rocks and is distinguished from it by its buff-brown weathering,
layers. Some laminated mudstone units are interbedded with sand- dolomitic composition of matrix and clasts with subordinate
stone or diamictite. chert. This member rapidly increases in thickness from ,4 m in
In the Gaessenjar’ga area, Hansen (1992) and Arnaud & Eyles the south to .10 m, approximately coincident with the southern
(2002) recognized essentially the same lithologies and units as end of Vestertana, giving two sub-members. The thin sub-member
598 A. H. N. RICE ET AL.

comprises 2– 4 m of stratified diamictite with primary and soft- Chemostratigraphy


sediment deformation structures. The thick sub-member comprises
five lithofacies; a blanket of massive purple to grey-green defor- d13C data from E. Finnmark were briefly reviewed by Halverson
mation diamictite, a zone of large tabular blocks (of the Nyborg et al. (2005); although more data are now available the results
Fm., diamictite and white sandstone), a relatively rare stratified dolo- are essentially the same as those previously reported. All data
mitic diamictite, a prominent buff-brown diamictite with a sandy are standardized to VPDB; see Halverson et al. (2005) for
matrix, which thins from 20 to 8 m from south to north, and, at the analytical methods.
top, a bedded buff brown diamictite. The unconformably overlying The Grasdalen Fm. has variable values, passing from –2 to
upper member (,40 m) consists of dark grey massive diamictite –3‰ at the base, gradually rising to þ6‰ and then falling very
overlain north of Stappugied’di and Leirpollen by a 40-cm-thick rapidly to – 3‰; the two stratigraphically highest samples record
dolomitic matrix diamictite with small dolostone and occasional values of þ3 to þ4‰.
large basement clasts. A 20–30-cm-thick polymict conglomerate Ten profiles through the dolostone (or interlayered dolostones–
overlies the formation over a large area (Edwards 1984). shales/siltstones/sandstones) sequences lying directly above the
Smalfjord Fm. have been analysed over a wide geographical
area and include the c. 160-m-thick dolostone at Alduskaidi
Lillevannet Member, Stappogiedde Fm. (40 – 110 m) (Fig. 57.1; Føyn & Siedlecki 1980). Although few profiles
expose both the bottom and top of the dolomite unaffected by
The Mortensnes Fm. is overlain by the Lillevannet Member of the either tectonic or penecontemporaneous erosional processes, all
Stappogiedde Fm. (Reading & Walker 1966), which thins from samples show d13C values between –1.0 and – 6.0‰ (VPDB).
south to north, approximately coincident with the southern end Most have essentially constant d13C values or show either a
of Vestertana (Edwards 1984). The lower sub-member (3 –55 m) slight increase or decrease in concentration. Only the 14-m-thick
comprises grey, parallel laminated mudstones, silty to sandy in profile at Miel’kejåkka (Fig. 57.1) shows a major variation,
the north, grading upwards into a siltstone with some ripple cross- passing from – 2.55‰ at the base to –5.92‰ at the erosive
lamination and with fine- to medium-grained lenticular sandstones upper contact. These data show a poor concave upwards distri-
(Edwards 1984). The upper sub-member is a complex assemblage bution, comparable to the uppermost part of the slope-bank d13C
of sandstone and shale facies, including coarse arkosic sandstones, profile documented by Hoffman et al. (2007).
poorly cross-bedded and granule conglomerates, medium-grained Thin dolostones in Member E of the Nyborg Fm., lying c. 20 m
sub-arkosic sandstone, relatively well sorted and rounded sand- below the sub-Mortensnes unconformity in the Gulgofjord area
stones, thin to medium bedded fine to very fine lenticular, erosive- (Edwards 1984; Figs 57.1 & 57.2) gave d13C values between
based sandstones, dark grey, brown weathering rippled and finely –7.6 and – 9.9‰ (VPDB). Significantly, the thin carbonate-matrix
laminated micaceous silty-sandy mudstone, sometimes in diamictite forming the uppermost part of the Mortensnes Fm. on
coarsening-upwards cycles and finely parallel laminated grey the NW side of Tanafjord (Edwards 1984) also gave extreme nega-
mudstones (Edwards 1984). tive values (down to – 10.44‰ VPDB), taken to reflect erosion of
the underlying dolostones of Member E in the Nyborg Fm. (Rice &
Halverson, unpublished data).
Boundary relations with overlying and underlying
non-glacial units
Palaeolatitude and palaeogeography
Southwards, the Smalfjord Fm. cuts down through up to 2.5 km of
the Tanafjorden, Ekkerøya and Vadsø Groups, to lie on the base- Palaeomagnetic analyses on the Nyborg Fm. have yielded palaeo-
ment around Varangerfjord. Minor outliers of Neoproterozoic latitudes for this area of 338S to 418S (Torsvik et al. 1995),
sediments overlie the basement, and minor inliers of basement although no robust geochronological constraint is associated
occur within the Smalfjord Fm., together preserving an irregular with this palaeolatitude. Samples were subjected to thermal
unconformity, locally palaeo-frost-shattered (Bjørlykke 1967; demagnetization and some also to alternating field demagnetiza-
Siedlecka 1990; Rice & Hofmann 2000; Rice et al. 2001; Laajokki tion. High-temperature components from two field areas (55
2001, 2002; Edwards 1984, 1975). The base of the Smalfjord Fm. samples) were identified with a positive fold test at 95% signifi-
is characterized by rare east –west to WNW –ESE oriented sub- cance level (k ¼ 4.9 in situ and 14.6 100% unfolded; a95 ¼
glacial striations; the best developed occur at Oaibaččanjar’ga 33.8 in situ and 18.1 100% unfolded, Torsvik et al. 1995).
(Bigganjar’ga) in Varangerfjord (Reusch 1891; Strahan 1897), The palaeogeography of Baltica as a whole has been constrained
although others have been reported (Bjørlykke 1967; Rice & from other sites in Baltica, although well-dated palaeopoles remain
Hoffman 2000; Laajoki 2002). Rice & Hofmann (2000) and Best- few (see reviews in Bingen et al. 2005, Cocks & Torsvik 2005). At
mann et al. (2006) documented a thin glacially formed breccia c. 750 Ma, a southwards facing Finnmark-Kola region at c. 158S
within and as ridges beside the striations at Oaibaččanjar’ga. formed the southern margin of Baltica (Hartz & Torsvik 2002).
Regionally, the upper contact of the Smalfjord Fm. with the This region lay adjacent to a roughly east –west-trending rift
overlying Nyborg Fm. is also an unconformity, because the latter (Timanian margin) that to the west linked with the north –south
rests on different parts of the former in different areas; specifi- trending rift/spreading axis developing between Baltica and Laur-
cally, the upper part of the Smalfjord Fm. in the Vaððasbak’te entia (Greenland). At 616 Ma, Baltica is thought to have been at
area of the Krokvatn Palaeovalley and in the Varanger palaeoval- polar latitudes (758S) based on the well-dated Egersund dolerite
ley are absent (Fig. 57.1; Føyn & Siedlecki 1980; Edwards 1984). dykes in southern Norway (Bingen et al. 2005). By c. 550 Ma,
In the Ruos’soai’vi-Lap’paluokoai’vi area (Fig. 57.1), the Nyborg E. Finnmark was lying at c. 508S (Cocks & Torsvik 2005). In con-
Fm. rests directly on the basement (Siedlecka 1990; Rice unpub- trast, Cawood & Pisarevsky (2006) and Pisarevsky et al. (2008)
lished data). place Baltica in a more equatorial position (c. 308) although
The base of the Mortensnes Fm. is an invariably planar surface whether it lay in the northern or southern hemisphere is unclear.
at outcrop-scale that cuts down-section towards the south at a very
low angle. The contact often exhibits brecciation or homogeniz-
ation of the underlying sediments (deformation diamictite; Other characteristics
Edwards 1984). At its upper contact, the Mortensnes Fm. is over-
lain with a sharp or rapid transition by the Lillevannet Member No economic deposits have been reported in these rocks. Acri-
(Fig. 57.2; Reading & Walker 1966; Edwards 1984). tarchs have been documented from the Vadsø, Ekkerøya,
NEOPROTEROZOIC SMALFJORD AND MORTENSNES FORMATIONS 599

Tanafjorden and Vestertana Groups, although only reworked acri- resulted in the formation of deformation tillites, comprised solely
tarchs and vase-shaped fossils have so far been found in the Smal- of deformed substrate sediments, while mixing of local and far-
fjord and Mortensnes formations (Vidal 1981; Vidal & Siedlecka travelled material formed banded tillites. Erosional bases of the
1983). A limited Ediacaran fauna (Farmer et al. 1992; Crimes & cycles are recorded at the outcrop-scale by deformation structures
McIlroy 1999), dominated by discoidal forms (cf Gehling et al. and mixing with subjacent material and on a regional scale by the
2000) and including Aspidella (Narbonne, pers. comm. 2008), absence of stratigraphy below diamictites. During retreat, poorly
has been documented, the oldest forms occurring in the Innerelv sorted sand and silt with pebbles were deposited at the ice margin
Member (Fig 57.2). by tractional underflows and gravity flows. In interglacial periods,
finer-grained sediments accumulated, occasionally with dropstones,
attesting to a glaciomarine environment. The dark silt-dominated
Geochronological constraints deposit at Grasdal was interpreted to be a loessite (Edwards 1979).
Hansen (1992), while agreeing that most of the diamictites
No robust and high-resolution geochronological constraints are were of glacial origin, argued for a shelf depositional environment
available. Rb –Sr dating of shales associated with the glaciogenic dominated by undermelt- and flow-tillites, some heavily deformed
strata provides broad constraints for the deposition of the Smal- by an advancing glacier. Rapid vertical and lateral thickness vari-
fjord and Mortensnes formations (630 – 560 Ma; Pringle 1973; ations and terminations of both tillites and associated facies were
Gorohkov et al. 2001). interpreted to have been mostly associated with changes in glacial
movement directions and local bathymetrical conditions rather
than glacial erosion and lodgement. Glaciogenic mass-movement
Discussion tillites (flow-tillite) laterally grading into sandstones, and at other
stratigraphic levels grading into rhythmites, indicates transport
Føyn & Siedlecki (1980) interpreted the Lower Krokvatn Diamic- directions and the relative location of the basin/continent.
tite as a massive indurated ground moraine and the Middle and In contrast, Arnaud & Eyles (2002) suggested that the diamic-
Upper Krokvatn diamictites as partly ground moraine and partly tites in both the Gaessenjar’ga and the southern and western
deposited under water. Aeolian or fluvial processes deposited shores of Varangerfjord accumulated from subaqueous gravity
the intervening sandstones into quiet lakes or marine basins in flows from the basin margin, with the latter forming part of a
the palaeovalley, while the red sediments above the middle diamic- debris apron (see also Crowell 1999; Schermerhorn 1974; Jensen
tite are likely loessites. Essentially, Føyn & Siedlecki (1980) inter- & Wulff-Pedersen 1996). Thus, although the slumped material
preted the sequence as alternating glacial stadial and interstadial may have been of glacial origin, seen in rare faceted and striated
events. However, Eyles (1993) argued that this model was pre- clasts and lonestones, the diamictites were not directly deposited
sumptive, with the interpretation that the rocks are glacial in by ice and so cannot be considered to be tillites. Instead, Arnaud
origin being based on their stratigraphic position rather than any & Eyles (2002) suggested that they record deposition of unstable
unequivocal diagnostic glacial criteria, and thus the alternations sediments on the edge of an active extensional basin (but see
were simply presumed to represent cyclical glacial advances and Edwards 2004; Arnaud & Eyles 2004), in which icebergs or
retreats, without substantive proof. sea-ice contributed dropstones. Ice proximal settings were inferred
Bjørlykke (1967) and Edwards (1984) interpreted the Varan- only for the eastern shore of Skjåholmen, where sediment gravity
gerfjord succession as an infill of a glacially scoured valley, the flow diamictite is closely associated with well stratified glacioflu-
ice moving towards the NW (Laajoki 2003; Baarli et al. 2006). vial sandstones and conglomerates and at Handelsneset, where
Irregular glacier retreat left ice-cored moraines and basement complex deformation observed in conglomerates and sandstones
highs that were submerged by rising water levels (Levine et al. suggests active deformation by overriding ice (Arnaud 2008).
2006). Bathymetric lows were filled first by sediment gravity The proposed source areas, and thus sediment transport (ice or
flows and overlain subsequently by rapidly prograding deltas and otherwise) directions for the Smalfjord Fm., also vary. Føyn & Sie-
sandur plains. Later diamictites may represent glacial advances dlecki (1980) argue that the Middle and probably also the Lower
and/or sediment slumping. Krokvatn Diamictite were derived from the south. Arnaud &
At Oaibaččanjar’ga, Edwards (1975) suggested that the diamic- Eyles (2002) also proposed a southerly source area for the probable
tite overlying the striations formed as a meltout diamictite from equivalent rocks at Vieranjarg’a in the Varangerfjord area.
sediment-laden dead-ice, with the margins slumping and being Edwards (1984) inferred a westward ice flow along the Varanger
eroded as the ice-entrained sediment lost cohesion. In contrast, palaeovalley, with two dominant striation directions documented:
Schermerhorn (1974), Jensen & Wulff-Pedersen (1996), Crowell NW –SE and east – west (Bjørlykke 1967; Edwards 1975; Jensen &
(1999) and Arnaud & Eyles (2002) argued that the diamictite Pedersen 1996; Rice & Hofmann 2000; Laajoki 2002). Edwards
formed during slumping, possibly of glaciogenic deposits, with (1984) also proposed that of the five advance/retreat cycles ident-
some suggestions that the striations also formed during slumping ified in the Tanafjord area, the first was southerly derived, the sub-
as the overriding diamictite was deposited; some of these authors sequent three northerly derived and the last of unknown origin.
did not realize that the pavement is part of a major, deep regional Hansen (1992), in contrast, suggested that the lower part of the suc-
angular unconformity. The very thin breccia preserved in a number cession was NE-derived and the latter part was SE-derived. These
of striations and as ridges beside striations on the pavement has differences are not simply due to the different order of the units
been ascribed to glacio-tectonic brittle deformation processes, pre- proposed by the two authors.
served by the instantaneous recrystallization of the highly strained Member A of the Nyborg Fm. represents post-glacial transgres-
comminuted material, associated with flash-heating of the sub- sion. Although Edwards (1984) suggested that the dolostones
strate, possibly up to c. 1700 8C, during rapid glacial movement. formed around topographically high areas, correlation of the carbon-
This was likely associated with glacial earthquakes (Rice & ates with Marinoan cap dolostones (see below) suggests that depo-
Hofmann 2000; Bestmann et al. 2006). The comminuted material sition was not directly related to palaeodepth (cf Hoffman et al.
represents proto-rock-flour. 2007). Variations in the clastic input certainly controlled which
In the Tanafjord area, Edwards (1984) interpreted the Smalfjord lithology formed. Subsequently, Members B–D reflect a rapid dee-
Fm. (equivalent to the Upper Krokvatn Diamictite of Føyn & pening of the basin (Fig. 57.3), followed by a more gradual shallow-
Siedlecki 1980; Fig. 57.2) as a cyclic deposit of glacial advances ing, with Member E representing a barrier lagoon facies, with
and retreats, with the former yielding a basal massive lodgement intermittent carbonate deposition (Edwards 1984).
tillite, locally with sandstones deposited in situ by sub-glacial The Mortensnes Fm. represents two advances and retreats of ice
meltwater at the ice margin. At the base of cycles, ice movement (Edwards 1984); the first was initially derived both from the south
600 A. H. N. RICE ET AL.

(lower member) and later from the north (middle member). separated by a major angular unconformity (Johnson et al. 1978;
The source area of the second advance (upper member), which Siedlecki & Levell 1978); no glacial deposits occur in these
had a major change in provenance, is not known, although the sequences. Biostratigraphic data (Vidal & Siedlecka 1983) are
extreme negative d13C values of the dolomitic diamictite at too imprecise to estimate the age of the youngest rocks in the
the top of the formation are thought to indicate derivation from Barents Sea Group (Moczydlowska-Vidal, pers. comm. 2007),
the underlying Member E of the Nyborg Fm. For both cycles, lod- although it is younger than 800 Ma (see above). Mafic dykes
gement tillite was followed by floating ice, giving finer-grained cutting the unconformity and basal part of the Løkvikfjell Group
sedimentation and dropstones. The polymict conglomerate were intruded at c. 577 + 14 Ma (Rice et al. 2004). Thus evi-
draping the formation has been interpreted as a lag-conglomerate, dence for glacial conditions recorded by the Smalfjord and Mor-
formed during local isostatic uplift, after glacial retreat (Edwards tensnes formations may not have been preserved in this basin.
1984). The succession passed essentially conformably upwards
into the post-glacial Lillevannet Member, representing fluvio- A. H. N. R. thanks A. and J. Pettersen, G. Stensvold, P. Sørflaten, Ø. Hauge
deltaic and subsequent marine deposition. and the extended Larsen family (Vestertana) for hospitality and boating assistance
Halverson et al. (2005) gave probable correlations of the glacio- during fieldwork in E. Finnmark, and C. and R. Hofmann and M. Ebner
genic units in E. Finnmark with those documented elsewhere, for assistance in the field. M. B. E. thanks A. and J. Pettersen for their
based on chemostratigraphy and carbonate lithologies. Dolostones hospitality; H. Reading, S. Føyn, S. Siedlecki, A. Siedlecka, S.-L. Røe,
at the base of the Nyborg Fm. have negative d13C values (–1 to T. Spencer, K. Bjørlykke and G. Boulton for scientific assistance in the field
–5.9‰), and, where massive, have sheet cracks, pseudo-tepees and/or discussions; the New York State Higher Education Assistance Program
and family for financial and moral support. Other acknowledgements are given
and, at one locality, sulphate crystal fans. This, in combination
in Edwards (1984). E.A. thanks the Natural Sciences and Engineering Research
with their buff-weathering colour strongly suggests that this is a
Council of Canada for funding, and C. Trotter and S. Aspden for assistance in
Marinoan-type cap dolostone (cf Kennedy et al. 1998). d13C data
the field. This represents a contribution of the IUGS- and UNESCO-funded
from the 230-m-thick Grasdalen Fm. in the Gulgofjord area IGCP (International Geoscience Programme) Project #512.
(Fig. 57.1) show positive values (6‰), falling rapidly to negative
values and then recovering to positive values below the
c. 20-m-thick Fadnuvag’gi Fm., which lies directly under the Smalf- References
jord Fm. in the Grasdal area (Fig. 57.1). Halverson et al. (2005)
equated this d13C pattern with the Trezona anomaly, but since a Arnaud, E. 2008. Deformation in the Neoproterozoic Smalfjord For-
broadly similar anomaly occurs under the Port Askaig Fm., which mation, Northern Norway: an indicator of glacial depositional con-
is no longer considered to be a Marinoan equivalent (McCay et al. ditions? Sedimentology, 55, 335– 356.
2006), this is not a robust constraint. Nevertheless, the character- Arnaud, E. & Eyles, C. H. 2002. Glacial influence on Neoproterozoic
istics of the cap dolostone secure the Smalfjord Fm. as an equivalent sedimentation: the Smalfjord Formation, northern Norway. Sedimen-
of the 635 Ma (Hoffmann et al. 2004) Marinoan glaciogenic tology, 49, 765– 788.
succession. Arnaud, E. & Eyles, C. H. 2004. Glacial influence on Neoproterozoic
Thin dolostones in the upper part of the Nyborg Fm. (Member sedimentation: the Smalfjord Formation, northern Norway — reply.
E), with d13C values of –7.6 to –9.9‰, are likely correlatives Sedimentology, 51, 1423– 1430.
of the Wonoka anomaly, which elsewhere has d13C values as Baarli, B. G., Levine, R. & Johnson, M. E. 2006. The Late Neoproter-
low as –12.81‰ (VPDB: Le Guerroué 2006); this implies that ozoic Smalfjord Formation of the Varanger peninsula in northern
Norway: a shallow fjord deposit. Norwegian Journal of Geology,
the almost immediately overlying Mortensnes Fm. is broadly a
86, 133– 150.
correlative of the 580 Ma Gaskiers diamictite (cf Halverson
Banks, N. L., Edwards, M. B., Geddes, W. P., Hobday, D. K. &
et al. 2005). Reading, H. G. 1971. Late Precambrian and Cambro-Ordovician
Kulling (1951) introduced the term Varangeristiden in a discus- sedimentation in East Finnmark. Norges geologiske Undersøkelse,
sion of diamictites in northern Sweden. Although those diamictites 269, 197– 236.
are now correlated solely with the Mortensnes Fm. (cf. Stodt et al. Bestmann, M., Rice, A. H. N., Langenhorst, F., Grasemann, B. &
2011), the term originally included all the Neoproterozoic glacio- Heidelbach, F. 2006. Subglacial bedrock welding associated with
genic rocks in E. Finnmark and implied one glacial period. glacial earthquakes. Journal of the Geological Society, 163, 417– 420.
However, as the Smalfjord and Mortensnes formations are now cor- Bingen, B., Griffin, W. L., Torsvik, T. H. & Saeed, A. 2005. Timing of
related with two distinct glacial events (Halverson et al. 2005), one Late Neoproterozoic glaciation on Baltica constrained by detrital
of which was worldwide and the other only localized, the concept of zircon geochronology in the Hedmark Group, south-west Norway.
a Varangeristiden (Varangian/Varangerian ice age) is no longer Terra Nova, 17, 250–258.
valid. Further, taking a 12 Ma duration for the Marinoan glaciation Bjørlykke, K. O. 1967. The Eocambrian ‘Reusch Moraine’ at Biggan-
(Bodiselitsch et al. 2005) and 1 Ma for the Gaskiers (Bowring et al. jargga and the geology around Varangerfjord, Northern Norway.
2003) implies a total time span of 647–579 Ma for the two glacial Norges geologiske Undersøkelse, 251, 18 –44.
events, of which only 13 Ma (18%) was actually spent under ice; Bodiselitsch, B., Koeberl, C., Master, S. & Reimold, W. U. 2005.
scarcely enough to justify the term ‘ice age’. For these reasons, Estimating duration and intensity of Neoproterozoic snowball glacia-
use of the term Varang(er)ian ice age is here discouraged. tions from Ir anomalies. Science, 308, 239–242.
No evidence of an older, Sturtian equivalent, glaciation has been Bowring, S., Myrow, P., Landing, E., Ramezani, J. & Grotzinger, J.
recorded in this area. The Lille Molvika Fm. (Ekkerøya Group; 2003. Geochronological constraints on terminal Neoproterozoic
events and the rise of metazoans. Geophysical Research Abstracts,
Fig. 57.2) is bounded by two regional unconformities (Rice 1994;
5, 13219.
Siedlecka 1995), either of which may ‘hide’ a pre-Marinoan
Cawood, P. A. & Pisarevsk, S. A. 2006. Was Baltica right-way-up or
glacial event. These unconformities are of post-800 Ma age, based upside-down in the Neoprotereozoic? Journal of the Geological
on correlations in the Manjunnas area (Fig. 57.1), where the Society, London, 163, 753–759.
Ekkerøya Group overlies the Båtsfjord Fm. of the Barents Sea Chapman, T. J., Gayer, R. A. & Williams, G. D. 1985. Structural cross-
Group (Rice 1994). d13C data from the 1400–1600-m-thick Båts- sections through the Finnmark Caledonides and timing of the Finn-
fjord Fm. in the North Varanger Region (Fig. 57.1) are consistently markian event. In: Gee, D. G. & Sturt, B. A. (eds) The Caledonide
negative (Rice & Halverson, unpublished data), suggesting a corre- Orogen — Scandinavia and Related Areas. J. Wiley & Sons, Chi-
lation with the c. 800 Ma old Bitter Springs anomaly (cf Halverson chester, 593–610.
et al. 2005). Cocks, L. R. M. & Torsvik, T. H. 2005. Baltica from the late Precambrian
In the North Varanger Region (Fig. 57.1), the Barents Sea (max. to mid-Palaeozoic times: the gain and loss of a terranes identity. Earth
9 km thick) and overlying Løkvikfjell (c. 5.7 km) Groups are Science Reviews, 72, 39 –66.
NEOPROTEROZOIC SMALFJORD AND MORTENSNES FORMATIONS 601

Crimes, T. P. & McIlroy, D. 1999. A biota of Ediacaran aspect from the Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M.,
Lower Cambrian strata on the Digermul Peninsula, Arctic Norway. Higgins, J. A. & Schrag, D. P. 2007. Are basal Ediacaran (635
Geological Magazine, 136, 633– 642. Ma) post-glacial ‘cap dolostones’ diachronous. Earth and Planetary
Crowell, J. C. 1999. Pre-Mesozoic Ice-ages. Their bearing on Under- Science Letters, 258, 114–131.
standing the Climate System. Geological Society of America Hoffmann, K.-H., Condon, D. J., Bowring, S. A. & Crowley, J. L.
Memoir, 192, 106. 2004. U –Pb zircon date from the Neoproterozoic Ghaub
Edwards, M. B. 1972. Glacial, interglacial and postglacial sedimentation Formation, Namibia: constraints on Marinoan Glaciation. Geology,
in a late Precambrian shelf environment. DPhil, Oxford University. 32, 817– 820.
Available at http://www.marcedwards.com/publications_nnorway. Jensen, P. A. & Wulff-Pedersen, E. 1996. Glacial or non-glacial origin
htm. for the Bigganjarga tillite, Finnmark, northern Norway. Geological
Edwards, M. B. 1975. Glacial retreat sedimentation in the Smalfjord Magazine, 133, 137– 145.
Formation, Late Precambrian, North Norway. Sedimentology, 22, Johnson, H. D., Levell, B. K. & Siedlecki, S. 1978. Late Precambrian
75 – 94. sedimentary rocks in East Finnmark, North Norway and their
Edwards, M. B. 1979. Late Precambrian glacial loessites from North relationship to the Trollfjord-Komagelva Fault. Journal of the
Norway and Svalbard. Journal of Sedimentary Petrology, 49, 85– 92. Geological Society, London, 135, 517– 533.
Edwards, M. B. 1984. Sedimentology of the Upper Proterozoic glacial Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffman, K. H. &
record, Vestertana Group, Finnmark, North Norway. Norges geolo- Arthur, M. A. 1998. Two or four Neoproterozoic glaciations?
giske Undersøkelse Bulletin, 394, 1– 76. Geology, 26, 1059– 1063.
Edwards, M. B. 2004. Glacial influence on Neoproterozoic sedimen- Kulling, O. 1951. Spår av Varangeristiden i Norbotten. Sveriges geolo-
tation: the Smalfjord Formation, northern Norway — discusssion. giska Undersökning Årsbok 43 (1949), C 503, 1 –45.
Sedimentology, 51, 1409– 1417. Kumpulainen, R. A. 2011. The Neoproterozoic Lillfjället Formation,
Edwards, M. B. & Føyn, S. 1981. Late Precambrian tillites in Finnmark, southern Swedish Caledonides. In: Arnaud, E., Halverson, G. P.
North Norway. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
Pre-Pleistocene Glacial Record. Cambridge University Press, Cam- Glaciations. Geological Society, London, Memoirs, 36, 629– 634.
bridge, 606– 610. Kumpulainen, R. A. & Greiling, R. O. 2011. Evidence for late Neopro-
Edwards, M. B., Baylis, P., Gibling, M., Goffe, W., Potter, M. & terozoic glaciation in the central Scandinavian Caledonides. In:
Suthren, R. J. 1973. Stratigraphy of the ‘Older Sandstone Series’ Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
(Tanafjord Group) and Vestertana Group North of Stallogaissa, Lak- logical Record of Neoproterozoic Glaciations. Geological Society,
sefjord District, Finnmark. Norges geologiske Undersøkelse, 294, London, Memoirs, 36, 623–628.
25 – 41. Laajoki, K. 2001. Additional observations on the late Proterozoic Varan-
Eyles, N. 1993. Earth’s glacial record and its tectonic setting. Earth gerfjorden unconformity, Finnmark, northern Norway. Bulletin of the
Science Reviews, 35, 1 –248. Geological Survey of Finland, 73, 17 –34.
Farmer, J., Vidal, G., Moczydłowska, A., Strauss, H., Ahlberg, P. & Laajoki, K. 2002. New evidence of glacial abrasion of the Late Protero-
Siedlecka, A. 1992. Ediacaran fossils from the Innerelv Member zoic unconformity around Varangerfjorden, northern Norway.
(late Proterozoic) of the Tanafjorden area, northeastern Finnmark. Special Publication of the International Association of Sedimentolo-
Geological Magazine, 129, 181– 195. gists, 33, 405– 436.
Føyn, S. 1937. The Eo-Cambrian series of the Tana District, northern Laajoki, K. 2003. The Larajæg’gi outcrop — a large combined Neopro-
Norway. Norsk Geologisk Tidsskrift, 17, 65– 164. terozoic/Pleistocene roche moutonnée at Karlebotn, Finnmark,
Føyn, S. & Glaessner, M. F. 1979. Platysolenites, other animal fossils, North Norway. Norwegian Journal of Geology, 84, 107– 115.
and the Precambrian – Cambrian transition in Norway. Norsk Geolo- Le Guerroué, E., Allen, P. A. & Cozzi, A. 2006. Chemostratigraphic
gisk Tiddskrift, 59, 25 – 46. and sedimentological framework of the largest negative carbon isoto-
Føyn, S. & Siedlecki, S. 1980. Glacial Stadials and Interstadials in pic excursion in Earth history: the Neoproterozoic Shuram Formation
the Late Precambrian Smalfjord Tillite on Laksefjordvidda, (Nafun Group, Oman). Precambrian Research, 146, 68 –92.
Finnmark, North Norway. Norges geologiske Undersøkelse, 358, Levine, R., Baarli, G. & Johnson, M. E. 2006. Glacial and rocky-shore
31 – 45. dynamics of the northern Karlebotn monadnocks: late Neoprotero-
Gayer, R. A. & Rice, A. H. N. 1989. Palaeogeographic reconstruction of zoic of northern Norway. Canadian Journal of Earth Sciences, 43,
the pre- to syn-Iapetus rifting sediments in the Caledonides of 1215–1228.
Finnmark, N. Norway. In: Gayer, R. A. (ed.) The Caledonide McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial
Geology of Scandinavia. Graham & Trotman, London, 127– 139. trinity: Neoproterozoic Earth history within the British –Irish Caledo-
Gehling, J. G., Narbonne, G. M. & Anderson, M. M. 2000. The first nides. Geology, 34, 909– 912.
named Ediacaran body fossil, Aspidella terranovic. Palaeontology, Nystuen, J. P. & Lamminen, J. T. 2011 Neoproterozoic glaciation of
43, 427– 456. South Norway: from continental interior to rift and pericratonic
Gorokhov, I. M., Siedlecka, A., Roberts, D., Melnikov, N. N. & basins in western Baltica. In: Arnaud, E., Halverson, G. P. &
Turchenko, T. L. 2001. Rb –Sr dating of diagenetic illite in Neopro- Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
terozoic shales, Varanger Peninsula, northern Norway. Geological Glaciations. Geological Society, London, Memoirs, 36, 613–622.
Magazine, 138, 541– 562. Pisarevsky, S. A., Murphy, J. B., Cawood, P. A. & Collins, A. S. 2008.
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Late Neoproterozoic and Early Cambrian palaeogeography: models
Rice, A. H. N. 2005. Towards a Neoproterozoic composite carbon and problems. In: Pankhurst, R. J., Trouw, R. A. J., de Brito
isotope record. Geological Society of America Bulletin, 117, Neves, B. B. & de Wit, M. J. (eds) West Gondwana: Pre-Cenozoic
1181– 1207. Correlations Across the South Atlantic Region. Geological Society,
Hambrey, M. J. 1983. Correlation of Late Proterozoic tillites in London, Special Publication, 294, 9–31.
the North Atlantic region and Europe. Geological Magazine, 120, Pringle, I. R. 1973. Rb –Sr age determinations on shales associated with
209–232. the Varanger ice age. Geological Magazine, 109, 465– 472.
Hansen, T. A. 1992. Sedimentologiske og Stratigrafiske Undersøkelser av Reading, H. G. 1965. Eocambrian and Lower Palaeozoic geology of the
den Sen-Prekambriske Smalfjord formasjonen, Øst-Finnmark. Cand. Digermul Peninsula, Tanafjord, Finnmark. Norges geologiske Under-
Real thesis, Universitet i Troms. søkelse, 234, 167– 191.
Harland, W. B., Armstrong, R. L., Cox, A. V., Craig, L. E., Smith, Reading, H. G. & Walker, R. G. 1966. Sedimentation of Eocambrian
A. G. & Smith, D. G. 1989. A Geological Time Scale 1989. tillites and associated sediments in Finnmark, North Norway. Palaeo-
Cambridge University Press. geography, Palaeoclimatology, Palaeoecology, 2, 177– 212.
Hartz, E. H. & Torsvik, T. H. 2002. Baltica upside down: a new plate Reusch, H. 1891. Skuringsmærker og morængrus eftervist i Finnmarken
tectonic model for Rodinia and the Iapetus Ocean. Geology, 30, fra en periode meget ældre end “istiden”. Norges geologiske Under-
255– 258. søkelse, 1, 78 –85 (97–100 English summary).
602 A. H. N. RICE ET AL.

Rice, A. H. N. 1994. Stratigraphic overlap of the late Proterozoic Vadsø Siedlecka, A. & Siedlecki, S. 1971. Late Precambrian sedimentary
and Barents Sea Groups and correlation across the Trollfjorden– rocks of the Tanafjord-Varangerfjord region of Varanger Peninsula,
Komagelva Fault, Finnmark, North Norway. Norsk Geologisk Tids- Northern Norway. Norges geologiske Undersøkelse, 269, 246–294.
skrift, 74, 48 –57. Siedlecka, A., Lyubtsov, V. V. & Negrutsa, V. Z. 1995. Correlations
Rice, A. H. N. & Townsend, C. 1996. Correlation of the late Precambrian between Upper Proterozoic successions in the Tanafjorden-
Ekkerøya Formation (Vadsø Group; E. Finnmark) and the Brennelv- Varangerfjorden Region of Varanger Peninsula, northern Norway,
fjord Interbedded member (Porsangerfjord Group; W. Finnmark), and on Sredni Peninsula and Kildin Island in the northern coastal
N. Norwegian Caledonides. Norsk Geologisk Tidsskrift, 76, 55– 61. area of Kola Peninsula in Russia. Norges geologiske Undersøkelse
Rice, A. H. N. & Hofmann, Ch.-Ch. 2000. Evidence for a glacial origin Special Publication, 7, 217– 232.
of the Neoproterozoic III striations at Oaibaččanjar’ga, Finnmark, Siedlecka, A., Roberts, D., Nystuen, J. P. & Olovyanishnikov, V. G.
northern Norway. Geological Magazine, 137, 355– 366. 2004. Northeastern and northwestern margins of Baltica in Neopro-
Rice, A. H. N. & Hofmann, Ch.-Ch. 2001. The transition from Neo- terozoic time: evidence from the Timanian and Caledonian
proterozoic glacial to interglacial sedimentation near Hammarnes, Orogens. In: Gee, D. G. & Pease, V. (eds) The Neoproterozoic Tima-
East Finnmark, North Norway. Norwegian Journal of Geology, 81, nide Orogen of Eastern Baltic. Geological Society, London,
257– 262. Memoirs, 30, 169–190.
Rice, A. H. N., Gayer, R. A., Robinson, D. & Bevins, R. E. 1989. Strike- Siedlecki, S. & Levell, B. K. 1978. Lithostratigraphy of the Late
slip restoration of the Barents Sea Caledonides Terrane, Finnmark, Precambrian Løkvikfjell Group on Varanger Peninsula, East
North Norway. Tectonics, 8, 247– 264. Finnmark, North Norway. Norges geologiske Undersøkelse, 343,
Rice, A. H. N., Hofmann, Ch.-Ch. & Pettersen, A. 2001. A new sedi- 73 – 85.
mentary succession from the southern margin of the Neoproterozoic Stodt, F., Rice, A. H. N., Björklund, L., Bax, G. & Pharaoh, T. C.
Gaissa Basin, south Varangerfjord, North Norway; the Lattanjaŕga 2011. Evidence of late Neoproterozoic glaciation in the Caledonides
unit. Norsk Geologisk Tidsskrift, 81, 41 –48. of NW Scandinavia. In: Arnaud, E., Halverson, G. P. & Shields-
Rice, A. H. N., Ntaflos, T., Gayer, R. A. & Beckinsale, R. D. 2004. Zhou, G. (eds) The Geological Record of Neoproterozoic Glacia-
Metadolerite geochronology and dolerite geochemistry from East tions. Geological Society, London, Memoirs, 36, 603–611.
Finnmark, Northern Scandinavian Caledonides. Geological Maga- Strahan, A. 1897. On Glacial Phenomena of Palaeozoic age in the
zine, 141, 301– 318. Varanger Fjord. Quarterly Journal of the Geological Society, 53,
Røe, S.-L. 2003. Neoproterozoic peripheral-basin deposits in eastern 137– 146.
Finnmark, N. Norway: stratigraphic revision and palaeotectonic Torsvik, T., Lohmann, K. C. & Sturt, B. A. 1995. Vendian glaciations
implications. Norwegian Journal of Geology, 83, 259– 274. and their relation to the dispersal of Rodinia: palaeomagnetic con-
Schermerhorn, L. J. G. 1974. Late Precambrian mixtites: glacial or straints. Geology, 23, 727– 730.
non-glacial. American Journal of Science, 274, 673–824. Townsend, C., Roberts, D., Rice, A. H. N. & Gayer, R. A. 1986. The
Siedlecka, A. 1985. Development of the Upper Proterozoic sedimentary Gaissa Nappe, Finnmark, North Norway: an example of a deeply
basins of the Varanger Peninsula, East Finnmark. Geological Survey eroded external imbricate zone within the Scandinavian Caledonides.
of Finland Bulletin, 331, 175–185. Journal of Structural Geology, 8, 431–440.
Siedlecka, A. 1987. Trollfjorden berggrunnskart 2336 3, 1:50,000, Vidal, G. 1981. Micropalaeontology and biostratigraphy of the Upper
foreløpig utgave. Norges geologiske Undersøkelse. Proterozoic and Lower Cambrian sequences in East Finnmark, north-
Siedlecka, A. 1990. Varangerbotn berggrunnskart 2335 3, 1:50,000, fore- ern Norway. Norges geologiske Undersøkelse, 362, 1 –53.
løpig utgave. Norges geologisk Undersøkelse. Vidal, G. & Siedlecka, A. 1983. Planktonic, acid-resistant microfossils
Siedlecka, A. 1995. Neoproterozoic sedimentation on the Rybachi and from the Upper Proterozoic strata of the Barents Sea region of Varan-
Sredni Peninsulas and Kildin Island, NW Kola, Russia. Norges ger Peninsula, East Finnmark, Northern Norway. Norges geologiske
geologiske Undersøkelse Bulletin, 427, 52– 55. Undersøkelse, 382, 45– 79.
Chapter 58

Evidence of late Neoproterozoic glaciation in the Caledonides of NW Scandinavia

F. STODT1, A. H. N. RICE2*, L. BJÖRKLUND3, G. BAX4, G. P. HALVERSON5,6 & T. C. PHARAOH7


1
Savignystrasse 9, 35037 Marburg, Germany
2
Vienna University, Structural Processes Group, Department of Geodynamics and Sedimentology, Althanstrasse 14,
1090 Vienna, Austria, Europe
3
Göteborg University, Earth Sciences Centre, Box 460, SE-405 30 Göteborg, Sweden
4
Department of Earth Sciences & Centre for Image Analysis, Uppsala University, SE-752 36, Uppsala, Sweden
5
Department of Geology & Geophysics, School of Earth & Environmental Sciences, University of Adelaide, North Terrace,
Adelaide, SA 5005, Australia
6
Present address: Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montreal,
Quebec, H3A 2A7, Canada
7
Geophysics & Marine Geosciences, British Geological Survey, Keyworth, Nottinghamshire, NG12 5GG, UK
*Corresponding author (e-mail: alexander.hugh.rice@univie.ac.at)

Abstract: The northwestern part of the Scandinavian Caledonides, formed by SE- to ESE-directed thrusting through the Neoproterozoic
W. Baltica continental shelf, contains numerous small and often isolated outcrops of diamictite and associated strata. No precise bio-
stratigraphic or isotopic data are available to constrain the age of these sediments, but, on the basis of their stratigraphic position,
most are correlated with the Mortensnes Formation (Fm.) in E. Finnmark and also presumed to be of glaciogenic origin. The Mortensnes
Fm. has been correlated with the 580 Ma Gaskiers glacial event on the basis of d13C isotope studies. Structurally, the deposits occur in the
Autochthon (below the Torneträsk Fm.), within an external imbricate zone (Lower Allochthon), within cover successions lying uncon-
formably on allochthonous basement (Window Allochthon) palaeogeographically derived from below or outboard of the Lower Alloch-
thon and, more rarely, within the Middle Allochthon, derived from outboard of the Window Allochthon. Evidence for a glaciogenic
origin is typically poor or lacking. Only in the Komagfjord Antiformal Stack (Window Allochthon), where an up to 40-m-thick succes-
sion of three fining upwards cycles has been mapped, are the deposits comparable in thickness and complexity to the Mortensnes Fm.
Other sequences are sometimes ,1 m thick and unconformably overlain by post-‘glacial’ deposits. The Vakkejokk Breccia, a submarine
slump in the Torneträsk area of the Autochthon closely underlies the correlative Precambrian –Cambrian lithostratigraphic boundary in
E. Finnmark but overlies the first appearance of the boundary marker fossil Treptichnus pedum. Although sometimes interpreted as peri-
glacial, this seems unlikely in view of the 30–508 palaeolatitude during deposition. Calcite nodules (,1 cm size) in the Vakkejokk
Breccia have previously been interpreted as glendonite, but the microstructure and palaeolatitude makes this unlikely; they are likely
a replacement of gypsum. Diamictites of uncertain origin have also been found in the Ediacaran Lower Siltstone Member of the Torne-
träsk Fm. and unconformably under the ?Lower Cambrian Lomvatn Fm. in the Komagfjord Antiformal Stack.

This chapter primarily covers (meta-)sediments correlated with the (Fig. 58.2; Gayer et al. 1987; Gayer & Greiling 1989; Rice
Mortensnes Fm., the younger of the two glaciogenic units of 1998, 2001, 2006). In NE Scandinavia, this continental shelf
E. Finnmark, N. Norway (cf Rice et al. 2011). Such deposits linked with the WNW –ESE-trending Timanian Basin (Siedlecka
typically occur as small and isolated outcrops within both the 1985; Gayer & Rice 1989; Siedlecka et al. 2004; cf Rice et al.
Autochthon and nappes of the Scandinavian Caledonides, often 2011) that formed after the accretion of several minor terranes to
resting unconformably on basement rocks (Kumpulainen 2011; Baltica in the middle Neoproterozoic (cf Cocks & Torsvik
Kumpulainen & Greiling 2011; Nystuen & Lamminen 2011) 2005). The Timanian Basin has been divided into northerly deep
After restoration of the Caledonian nappes, these diamictite out- water and southerly shelf parts (Siedlecka et al. 1995, 2004), sep-
crops indicate that glacial deposits covered an area of some arated by the Trollfjorden – Komagelva Fault (cf Rice et al. 1989).
140 000 km2 in N. Scandinavia (including E. Finnmark). In Scandinavia, the shelf part comprises the Gaissa Basin, which,
Studies in Norrbotten, Sweden, in the southernmost part of the in the west, was also an inboard basin of the NNE –SSW-trending
area covered here (Fig. 58.1), led Kulling (1951) to introduce the W. Baltica shelf.
term Varangeristiden (Varang(er)ian glaciation) to encompass Predominantly in-sequence ESE- to SE-directed Caledonian
both the glacial events described here and the earlier Smalfjord shortening occurred mostly in Silurian – Devonian times. This
Fm. (Marinoan). However, for reasons summarized below, this passed through the outboard basin (Middle Allochthon), then a
term should no longer be used. palaeogeographic basement-high (Window Allochthon, Rice
2001) and then the inboard basin, forming an external imbricate
zone (external Lower Allochthon), finally emplacing the nappe
Structural framework pile onto the Autochthon. In some areas, the internal part of the
Lower Allochthon restores to above the Window Allochthon and
Restoration of the Scandinavian Caledonides, using balanced was imbricated earlier (Fig. 58.2). Both basins have essentially
cross-sections, demonstrates that the NNE –SSW-trending the same lithostratigraphy; in most areas, the basement-high was
(present orientation) continental shelf of W. Baltica comprised finally drowned in late Ediacaran to mid-Cambrian –Tremadocian
an outboard basin lying adjacent to the continental edge and an times, essentially contemporary with the mainland to the SE. In the
intermittently developed, somewhat shallower inboard basin Tysfjord-Akkajaure area, both the Middle and Lower Allochthons

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 603– 611. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.58
604 F. STODT ET AL.

Fig. 58.1. Map of north Scandinavian Caledonides showing the distribution of glaciogenic lithologies and non-glaciogenic diamictites outside the E. Finnmark
(Laksefjordvidda-Tanafjord-Mortensnes) area. A, Altenes Window; An, Andabakoaivi; Ak, Akkajaure; AK, Alta-Kvænangen Window; Au, Autajaure Window;
B, Bulljovagge; G, Gæv’dnjajav’ri; H, Halkavarre; J, Jerta; K, Komagfjord Window; Ka, Karnjelajåkka; Ko, Kuokkel Window; Lv, Laksefjordvidda; Mn, Mauken
Window; P, Porsavatn; Po, Porsangerfjord; Pa, Paittasjärvi; Ri, Ritsemjaure Window; RS, Rombak-Sjangeli Window; Ru, Ruoddojokka; Si, Sitojaure; Sj, Stora
Sjöfallet; Sk, Skoganvarre; Sv, Sarvapakte; T, Torneträsk; Ty, Tysfjord; Vk, Vakkejåkka (now called Sarvájohka); Vu, Vuojtasrika.

are dominated by basement rocks, with only a thin cover succes- adjacent Kuokkel Windows, and to the south, the relatively
sion (Björklund 1985, 1989), indicating that very restricted basin small Autajaure and Ritsemjaure Windows (Fig. 58.1). Although
development occurred in this region. a basal thrust is typically not exposed, their allochthonous status
In the area covered here, much of the Middle Allochthon com- has been inferred from structural and metamorphic criteria
prises pre-Caledonian basement (Björklund 1985, 1989; Kirkland (Gayer et al. 1987; Anderson 1989; Bax 1989; Björklund 1989;
et al. 2006). Window Allochthon rocks are exposed in the cores Rice 2001).
of large domal/periclinal structures, with the major units being The Gaissa Thrust Belt (derived from the inboard, Gaissa
the Komagfjord Antiformal Stack (Komagfjord, Altenes & Alta- Basin), with 50% shortening in the west (Townsend et al. 1986,
Kvaenangen Windows combined), the Rombak-Sjangeli, and 1989), is the main development of the Lower Allochthon. On

Fig. 58.2. Schematic (V = H) WNW– ESE-oriented palinspastic cross-section showing the development of inboard and outboard basins in pre-diamictite times,
separated by a basement high. The basins and basement high were then buried under glacial and post-glacial deposits. The main tectonic units are numbered in the order of
their imbrication, with schematic main and minor thrusts. Localities along the top are given in the caption to Figure 58.1. Note that in many areas only one basin formed
and in some areas neither basin formed.
THE CALEDONIDES OF NW SCANDINAVIA 605

Laksefjordvidda and to the east (Fig. 58.1), shortening decreases to Stratigraphy


c. 15% (Chapman et al. 1985). SW of Ruoddojokka, most of the
Lower Allochthon has been eroded away (cf. Hossack & Cooper The deposits documented here are associated with condensed
1986; Anderson 1989), although several relicts have been pre- sequences correlated directly or indirectly with the late Precam-
served (Fig. 58.1; Jerta Nappe, Rautas Complex, Lower Thrust brian sediments exposed in E. Finnmark (predominantly the
Complex – see below). Vestertana Group; cf. Rice et al. 2011; Fig. 58.3).

Fig. 58.3. Regional stratigraphic profiles. (a) Alta-Kvaenangen, S. Komagfjord and N. Komagfjord refer to where each part of the log-section was measured. In the
S. Komagfjord zone, Borras Group, Rafsbotn Fm. and Slettfjell Fm. are the three lithostratigraphic equivalents from the three tectonic windows (from south to north) in
the Komagfjord Antiformal Stack. The section shown is of the Slettfjell Fm. (Pharaoh 1985). (b) Section from Torneträsk (Thelander 1982; modified after Bax 1984 and
Stodt 1987). (c) E. Finnmark and W. Finnmark refer to the successions in the eastern (Johnson et al. 1978) and western (Williams 1976; Rice & Townsend 1996) parts of
the Gaissa Basin. Pl, Platysolenties antiquissimus; T, Treptichnus pedum; UE, LE, upper and lower limit of reported Ediacaran fauna; C, calcite nodules.
606 F. STODT ET AL.

West of Andabakoaivi (Fig. 58.1), the Caledonian Autochthon 1–50 cm clasts of amphibolite, gneiss and quartzite, lying below
comprises the Dividal Group, correlated with the post-Mortensnes the distinctive basal conglomerate of the Torneträsk Fm. To the
Fm. part of the Vestertana Group and overlying Digermul Group SW of Ruoddojokka, Holmsen (1956) noted a very variable stra-
in E. Finnmark (Fig. 58.3; Kulling 1964; Føyn 1967; Vogt 1967; tigraphy under the Torneträsk Fm.; in some places only occasional
Thelander 1982). The constituent Torneträsk (c. 100– 260 m) and lonestones are present at the base of the sequence, essentially lying
overlying Alum Shale (c. 80 m) formations are locally separated on the basement, whereas elsewhere 2–4 m of grey diamictite,
from the underlying pre-Caledonian basement (often a regolith; with 1 m clasts, is overlain by 0.5– 1m of carbonate-bearing
Bax 1984; Stodt 1987) by small patches of diamictite (Fig. 58.1); sandstone. At Bulljovagge, c. 2 m of red brown and grey diamictite
typically these are unnamed. The locally developed, but impor- overlie quartzitic sandstone, but the contact with the basement
tant Vakkejokk Breccia (cf. Kulling 1964; Thelander 1982; Stodt is not exposed (Holmsen 1957; Mathiesen in Skjerlie & Tan
1987) cuts across the lower part of the Torneträsk Fm. 1960). In these outcrops, the diamictite is often dominated
(Fig. 58.3). Although Kulling (1964) defined a Middle Sandstone by clasts derived from the immediately underlying rocks. In the
Fm. (; unit C of Vogt 1967), Thelander (1982), who downgraded Paittasjärvi – Sitojaure area and further south (Fig. 58.1), Kautsky
the rank of all the units from formations to members, merged this (1949) described undeformed chloritic diamictites (Sito diamic-
into the Lower Siltstone Member (Fig. 58.3). (Note that the Vak- tite; Stromberg 1981) up to 4 m thick, with a dark matrix. The
kejokk Breccia type locality, Vakkejåkka, or Orddajohka on unsorted, angular to poorly rounded blocks are up to cubic
some maps, is now called Sarvájohka; but this is not the same as metres in size and were derived from the underlying basement,
the Sarvajåkka, now Sarvágorsa, described by Thelander, 1982). which shows no signs of pre-diamictite weathering. The matrix
In the western part of the Gaissa Thrust Belt (external Lower is siliciclastic, comprising small quartz and felspar fragments; no
Allochthon; imbricated Gaissa Basin; Fig. 58.2; Townsend et al. carbonate is present. Near Sitojaure, the diamictite is overlain by
1986, 1989), the Tanafjord Group is thinner than in the type .30 m of green mudstones with frequent blue quartzites and
area, although correlatives of the Grasdal and Lille Molvika for- these are probably overlain by the Dividal Group, which directly
mations (Stabbursdal plus Porsanger formations and Brennelvfjord overlies diamictites in nearby areas (Kulling 1951). Kulling
Fm., respectively; Fig. 58.3) are thicker (cf. Rice & Townsend (1951) also recorded banded silts and clays in the Sitojaure area.
1996). The poorly studied Airoaivi Group is thinner and lithologi- Other thin diamictites in this region were documented by
cally different to the Vadsø Group, its possible chronostratigraphic Kulling (1951; Fig. 58.1), some with faceted clasts, although it is
equivalent (Townsend et al. 1989; Rice & Townsend 1996). No often unclear whether the outcrops documented belong to the
diamictite has been preserved from this part of the Gaissa Basin. Autochthon or overlying Lower Allochthon (cf Strömberg 1981).
In the Jerta Fm. (Jerta Nappe; Lower Allochthon), Skjerlie & The Vakkejokk Breccia lies in the upper part of the Lower Silt-
Tan (1961; Fig. 58.1) recorded a predominantly clastic sequence stone Member of the Torneträsk Fm. on the NE side of Torneträsk
with a basal diamictite. This whole sequence, estimated at (Kulling 1964; Thelander 1982; Stodt 1987; Figs 58.1 & 58.3),
c. 500 m thick, is likely a correlative of the Vestertana Group, thinning from 3– 10 m (authors disagree on the thickness) in the
although the presence of pyritiferous black shales suggest it may Vakkejåkka (Sarvágorsa) area to c. 0.5 m some 7.5 km to the
be partially correlated with the Digermul Group/Alum Shale ESE. This locally cuts down through the Lower Sandstone
Fm. (Reading 1965; Thelander 1982). The successions in both Member to the basement (Stodt 1987). Where the breccia is thin,
the Rautas Complex and the Lower Thrust Complex (near Torne- it may be overlain by sandstones of the uppermost part of the
träsk and Stora Sjöfallet, respectively; Fig. 58.1) have been Lower Siltstone Member (Thelander 1982). Stodt (1987) docu-
reliably correlated with the tectonically underlying autochthonous mented four breccia types.
Dividal Group (Björklund 1985; Bax 1989).
In the Komagfjord Antiformal Stack (Window Allochthon), the † Type A (recorded only at Gaev’dnjajav’ri; Fig. 58.1): conglom-
cover sequence youngs northwards, with ,185 m of Bossekop erate with ,4-cm-sized, well-rounded dark-grey clay-silt clasts
Group (; Tanafjord Group; Fig. 58.3; Føyn 1964, 1985) lying and some quartz fragments, forming the base of a 12-cm-thick
unconformably on the basement in the Alta-Kvaenangen BoumaABCD sequence lacking a structureless sandstone and
Window (Fig. 58.1). The Bossekop Group is unconformably over- dewatering structures. This lies with an erosional contact on
lain by the Borras Group, correlated with the post-Nyborg Fm. part the underlying thin-bedded sandstones of the Lower Siltsone
of the Vestertana Group, including c. 10 m of basal diamictite Member, c. 1 m below the base of the Middle Siltstone Member.
(Fig. 58.3; Føyn 1964, 1985). In the Altenes and southern part of † Type B: angular to rounded granitic, vein-quartz, silt and shale
the Komagfjord Window, patchy outcrops of diamictite up to matrix-supported clasts up to 15 cm in size in sandstone beds up
40 m thick rest directly on the basement and are overlain by a to 30 cm thick. Platy clasts are bedding parallel. In some cases,
,150-m-thick sequence again directly comparable to the lower the beds show normal grading. Matrix sandstone grains are
part of the Torneträsk Fm. (Rafsbotn & Slettfjell formations; rounded to well-rounded and predominantly quartz. The car-
Roberts & Fareth 1974; Pharaoh 1985) whereas in the northern bonate cement may have been derived from a carbonate-mud
part of the Komagfjord Window a small patch of diamictite matrix.
(here termed the Porsavatn Diamictite Bed) is overlain by † Type C: normally .90% granitic clasts with minor elongate,
,160 m of the Lomvatn Fm. (Pharoah 1985), broadly correlated randomly oriented sedimentary clasts, often folded. The
in this chapter with the upper parts of the Torneträsk Fm. former are predominantly clast-supported, equidimensional to
(Fig. 58.3). slightly elongate, mostly angular but sometimes well-rounded.
Other successions within the area, from the Lower, Window and Intraclasts are platy and show abundant fold structures. The
Middle Allochthons, that overlie diamictites have been directly matrix comprises grey, green and red silt/sandstone, often
correlated with the Dividal Group and have not been given local internally brecciated. Two such flows have been recognized,
stratigraphic names. except at Vakkejåkka, where five flows occur. South of Sarva-
pakte (Sarvabakti; Fig. 58.1) one flow contains up to
1-cm-sized calcite nodules interpreted as glendonite by Stodt
Glaciogenic and associated deposits (1987; but see ‘Discussion’). The flows usually lie conformably
or slightly erosively on the Lower Siltstone Member.
Autochthon † Type D: granitic blocks 2– 100 m in size, lying either within
and above siltstones that are often folded and disturbed or on
Near Skoganvarre, Siedlecka (1987; Fig. 58.1) recorded three out- type C breccias, along sharp boundaries, exposed for c. 4 km
crops of polymict conglomerate within a 1km distance, comprising east of Vakkejåkka. SE of Vakkejåkka, 2– 3 m clasts lie directly
THE CALEDONIDES OF NW SCANDINAVIA 607

on the basement. Some breccia clasts, derived from earlier † Facies D: intercalated sandstones and mudstone in bands 1–
flows, are also present (breccia-in-breccia structure). Granitic 5 cm thick, usually reddish brown, although the sandstones
blocks and clasts are basement-derived. may be paler and greyish-green. Sandstones are frequently
graded, with microconglomeratic erosive bases, ripples, load
casts and intraformational mud-flake breccias. Impact pits are
Lower Allochthon present, as is evidence of soft-sediment folding and both
thrust and normal soft-sediment faulting.
No outcrops of diamictites have been recorded within the Gaissa
Thrust Belt in the Porsangerfjord area; exposures from areas Bedding in the diamictite is parallel to that in the overlying basal
further east are reviewed in Rice et al. (2011). In the Jerta Fm., conglomerate of the Vargelv Member (Fig. 58.3), indicating that
Skerlie & Tan (1960; Fig. 58.1) described a diamictite with diamictite thickness variations reflect a syn-sedimentary uneven
5–30 cm angular to rounded, unsorted clasts of light grey, yellow- basement surface; a palaeorelief, up to 3 m high, has been pre-
weathering dolomite, as well as greenstones and quartzites. The served at outcrops and an overall pre-diamictite topography of
matrix is dark to brownish-grey, with angular dolomite and 40 m has been mapped. Pharaoh (1980) suggested that the
smaller poorly rounded quartz grains. Some finer, stratified and Nyvoll Member was deposited in a NW – SE-trending palaeovalley
possibly graded beds are also present. No thicknesses were c. 1km wide and up to 40 m deep, accounting for the sequence
given, but photographs indicate a c. 2 m or greater thickness. having a thickness atypical outside the E. Finnmark area (cf Rice
To the north of Sitojaure, in the Stora Sjöfället area, Kulling et al. 2011). The northern margin of the valley is poorly defined
(1948, 1951, 1982) reported a close association of thinly bedded due to scarce exposure. Clasts in the diamictites comprise felds-
calcareous shales and diamictite or diamictite-like deposits with pathic metasandstones, jasper, vein quartz, metagabbro, serpenti-
lithologies similar to the diamictites in the Ritsemjaure Window nite and trondjemite, all locally derived; sometimes they can be
(see below). linked to nearby palaeo-topographic highs. These sediments
are overlain by a quartz conglomerate, with little topographic
Window Allochthon relief at the contact, and then by red and green siltstones/shales,
comparable with the Borras Group/Rafsbotn Fm. to the south
In the Alta-Kvaenangen window, the most southerly part of the (Fig. 58.3).
Komagfjord Antiformal Stack, 185 m of quartzites and shales of In the northern part of the Komagfjord Window, a
inferred shallow marine origin (Bossekop Group) are unconform- ,0.5-m-thick, very poorly sorted matrix-supported diamictite
ably overlain by c. 10 m of red-brown diamictite at the base of the (Pharaoh 1985), here termed the Porsavatn Diamictite Bed, com-
Borras Group, passing up gradually to conglomerates correlated prising 0.5 –3 cm angular clasts of vein quartz and quartzite in a
with the base of the Torneträsk Fm. and thence to sandstones grey muddy matrix, underlies the 180-m-thick Lomvatn Fm. at
and shales (Føyn 1964, 1985; Fig. 58.3). Further north, in the one area (Figs 58.1 & 58.3). Elsewhere, the 0.6–2-m-thick basal
Altenes Window, Roberts & Fareth (1974) very locally found conglomerate of this formation, consisting of well-rounded vein
c. 2 m of reddish-brown diamictite with clasts (,30 cm size; quartz and quartzite pebbles, rests unconformably on the base-
30% of the rock) of mixed cover and basement lithologies, many ment and is overlain by sandstones and shales (Pharaoh 1985),
derived from the immediate substrate, lying on the basement. probably correlatives of the upper part of the Torneträsk Fm.
Nearby, the Rafsbotn Fm., comprising a basal well-rounded (Fig. 58.3).
pebble-sized quartz conglomerate, c. 1 m thick, and then by The Gearbeljávri Fm., a thin sedimentary veneer that crops out
green and red mudstones, slates and siltstones, directly overlies around the rim of the Rombak-Sjangeli and Kuokkel Windows and
the basement. The Rafsbotn Fm. is equivalent to the basal part on minor tectonic klippen within them, has also been correlated
of the Torneträsk Fm. with the basal part of the Torneträsk Fm. (Fig. 58.1, Brown &
Along the southern rim of the Komagfjord Window (northern Wells 1966; Tull et al. 1985; Bax 1989, 2001). In places, a basal
part of the Komagfjord Antiformal Stack), outcrops of the essen- conglomerate lies at the basement –cover contact, but elsewhere
tially undeformed Nyvoll Tillite Member (Slettfjell Fm.) are pre- the contact grades from unweathered basement, through regolith
served as lenses up to 1km long and 44 m thick (although this into quartzites. At Vuojtasrika, Brown & Wells (1966) documen-
particularly thick outcrop thins to nothing in 300 m along strike; ted 0.5-m-deep, diamictite-filled fissures in the basement, but
Pharaoh 1980, 1985). Pharaoh (1980) described four facies, with Bax (unpublished data) reinterpreted these as fault breccias.
up to three fining-upwards cycles: Along the northern margin of the Rombak-Sjangeli Window,
Tull et al. (1985) described grey to white arkosic cross-bedded
† Facies A: poorly sorted, unstratified polymict diamictites with sandstones, interlayered with conglomerates and pelitic schists,
.20% clasts in a reddish-brown matrix, locally clast supported locally overlain by diamictites with clasts of granitic, ?dioritic
at the base, where the clast content may rise to 40%. Most clasts and pelitic lithologies up to 25 cm in diameter. Coarsely graded
are ,5 cm across, but range up to 60 cm in size and vary from sequences are common. Current directions indicate flow to the
angular to sub-rounded, and are sometimes facetted. A large south to SSW, although the basement –cover contact is a palaeo-
percentage was locally derived, although gneiss, granite and peneplain.
quartzite clasts might have been externally derived. This At one outcrop in the SE part of the Kuokkel Window, a
facies is up to 5 m thick and always forms the base of the 1-m-thick diamictite is underlain by quartzites correlated with
sequence, grading up into Facies B, although it reappears Lower Sandstone Member (Torneträsk Fm.). This diamictite, which
higher in the sequence. lies c. 2 m above the basement –cover unconformity, although the
† Facies B: poorly sorted polymict diamictite similar to facies A, contact is not exposed directly here (Bax 1984), comprises angular,
but with significantly fewer (,20%) and smaller clasts unsorted, carbonate and sandstone clasts (no basement clasts) up to
(c. 10 cm max.), locally with a carbonate matrix/cement 20 cm in size, surrounded by a fine-grained matrix. No clast sorting
towards the top. Thickness varies from 1 to over 20 m. or preferred orientation has been observed. Upwards in the section,
† Facies C: poorly sorted, poorly stratified diamictite, gradational above the overlying cross-bedded sandstones, ripple marks
from facies B by a continued decline in clast size (,1cm) and are common in sandy layers interbedded with brown shales
content (,5%). ‘Bedding’ is defined by thin laminae of quart (Bax 1984).
grains in a reddish brown muddy matrix. This facies is gener- In the Ritsemjaure Window, within the Middle Allochthon
ally thinly developed, although a thickness of 23 m occurs in (Akkajaure Nappe Complex), vertical fissure fillings trending
one section. 0108 occur in mesoperthitic granite. The fissures have highly
608 F. STODT ET AL.

irregular margins and widths of up to 5 m, but are of unknown In contrast, the diamictite in the Kuokkel Window lies within
depths (Fig. 58.1; Björklund 1989). The fissures are filled with a sandstone/quartzite succession correlated with the Lower
an unsorted sedimentary breccia, consisting of a dark grey, silty Sandstone Member of the Torneträsk Fm., above the basal con-
matrix, chaotically mixed with all sizes (,1m) of angular clasts glomerate. Rather similarly, the Vakkejokk Breccia rests with
from the host granite and very subordinate shale and dolomite a marked erosive unconformity on the underlying sediments of
clasts. The same type of breccia overlies the basement granite, the Torneträsk Fm., cutting down-section to, very locally, the
but with smaller (,10 cm) clasts. Microscopic clasts reflect the basement, and is conformably overlain by younger sediments
mesoperthitic feldspars of the subjacent granite. The breccias (Stodt 1987).
grade upwards into dark conglomerates with up to pebble-sized
clasts and a weak clast size stratification and thence into dark
shales with ,5 cm lonestones. Local shearing within the diamic- Chemostratigraphy
tite sequence makes thickness estimates uncertain, but the diamic-
tite lying on the granite is 5 –8 m thick and the overlying No stable isotope data are available due to the general lack of car-
conglomeratic to shaly part is c. 5 m thick in the least disturbed bonates in the sequence. d13C values from calcite nodules in the
areas. Overlying this, along a sharp contact, is a ,3-m-thick, typi- Vakkejokk Breccia range between –4.03 and 0.79‰ (VPDB;
cally blue-grey orthoquartzite with a basal quartz conglomerate mean, – 1.57‰, +1.38, N ¼ 11; crystals were drilled out and
including porphyry clasts, correlated with the base of the Torne- analysed using the technique of Halverson et al. 2005).
träsk Fm. (Björklund 1989).
Further west, on the NE side of the Tysfjord Culmination,
several local pockets of strongly sheared ,0.5-m-thick schistose Palaeolatitude and palaeogeography
unsorted diamictites with granite clasts in a dark grey matrix have
been mapped with irregular contacts on the basement (Fig. 58.1; Well-dated palaeopoles that can be used to constrain the Ediacaran
Björklund 1989). These are overlain by c. 1-m-thick white to palaeogeogeography of Baltica are scarce (cf Torsvik et al. 1996;
grey quartzites and thence by overthrust rocks of the Lower Bingen et al. 2005; Cocks & Torsvik 2005). The southern rim of an
Allochthon. inverted Baltica lay at c. 158S at c. 750 Ma (Hartz & Torsvik 2002).
At 616 Ma, data from mafic dykes in southern Scandinvia indicate
that Baltica lay at 758S although by c. 550 Ma, the area covered
Middle Allochthon here lay at c. 45–508S (Cocks & Torsvik 2005). In contrast,
Cawood & Pisarevsky (2006) and Pisarevsky et al. (2008) place
In the eastern part of the lowest imbricate of the Akkajaure Nappe Baltica in a more equatorial position (c. 308), although whether
Complex (in Karnjelajåkka; Fig. 58.1), a 1–2-m-thick diamictite it lay in the northern or southern hemisphere is unclear.
with centimetre-sized angular mesoperthitic granite clasts in a
carbonate-bearing matrix rests on an uneven granitic basement
surface. The diamictite is sharply overlain by a c. 5-m-thick Geochronological constraints
conglomerate with sub-rounded to sub-angular quartz and less
common mesoperthite clasts fining upwards to gravelly quartzite, No robust isotopic age constraints are available from this area.
followed by grey quartz phyllite, all correlated with the base of the Correlation of the sub-Dividal Group sediments with the
Torneträsk Fm. Westwards within this imbricate, a transition from Mortensnes Fm. gives a broad age constraint of c. 580 Ma, by cor-
mesoscopic clast-rich diamictites to strongly weathered calcite- relation with the Gaskiers glaciation (Bowring et al. 2003).
dolomite mica schists occurs, overlain by quartz mylonites. Clast Although it is likely that the Gaskiers event was diachronous,
size diminishes westwards. The schists have microscopic sub- the close association of the Mortensnes Fm. with extremely nega-
angular clasts of mesoperthite, quartz and albite similar to those tive d13C values in Finnmark, correlated with the Wonoka anomaly
in the diamictite; chemically, the calcite-dolomite mica schists and (Halverson et al. 2005; Rice et al. 2011), probably makes the effect
the diamictite matrix are very similar (Björklund 1989). On the of diachroneity on the timing of glaciation relatively small.
higher, more westerly derived thrust sheets, micaceous marbles The Vakkejokk Breccia is closely underlain (within a few
are interlayered with these schists. metres) by the fossil-rich Kullingia Beds, within which Treptich-
nus pedum has been recorded (Stodt 1987). Although Treptichnus
pedum was established as the index trace fossil for the base of the
Boundary relations with overlying and underlying Cambrian (Brasier et al. 1994), it has subsequently been found in
non-glacial units the youngest Ediacaran successions (Gehling et al. 2001). Thus the
Vakkejokk Breccia is constrained to very latest Precambrian to
The sediments (mostly diamictites) generally lie unconformably early Cambrian times. This boundary lies somewhat below the
on autochthonous or allochthonous Baltic Shield-derived base- lithostratigraphic correlation with the proposed Precambrian –
ment rocks and are unconformably overlain by the distinctive Cambrian boundary from E. Finnmark (Fig. 58.3).
quartz-rich basal conglomerate of the Torneträsk Fm. or
(deformed) correlatives. In some cases, a few metres of finer-
grained sediments lie between the diamictites and the basal con- Discussion
glomerate. The succession in the Komagfjord Antiformal Stack
differs in that in the south, around the Alta-Kvaengen Window, In East Finnmark, the glaciogenic Smalfjord Fm. has a typical
the diamictites lie on a condensed sequence correlated with the Marinoan-type cap dolostone, forming the base of the Nyborg
mid- to upper part of the Tanafjord Group (Fig. 58.3; Føyn 1964, Fm. (cf Halverson et al. 2005; Rice et al. 2011). Thin dolostones
1985), whilst further north they lie on allochthonous basement in the upper part of the Nyborg Fm. (Member E; Edwards 1984;
(Roberts & Fareth 1974; Pharoah 1985). The upper contact in Fig. 58.3) have d13C values of –7.6 and –9.9‰ (VPDB); these
the Komagfjord Antiformal Stack is also atypical, in that it is have been correlated with the extreme negative d13C values
gradational with the Borras Group and Slettfjell formations, recorded in the Wonoka anomaly in other parts of the world
correlatives of the Lillevatn Fm. (¼ base Torneträsk Fm.) (down to – 12‰; cf. Halverson et al. 2005, Le Guerroué et al.
in E. Finnmark (Føyn 1964; Pharaoh 1980, 1985). The upper 2006) to suggest that the almost immediately overlying Mort-
contact of the Porsavatn Diamictite Bed is erosively overlain by nesnes Fm. is broadly a correlative of the 580 Ma Gaskiers diamic-
the Lomvatn Fm. (Fig. 58.3). tite (cf Halverson et al. 2005). Correlation of the sub-Tornetäsk
THE CALEDONIDES OF NW SCANDINAVIA 609

Fm. diamictites and other ‘glaciogenic’ sediments in the present internal brecciation is indicative of rigid plugs, both typical of
area with the Mortensnes Fm. is based on the robust lithostrati- debris flows (Stodt 1987). Stodt (1987) proposed that the breccia
graphic correlation of the Stappogiedde Fm. in the Vestertana formed due to the uplift and collapse of a palaeotopographic
area with the Dividal Group in the Andabakaoaivi-Halkavarre high; based on the asymmentry of soft-sediment folds within the
area (Føyn 1967) and from there to regions to the SW (Føyn breccia, this lay to the west, with east-directed slumping.
1964; Vogt 1967; Thelander 1982, and references cited above). Similar, but slightly younger normal faulting has been documented
The limited palaeontological data (Platysolenites antiquissimus at the Caledonian front SW of Akkajaure (Hansen 1989).
Eichwald, Treptichnus pedum) support these correlations The diamictites at the northern margin of the Rombak-Sjangeli
(Kulling 1964; Hamar 1967; Føyn & Glaessner 1979; Stodt Window and in the SE corner of the Kuokkel Window also lie
1987; Crimes & McIlroy 1999; Fig. 58.3). within, rather than under, sediments correlated with the lower
The outcrops documented here are often small and rather unin- part of the Torneträsk Fm. (Bax 1984; Tull et al. 1985). Although
spiring. Some have little or no direct (reported) evidence of glacial it cannot be wholly discounted, it seems unlikely that these are
activity and it is their stratigraphic position, underlying the distinc- tillites. A more likely origin is that they are debris-flows, possibly
tive quartz-conglomerate forming the base of the Dividal Group, reworked nearby sub-Torneträsk Fm. tills or the commonly
that has been used to infer a glacial (direct or very indirect) reported regolith in the area (Stodt 1987; Bax 1989).
origin. Exposures such as the polymict conglomerates near Sko- The most northerly part of the Komagfjord Antiformal Stack, a
ganvarre (Siedlecka 1987) are clearly of this type. Similarly, the palaeo-basement high, lay well above the Mortensnes post-glacial
finer-grained, stratified deposits associated with diamictites in sea level, with drowning occurring some time after it occurred in
the Jerta Fm. (Skerlie & Tan 1960) were inferred to be fluvioglacial more southerly parts of the high; glacial sediments deposited in
solely based on the presence of diamictites; however, these are also this area would thus have been strongly affected by erosion, with
an unreliable indicator of glaciogenic facies. a low preservation potential. The Porsavatn Diamictite Bed,
Criteria used for suggesting a glacial origin include faceted lying under a sharp erosional contact with the basal conglomerates
clasts, impact pits (reflecting dropstones), soft-sediment folding of the Lomvatn Fm. is the only potential relict of such deposits
and faulting, taken to reflect glacial shear stresses, and the diamic- found (Fig. 58.3) and, if of glaciogenic origin (sensu lato) is
titic composition and texture of the rocks. In the Komagfjord most likely reworked (debris-flow), rather than primary.
Window, Pharoah (1980) recognized small roches moutonées Kulling (1951) introduced the term Varangeristiden (Varan-
forming part of the 3-m-high palaeorelief within the 40-m-deep g(er)ian ice age) in a discussion of diamictites in northern
palaeovalley. The presence of regular ‘varved’ sequences associ- Sweden. Although these deposits are now correlated solely with
ated with diamictite has also been cited as evidence of glacial the Mortensnes Fm., the term originally encompassed all the Neo-
activity (Kulling 1951). Although most of these criteria are no proterozoic glaciogenic rocks of the Smalfjord and Mortensnes
longer regarded as being unequivocally diagnostic of glacial formations in E. Finnmark (and elsewhere in Scandinavia). The
activity, their co-occurrence, together with their stratigraphic Smalfjord and Mortensnes formations have now been correlated
position, forms a reasonably compelling argument for a glacial with the Marinoan and Gaskiers glacial events based on, respect-
(sensu lato) origin; whether the diamictites are tillites or slightly ively, the development of a typical Marinoan cap dolostone and
reworked material is unknown. a close association with extreme negative d13C values taken to
Within the lowest imbricate of the Akkajaure Nappe Complex reflect the Wonoka anomaly (Halverson et al. 2005; Rice et al.
there is some evidence for a transition from coarse diamictites in 2011). One glacial event was worldwide in scope, while the
the ESE to finer-grained diamictites (now schists) in the WNW. other was much more localized. Taking a 12 Ma duration of the
Both lithologies have comparable angular mesoperthitic clasts, Marinoan glaciation (Bodiselitsch et al. 2005) and 1 Ma for the
with clast sizes decreasing to the west, and similar geochemistries. Gaskiers (Bowring et al. 2003) implies a total time span of
This has been inferred to reflect a transition from a proximal glacial 647– 579 Ma for the two glacial events, of which only 13 Ma
to a more distal, possibly glaciomarine, depositional environment (18%) were actually spent under ice, scarcely enough to justify
(Björklund 1989). the term ‘ice age’. Further, Harland et al. (1989) used the term
Kulling (1951) suggested that the Vakkejokk Breccia is a tillite, for a Precambrian epoch. In view of the very confused implications
but later studies indicated a mass-flow origin (cf Stodt 1987). A given by the term, we discourage the use of the term Varang(er)ian
peri-glacial environment has been proposed (Strömberg 1981; in any sense.
Thelander 1982) and this seemed to be confirmed by the discovery In summary, the many isolated, but regionally persistent sub-
of glendonite in the breccia (Stodt 1987). Similarly, the breccia Dividal Group diamictites in northern Scandinavia, occurring in
clasts found in the Type D breccia (breccia-in-breccia structure) the Autochthon and at all tectonic levels of the Caledonian
were interpreted to indicate that the material was frozen during nappes derived from Baltica, testify to a mid-Ediacaran glaciation
transportation. Stodt (1987) also reported up to 20-cm-diameter broadly equated with the 580 Ma Gaskiers event (which was
lonestones in the Lower Siltstone Member, under the Vakkejokk probably diachronous). Lithological, structural and mineralogi-
Breccia, although no dropstones were seen. However, detailed cal characteristics suggest terrestrial as well as glaciomarine
examination of the ‘glendonite’ does not show any evidence for depositional environments. These deposits covered an area of
the 30% volume loss associated with the breakdown of ikaite. 140 000 km2 in the region considered. Applying the same struc-
Instead, each radial ‘arm’ comprises a single, slightly distorted tural model as used here to the whole orogen (cf Gayer & Greiling
calcite crystal, with curved (strained) twin planes and radially 1989; Rice 2006) indicates that patches of Mortensnes Fm. equiva-
varying (undulose) extinction orientations. Essentially, the internal lents (Kumpulainen 2011; Kumpulainen & Greiling 2011; Nystuen
microstructure seems too regular to be a replacement of ikaite and & Lamminen 2011) cover a restored area of c. 780 000 km2.
is instead thought to represent replacement of an initial gypsum
nodule (Peckmann, pers. comm. 2011). The nodules are too irregu- A. H. N. R. thanks A. and J. Pettersen and U. and P.-Ø. Gjøvik for many years of
lar, partly due to bedding parallel pressure solution, to make a diag- hospitality during fieldwork in Finnmark, and C. and R. Hofmann for assistance in
nostic determination of the crystal form. This non-glacial the field. T.C.P.’s contribution appears with the permission of the Director, British
reinterpretation is consistent with the palaeomagnetic data and Geological Survey (Natural Environment Research Council). D. Roberts and R.
the lack of widespread glacial deposits in places nearer the poles Kumpulainen are thanked for providing information. The latter and G. Shields
at that time. are thanked for their reviews. G. Narbonne is thanked for confirming the identi-
The type A conglomerate in the Vakkejokk Breccia is likely a fication of Treptichnus pedum in the Torneträsk Fm. This represents a con-
rapidly accumulated mid-fan deposit. In the overlying breccias, tribution of the IUGS- and UNESCO-funded IGCP (International Geoscience
the parallelism of platy clasts indicates laminar flow and the Programme) Project #512.
610 F. STODT ET AL.

References Gayer, R. A. & Rice, A. H. N. 1989. Palaeogeographic reconstruction


of the pre- to syn-Iapetus rifting sediments in the Caledonides
Anderson, M. W. 1989. Basement-cover evolution during Caledonian of Finnmark, N. Norway. In: Gayer, R. A. (ed.) The Caledonide
Orogenesis, Troms, N. Norway. In: Gayer, R. A. (ed.) The Caledo- Geology of Scandinavia. Graham & Trotman, London, 127–139.
nide Geology of Scandinavia. Graham & Trotman, London, Gayer, R. A., Rice, A. H. N., Roberts, D., Townsend, C. & Welbon, A.
101– 110. 1987. Restoration of the Caledonian Baltoscandian margin from
Bax, G. 1984. Geologie des Tornehamngebietes am Westufer des Torne- balanced cross-sections: the problem of excess continental crust.
träsk, Schwedisch-Lappland. Diploma thesis, University of Marburg. Transactions of the Royal Society of Edinburgh: Earth Sciences,
Bax, G. 1989. Caledonian structural evolution and tectonostratigraphy in 78, 197– 217.
the Rombak-Sjangeli Window and its covering sequences, northern Gehling, J. G., Jensen, S., Droser, M. L., Myrow, P. M. & Narbonne,
Scandinavian Caledonides. Norges geologiske Undersøkelse Bulle- G. M. 2001. Burrrowing below the basal Cambrian GSSP, Fortune
tin, 415, 87 –104. Head, Newfoundland. Geological Magazine, 138, 213– 218.
Bax, G. 2001. Application of remote sensing techniques for the geologi- Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice,
cal mapping of tectonic klippen in the northern Scandinavian Caledo- A. H. N. 2005. Towards a Neoproterozoic composite carbon isotope
nides. In: Buchroithner, M. F. (ed.) A Decade of Trans-European record. Geological Society of America Bulletin, 117, 1181–1207.
Remote Sensing Cooperation. Balkema, Rotterdam, 245– 249. Hamar, G. 1967. Platysolenites antiquissimus Eich. (Vermes) from the
Bingen, B., Griffin, W. L., Torsvik, T. H. & Saeed, A. 2005. Timing of Lower Cambrian of northern Norway. Norges geologiske Undersø-
Late Neoproterozoic glaciation on Baltica constrained by detrital kelse, 249, 87 – 95.
zircon geochronology in the Hedmark Group, SE Norway. Terra Hansen, L. 1989. Age relationships between normal and thrust faults near
Nova, 17, 250– 258. the Caledonian front at the Vietas hydropower station, northern
Björklund, L. 1985. The Middle and Lower Allochthons in the Sweden. In: Gayer, R. A. (ed.) The Caledonide Geology of Scandi-
Akkajaure-Tysfjord area, northern Scandinavian Caledonides. In: navia. Graham & Trotman, London, 91 – 100.
Gee, D. G. & Sturt, B. A. (eds) The Caledonide Orogen — Harland, W. B., Armstrong, R. L., Cox, A. V., Craig, L. E., Smith,
Scandinavia and Related Areas. J. Wiley and Sons, Chichester, A. G. & Smith, D. G. 1989. A Geological Time Scale 1989.
515– 528. Cambridge University Press, Cambridge.
Björklund, L. 1989. Geology of the Akkajaure-Tysfjord-Lofoten tra- Hartz, E. H. & Torsvik, T. H. 2002. Baltica upside down: a new plate
verse, N. Scandinavian Caledonides. PhD thesis, Göteborg Univer- tectonic model for Rodinia and the Iapetus Ocean. Geology, 30,
sity Publ. A 59, ISSN 0348-2367. 255– 258.
Bodiselitsch, B., Koeberl, C., Master, S. & Reimold, W. U. 2005. Holmsen, P. 1956. Hyolithus-sonens basale lag i Vest-Finnmark. Norges
Estimating duration and intensity of Neoproterozoic snowball glacia- geologiske Undersøkelse, 195, 65 –72.
tions from Ir anomalies. Science, 308, 239– 242. Holmsen, P. 1957. De eokambriske lag under hyolithussonen mellen
Bowring, S., Myrow, P., Landing, E., Ramezani, J. & Grotzinger, J. Čarajavvre og Časkias, Vestfinnmark. Norges geologiske Undersø-
2003. Geochronological constraints on terminal Neoproterozoic kelse, 200, 47 – 50.
events and the rise of metazoans. Geophysical Research Abstracts, Hossack, J. R. & Cooper, M. A. 1986. Collision Tectonics in the
5, 13219. Scandinavian Caledonides. Geological Society of London, Special
Brasier, M., Cowie, J. & Taylor, M. 1994. Decision on the Precam- Publications, 19, 287–304.
brian –Cambrian boundary stratotype. Episodes, 17, 95 –100. Johnson, H. D., Levell, B. K. & Siedlecki, S. 1978. Late Precambrian
Brown, B. R. & Wells, M. L. 1966. A contribution to the geology of the sedimentary rocks in East Finnmark, North Norway and their
Vassijaure-Sjangel area of Swedish Lapland. Geologiska Förenin- relationship to the Trollfjord–Komagelva Fault. Journal of the
gens i Stockholm Förhandlingar, 87, 527– 547. Geological Society, London, 135, 517–533.
Cawood, P. A. & Pisarevsky, S. A. 2006. Was Baltica right-way-up or Kautsky, G. 1949. Eokambrische Tillitvorkommen in Norbotten, Schwe-
upside-down in the Neoprotereozoic? Journal of the Geological den. Geologiska Föreningens i Stockholm Förhandlingar, 71,
Society, London, 163, 753–759. 595– 603.
Chapman, T. J., Gayer, R. A. & Williams, G. D. 1985. Structural cross- Kirkland, C. L., Daly, S. J. & Whitehouse, M. J. 2006. Provenance and
sections through the Finnmark Caledonides and timing of the Terrane Evolution of the Kalak Nappe Complex, Norwegian Caledo-
Finnmarkian event. In: Gee, D. G. & Sturt, B. A. (eds) The Caledo- nides: Implications for Neoproterozoic Palaeogeography and Tec-
nide Orogen — Scandinavia and Related Areas. J. Wiley and Sons, tonics. Journal of Geology, 115, 21– 41.
Chichester, 593–610. Kulling, O. 1948. Om berggrunden i Sareks randområden. Geologiska
Cocks, L. R. M. & Torsvik, T. H. 2005. Baltica from the late Precambrian Föreningens i Stockholm Förhandlingar, 70, 661–672.
to mid-Palaeozoic times: the gain and loss of a terranes identity. Earth Kulling, O. 1951. Spår av Varangeristiden i Norbotten. Sveriges geolo-
Science Reviews, 72, 39 –66. giska Undersökning Årsbok 43 (1949), C 503, 1 –45.
Crimes, T. P. & McIlroy, D. 1999. A biota of Ediacaran aspect from the Kulling, O. 1964. Översikt over norra Norrbottensfjällens kaledonberg-
Lower Cambrian strata on the Digermul Peninsula, Arctic Norway. grund (with English summary). Sveriges geologiska Undersökning,
Geological Magazine, 136, 633–642. Ba 19, 166.
Edwards, M. B. 1984. Sedimentology of the Upper Proterozoic glacial Kulling, O. 1982. Översikt över södra Norrbottenfjällens Kaledonberg-
record, Vestertana Group, Finnmark, North Norway. Norges geolo- grund. Sveriges Geologiska Undersökning, Ba 26, 295.
giske Undersøkelse Bulletin, 394, 1– 76. Kumpulainen, R. A. 2011 The Neoproterozoic Lillfjället Formation,
Føyn, S. 1964. Den tillitførende formasjonsgruppe i Alta — en jevnføring southern Swedish Caledonides. In: Arnaud, E., Halverson, G. P.
med Øst-Finnmark og med indre Finnmark. Norges geologiske & Shields-Zhou, G. (eds) The Geological Record of Neoprote-
Undersøkelse, 228, 139– 150. rozoic Glaciations. Geological Society, London, Memoirs, 36,
Føyn, S. 1967. Dividal-gruppen (HYOLITHUS-sonen) i Finnmarke 629– 634.
og dens forhold til de eokambrisk-kambriske formasjoner. Norges Kumpulainen, R. A. & Greiling, R. O. 2011. Evidence for late Neopro-
geologiske Undersøkelse, 249, 1– 84. terozoic glaciation in the central Scandinavian Caledonides. In:
Føyn, S. 1985. The Late Precambrian in northern Scandinavia. In: Gee, Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
D. G. & Sturt, B. A. (eds) The Caledonide Orogen — Scandinavia logical Record of Neoproterozoic Glaciations. Geological Society,
and related areas. J. Wiley and Sons, Chichester, 233– 245. London, Memoirs, 36, 623–628.
Føyn, S. & Glaessner, M. F. 1979. Platysolenites, other animal fossils, Le Guerroué, E., Allen, P. A. & Cozzi, A. 2006. Chemostratigraphic
and the Precambrian –Cambrian transition in Norway. Norsk Geolo- and sedimentological framework of the largest negative carbon isoto-
gisk Tidsskrift, 59, 25– 46. pic excursion in Earth history: the Neoproterozoic Shuram Formation
Gayer, R. A. & Greiling, R. O. 1989. Caledonian nappe geometry in (Nafun Group, Oman). Precambrian Research, 146, 68– 92.
north-central Sweden and basin evolution on the Baltoscandian Nystuen, J. P. & Lamminen, J. T. 2011. Neoproterozoic glaciation of
margin. Geological Magazine, 126, 499– 513. South Norway: from continental interior to rift and pericratonic
THE CALEDONIDES OF NW SCANDINAVIA 611

basins in western Baltica. In: Arnaud, E., Halverson, G. P. & Siedlecka, A. 1987. Skoganvarre berggrunnskart 2034 4, 1:50,000.
Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Foreløpig utgave. Norges geologiske Undersøkelse.
Glaciations. Geological Society, London, Memoirs, 36, 613– 622. Siedlecka, A., Lyubtsov, V. V. & Negrutsa, V. Z. 1995. Correlations
Pharaoh, T. C. 1980. The geological history of the Komagfjord Tectonic between Upper Proterozoic successions in the Tanafjorden-
Window, Finnmark, Northern Norway. PhD thesis, University of Varangerfjorden Region of Varanger Peninsula, northern Norway,
Dundee. and on Sredni Peninsula and Kildin Island in the northern coastal
Pharaoh, T. C. 1985. The stratigraphy and sedimentology of autoch- area of Kola Peninsula in Russia. Norges geologiske Undersøkelse,
thonous metasediments in the Repparfjord-Komagfjord Tectonic Special Publicationi, 7, 217– 232.
Window, west Finnmark. In: Gee, D. G. & Sturt, B. A. (eds) The Siedlecka, A., Roberts, D., Nystuen, J. P. & Olovyanishnikov, V. G.
Caledonide Orogen — Scandinavia and Related Areas. J. Wiley 2004. Northeastern and northwestern margins of Baltica in Neopro-
and Sons, Chichester, 347– 357. terozoic time: evidence from the Timanian and Caledonian Orogens.
Pisarevsky, S. A., Murphy, J. B., Cawood, P. A. & Collins, A. S. 2008. In: Gee, D. G. & Pease, V. (eds) The Neoproterozoic Timanide
Late Neoproterozoic and Early Cambrian palaeogeography: models Orogen of Eastern Baltic. Geological Society, London, Memoirs,
and problems. In: Pankhurst, R. J., Trouw, R. A. J., De Brito 30, 169– 190.
Neves, B. B. & De Wit, M. J. (eds) West Canadian: Pro-Cenozoic Skjerlie, F. J. & Tan, T. H. 1960. The geology of the Caledonides of
Correlations Across the South Atlantic Region. Geological Society, the Reisa Valley area, Troms-Finnmark, Northern Norway. Norges
London, Special Publication, 294, 9 – 31. geologiske Undersøkelse, 213, 175–196.
Reading, H. G. 1965. Eocambrian and Lower Palaeozoic geology of the Stodt, F. 1987. Sedimentologie, Spurenfossilien und Weichkörper-
Digermul Peninsula, Tanafjord, Finnmark. Norges geologiske Under- Metazoan der Dividal Gruppe (Wendium-Unterkambrium) im
søkelsei, 234, 167– 191. Torneträskgebiet/Nordschweden. PhD thesis, Phillips-University
Rice, A. H. N. 1998. Stretching lineations and structural evolution of the Marburg-Lahn.
Kalak Nappe Complex (Middle Allochthon) in the Repparfjord- Stromberg, A. G. B. 1981. The Late Precambrian Sito tillite and the Vak-
Fægfjord area, Finnmark, N. Norway. Norsk Geologisk Tidsskrift, kejokk breccia in the northern Swedish Caledonides. In: Hambrey,
78, 277– 289. M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial
Rice, A. H. N. 2001. Field evidence for thrusting of the basement rocks Record. Cambridge University Press, Cambridge, 611– 614.
coring tectonic windows in the Scandinavian Caledonides; an Thelander, T. 1982. The Torneträsk Formation of the Dividal Group,
insight from the Kunes Nappe, Finnmark, Norway. Norsk Geologisk northern Swedish Caledonides. Sveriges geologiska Undersökning,
Tidsskrift, 81, 321– 328. C 789, 1– 41.
Rice, A. H. N. 2006. Quantifying the exhumation of UHP-rocks in the Torsvik, T. H., Smethurst, M. A. et al. 1996. Continental break-up and
Western Gneiss Region, S. W. Norway: a branch-line — balanced collision in the Neoproterozoic and Palaeozoic — a tale of Baltica and
cross-section model. Austrian Journal of Earth Sciences, 98, 2 –21. Laurentia. Earth Science Reviews, 40, 229– 258.
Rice, A. H. N. & Townsend, C. 1996. Correlation of the late Precambrian Townsend, C., Roberts, D., Rice, A. H. N. & Gayer, R. A. 1986. The
Ekkerøya Formation (Vadsø Group; E. Finnmark) and the Brennelvf- Gaissa Nappe, Finnmark, North Norway: an example of a deeply
jord Interbedded member (Porsangerfjord Group; W. Finnmark, eroded external imbricate zone within the Scandinavian Caledonides.
N. Norwegian Caledonides. Norsk Geologisk Tidsskrift, 76, 55 –61. Journal of Structural Geology, 8, 431– 440.
Rice, A. H. N., Edwards, M. B., Hansen, T. A., Arnaud, E. & Halver- Townsend, C., Rice, A. H. N. & Mackay, A. 1989. The structure and stra-
son, G. P. 2011. Glaciogenic rocks of the Smalfjord and Mortensnes tigraphy of the southwestern portion of the Gaissa Thrust Belt and
Formations, Vestertana Group, E. Finnmark, Norway. In: Arnaud, adjacent Kalak Nappe Complex, Finnmark, N Norway. In: Gayer,
E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological R. A. (ed.) The Caledonide Geology of Scandinavia. Graham &
Record of Neoproterozoic Glaciations. Geological Society, London, Trotman, London, 111–126.
Memoirs, 36, 593–602. Tull, J. F., Bartley, J. M., Hodges, K. V., Andresen, A., Stelten-
Rice, A. H. N., Gayer, R. A., Robinson, D. & Bevins, R. E. 1989. Strike- pohl, M. G. & White, J. M. 1985. The Caledonides in the
slip restoration of the Barents Sea Caledonides Terrane, Finnmark, Ofoten region (68 –698N), north Norway: key aspects of tectonic
North Norway. Tectonics, 8, 247– 264. evolution. In: Gee, D. G. & Sturt, B. A. (eds) The Caledonide
Roberts, D. & Fareth, E. 1974. Correlations of autochthonous stratigra- Orogen — Scandinavia and Related Areas. J. Wiley and Sons,
phical sequences in the Alta-Reppparfjord Region, West Finnmark. Chichester, 553– 568.
Norsk Geologisk Tidsskrift, 54, 123– 129. Vogt, T. 1967. Fjellkjedestudier in den ostlige del av Troms. Norge
Siedlecka, A. 1985. Development of the Upper Proterozoic sedimentary geologiske Undersøkelse, 248, 1 –59.
basins of the Varanger Peninsula, East Finnmark. Geological Survey Williams, D. M. 1976. A revised stratigraphy of the Gaissa Nappe,
of Finland Bulletin, 331, 175–185. Finnmark. Norges geologiske Undersøkelse, 324, 63 –78.
Chapter 59

Neoproterozoic glaciation of South Norway: from continental interior to rift


and pericratonic basins in western Baltica

JOHAN P. NYSTUEN* & JARKKO T. LAMMINEN


Department of Geosciences, University of Oslo, P.O. Box 1047 Blindern, NO-0316 Oslo, Norway
*Corresponding author (e-mail: j.p.nystuen@geo.uio.no)

Abstract: Neoproterozoic glacial deposits of South Norway comprise the Moelv and Koppang formations. The former occurs on
Baltican crystalline basement in autochthonous position at the Caledonian erosional nappe front, on basement windows and basement
thrust sheets in the Caledonian nappe region, and in thick sedimentary successions in the allochthonous Hedmark and Valdres rift
basins. The Koppang Formation (Fm.) occurs on top of platform carbonates in the allochthonous pericratonic Engerdalen Basin. The
glacial deposits are dominated by diamictite interpreted as basal till from warm-based grounded ice, whereas stratified successions of
diamictite beds, sandstone and laminated siltstone with outsized stones represent local ice-margin deposits and/or subglacially infilled
water bodies, and the final glaciomarine stage. Palinspastic reconstruction of Caledonian nappe complexes carrying the glacial for-
mations indicates that the glacial deposits were deposited over a wide area by a large western Baltoscandian ice sheet, probably
during the Gaskiers (c. 580 Ma) glacial event (or events), but the age of the glaciation in South Norway needs to be better constrained.

Neoproterozoic glacial deposits in South Norway have been In the present review of the Neoproterozoic glacial deposits in
recognized since Holtedahl (1922) interpreted the ‘Moelv con- South Norway, main emphasis will be given to facies, stratigraphic
glomerate’ by Lake Mjøsa as an ‘Eo-Cambrian’ tillite. The and structural position and implication for depositional extent,
‘Moelv tillite’, formally termed the Moelv Fm. (Bjørlykke et al. palaeogeography, character of glaciation, correlation and age.
1967), has since been recorded in a series of outcrops in autochtho- Geochronological and geochemical data given without any refer-
nous position and in the Caledonian Osen-Røa Nappe Complex ences correspond to the authors’ own yet unpublished results
and the Valdres Nappe Complex in the Sparagmite region of from provenance studies.
central and eastern South Norway (the term sparagmite, from
Greek ‘sparagma’ meaning fragment, was introduced by Esmark
(1829) for the arkosic Neoproterozoic sandstones that dominate Structural framework
the region). The Koppang Fm., a diamictite unit that has been cor-
related with the Moelv Fm., occurs in the Kvitvola Nappe Complex Structural setting of the Moelv Fm.
of the northern part of the Sparagmite region (Bjørlykke &
Nystuen 1981) (Figs 59.1 & 59.2). The outcrops of both formations The Moelv Fm. rests with depositional contact on crystalline base-
are scattered and generally small, from a few to several hundreds of ment rocks in three types of structural settings: (i) in autochthonous
square metres. The best exposures are present in road and railway position at the erosional nappe front at the eastern side of lake
sections and along very steep cliff sides. The largest outcrop of the Storsjøen, (ii) in thrust sheets within the Osen-Røa Nappe
Moelv Fm. measures c. 3 km in length along the western side of the Complex east of lake Storsjøen and along the river Mistra at the
mountain Fonnåsfjellet (Fig. 59.2). In its type area, the Moelv Fm. northern end of lake Storsjøen and in the Valdres Nappe
has recently been exposed in a series of new road sections and Complex at Ormtjernskampen and (iii) in the Snødøla, Atnsjøen,
bedrock surfaces in a residential area within the town of Moelv Øversjødalen and Tufsingdalen basement windows in the northern
(Nystuen 2008). part of the Sparagmite region (Siedlecka et al. 2004) (Figs 59.2 &
The Koppang Fm. (Törnebohm 1896, p. 31), being poorly 59.3). Basement underlying the diamictite in the Valdres Nappe
exposed in the Koppang type area, is best exposed in its hypostra- Complex consists of gabbroic rocks, whereas in the other settings
totype in the Engerdalen area further to the east (Nystuen 1980). the crystalline basement rocks are granitic with minor bodies of
The Moelv and Koppang formations were recorded during gabbro and dolerite (Siedlecka et al. 1987).
regional mapping in the 1960s–1980s (e.g. Loesche & Nickelsen In the window structures, Precambrian basement rocks and
1968; Nickelsen 1974; Siedlecka et al. 1987). Sedimentological a succession of late Neoproterozoic –Ordovician rocks up to
and stratigraphic studies of the Moelv Fm. were also performed c. 130 m in thickness occur beneath the Osen-Røa and Kvitvola
during this period (Bjørlykke 1966, 1969, 1974; Englund 1966, nappe complexes; the Moelv Fm. is the oldest unit in this cover
1973a; Løberg 1970; Bjørlykke et al. 1976; Nystuen 1976a, b, succession. The granitic rocks of the structural windows yield
1985; Nystuen & Sæther 1979; Nystuen & Ilebekk 1981; Sæther U –Pb zircon ages of c. 1650 Ma, whereas similar rocks at the
& Nystuen 1981; Siedlecka & Ilebekk 1982). The glaciogenic for- nappe front have been dated to 1680 Ma. The granitic rocks
mations were shown to have been deposited in four major principal belong to the younger intrusives of the Transscandinavian
depositional settings: directly upon crystalline basement and in Igneous Belt (TIB) in central southern Fennoscandia (e.g. Heim
three palaeobasins derived from the western marginal zone of Neo- et al. 1996; Högdahl et al. 2004; Andersen et al. 2009). The
proterozoic Baltica, represented by the Moelv Fm. in the Hedmark window structures in the northern part of the Sparagmite region
and Valdres basins and the Koppang Fm. in the Engerdalen Basin reveal the same lithology and stratigraphy as the autochthon at
(Nystuen 1985; Siedlecka et al. 2004). These palaeobasins, the nappe front. For this reason the window structures have been
together with similar allochthonous basins with Neoproterozoic interpreted as domes of the underlying Baltica basement and its
glacial formations in Sweden (Kumpulainen 2011; Kumpulainen sedimentary cover (Nystuen 1981). Morley (1986) suggested the
& Greiling 2011) are referred to the western Baltoscandian window structures to be allochthonous. This is supported by
basins (Kumpulainen & Nystuen 1985). regional studies of similar basement windows in the Scandinavian

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 613– 622. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.59
614 J. P. NYSTUEN & J. T. LAMMINEN

Fig. 59.1. Location of outcrop area (framed area) for the Neoprotereozoic glacial Moelv and Koppang formations in South Norway, with outcrops shown in the map
of Figure 59.2. Major geological crustal domains within the pre-glacial Fennoscandian basement are shown (modified from Bingen et al. 2008). The Caledonian
crustal segment Western Gneiss Region (WGR) comprises the northwesternmost coastal part of South Norway. The Cambrian– Devonian Caledonian domain and small
areas with Phanerozoic cover rocks in the Fennoscandian shield are not shown.

Caledonides (e.g. Gayer et al. 1987; Gayer & Greiling 1989; Gee Complex. The windows may thus represent structural elements
et al. 2010). In accordance with Morley (1986) and Rice (2005), (horsts) of large duplex structures in nappe complexes of the
the Snødøla, Atnsjøen, Spekedalen, Øversjødalen, Tufsingdalen Lower and Middle Allochthon, with the Caledonian basal thrust
and other window structures are here considered allochthonous, as the lower bounding surface, a thrust not being exposed in the
in a tectonostratigraphic position beneath the Osen-Røa Nappe northern part of the Sparagmite region.

Fig. 59.2. Simplified geological map of the


Sparagmite region in South Norway,
showing outcrops of the Neoproterozoic
glacial Moelv and Koppang formations.
The Western Gneiss Region (WGR)
mentioned in the text is located NW of the
area covered by the map (Figs 59.1 & 59.4).
Map modified from Nystuen (1987).
NEOPROTEROZOIC GLACIATION OF SOUTH NORWAY 615

Fig. 59.3. Representative examples of facies and stratigraphic context of the glacial Moelv and Koppang formations. The inset map covers the same area as the map of
Figure 59.2, in which localities are also shown. The Moelv Fm. in the Hedmark Group is in an autochthonous position in the Osen-Røa Nappe Complex and tectonic
windows: 1, Slemdalen (Sl); 2, Osdalen (Os); 3, Sjølisand (Sj); 4, Andrå (An); 5, Rena; 6, Moelv; 7, Åstdalen (Ås); 8, Kjølsjøberget (Kj); 9, Øvre Rendal (ØR);
10, Atnsjøen window (At); 11, Famphøgdene (Fa); 12, Prestkampen (Pr). Koppang Fm. in the Engerdalen Group of the Kvitvola Nappe Complex: 13, Hylleråsen (Hy);
14, Sjoa (Sjo). Equivalent to the Moelv Fm. in the Valdres Group of the Valdres Nappe Complex: 15, Ormtjernskampen East (OrE); 16, Ormtjernskampen West (OrW);
17, Mellene (Mel).

The second major structural setting of the Moelv Fm. is within basin, the Engerdalen Basin (Siedlecka et al. 2004). In this
the thick rift basin successions of the Hedmark and Valdres basins basin, sandstone, carbonate and diamictite formations have a
of the Osen-Røa and Valdres nappe complexes, respectively wide lateral extent. The Kvitvola Nappe Complex is located
(Fig. 59.2). The rift character of these allochthonous palaeobasins above the Osen-Røa Nappe Complex and beneath the Upper
is reflected by at least 5000 –6000-m-thick successions with large Allochthonous Trondheim Nappe Complex (Fig. 59.2).
lateral facies variations, from marginal alluvial or deep-marine Compared to the Moelv Fm., the Koppang Fm. is generally more
conglomerate fans to central basinal facies of either deep-marine deformed. The formation is located very close to the thrust boundary
sandstone and shale or alluvial plain sandstone, and in the of the Kvitvola Nappe Complex. Together with the underlying
Hedmark Basin also by tholeiitic basalt flows. The Moelv Fm. carbonate-shale formation, the diamictite of the Koppang Fm. has
occurs both in marginal and distal positions of the basin succes- acted as a detachment horizon during the Caledonian thrust move-
sions (Siedlecka et al. 2004). ments. In some areas, stretching and flattening of clasts makes the
In most settings where the Moelv Fm. rests on crystalline base- identification of clast lithologies very difficult; in some thrust
ment, the primary features are well preserved with little distortion zones the diamictite lithology is camouflaged by intense shear defor-
of clasts and matrix of the diamictite, except close to overlying mation. The metamorphism is of greenschist facies with chlorite and
thrust boundaries. In the Hedmark and Valdres basin successions, pale mica formed in the matrix. The diamictite of the Koppang Fm.
deformation and metamorphism generally increase from south to is best preserved in the eastern part of the Kvitvola Nappe Complex
north. In high-strain zones along thrusts and shear zones, clasts in the Engerdalen district (Nystuen 1980) (Figs 59.2 & 59.3).
in the diamictite can be flattened and elongated parallel to foliation
in the matrix. The metamorphism is of lower greenschist facies.
Non-deformed diamictite is, however, very common in most
parts of the outcrop area of the Moelv Fm.
Stratigraphy and sedimentology of the Moelv Fm.

Moelv Fm. deposited directly on Precambrian basement


Structural setting of the Koppang Fm.
The Moelv Fm. is developed as brown to grey diamictite in all
The Koppang Fm., occurring in the Kvitvola Nappe Complex, is localities where the formation rests directly on granitic basement
interpreted to have been deposited in a wide pericratonic shelf rocks, in autochthonous position, in basement thrust sheets or in
616 J. P. NYSTUEN & J. T. LAMMINEN

structural windows (Figs 59.2 & 59.3). In these settings, the dia- Englund (1973a), the gabbroic basement rock is overlain by a
mictite unit is up to c. 20 m in thickness, structureless, unsorted thin quartz- and quartzite pebble conglomerate, followed by 5–
and with clasts up to about 1 m in diameter. The matrix is 8-m-thick diamictite with clasts up to 1 m across of quartz diorite,
mud-rich sandstone. The clasts are dominated by different types granite and anorthosite. The present authors visited the Feforkam-
of granitic rocks and rhyolitic porphyries; grey and red quartzite pen locality during the summer of 2008 and found this unit to be a
and sandstone, dolerite and vein quartz are also common matrix- to clast-supported conglomerate dominated by quartzite
(Nystuen & Sæther 1979; Nystuen & Ilebekk 1981). On the north- boulders of up to c. 1.5 m.
ern side of the Atnsjøen window, the 20-m-thick diamictite is The basement –diamictite boundary in outcrops east of Lake
dominated by granitic clasts of the same type as in the basement Storsjøen and in the windows in the northern part of the Sparag-
below; however, gabbro, which is also a prominent basement mite region is a smooth unconformity with relief of up to some
rock in this window structure, is not recorded among the clast few metres. Parallel striations with direction from ESE to WNW
population (Siedlecka & Ilebekk 1982). The diamictite resting are present on the ‘polished’ granite surface below the diamictite
on gabbro in Ormtjernskampen in the Valdres Nappe Complex in the hill at Andrå east of Lake Storsjøen. The diamictite is pre-
(Ormtjernskampen West, Fig. 59.2) is dominated by gabbro served as erosional remnants, in thickness varying from some
clasts, but granite and quartzite clasts also occur in this few centimetres (Tufsingdalen) to 5–20 m (Atnsjøen, Spekeda-
2-m-thick diamictite that is succeeded by laminated siltstone and len and Øversjødalen) (Figs 59.2 & 59.3). At some localities,
quartz arenite formations correlated with the Ekre and Vangsås diamictite is overlain by lower Cambrian quartz arenite (Vangsås
formations, respectively, in the Hedmark Group (Nickelsen Fm.) and greenish grey shale, followed by Middle Cambrian
1974; Nickelsen et al. 1985). black alum shale.
A conglomerate rests on gabbro basement in the Feforkampen The uniform lithology of the diamictite, the polished and striated
outlier of the Valdres Nappe Complex (Fig. 59.2). According to crystalline substratum and the wide regional extent of the

Fig. 59.4. Palaeogeographical


reconstruction of South Norway at the time
of the Neoproterozoic glaciation, showing
the Western Baltoscandian Ice Sheet and the
Hedmark and Valdres rift basins and the
pericratonic Engerdalen Basin, together
with the present outcrop area of the Moelv
and Koppang glacial formations in the
Sparagmite region. Note the reference
locality of autochthonous diamictite of the
Moelv Fm. at the Baltica basement east of
Lake Storsjøen. The TIB domain and the
Gothian and Sveconorwegian domain are
explained in terms of geological age in
Figure 59.1. The Western Gneiss Region
(WGR) also discussed in the text comprises
the present northwestern coastal regions of
South Norway. The Engerdalen Basin was
located west of the pre-Caledonian position
of WGR.
NEOPROTEROZOIC GLACIATION OF SOUTH NORWAY 617

diamictite unit are features favouring an origin as basal till of a rocks. Recent radiometric datings and isotopic studies (U – Pb
continental ice sheet. The clast assemblage is dominated by grani- and Lu –Hf in zircons) of the igneous clast material have confirmed
tic and porphyritic rock types common for the TIB and associated that the TIB, or its equivalents in terms of age and petrogenetic
sandstones, quartzites and dolerites to the east and NE of the history, contributed significantly to the clastic material in the
present outcrop region for the Moelv diamictites. In the Valdres eastern Hedmark Basin with ages ranging from 1770 Ma to 1010
Nappe Complex the grounded ice sheet has picked up local Ma, most being c. 1670 Ma. Only one clast with an age of
bedrock types. The clast-supported conglomerate resting on top c. 1540 Ma has been found, which means that clasts of Late
of the gabbro in Feforkampen is lithologically quite different Gothian (c. 1590 –1520 Ma; Åhäll & Connelly 2008) age are
from the other diamictites and a glacial origin is here rare in the Moelv Fm. In the western part of the Hedmark Basin,
considered uncertain. the clast assemblage also includes fine-grained quartzite, grey
and pale granite, gneisses and amphibolite (Løberg 1970;
Englund 1973a). Many Sveconorwegian (Grenvillian) c.960 Ma
Moelv Fm. in the Hedmark Basin granitoid clasts occur in the western and southwestern parts of
the basin (Lamminen et al. 2009). Clasts corresponding to
The Moelv Fm. rests with depositional contact within the Hedmark typical rock types dominating in the Gothian-Sveconorwegian
Group on (1) coarse-grained fan-delta sandstones of the Ring Fm., basement adjacent to the southernmost outcrops of the Moelv
(2) carbonate platform and shale beds of the Biri Fm., (3) fluvial Fm., such as coarse-grained red granodiorites, various types of
sandstones of the Rendalen Fm., (4) shallow-marine sandstones gneisses, amphibolites and gabbros (Nordgulen 1999), are not
of the Atna Fm. and (5) basalt flows of the Svarttjørnkampen present in the diamictite. Most clasts of crystalline rock types are
Fm. (Figs 59.2 & 59.3). Except for some deep marine sub-basins, angular to sub-rounded, whereas sandstone, quartzite and vein
most of the Hedmark Basin was filled up before the onset of the quartz clasts of pebble size are skewed towards the sub-rounded
glaciation, in the east with at least 4000-m-thick alluvial conglom- to well-rounded range (Nystuen 1976a).
erates and fluvial sandstones (Rendalen Fm.) and in the west by The lithology of formations in the Hedmark Group directly
more than 4000 m of marine turbidites and black shale (Brøttum underlying the Moelv Fm. is reflected in the composition of the
Fm.), submarine conglomerate fans (Biskopåsen Fm.) and coarse- diamictite close to the lower boundary. The matrix is enriched in
grained fan-delta accumulations (Ring Fm.). Local basalt flows of coarse-grained sand particles, granules and pebbles of quartz and
the Svarttjørnkampen Fm. on top of the Rendalen Fm. and within feldspar when resting upon the coarse-clastic Rendalen and Ring
and on top of the Atna Fm. in the northern and eastern part of the formations. The basal zone of the diamictite carries small clasts
basin are interpreted as being related to fissure eruptions (Nystuen of limestone and basalt at sites where Biri limestone and basalt
1982, 1987; Furnes et al. 1983). Basalt clasts in marine canyon are the underlying stratigraphic units, respectively. Limestone
infill conglomerates of the Biskopåsen Formation located at the clasts are also abundant at localities where limestone is not directly
top of the Brøttum Fm. in the western part of the basin indicate underlying the diamictite, as at the type locality in Moelv. This
that local volcanic eruptions may have taken place repeatedly at may be due to the presence of a carbonate platform of the Biri
several places in the rift basin (Siedlecka et al. 2004). The Fm. that was located along the western margin of the Hedmark
pre-glacial basin infill culminated with the formation of carbonate Basin at the time when the diamictite was formed (Løberg 1970;
platforms, and black mud in adjacent deep sub-basins (Biri Fm.). Bjørlykke et al. 1976). Clasts of siltstone and sandstone show
Acritarchs in the upper part of the Brøttum Fm. and in the Biri glacial striations (Bjørlykke 1974; Nystuen & Sæther 1979).
Fm. indicate a late Riphean to early Vendian age (Cryogenian to The diamictite deposits of the Moelv Fm. are unsorted to poorly
Ediacaran) (Vidal & Nystuen 1990). sorted and display grain size curves very similar to those of Pleis-
The lower boundary of the Moelv Fm. is a distinct erosional tocene diamictons. Fabric studies in the eastern part of the
unconformity where the formation’s basal lithology is diamictite. Hedmark Basin reveal an east – west-oriented peak of preferred
A sand-filled wedge in underlying fluvial Rendalen sandstone is long-axis orientation that has been inferred to represent an original
interpreted as a fossil ice-wedge (Nystuen 1976b). At localities depositional fabric formed by subglacial westward transport of
where the formation rests on thick marine shales of the Biri Fm., debris (Nystuen 1976a).
the lower boundary appears gradual from shale to laminated grey The structureless diamictite is interpreted as tillite sensu
silt-rich mudstone with scattered outsized stones of the Moelv Fm. stricto, formed as basal till beneath a grounded ice sheet, as also
The Moelv Fm. usually varies in thickness from 1 to 2 m up to suggested for the similar type of diamictite in settings directly
150 m. The laminated facies with outsized stones overlying the on basement. The siliciclastic beds in the Rendalen and Ring for-
Biri shale is thinnest. The most frequently encountered thickness mations, directly underlying the Moelv Fm., may have been poorly
is in the order of 10 –15 m, and the dominant lithology is brownish, consolidated at the time of glaciation, thus supplying clastic grains
greenish or grey diamictite. In many localities, the diamictite to the till matrix.
passes upwards into laminated mudstone with outsized stones, a Successions consisting of structureless and faintly stratified dia-
facies interval that may be up to c. 2 m in thickness. An upwards mictite beds, stratified sandstone and laminated siltstone with out-
decrease in clast content marks the transition of this facies into sized stones have been interpreted as ice-margin deposits, formed
the overlying Ekre Fm., a laminated green or red siltstone at sites where water depth was close to the buoyancy depth of the
without any outsized stones. ice sheet (Nystuen 1976a). Another possible depositional environ-
The thickness increases to 100–150 m in Fonnåsfjell, Slemda- ment for these localized stratified successions may have been small
len and Brennåsen (Figs 59.2 & 59.3). In the eastern hillside of subglacial water bodies. Laminated siltstone with outsized stones
the mountain Fonnåsfjell, the formation consists of structureless within successions of mixed lithology at the top towards the over-
to faintly stratified light grey diamictite with clasts up to c. 1 m lying Ekre Fm. is thought to represent glacial mud with ice-
in diameter. In Slemdalen and Brennåsen, the c. 150-m-thick dropped debris formed in an iceberg zone. This facies probably
Moelv Fm. consists of structureless and poorly stratified brown formed during the final stage of glaciation, maybe during eustatic
diamictite beds, stratified sandstone beds with scattered pebbles sea-level rise.
and boulders, and laminated siltstone with outsized stones up to
c. 1 m in diameter.
The clast (larger than 2 cm) content of the diamictite facies Moelv Fm. in the Valdres Basin
ranges from c. 7.5 to 30% (Nystuen 1976a). The extrabasinal
clast suite in eastern and northern outcrops is very similar to that In the Valdres Basin, the best outcrops of diamictite correlated
of the Moelv diamictite located directly on crystalline basement with the Moelv Fm. are in the Ormtjernskampen area (Figs 59.2 &
618 J. P. NYSTUEN & J. T. LAMMINEN

59.3). In addition to the diamictite resting on gabbroic basement extent and setting within a pericratonic basin (Nystuen 1980;
rock, described above from the outcrop in the western part of the Kumpulainen & Nystuen 1985; Siedlecka et al. 2004). The
Ormtjernskampen area, a phyllitic diamictite unit also occurs in facies association in the underlying Hylleråsen Fm. implies that
the eastern part of the area on top of the 800-m-thick Ormtjerns- this carbonate unit was formed during warm-water to tropical con-
kampen Fm. of alluvial conglomerate and below a sandstone ditions in a carbonate platform setting including evaporitic lagoo-
unit correlated with the Vangsås Fm. in the Hedmark Group (Nick- nal sub-environments favourable for magnesite formation, anoxic
elsen 1974; Nickelsen et al. 1985) (Ormtjernskampen East, sub-basins and sabkhas.
Fig. 59.2). This diamictite contains the same clast lithologies as
the diamictite resting on the gabbroic crystalline basement in the
outcrop Ormtjernskampen West (Fig. 59.2). Chemostratigraphy
At Mellene (Figs 59.2 & 59.3), a 0.5–3-m-thick diamictite, cor-
related with the Moelv Fm., occurs on top of the 3000-m-thick No data are available from the glacial units themselves. The
Olefjell Fm. of arkosic sandstone. As with the Moelv Fm. in the distribution of major and some trace elements through the
Hedmark Basin, the diamictite is succeeded by siltstone and sand- Hedmark Group and the overlying Cambro –Silurian succession
stone, possible correlatives to the Ekre and Vangsås formations, was presented by Englund (1973b) and Bjørlykke & Englund
and Cambro-Ordovician strata (Loeschke & Nickelsen 1968). (1979). The stratigraphic resolution for the interval containing
The diamictite unit recorded within the rift basin succession the Moelv Fm. is too low for chemostratigraphic correlation.
in the Valdres Basin reveals similar lithologies and stratigraphic Tucker (1983), in his study of the Biri Fm., published stable
characteristics to those of the Moelv Fm. in the Hedmark Basin, isotope analyses with an average d18O of –11.4‰ and d13C of
which supports the interpretation of this diamictite unit as a þ1.4‰ relative to PDB, indicating that the limestone was
glacial deposit. The stratigraphic position of the Moelv Fm. in formed primarily from aragonite in seawater of normal compo-
the Valdres Basin reflects the geological setting of glaciation in sition. The data do not give any stratigraphic trend that can be
a continental rift basin with crystalline basement forming the correlated by isotope curves obtained from other Neoproterozoic
basin margins, together with coarse-clastic marginal alluvial carbonate successions.
fans, and very thick fluvial accumulations in the central part, all
settings in which glacial diamictite was deposited. Thus, in the
Ormtjernskampen area, the glacial diamictite was deposited on Palaeolatitude
top of an alluvial fan succession and on crystalline basement,
and these two sites were likely separated by a fault. In the No data are available.
Mellene area, the glacial formation was deposited in a more
central position of the rift basin.
Palaeogeography

Stratigraphy and sedimentology of the Koppang Fm. Pre-Caledonian position of the Hedmark, Valdres and
Engerdalen basins
The Koppang Fm. occurs in the lower part of the Engerdalen
Group in the Kvitvola Nappe Complex. The best outcrops of the The Hedmark and Valdres rift basins and the pericratonic
succession are in the Engerdalen area (Fig. 59.2). Engerdalen Basin were formed at the western Baltoscandian
The formation is a 5–15-m-thick diamictite and rests with margin (present coordinates) (Siedlecka et al. 2004; Nystuen
primary depositional contact on a carbonate unit, the Hylleråsen et al. 2008). Their pre-Caledonian structural setting and relative
Fm. (Fig. 59.3). The Hylleråsen Fm. is 15–30 m thick and consists position to the Baltoscandian craton is crucial for the interpretation
predominantly of dolomite with thin siltstone laminae, but also con- of the extent and character of the Neoproterozoic glaciation of
tains magnesite with chert nodules (Nystuen 1969), black shale and western Baltica. Palaeogeography of the glaciated and deposi-
laminated fine-grained sandstone, locally with dolomite pseudo- tional area during the Neoproterozoic glaciation in South
morphs after gypsum crystals (Nystuen 1980). A unit of shallow- Norway can first be inferred after palinspastic restoration of the
marine sandstone locally occurs below the carbonate formation. allochthonous successions in the Osen-Røa, Kvitvola and
The diamictite in the Koppang Fm. is structureless, except close Valdres nappe complexes.
to its upper boundary at some localities, where it may be weakly Oftedahl (1943) first quantified the shortening of the southern-
laminated. Its lower boundary is erosional, and the diamictite is most part of the Osen-Røa Nappe Complex, represented by the
enriched with dolomite clasts (up to 90% in the basal zone), with Cambro –Silurian decollement succession now preserved in the
the largest clast having a diameter of 80 cm. Extrabasinal clasts Oslo Graben, and concluded that the succession in the northern
are sub-angular to sub-rounded red and grey granite, with boulders part of the lake Mjøsa area had been dislocated c. 150 km to the
up to 1 m in diameter, rhyolitic porphyry, red sandstone, light grey SSE during the Caledonian orogeny. Nystuen (1981) showed
quartzite and vein quartz pebbles. The clast assemblage resembles that the Osen-Røa Nappe Complex, riding on top of the window
that of the Moelv Fm. in the eastern part of the Hedmark Basin. The structures in northern part of the Sparagmite region, had to be
diamictite is overlain by medium-grained, well-sorted sandstone restored to a position NW of the windows, giving a minimum
inferred to be of shallow-marine origin (Nystuen 1980). The thrust distance of Osen-Røa Nappe Complex rocks at the erosional
upper contact towards the overlying sandstone formation is nappe front within the Mjøsa-Trysil sector (Fig. 59.2) of c. 140 km.
lithologically sharp and often tectonically deformed. Morley (1986) restored hanging- and footwall cut-offs within
The carbonate–diamictite–sandstone succession can be traced the Osen-Røa Nappe Complex and obtained a minimum thrust dis-
along the sole of the Kvitvola Nappe Complex for c.140 km from tance of 130 km of rocks in the nappe complex at the northern part
west to east (Figs 59.2 & 59.3). The Hylleråsen Fm. at its base is of Lake Mjøsa, about the same shortening as quantified by Ofte-
commonly cut by the sole thrust of the nappe complex and has dahl (1943). Morley (1986) restored the window structures to a
obviously acted as a detachment horizon during emplacement of position NW of the restored position of the Hedmark Basin,
the Kvitvola Nappe Complex. The carbonate–diamictite–sandstone outside the present coastline of mid-Norway. This restoration
tripartite succession is frequently strongly tectonized and deformed. model was further developed by Rice (2005), who included the
A glacial origin is inferred for the Koppang Fm. from its diamic- window structures in a hypothetical ‘Jostedal platform’ of
tite lithology and clast content, erosional lower boundary with Baltica basement northwestward of the Hedmark Basin.
clasts derived from underlying formation, its laterally wide However, the thrust position of the Osen-Røa Nappe Complex
NEOPROTEROZOIC GLACIATION OF SOUTH NORWAY 619

above the structural windows and their sedimentary cover succes- processes (Hannah et al. 2008). Because the Re –Os system is sus-
sion shows that the windows must be restored to a position craton- ceptible to disturbances that are difficult to detect, unlike the U – Pb
ward of the Hedmark Basin, and not the opposite as proposed in the system in zircon, the present authors think that a robust age for the
models of Morley (1986) and Rice (2005). Moelv Fm. still remains to be established (see below).
Hossack et al. (1985) restored the package of thrust sheets in the
Valdres area to the NW by c. 315 km, west and NW of the original
position of the Hedmark Basin. The Kvitvola Nappe Complex with Discussion
the Engerdalen Basin, located on top of the Osen-Røa Nappe
Complex, has been displaced an even greater distance, probably The western Baltoscandian ice sheet
in the order of 400 –600 km, from the NW to the SE (Siedlecka
et al. 2004). The Caledonian nappe pile in South Norway is According to the tectonostratigraphic setting of the Moelv and
located on top of the Western Gneiss Region of the southern Scan- Koppang glacial formations and the palinspastic restoration of
dinavian Caledonides (Fig. 59.1). Consequently, the nappe com- the Caledonian nappe complexes carrying the Neoproterozoic
plexes in discussion were likely derived from positions west of basins (see above), the area covered with glacial ice must have
the Western Gneiss Region. In the gneiss region, Precambrian had a width of at least 400–600 km from SE to NW, measured
basement rocks with overlying and infolded Neoproterozoic to from the autochthonous reference outcrop of the Moelv Fm. at
lower Palaeozoic sedimentary cover successions and Caledonian Storsjøen to the restored position of the Engerdalen Basin
thrust sheets are strongly deformed and metamorphosed. Rice at the Baltoscandian margin (Figs 59.2 –59.5). This western
(2005) restored the rocks of the Western Gneiss Region to the Baltoscandian ice sheet also covered carbonate platforms of the
NW as one single crustal block close to the position of the Risbäck rift basin and the pericratonic Tossåsfjället Basin along
Valdres Basin and west of the Hedmark Basin. However, the het- the Baltoscandian margin (Kumpulainen 2011; Kumpulainen &
erogeneous structural architecture of the Western Gneiss Region Greiling 2011).
indicates that the displacement history of this region and its Clast lithologies of the Moelv and Engerdalen formations indi-
cover of sedimentary successions and thrust sheets are very com- cate that the glacial ice streams were sourced from areas located
plicated (Tucker et al. 2004), and accordingly difficult to within the TIB and the Sveconorwegian domain in the central,
restore palinspastically. and the western to south-western part of Fennoscandia, respect-
It is beyond the scope of this paper to discuss various palinspas- ively. The Hedmark and Valdres rift basins probably formed
tic models of the southern Scandinavian Caledonides. However, along zones of weakness between these crustal blocks (Lamminen
irrespective of models, it must be concluded that during Neoproter- et al. 2009). Uplifted shoulders of the rift basins may have given
ozoic times when the Moelv and Koppang formations were depos- rise to locally glaciated mountain ridges (Figs 59.4 & 59.5).
ited, a rather well-denudated Baltica craton extended several Ice streams that originated in highland areas must have coalesced
hundred kilometres further west of the outcrop east of lake Stors- into extensive ice sheets that moved across lowland regions, now
jøen where the Moelv diamictite is located autochthonously on represented by denuded Baltica basement and the carbonate
striated basement rocks, a reference locality of crucial importance platform of the Engerdalen Basin, both types of substratum
for any reconstruction of the palaeogeography of Neoproterozoic being covered by glaciogenic diamictite. Striations on the granitic
glaciation in South Norway. The Hedmark and Valdres rift surface beneath diamictite east of lake Storsjøen demonstrate the
basins and the Engerdalen shelf basin were located in outboard warm-based character of the glacier ice at the time. Lithoclasts
positions of the denuded Baltica. and the textural and mineralogical composition of basal tills are
generally dominated by the local substratum. This is also revealed
by the concentration of carbonate clasts in the lower part of the
Geochronological constraints Koppang Fm. where the diamictite rests directly on dolomite,
and by basalt clasts in diamictite of the Moelv Fm. at sites where
No radiometric age data are available to unequivocally constrain this formation has been deposited upon basalt of the Svarttjørn-
the ages of the Moelv and Koppang formations. Rankama (1973) kampen Fm. (Figs 59.2 & 59.3). The absence of Gothian gneisses,
reported a Rb – Sr whole-rock age of 612 + 18 Ma on an argilla- which are typical in the autochthonous basement, among clasts in
ceous sample from the Ekre Fm., overlying the Moelv Fm. in the the Moelv Fm. at Mjøsa is thus in accordance with the allochtho-
western part of the Hedmark Basin. Bingen et al. (2005) identified nous position of the glacial formation in the area.
a 677 + 15 to 620 + 14 Ma old detrital zircon U – Pb age popu- The sedimentological character of the carbonate units under-
lation in the Rendalen Fm. at Hanestad in the eastern part of the lying the Moelv and Koppang formations, the Biri and Hylleråsen
Hedmark Basin (Fig. 59.2). This is a very uncommon age range formations, respectively, indicates dominating warm and arid
within Fennoscandian crystalline rocks, and the source for these climatic conditions before the onset of glaciation. The presence
zircon grains remains enigmatic. Within the range of this age popu- of fossil ice wedges in alluvial sediments beneath the Moelv Fm.
lation, a granite in the allochthonous Seve Nappe in northern shows that periglacial climatic conditions prevailed in the area in
Sweden gave a U –Pb titanite age of 637 + 3 Ma; the granite front of the advancing ice sheet.
was considered to be related to terrane accretion in the western-
most part of Baltica during the Timanian orogeny (Rehnström
et al. 2002). Recent detrital zircon U –Pb analyses from the upper- Correlation and age
most parts of the Brøttum Fm. have yielded a wide range of Neo-
proterozoic ages. The same c. 620 Ma population also shows up in The Moelv and Koppang formations are generally correlated with
the Brøttum Fm. Because the Brøttum Fm. underlies the Moelv the Mortensnes Fm. in Finnmark, the younger of the two Neopro-
Fm., this confirms the findings of Bingen et al. (2005) that the terozoic glacial units traditionally referred to the ‘Varangerian
Moelv Fm. appears to be younger than c. 620 Ma. glacial period’ in northern Norway, and to the Långmarkberg
Hannah et al. (2007) reported a c. 560 Ma Re –Os age for a black and Lillfjället glacial formations in Sweden (e.g. Bjørlykke &
shale in the Biri Fm., beneath the Moelv Fm. in the Mjøsa area. Nystuen 1981; Kumpulainen & Nystuen 1985; Siedlecka et al.
However, the same authors also reported a c. 300 Ma age for the 2004; Nystuen et al. 2008). Halverson et al. (2005) correlated
same formation at another locality further NE. The latter age the Mortensnes Fm. with the Gaskiers glaciation and the older
was explained by thermal disturbance in the Re – Os isotopic glacial Smalfjord Fm. with the Marinoan glaciation, from d13C
system due to the Permian Oslo rift. These authors later investi- profiles obtained from carbonate deposits associated with
gated scatter in their Re –Os data and related it to oxidation the two formations. The Marinoan glacial units are dated to
620 J. P. NYSTUEN & J. T. LAMMINEN

Fig. 59.5. Simplified and composite section across the western Baltoscandian craton and the Hedmark and Valdres rift basins and the pericratonic Engerdalen Basin,
showing the Western Baltoscandian Ice Sheet from the cratonic interior in the SE to the Baltoscandian Margin in the NW, with local areas of glaciation along mountain
ridgers along the rift basins. Note the reference locality of autochthonous diamictite of the Moelv Fm. at the Baltica basement east of Lake Storsjøen.

c. 635 Ma (Hoffmann et al. 2004) and the Gaskiers and Squantum basin successions, at least not as glacial diamictites. In this
glacial deposits of Avalonia to c. 580 Ma (Myrow & Kaufman connection, it should be emphasized that the various glacial for-
1999; Thompson & Bowring 2000; Bowring et al. 2003; Thomp- mations in South Norway, Finnmark of North Norway, and
son et al. 2007). The Marinoan glacial episode may have lasted Sweden are lithostratigraphically, not chronostratigraphically, cor-
at least 3 million, and most likely 12 million, years (Bodiselitsch related. A Gaskiers age is not presently unequivocally constrained
et al. 2005). Duration and variation in age for the Gaskiers for the formations, neither by radiometric age determinations,
glacial event or events are also very uncertain. nor by d13C studies on carbonate units beneath the glacial units.
On the basis of d13C values in cap carbonates of the youngest of If the Ottfjället dolerite dykes were shown to have about the
two diamictite units in Svalbard, Halverson et al. (2004) correlated same age as the Egersund dyke swarm (c. 616 Ma), this would
the Wilsonbreen Fm. with the Marinoan glacial event and the put the age of the glacial Lillfjället Fm., and maybe also the
inferred global extent of this glaciation, in accordance with Moelv Fm., to be older than Gaskiers. It should also be emphasized
the Snowball Earth hypothesis (Hoffman & Schrag 2002). In the that the detrital zircon U –Pb age of c. 620 Ma is problematic in the
British-Irish Caledonides, three stratigraphic levels of diamictite sense that no likely source of these zircons is known, and this age
have been correlated as equivalents to the c. 700 Ma Sturtian, appears to belong to a zircon U – Pb age population with a wide
c. 635 Ma Marinoan and c. 580 Ma Gaskiers glacials, respectively time range (Bingen et al. 2005). The meaning of this is not known.
(McCay et al. 2006). Improved constraining of the depositional age for Neoprotero-
Bingen et al. (2005) suggested that the Moelv Fm. corresponds zoic glacial formations in central and southern Scandinavia may
to the Gaskiers glacial event, constrained by the 620 + 14 Ma det- be obtained by radiometric age determination of the Svarttjørn-
rital zircon age obtained from the Rendalen Fm. at Hanestad. The kampen basalt directly underlying the Moelv Fm. and the Ottfjället
Lillfjället Fm. in the Tossåsfjället Basin, Sweden, an alleged cor- dolerites cutting the Lillfjället Fm. Such studies are in progress. In
relative to the Moelv Fm. (Nystuen et al. 2008), is penetrated by addition, d13C studies of carbonate formations underlying the
the Ottfjället dolerite dyke swarm (Kumpulainen 1980, 2011). glacial formations (cf. Siedlecka et al. 2004; Nystuen et al.
This dyke swarm is generally considered to be the result of pre- 2008) should be performed.
Iapetus rifting at the western Baltoscandian craton, contempora-
neous with the intrusion of the 616 + 3 Ma Egersund dolerite
dyke swarm (Bingen et al. 1998) in southwestern Norway (e.g. Conclusions
Siedlecka et al. 2004; Nystuen et al. 2008). There are presently
several data and arguments favouring a mid-Ediacaran (possible The Neoproterozoic Moelv and Koppang glacial formations were
Gaskiers equivalent) age for Neoproterozoic glaciation in South mainly formed as basal till from a warm-based, grounded, western
Norway and Sweden, represented by glacial diamictite formations. Baltoscandian ice sheet that covered large parts of the northwes-
The glacial formations are located in the uppermost part of tern Baltica, and the sedimentary Hedmark, Engerdalen and
Neoproterozoic successions, with the Moelv Fm. above a strati- Valdres basins. Accumulation areas of the glacial ice were in the
graphic level in the uppermost part of the Brøttum Fm. that con- zone of the TIB in the east and NE, and highland areas dominated
tains c. 620 Ma zircon grains, and are overlain by Ediacaran by rocks with metamorphic and igneous ages corresponding to the
and/or Lower Cambrian formations. The glacial deposits also Gothian and the Sveconorwegian (Grenvillian) orogeneses in the
lack cap carbonates, typical of most Gaskiers glacial deposits SE and SW. Local ice-marginal and/or subglacial waterbody suc-
according to Halverson et al. (2005). cessions were formed in deeper parts of the rift basins. Laminated
If all glacial diamictite units in central western Sweden and mudstone with ice-dropped stones originated towards the end of
in South Norway are Gaskiers equivalents, then deposits of the the glaciation due to a rise in sea level and the change from
allegedly global end-Cryogenian or Marinoan glacial period grounded to floating ice sheet margins and icebergs. The glaciation
appear not to be represented within these western Baltoscandian may correspond to the c. 580 Ma Gaskiers glacial event of
NEOPROTEROZOIC GLACIATION OF SOUTH NORWAY 621

Avalonia, but this interpretation still needs to be better constrained sequence, Southern Norway. Norges geologiske undersøkelse, 286,
by radiometric age determinations of igneous rocks associated 1– 60.
with the glacial successions and d13C studies on carbonate for- Esmark, J. 1829. Reise fra Christiania til Trondhjem. Christiania,
mations underlying the glacial formations. Oslo, 81.
Furnes, H., Nystuen, J. P., Brunfelt, A. O. & Solheim, S. 1983.
We thank R. Bäckmark and H. B. Totland for making drawings (Figs 59.2, 59.3 Geochemistry of Upper Riphean-Vendian basalts associated with
& 59.5). The manuscript has benefited from the criticism of an earlier version the ‘sparagmites’ of southern Norway. Geological Magazine, 120,
by H. Rice and by corrections of the English text by A. Read. We also thank 349– 361.
editor G. Shields for corrections and suggestions for improvements to this contri- Gayer, R. A. & Greiling, R. O. 1989. Caledonian nappe geometry in
bution. This represents a contribution of the IUGS- and UNESCO-funded IGCP north-central Sweden and basin evolution of the Baltoscandian
(International Geoscience Programme) Project #512. margin. Geological Magazine, 126, 499–513.
Gayer, R. A., Rice, A. H. N., Roberts, D., Townsend, C. & Welbon, A.
1987. Restoration of the Caledonian Baltoscandian margin from
References balanced cross-sections: the problem of excess continental crust.
Transactions of the Royal Society of Edinburgh, 78, 197– 217.
˚ häll, K.-I. & Connelly, J. N. 2008. Long-term convergence along
A Gee, D. G., Juhlin, C., Pascal, C. & Robinson, P. 2010. Collision
SW Fennoscandia: 330 m.y. of Proterozoic crustal growth. Pre- orogeny in the Scandinavian Caledonides (COSC). Geologiska Före-
cambrian Research, 163, 402–421. ningens i Stockholm Förhandlingar, 132, 29– 44.
Andersen, T., Andersson, U. B., Graham, S., A ˚ berg, G. & Simonsen, Halverson, G. P., Maloof, A. C. & Hoffman, P. F. 2004. The Marinoan
S. 2009. Granitic magmatism by melting juvenile continental crust: glaciation (Neoproterozoic) in northeast Svalbard. Basin Research,
new constraints on the source of Palaeoproterozoic granitoids in 16, 297– 324.
Fennoscandia from Hf isotopes in zircon. Journal of Geological Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
Society, London, 166, 1 –15. Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
Bodiselitsch, B., Koeberl, C., Master, S. & Reimold, W. U. 2005. isotope record. Geological Society of America Bulletin, 117,
Estimating duration and intensity of Neoproterozoic snowball glacia- 1181–1207.
tions from Ir anomalies. Science, 308, 239–242. Hannah, J. L., Yang, G., Bingen, B., Stein, H. J. & Zimmerman, A.
Bingen, B., Demaiffe, D. & Van Breemen, O. 1998. The 616 Ma old 2007.  560 Ma and 300 Ma Re –Os ages constrain Neoprotero-
Egersund basalt dike swarm, SW Norway, in the context of late- zoic glaciation and record Variscan hydrocarbon migration on exten-
Neoproterozoic opening of the Iapetus Ocean. Norsk geologisk tids- sion of Oslo rift. Goldschmidt Conference Abstracts 2007, A378.
skrift, 79, 69 – 86. Hannah, J. L., Yang, G., Xu, G., Zimmerman, A., Stein, H. J. & Egenh-
Bingen, B., Griffin, W. L., Torsvik, T. H. & Saeed, A. 2005. Timing of off, S. O. 2008. Re –Os Isotopic disturbances at unconformities:
Late Neoproterozoic glaciation on Baltica constrained by detrital challenges and opportunities. American Geophysical Union, Fall
zircon in the Hedmark Group, south-east Norway. Terra Nova, 17, Meeting 2008, abstract #PP31C-1519.
250– 258. Heim, M., Skiöld, T. & Wolff, F. C. 1996. Geology, geochemistry and
Bingen, B., Andersson, J., Söderlund, U. & Möller, C. 2008. The age of the ‘Tricolor’ granite and some other Proterozoic (TIB)
Mesoproterozoic in the Nordic countries. Episode, 31, 29 –34. granitoids at Trysil, southeast Trysil. Norsk geologisk tidsskrift, 76,
Bjørlykke, K. 1966. Studies on the Latest Precambrian and Eocambrian 45– 54.
rocks in Norway. 1. Sedimentary petrology of the Sparagmites of Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
the Rena district, S. Norway. Norges geologiske undersøkelse, 238, testing the limits of global change. Terra Nova, 14, 129– 155.
5 – 53. Hoffman, K.-H, Conodon, D. J., Bowring, S. A. & Crowley, J. L.
Bjørlykke, K. 1969. Geologien i sentrale deler av Österdalen. Norsk geo- 2004. U– Pb zircon date from the Neoproterozoic Ghaub Formation,
logisk tidsskrift, 49, 313– 318. Namibia: constraints on Marinoan glaciation. Geology, 32, 817– 820.
Bjørlykke, K. 1974. Glacial striations on clast from the Moelv Tillite Holtedahl, O. 1922. A tillite-like conglomerate in the ‘Eo-Cambrian’
of the Late Precambrian of Southern Norway. American Journal of Sparagmite of Southern Norway. American Journal of Science, 4,
Science, 274, 443– 448. 165– 173.
Bjørlykke, K. & Englund, J. O. 1979. Geochemical response to upper Hossack, J. R., Garton, M. R. & Nickelsen, R. P. 1985. The geological
Precambrian rift basin sedimentation and lower Palaeozoic epiconti- section from the foreland up to the Jotun thrust sheet in the Valdres
nental sedimentation in South Norway. Chemical Geology, 27, area, south Norway. In: Gee, D. G. & Sturt, B. A. (eds) The Cale-
271– 295. donide Orogen – Scandinavia and Related Areas. John Wiley &
Bjørlykke, K. & Nystuen, J. P. 1981. Late Precambrian tillites of South Sons, Chichester, 443–456.
Norway. In: Hambrey, M. J. & Harland, W. B. (eds) Earth’s Högdahl, K., Andersson, U. B. & Eklund, O. (eds) 2004. The Trans-
Pre-Pleistocene Glacial Record. Cambridge University Press, scandinavian Igneous Belt (TIB) in Sweden; A Review of its Character
Cambridge, 624–628. and Evolution. Geological Survey of Finland, Special Paper, 37, 123.
Bjørlykke, K., Englund, J.-O. & Kirkhusmo, L. 1967. Latest Kumpulainen, R. 1980. Upper Proterozoic stratigraphy and depositional
Precambrian and Eocambrian stratigraphy of Norway. Norges geo- environments of the Tossåsfjället Group, Särv Nappe, southern
logiske undersøkelse, 251, 5– 17. Swedish Caledonides. Geologiska Föreningen i Stockholm Förhand-
Bjørlykke, K., Elvsborg, A. & Høy, T. 1976. Late Precambrian sedi- lingar, 102, 531–550.
mentation in the central sparagmite basin of South Norway. Norsk Kumpulainen, R. A. 2011. The Neoproterozoic Lillfjället Formation,
geologisk tidsskrift, 56, 233–290. southern Swedish Caledonides. In: Arnaud, E., Halverson, G. P.
Bowring, S. A., Myrow, P. M., Landing, E., Ramezani, J. & Grotzin- & Shields-Zhou, G. (eds) The Geological Record of Neopro-
ger, J. 2003. Geochronological constraints on terminal Neoprotero- terozoic Glaciations. Geological Society, London, Memoirs, 36,
zoic events and the rise of metazoans. Geophysical Research 629– 634.
Abstract, 5, 13219. Kumpulainen, R. A. & Greiling, R. O. 2011. Evidence for Neoproter-
Englund, J.-O. 1966. Sparagmittgruppens bergarter ved Fåvang, ozoic glaciation in central Scandinavian Caledonides. In: Arnaud,
Gudbrandsdalen-En sedimentologisk og tektonisk undersøkelse. E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological
Norges geologiske undersøkelse, 238, 55 – 103. Record of Neoproterozoic Glaciations. Geological Society, London,
Englund, J.-O. 1973a. Stratigraphy and structure of the Ringebu-Vinstra Memoirs, 36, 623–628.
district, Gudbrandsdalen; with a short analysis of the western part of Kumpulainen, R. & Nystuen, J. P. 1985. Late Proterozoic basin evol-
the sparagmite region in Southern Norway. Norges geologiske ution and sedimentation in the westernmost part of Baltoscandia.
undersøkelse, 293, 1 –58. In: Gee, D. G. & Sturt, B. A. (eds) The Caledonide Orogen –
Englund, J.-O. 1973b. Geochemistry and mineralogy of pelitic Scandinavia and Related Areas. John Wiley & Sons, Chichester,
rocks from the Hedmark Group and the Cambro-Ordovician 213– 232.
622 J. P. NYSTUEN & J. T. LAMMINEN

Lamminen, J., Nystuen, J. P. & Andersen, T. 2009. U –Pb ages and International Geological Congress, Oslo, Norway, available at
Lu –Hf isotopes of granitoid clasts and basement rocks of the http://www.33igc.org/coco/LayoutPage.aspx, 59.
Hedmark Group: implications for structural setting of the Hedmark Nystuen, J. P. & Sæther, T. 1979. Clast studies in the Late Precambrian
Basin at the Baltoscandian margin. NGF Vinterkonferansen 2009, Moelv Tillite and Osdal Conglomerate, Sparagmite Region, south
Bergen 13 – 15 January 2009. Abstracts and Proceedings of the Geo- Norway. Norsk geologisk tidsskrift, 59, 239– 251.
logical Society of Norway, 60. Nystuen, J. P. & Ilebekk, S. 1981. Stratigraphy and Caledonian struc-
Loeschke, J. & Nickelsen, R. P. 1968. On the age and tectonic position tures in the area between the Atnsjøen and Spekedalen windows,
of the Valdres Sparagmite in Slidre (Southern Norway). Neues Jahr- Sparagmite Region, southern Norway. Norsk geologisk tidsskrift,
buch für Geologie und Paläontologie, 131, 337– 367. 61, 17 – 24.
Løberg, B. E. 1970. Investigations at the south-western border of the Nystuen, J. P., Andresen, A., Kumpulainen, R. & Siedlecka, A. 2008.
sparagmite basin (Gausdal Vestfjell and Fåberg Vestfjell), Southern Neoproterozoic basin evolution in Fennoscandia, East Greenland
Norway. Norges geologiske undersøkelse, 266, 160– 205. and Svalbard. Episodes, 31, 35– 43.
McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial Oftedahl, C. 1943. Overskyvninger i den norske fjellkjede. Naturen
trinity: Neoproterozoic Earth history within the British– Irish Caledo- (Oslo), 5, 143–150.
nides. Geology, 34, 909–912. Rankama, K. 1973. The Late Precambrian glaciation, with particular
Morley, C. K. 1986. The Caledonian thrust front and palinspastic restor- reference to the Southern Hemisphere. Journal of Proceedings of
ations in the southern Norwegian Caledonides. Journal of Structural the Royal Society of New South Wales, 106, 89– 97.
Geology, 8, 753– 765. Rehnström, E. F., Corfu, F. & Torsvik, T. H. 2002. Evidence of a Late
Myrow, P. M. & Kaufman, A. J. 1999. A newly discovered cap carbonate Precambrian (637 Ma) deformational event in the Caledonides of
above Varanger-age glacial deposits in Newfoundland, Canada. Northern Sweden. Journal of Geology, 110, 591– 601.
Journal of Sedimentary Research, 69, 784– 793. Rice, A. H. N. 2005. Quantifying the exhumation of UHP-rocks in the
Nickelsen, P. R. 1974. Geology of the Røssjøkollan – Dokkvatn area, Western Gneiss Region, S.W. Norway: a branch-line – balanced
Oppland. Norges geologiske undersøkelse, 314, 53– 100. cross section model. Austrian Journal of Earth Sciences, 98, 2– 21.
Nickelsen, P. R., Garton, M. & Hossack, J. R. 1985. Late Precambrian Sæther, T. & Nystuen, J. P. 1981. Tectonic framework, stratigraphy,
to Ordovician sedimentology and stratigraphic correlation of the sedimentation and volcanism of the late Precambrian Hedmark
Valdres and Synnfjell thrust sheets in the Valdres area, central Group, Østerdalen, south Norway. Norsk geologisk tidsskrift, 61,
Norway. In: Gee, D. G. & Sturt, B. A. (eds) The Caledonide 193– 211.
Orogen – Scandinavia and Related Areas. John Wiley & Sons, Siedlecka, A. & Ilebekk, S. 1982. Forekomster av tillitt på nordsiden av
Chichester, 369–378. Atnsjøvinduet, Syd-Norge. Norges geologiske undersøkelse, 373,
Nordgulen, Ø. 1999. Geologisk kart over Norge, berggrunnskart 33 – 37.
(bedrock map) HAMAR, M 1:250 000. Norges geologiske Siedlecka, A., Nystuen, J. P., Englund, J.-O. & Hossack, J. 1987.
undersøkelse. LILLEHAMMER – berggrunnskart M. 1:250 000. Norges geolo-
Nystuen, J. P. 1969. On the paragenesis of chert and carbonate minerals giske undersøkelse.
in chert-bearing magnesitic dolomite from the Kvitvola nappe, Siedlecka, A., Roberts, D., Nystuen, J. P. & Olovyanishnikov, V. G.
Southern Norway. Norges geologiske undersøkelse, 258, 66 –78. 2004. Northeastern and northwestern margins of Baltica in Neopro-
Nystuen, J. P. 1976a. Facies and sedimentation of the Late Precambrian terozoic time: evidence from the Timanian and Caledonian
Moelv Tillite in the eastern part of the Sparagmite Region, southern Orogens. In: Gee, D. G. & Pease, V. (eds) The Neoproterozoic Tima-
Norway. Norges geologiske undersøkelse, 329, 1 – 70. nide Orogen of Eastern Baltica. Geological Society, London,
Nystuen, J. P. 1976b. Late Precambrian Moelv Tillite deposited on a Memoirs, 30, 169–190.
discontinuity surface associated with a fossil ice wedge, Rendalen, Thompson, M. D. & Bowring, S. A. 2000. Age of the Squantum ‘tillite’,
southern Norway. Norsk geologisk tidsskrift, 56, 29 –56. Boston basin, Massachusetts: U–Pb zircon constraints on terminal
Nystuen, J. P. 1980. Stratigraphy of the Upper Proterozoic Engerdalen Neoproterozoic glaciation. American Journal of Science, 300,
Group, Kvitvola Nappe, southeastern Scandinavian Caledonides. 630– 655.
Geologiska Föreningen i Stockholm Förhandlingar, 102, 551– 560. Thompson, M. D., Grunow, A. M. & Ramezani, J. 2007. Late Neopro-
Nystuen, J. P. 1981. The Late Precambrian ‘Sparagmites’ of southern terozoic paleogeography of the Southeastern New England Avalon
Norway: a major Caledonian allochthon – The Osen-Røa Nappe Zone: insights from U –Pb geochronology and paleomagnetism.
Complex. American Journal of Science, 281, 69– 94. Geological Society of America Bulletin, 119, 681–696.
Nystuen, J. P. 1982. Late Proterozoic basin evolution on the Baltoscan- Tucker, M. E. 1983. Sedimentation of organic-rich limestones in the
dian Craton: the Hedmark Group, southern Norway. Norges geolo- late Precambrian of southern Norway. Precambrian Research, 22,
giske undersøkelse, 375, 1 –74. 295– 315.
Nystuen, J. P. 1985. Facies and preservation of glaciogenic sequences Tucker, R. D., Robinson, P. et al. 2004. Thrusting and extension in the
from the Varanger Ice Age in Scandinavia and other parts of the Scandian hinterland, Norway: new U– Pb ages and tectonostrati-
North Atlantic region. Palaeogeography, Palaeoclimatology, Palaeo- graphic evidence. American Journal of Science, 304, 477– 532.
ecology, 51, 209–229. Törnebohm, A. E.1896. Grunddragen af det centrala Skandinaviens
Nystuen, J. P. 1987. Synthesis of the tectonic and sedimentological bergbyggnad. Kongl. svenska vetenskaps-akademiens handlingar,
evolution of the late Proterozoic-early Cambrian Hedmark Basin, 28, 1 – 210.
the Caledonian Thrust Belt, southern Norway. Norsk geologisk tids- Vidal, G. & Nystuen, J. P. 1990. Micropaleontology, depositional
skrift, 67, 395– 418. environment, and biostratigraphy of the Upper Proterozoic
Nystuen, J. P. 2008. Neoproterozoic Moelv Tillite and the Hedmark Hedmark Group, southern Norway. American Journal of Science,
Basin, Mjøsa Area, South Norway. 33 IGC Excursion No 101, 33rd 290-A, 170–211.
Chapter 60

Evidence for late Neoproterozoic glaciation in the central Scandinavian Caledonides

RISTO A. KUMPULAINEN1* & REINHARD O. GREILING2*


1
Department of Geological Sciences, Stockholm University, SE-106 91 Stockholm, Sweden
2
Institut für Angewandte Geowissenschaften, Abteilung Strukturgeologie und Tektonophysik,
Karlsruher Institut für Technologie (KIT), Hertzstraße 16, D-76187 Karlsruhe, FR Germany
*Corresponding authors (e-mail: risto.kumpulainen@geo.su.se; er8@agk.uka.de)

Abstract: The Långmarkberg Formation (Fm). rests on either the Neoproterozoic rift-related Risbäck Group or Palaeoproterozoic base-
ment in central Scandinavia. It is always succeeded in stratigraphy by the marine Gärdsjön Fm. The Långmarkberg Fm. is a thin, gen-
erally less than 50-m-thick, discontinuous, but persistent unit, primarily in the Lower Allochthon; subordinately also in the Autochthon,
and Middle Allochthon of the central Scandinavian Caledonides. The formation is composed of a lower diamictite and an upper lami-
nated, lonestone-bearing mudstone.
Criteria supporting the glaciogenic origin of the formation are (i) a large regional extension of the formation (.5000 km2) and (ii) its
fixed position in the regional stratigraphy, (iii) the ‘tillite-like’ appearance of the lower part of the formation, (iv) the lonestone-bearing
mudstone in the upper part, (v) the presence of striated clasts and (vi) striated pavement. It is interpreted as having formed during one
single glacial–deglacial cycle.
Information on chemostratigraphy and realistic geochronology from this region is missing. The Långmarkberg Fm. has commonly
been correlated with the Moelv Fm. in southern Norway and the Mortensnes Fm. in northernmost Norway. The Lower Allochthon in
this region also contains an older, potentially glacially influenced diamictite unit at the base of the Risbäck Group, which is also
briefly discussed.

The Långmarkberg Fm. composed of diamictite and mudstone is in southern Norway. Subsequently, Kulling (1942) described about
part of the Jämtland Supergroup (Gee 1975), which comprises 20 diamictite sections in this region. Although unable to produce a
the Neoproterozoic to Lower Palaeozoic sedimentary succession proper description of the section at Långmarkberget (Fig. 60.1b;
of the Caledonian Lower Allochthon (Fig. 60.1a) in central Scan- 648320 N/158210 E), Kulling (1942) selected it to be the type
dinavia. The formation occurs in small outcrops, which typically section and named the unit the Långmarkberg Fm. Recent
may be up to a few hundred square metres in size and extend up mapping in the area reveals that this type section is complicated
to a few hundred metres along strike. The formation is a thin, dis- tectonically (Zachrisson 1997d ). Work in the late 1960s to
continuous, but regionally persistent unit resting on either the mid-1990s revealed more outcrops, which amount today to more
Risbäck Group or the Palaeoproterozoic basement and is always than 60. Some of the other sections are much better exposed
overlain by the Gärdsjön Fm. (Fig. 60.2). The preserved thickness than Långmarkberget. For this reason the unit was renamed the
of the unit varies from less than a metre commonly to less than Dabbsjön Fm. (Gee et al. 1974) after a new type section at Dabbs-
20 m, but ranges locally to more than 50 m. The majority of the jön (Fig. 60.1b; 648410 N/158180 E). However, encouraged by the
Långmarkberg Fm. outcrops occur within the imbricated Lower ‘rule of priority’, Gee et al. (1978) revived the name Långmark-
Allochthon and are indicated (by stars) in Figure 60.1. Palinspastic berg Fm. and the sections at Dabbsjön remained reference sections.
restoration of the imbricated thrust units to the original pre- The work up to about the mid-1970s was summarized by Thelan-
tectonic (pre-Caledonian) basin(s) suggests that the minimum der (1981). Some additional sedimentological information derives
depositional area exceeded 5000 km2 (Kumpulainen & Nystuen from geological mapping of the areas north of Bijelite and Malgo-
1985; Gayer et al. 1987; Gayer & Greiling 1989, Siedlecka et al. maj (Greiling & Zachrisson 1999a, b; Greiling et al. 1999b).
2004; Nystuen et al. 2008). Only a few outcops of the formation
are known between the main study area (Fig. 60.1b) and Laisvall
in the north, where a diamictite unit of uncertain origin is described Structural framework
from the Autochthon (Ackerselet Fm., Willdén 1980; Greiling
et al. 1999a). The unit reappears again in the Akkajaure area The Scandinavian Caledonides are a typical thrust-and-fold belt
further north (Stodt et al. 2011). Towards the south, only a few out- and are subdivided into the Caledonian Autochthon, Lower,
crops of the formation are known in the Lower Allochthon Middle, Upper and Uppermost Allochthons (e.g. Gee et al.
(Fig. 60.1a), before reaching southeastern Norway (Nystuen & 1985). Only the Autochthon and Lower and Middle Allochthons
Lamminen 2011). The Middle Allochthon (Gayer & Greiling are relevant for the discussion in this chapter.
1989; Greiling & Zachrisson 1999a, b) hosts sections of the Lång- In the Autochthon of central Scandinavia, the sedimentary suc-
markberg Fm. NE of Stalon (Fig. 60.1b) and south of Östersund cession of sandstone and shale of Cambrian age was deposited on
(Fig. 60.1a). the passive Baltica margin (Willdén 1980; Greiling et al. 1999a).
This chapter focuses primarily on the Långmarkberg Fm. The succession is subhorizontal and undeformed. The alteration in
between Ströms Vattudal and Dikanäs (Fig. 60.1b). However, dia- the Autochthon is diagenetic with peak temperatures below 200 8C
mictite units of uncertain origin, which occur locally at the base of (Warr et al. 1996).
the Risbäck Group, are also discussed. The Lower Allochthon is composed of the Neoproterozoic to
Historically, a ‘tillite-like sedimentary breccia’ was described by Lower Palaeozoic sedimentary succession of the Jämtland Super-
Asklund & Thorslund (1934) in Sjoutälven (old spelling, Sjougdäl- group (Gee 1975) and slices of the Palaeoproterozoic basement.
ven; Fig. 60.1b) of the Lower Allochthon of west-central Sweden. The Jämtland Supergroup represents (Fig. 60.2) a gradual develop-
Later, Asklund (1938) named this unit ‘Tillitabteilung/Tillit Fm.’ ment from an initial continental rift setting (Risbäck Group) to
and correlated it with the Moelv Conglomerate of Holtedahl (1922) passive, Iapetus marginal setting (Tåsjön Group) along the

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 623– 628. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.60
624 R. A. KUMPULAINEN & R. O. GREILING

Fig. 60.1. (a) Simplified geological map of


the central Scandinavian Caledonides. (b)
Simplified geological map of the Ströms
Vattudal–Dikanäs areas, modified from
Kumpulainen & Thelander (1978), Gee
et al. (1985), Zachrisson (1997a, b, c, d) and
the sources referred to in the text.

western (present coordinates) Baltica margin (Kumpulainen & conglomerate, sandstone and shale of probable Neoproterozoic
Nystuen 1985; Gayer & Greiling 1989). Subsequent Caledonian age. The deformation is polyphase, including a pervasive cleavage
orogenic shortening formed hinterland dipping duplexes and in the pelitic rocks (Greiling 1985; Greiling & Zachrisson 1999a,
imbricate fans (Gee et al. 1978; Gayer & Greiling 1989; Zachris- b). Primary sedimentary features are preserved only in competent,
son & Greiling 1993a, b; Febbroni 1997; Greiling & Zachrisson psammitic rocks.
1999a, b; Greiling et al. 1999b, c). In the interior parts of the
fold belt, antiformal stacks of the Lower Allochthon are exposed
in tectonic windows, for example Børgefjellet, Bångonåive and Stratigraphy
Nasafjäll (Fig. 60.1), (Greiling 1988; Greiling et al. 1993). The
sedimentary sequence is faulted and sheared, with one or two clea- The Neoproterozoic to Lower Palaeozoic sedimentary succession,
vages in shaly rocks. Metamorphism is of sub-greenschist grade the Jämtland Supergroup (Gee 1975), is subdivided into the
(anchimetamorphism) in the marginal fold belt (Warr et al. Risbäck, Sjoutälven, Tåsjön and Änge Groups (Fig. 60.2).
1996) and up to biotite grade in the tectonic windows (Greiling The Neoproterozoic Risbäck Group (Gee et al. 1974) is domi-
et al. 1993; our data). Primary sedimentary structures and textures, nated by alluvial siliciclastic rocks. It may reach a thickness of
as well as fossil assemblages, are widely preserved. c. 1.5 km, but wedges out laterally to nothing. In its lower part,
Similar to the Lower Allochthon, the Middle Allochthon the Cryogenian to Cambrian Sjoutälven Group is composed of
is formed of imbricate units with crystalline basement of the glaciogenic Långmarkberg Fm. (Kulling 1942; Thelander
Palaeo- and Mesoproterozoic age and metamorphosed units of 1981), which is stratigraphically always succeeded by the
CENTRAL SCANDINAVIAN CALEDONIDES 625

Fig. 60.2. Jämtland Supergroup


stratigraphy (Gee 1975) with outlines of the
geological development of the area.

Gärdsjön Fm. (Gee et al. 1974). The overlying Tåsjön and Änge upper part is mudstone-dominated. The boundary from the lower
groups are not discussed here. unit to the upper unit may be gradational or sharp (Fig. 60.3).
Sedimentary facies types in the rift-related Risbäck Group The thickness of the formation varies from less than a metre gen-
suggest a general change from a proximal alluvial facies, domi- erally to c. 20 m (Fig. 60.3), but thicknesses exceeding 50 m have
nated by arkosic sandstone with minor (braided-stream-type) been encountered at Bergvattnet (Fig. 60.1b) and Jillesnåle
conglomerate, mudstone + occasional beds of poorly sorted (Fig. 60.1a); there are no systematic changes of the thickness in
matrix-supported conglomerate in the east to a marine facies of any particular direction.
sandstone and shale including a dolomite unit, the Kalvberget The coarse-clastic lower part of the formation may be composed
Fm., in the west and NW (Kumpulainen & Nystuen 1985; of up to three diamictite units that are poorly sorted, structureless
Nystuen et al. 2008). In its type area the dolomite is microcrystal- and matrix-supported, although changes in the grain size of the
line, but contains scattered euhedral dolomite crystals (,1 mm). In matrix may produce an indistinct stratification. The poorly sorted
other areas it is strongly recrystallized, and a network of quartz matrix in the diamictite varies from clayey to coarse sandy. The
veins occurs also in Kalvberget. proportion of matrix and framework clasts varies greatly. Stratified
The Långmarkberg Fm. forms a thin, discontinuous, but sandstone- or mudstone beds with lonestones up to c. 30 cm in
regionally persistent unit and rests on the Risbäck Group. diameter occur between diamictite beds.
Outside the Risbäck basin, the Långmarkberg Fm. was deposited The size of the clasts in the diamictite varies from pebble to
on top of the slightly undulating Palaeoproterozoic basement boulder, with the largest boulder measuring 2 m in diameter.
(Fig. 60.2). Arkose and other feldspar-bearing sandstone, quartzite, mudstone
The Långmarkberg Fm. is, in all outcrops if not disturbed by
fault surfaces, succeeded by the Gärdsjön Fm. formed during
several transgression –regression cycles in a shallow-marine
setting. Regressions produced quartz-arenite units, whereas trans-
gressions deposited mudstone-dominated units (e.g. Kumpulainen
& Nystuen 1985; Greiling et al. 1999a).
The Gärdsjön Fm. rocks are distinctly different, in lithology and
facies, from those in the Risbäck Group, so mistaking these strati-
graphic units in an outcrop is unlikely, also outside the Ströms
Vattudal –Dikanäs area (Fig. 60.1b), despite the fact that this sedi-
mentary succession is tectonically imbricated.

Glaciogenic deposits and associated strata

The Långmarkberg Fm.

The description of the Långmarkberg Fm. in this chapter is essen-


tially a summary of the general characteristics of the formation
based on Thelander (1981), Kulling (1942) and, north of Bijelite-
Malgomaj, on Greiling & Zachrisson (1999a, b). On the basis of
colour and clast content, Kulling (1942) subdivided the Långmark- Fig. 60.3. Sections through the Långmarkberg Fm. in the Dabbsjön –
berg Fm. into a lower diamictite and an upper diamictite. This subdi- Långvattnet areas. The locations are shown in Figure 60.1b: 1– 3, on the
vision has not been confirmed by later studies (Thelander 1981). The northeastern shore of Dabbsjön (648410 N/158180 E), redrawn from Thelander
stratigraphy and facies in the formation vary greatly from one section (1981); 4, south of Blaikliden, south of Marsån (65820 400 N/158470 700 E); 5, SE
to another (Fig. 60.3). However, generally (for exception, see Grytsjö, south of the Grytsjö–Blaikliden road (65800 2100 N/158360 4600 E); 6, east
Fig. 60.3, section 5), the lower part of the formation is dominated of Grytsjö, road section at north side of road Grytsjö– Blaikliden (65800 3400 N/
by coarse-clastic, matrix-supported diamictite beds, whereas the 158370 1000 E); 7, western shore of Långvattnet (658100 1100 N/168310 5300 E).
626 R. A. KUMPULAINEN & R. O. GREILING

and dolomite are the dominant clast types of sedimentary origin; dominates the group, whereas units of greenish grey to dark grey
gneiss clasts of sedimentary origin also occur. The majority of the mudstone are subordinate. Dolomites of the Kalvberget Fm. up
crystalline rocks are represented by various types of granite, felsic to c. 100 m thick are a discontinuous but significant rock unit in
porphyry and syenite. With few exceptions, the clast content in the upper part of the group.
the lower part of the diamictite sections reflects the composition A unit of diamictite in the Flakaträsk and Vojmsjön areas
of the substratum. The proportion of crystalline clasts increases (Fig. 60.1), described by Kulling (1942) as ‘red conglomerate’,
up-section and they always dominate in the lonestone-bearing, lami- underlies the Risbäck Group composed of arkose and feldspathic
nated upper mudstone unit. The proportion of crystalline clasts in the quartzite (Febbroni 1997; Greiling & Zachrisson 1999b; Greiling
formation varies from 10 to 100%, whereas the proportion of sedi- et al. 1999c). At Vojmsjön, a similar matrix-supported diamictite
mentary clasts varies from 0 to 90 %. Sub-angular to sub-rounded contains boulders of c. 5– 50 cm in diameter (Grambow 2001,
clasts dominate, but shapes range from angular to well-rounded. unpublished diploma thesis, Heidelberg University). The boulders
Some stones are faceted and a few display striations (Kulling 1942). are composed of granitoid gneiss (44%), rhyolite (36%), vein
Laminated mudstone with or without normal grading is a common quartz (10%), meta-arkose and meta-pelite (5%), epidote frag-
component in these diamictite sections, particularly in the upper part ments (4%) and amphibolite (1%). Most of these boulders are
of the formation. Occasionally, the mudstone unit contains lenses of angular or sub-angular; no striated boulders have been reported.
muddy diamictite and some stratified gravel beds. Thin sandy beds
are particularly common in the basal part of a mudstone unit. Lone-
stones range from sand-grade to boulders in size. The preserved The Gärdsjön Fm.
thickness of this facies type is generally in the order of 5–10 m.
The Gärdsjön Fm. always succeeds the Långmarkberg Fm. The
Gärdsjön Fm. has been divided (Gee et al. 1974) into ten informal
Vindelälven River subunits, which may be named M1 to M10; five are sandstone
members, representing sea-level lowstands, and the remaining five
Diamictite outcrops in the Lower Allochthon north of the Vindeläl- are mudstone-dominated members, representing sea-level high-
ven River at Jillesnåle rest on the Risbäck Group (Grambow 2001, stands. Tectonic deformation disturbs the preservation of sedimen-
unpublished diploma thesis, Heidelberg University). It is similar in tary structures, but in many sections they are common, and readily
composition to the Långmarkberg Fm. diamictite, although its discernible in both rock types. The Gärdsjön Fm. is a common com-
exact stratigraphic relationship remains uncertain. Basal, lami- ponent in the Lower Allochthon and certain units, particularly a
nated mudstone, c. 5 m thick, with lonestones is overlain by maroon mudstone member in the middle of the formation, and
more than 50 m of coarse, matrix-supported diamictite. This dia- may be traced long distances along most of the thrust belt (e.g. The-
mictite contains boulders of meta-arkose (26%), syenite (26%), lander 1982; Greiling et al. 1999a). Owing to the onlapping charac-
rhyolite (19%), quartzite (18%), granite (10%) and meta-pelite ter of the Gärdsjön Fm., only some of the upper members are present
(1%). Ten percent of the boulders are angular, 75% sub-angular in the Autochthon (Fig. 60.1b), such as the Storuman–Laisvall area,
and 15% rounded. No striated stones are reported. where they are referred to as the Såvvovare Fm. (Willdén 1980;
Eliasson et al. 2003; Greiling et al. 1999b).
Bångonåive window
Boundary relations with overlying and underlying
Diamictite of uncertain origin and stratigraphic position is exposed
non-glacial units
in the Lower Allochthon of the Bångonåive window (Fig. 60.1a)
NE of Tärnaby (648430 4700 N/158240 1400 E) at the base of the Oltok-
The Långmarkberg Fm.
ken Fm. (Stephens 1977) interpreted as an equivalent to the Lång-
markberg and Gärdsjön formations combined (Greiling et al.
The Långmarkberg Fm. rests on three different rock types: (i) Pro-
1993). The diamictite forms a lensoidal body, with a maximum
terozoic crystalline rocks, (ii) arkose and feldspathic sandstone of
thickness of c. 1 m, only a few metres stratigraphically above
the Risbäck Group and (iii) dolomite of the Kalvberget Fm. The
the crystalline substratum. The exposed lateral extent is in the
lower boundary of the Långmarkberg Fm. at a regional scale con-
order of 100 m. The diamictite is composed of a fine sandy
forms to bedding in the underlying Risbäck Group. At a local scale,
matrix and enclosed angular clasts of quartzite, feldspathic quart-
grooves and linear depressions and disconformities are observed
zite and subordinate felsic crystalline rocks. The clast size varies
(Thelander 1981; Greiling & Kumpulainen 2004). Långmarkber-
from a few centimetres to decimetres.
get (Fig. 60.1b) exposes a low-angle disconformity at the top of
the Risbäck Group. In a minor outcrop SE of Långmarkberget,
Børgefjellet window bedding in the Risbäck Group is cut at a high angle by the for-
mation boundary, which Thelander (1981) and Gee et al. (1990)
Close to the eastern margin of the Børgefjellet window (Fig. 60.1a), suggested was a major unconformity. However, soft-sediment
a diamictite unit of uncertain stratigraphic position overlies the deformation occurs frequently in this part of the Risbäck Group
Proterozoic basement at the northern slope of the Sipmekfjäll hill and it is possible that the erosional surface only incidentally trun-
(658430 4700 N/158240 1400 E; Greiling 1988). The unit, c. 1 m thick, cated an overturned soft-sediment fold in the Risbäck Group.
occurs in a tectonic slice and extends laterally for a few hundred Asklund (1960) reported a striated pavement below the Långmark-
metres. The diamictite is characterized by a quartz-rich, shaly berg Fm., but later that interpretation was questioned by Crowell
matrix, which is intensely cleaved and contains angular fragments (1964). However, still later, Thelander (1981) reported such struc-
of white vein quartz and quartzite up to c. 10 cm in diameter. The dia- tures in this region.
mictite is overlain by white, coarse feldspathic quartzite, interpreted In most cases and independent of the substratum, the boundary
as an equivalent of the Gärdsjön Fm. (Gayer & Greiling 1989). is sharp. A close correlation of clast lithologies in the basal parts of
the diamictites and their substratum additionally indicates that the
boundary must be erosional. Only locally, at Långvattnet
The Risbäck Group and diamictites older than the Risbäck Group (Fig. 60.3), the Kalvberget Fm. dolomites display a gradual tran-
sition up to a diamictite (of uncertain origin). The diamictite, in
In the Risbäck area, the type area of the Risbäck Group, a succes- turn, grades upwards into a finer-grained clastic sediment with
sion of maroon to grey, arkose and feldspathic sandstone coarse framework grains of quartz in a dolomite matrix, and then
CENTRAL SCANDINAVIAN CALEDONIDES 627

to the Gärdsjön Fm. quartzite (Fig. 60.3); the lonestone-bearing environment. Occasional lensoidal bodies of diamictite with
mudstone is missing here. With this exception, the upper contact muddy matrix within this upper laminated unit are probable mass-
of the Långmarkberg Fm. with the Gärdsjön Fm. is sharp. Ero- flow deposits or were dumped from icebergs.
sional features are encountered locally. The lower boundary of the formation is a regional, erosional,
low-angle disconformity, which is substantiated by the presence
Diamictite units older than the Risbäck Group (side views) of grooves and other minor linear irregularities.
The diamictite units and their substratum are so indurated that
The diamictite units at the base of the Risbäck Group rest on crys- the possible glacially striated contact surface is only rarely avail-
talline basement rocks; the exact contact is not exposed, but was able. This also limits the findings of striated lonestones. The
probably deposited in shallow depressions of the basement upper boundary is commonly sharp and erosional and overlain
surface. The upper boundary of the diamictite is gradational to by the basal quartz arenite (occasionally conglomerate-bearing)
red arkose of the Risbäck Group. This diamictite is distinctly of the Gärdsjön Fm.
red, which implies staining by iron oxides, mostly hematite and The Långmarkberg Fm. has been correlated stratigraphically
goethite, similar to the overlying arkose and feldspathic sandstone with the upper tillite horizon, the Mortensnes Fm. in northern
of the Risbäck Group. However, magnetic susceptibility in the dia- Norway (Siedlecka et al. 2004) and with the Moelv Fm. in south-
mictite is exceptionally high, in the order of 5–10  1023 SI eastern Norway (Asklund & Thosrslund 1934; Thelander 1981;
units (Febbroni 1997, our data). Such high susceptibilities point Siedlecka et al. 2004; Nystuen et al. 2008; Nystuen & Lamminen
to an additional magnetite content of up to 1%. The magnetite is 2011), but the scarce isotopic evidence derived from this particular
distributed both in the clasts (boulders) of crystalline basement region of Scandinavia precludes such correlations and remains
rocks and in the matrix of the diamictite. speculative.

Chemostratigraphy Possible glaciogenic diamictite at the base of the


Risbäck Group
No data are available.
The ‘red conglomerate’ at the base of the Risbäck Group was inter-
preted by Kulling (1942) as an ‘in water redeposited conglomerate,
Palaeolatitude and palaeogeography which originally may have been a till’. In addition, it is peculiar,
because of its characteristic magnetite content (Febbroni 1997)
Restorations by Greiling & Smith (2000), Hartz & Torsvik (2002) which has not been observed in the overlying sedimentary succes-
or Cawood & Pisarevsky (2006) imply high southern latitudes for sion. Magnetite is not stable under present atmospheric or shallow-
the present area at the time of Långmarkberg Fm. deposition. water conditions (e.g. Piper 1987; Chan et al. 2005). It is therefore
speculated that magnetite formation and/or stability in this sedi-
mentary environment may be due to exceptional atmospheric/cli-
Geochronological constraints
matic conditions and perhaps related to a Snowball Earth situation
(e.g. Hoffman & Schrag 2002). Therefore, the magnetite contents,
No data are available.
if related to a Snowball Earth situation, may also be a further indi-
cation for a diamictite horizon at the base of the Risbäck Group in
Discussion other areas. The red conglomerates at the base of the Risbäck
Group may then correspond to the lower tillite horizon (Smalfjord
There are potentially two different glaciogenic diamictite units in Fm.) in northern Norway.
this region, one on top of the Risbäck Group and the other below
the Risbäck Group. R.O.G. acknowledges support from the Geological Survey of Sweden. Some
of the data presented here were acquired during the PNASTINA project.
Sorsele Kommun supported the work in the Jillesnåle/Vindelälven areas.
Långmarkberg Fm. R.A.K. acknowledges support from Stockholm University. The manuscript was
improved by the comments of E. Arnaud, an anonymous reviewer and
Although thin and discontinuous at a regional scale, the Långmark- G. A. Shields. This represents a contribution of the IUGS- and UNESCO-
berg Fm. can be traced all across the Cryogenian to Cambrian suc- funded IGCP (International Geoscience Programme) Project #512.
cession from the Autochthon, Lower and Middle Allochthons of
the central Scandinavia Caledonides and further north and south.
The interpretations of the various facies types within the Lång- References
markberg Fm. rely essentially on Hambrey & Harland (1981)
and Hambrey (1994). Asklund, B. 1938. Hauptzüge der Tektonik und Stratigraphie der mittle-
The deposition of the Långmarkberg Fm. has been interpreted ren Kaledoniden in Schweden. Sveriges geologiska undersökning C,
(Thelander 1981) in the framework of one glacial–deglacial 417, 99.
cycle. The various sections through the formation vary greatly in Asklund, B. 1960. Studies of the thrust region of the southern part of the
facies, although generally poorly sorted matrix-supported diamic- Swedish mountain chain. International Geological Congress, XXI
Session Norden, Guide to the Excursions, A24, C19.
tite occupies the basal part of the sections. The massive units may
Asklund, B. & Thorslund, P. 1934. Fjällkedjerandens bergbyggnad i
have been deposited by meltout, dumping or redeposition subgla-
norra Jämtland och Ångermanland. Sveriges geologiska undersökn-
cially or in a proglacial marine setting. The poorly stratified dia- ing C, 382, 110.
mictite may have been deposited subglacially from a grounded Cawood, P. A. & Pisarevsky, S. A. 2006. Was Baltica right-way-up or
ice-sheet. The stratified, sandy to gravelly beds in or between dia- upside down in the Neoproterozoic? Journal of the Geological
mictite beds could have been deposited from glaciofluvial traction Society London, 163, 753–759.
currents beneath temporarily floating ice or close to the sub- Chan, M. A., Bowen, B. B., Parry, W. T., Ormö, J. & Komatsu, G.
aqueous grounding line of a glacier. 2005. Red rock and red planet diagenesis: comparisons of Earth
The upper, mudstone-dominated part displays distinct lami- and Mars concretions. GSA Today, 15, 4 – 10.
nation originating from subaqueous deposition at some distance Crowell, J. C. 1964. Climate significance of sedimentary deposits
from floating ice or a grounded ice margin. Melting icebergs trans- containing disperced megaclasts. In: Nairn, A. E. M. (ed.) Problems
ported clastic debris, dropstones up to boulder size, to this muddy in Palaeoclimatology. Interscience, London, 86– 99.
628 R. A. KUMPULAINEN & R. O. GREILING

Eliasson, T., Greiling, R. O. & Triumf, C.-A. 2003. Bedrock map 24H Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis:
Sorsele SV, 1:50,000. Sveriges geologiska undersökning Ai, 188. testing the limits of global change. Terra Nova, 14, 129–155.
Febbroni, S. 1997. Rilevamento geologico e dei valori della suscettivita Holtedahl, O. 1922. A tillite-like conglomerate in the ‘Eo-Cambrian’
magnetica nelle Caledonidi della Lapponia svedese. AZ Marmi, Sparagmite of Southern Norway. American Journal of Science, 4,
131, 50. 165– 173.
Gayer, R. A. & Greiling, R. O. 1989. Caledonian nappe geometry in Kulling, O. 1942. Grunddragen av fjällkedjerandens bergbyggnad
north-central Sweden and basin evolution on the Baltoscandian inom Västerbottens län. Sveriges geologiska undersökning C,
margin. Geological Magazine, 126, 499– 513. 445, 320.
Gayer, R. A., Rice, A. H. N., Roberts, D., Townsend, C. & Welbon, A. Kumpulainen, R. & Thelander, T. 1978. Geological map of the
1987. Restoration of the Caledonian Baltoscandian margin from area between Ströms Vattudal and Malgomaj, northern Jämtland–
balanced cross-sections: the problem of excess continental crust. southern Västerbotten, central Swedish Caledonides. Sveriges geo-
Transactions of the Royal Society of Edinburgh: Earth Sciences, logiska undersökning BRAP, 97001 (unpublished map 1:100,000).
78, 197– 217. Kumpulainen, R. & Nystuen, J. P. 1985. Late Proterozoic basin
Gee, D. G. 1975. A geotraverse through the Scandinavian Caledonides – evolution and sedimentation in the westermost part of Baltoscandia.
Östersund to Trondheim. Sveriges geologiska undersökning C, In: Gee, D. G. & Sturt, B. A. (eds) The Caledonide Orogen – Scan-
717, 66. dinavia and Related Areas. John Wiley & Sons, Chichester,
Gee, D. G., Karis, L., Kumpulainen, R. & Thelander, T. 1974. A 213– 232.
summary of the Caledonian front stratigraphy, northern Jämtland/ Nystuen, J. P. & Lamminen, J. T. 2011. Neoproterozoic glaciation of
southern Västerbotten, central Swedish Caledonides. Geologiska South Norway: from continental interior to rift and pericratonic
Föreningens i Stockholm Förhandlingar, 96, 389– 397. basins in western Baltica. In: Arnaud, E., Halverson, G. P. &
Gee, D. G., Kumpulainen, R. & Thelander, T. 1978. The Tåsjön Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
Décollement, central Swedish Caledonides. Sveriges geologiska Glaciations. Geological Society, London, Memoirs, 36, 613–622.
undersökning C, 742, 35. Nystuen, J. P., Andresen, A., Kumpulainen, R. & Siedlecka, A. 2008.
Gee, D. G., Kumpulainen, R., Roberts, D., Stephens, M. B., Thon, A. Neoproterozoic basin evolution in Fennoscandia, East Greenland and
& Zachrisson, E. 1985. Scandinavian Caledonides –Tectonostrati- Svalbard. Episodes, 31, 35 –43.
graphic Map, Scale 1:2,000,000. Sveriges geologiska undersökning Piper, J. D. A. 1987. Palaeomagnetism and the Continental Crust. Open
Ba, 35. University Press, Milton Keynes.
Gee, D. G., Kumpulainen, R. & Karis, L. 1990. Fjällrandens berggrund. Siedlecka, A., Roberts, D., Nystuen, J. P. & Olovyanishnikov, V. G.
In: Lundqvist, T., Gee, D. G., Kumpulainen, R., Karis, L. & 2004. Northeastern and northwestern margins of Baltica in Neopro-
Kresten, P. Beskrivning till berggrundskartan över Västernorrlands terozoic time: evidence from the Timanian and Caledonian
län. Sveriges geologiska undersökning Ba, 31, 206–237. Orogens. In: Gee, D. G. & Pease, V. (eds) The Neoproterozoic Tima-
Greiling, R. O. 1985. Strukturelle und metamorphe Entwicklung an der nide Orogen of Eastern Baltica. Geological Society, London,
Basis grosser, weittransportierter Deckeneinheiten am Beispiel des Memoir, 30, 169– 190.
Mittleren Allochthons in den zentralen Skandinavischen Kaledoniden Stephens, M. B. 1977. Stratigraphy and relationship between folding,
(Stalon-Deckenkomplex in Västerbotten, Schweden). Geotekto- metamorphism and thrusting in the Tärna-Björkvattnet area, northern
nische Forschungen, 69, 129. Swedish Caledonides. Sveriges geologiska undersökning C, 726,
Greiling, R. O. 1988. Ranseren, berggrunnskart 2025/3, 1:50,000, 1 –146.
foreløpig utgave. Norges geologiske undersökelse. Stodt, F., Rice, A. H. N., Björklund, L., Bax, G., Halverson, G. P. &
Greiling, R. O. & Zachrisson, E. 1999a. Bedrock map 23 G Dikanäs Pharaoh, T. C. 2011. Evidence of Neoproterozoic glaciation in the
NW, 1:50,000. Sveriges geologiska undersökning Ai, 122. Caledonides of NW Scandinavia. In: Arnaud, E., Halverson,
Greiling, R. O. & Zachrisson, E. 1999b. Bedrock map 23 G Dikanäs G. P. & Shields-Zhou, G. (eds) The Geological Record of Neopro-
SW, 1:50,000. Sveriges geologiska undersökning Ai, 123. terozoic Glaciations. Geological Society, London, Memoirs, 36,
Greiling, R. O. & Smith, A. G. 2000. The Dalradian of Scotland: missing 603– 611.
link between the Vendian of northern and southern Scandinavia? Thelander, T. 1981. The late Precambrian Långmarkberg Formation,
Physics and Chemistry of the Earth, 25, 495– 498. central Swedish Caledonides. In: Hambrey, M. J. & Harland, W.
Greiling, R. O. & Kumpulainen, R. 2004. Spår av världens största B. (eds) Earth’s Pre-Pleistocene Glacial Record. Cambridge Univer-
inlandsis i vilhelminafjällen? Geologiskt Forum, 43, 8– 11. sity Press, Cambridge, 615–619.
Greiling, R. O., Gayer, R. A. & Stephens, M. B. 1993. A basement Thelander, T. 1982. The Torneträsk Formation of the Dividal Group,
culmination in the Scandinavian Caledonides formed by antiformal northern Swedish Caledonides. Sveriges geologiska undersökning
stacking (Bångonåive, northern Sweden). Geological Magazine, C, 789, 1 –41.
130, 471– 482. Warr, L. N., Greiling, R. O. & Zachrisson, E. 1996. Thrust-
Greiling, R. O., Jensen, S. & Smith, A. G. 1999a. Vendian– Cambrian related, very low-grade metamorphism in the marginal part of
subsidence of the passive margin of western Baltica – application of an orogenic wedge, Scandinavian Caledonides. Tectonics, 15,
new stratigraphic data from the Scandinavian Caledonian margin. 1213– 1229.
Norsk Geologisk Tidsskrift, 79, 133– 144. Willdén, M. Y. 1980. Paleoenvironment of the autochthonous sedimen-
Greiling, R. O., Zachrisson, E., Thelander, T. & Sträng, T. 1999b. tary rock sequence at Laisvall, Swedish Caledonides. Stockholm Con-
Bedrock map 23 G Dikanäs NE, 1:50,000. Sveriges geologiska under- tributions in Geology, 33, 1– 100.
sökning Ai, 124. Zachrisson, E. 1997a. Bedrock map 22F Risbäck NW, 1:50,000. Sveriges
Greiling, R. O., Zachrisson, E., Thelander, T. & Sträng, T. 1999c. geologiska undersökning Ai, 102.
Bedrock map 23 G Dikanäs SE, 1:50,000. Sveriges geologiska under- Zachrisson, E. 1997b. Bedrock map 22F Risbäck SW, 1:50,000. Sveriges
sökning Ai, 125. geologiska undersökning Ai, 103.
Hambrey, M. J. 1994. Glacial Environments. University College London Zachrisson, E. 1997c. Bedrock map 22F Risbäck NE, 1:50,000. Sveriges
Press, London. geologiska undersökning Ai, 104.
Hambrey, M. J. & Harland, W. B. 1981. Part I Introduction. In: Zachrisson, E. 1997d. Bedrock map 22F Risbäck SE, 1:50,000. Sveriges
Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene geologiska undersökning Ai, 105.
Glacial Record. Cambridge University Press, Cambridge, 3 – 28. Zachrisson, E. & Greiling, R. O. 1993a. Bedrock map 23F Fatmomakke
Hartz, E. H. & Torsvik, T. H. 2002. Baltica upside down: a new plate NE, 1:50,000. Sveriges geologiska undersökning Ai, 77.
tectonic model for Rodinia and the Iapetus Ocean. Geology, 30, Zachrisson, E. & Greiling, R. O. 1993b. Bedrock map 23F Fatmomakke
255– 258. SE, 1:50,000. Sveriges geologiska undersökning Ai, 78.
Chapter 61

The Neoproterozoic glaciogenic Lillfjället Formation, southern Swedish Caledonides

RISTO A. KUMPULAINEN
Department of Geological Sciences, Stockholm University, SE-106 91 Stockholm, Sweden (e-mail: risto.kumpulainen@geo.su.se)

Abstract: The Tossåsfjället Group, which crops out in the Middle Allochthon of the Scandinavian Caledonides, contains the glaciogenic
Lillfjället Fm. Evidence that supports the interpretation of the Lillfjället formations as glacially related include the presence of diamictite
beds, the presence of lonestones, although few, in laminated units, one possible faceted clast, involution-like deformation structures, and
the presence of sandstone wedges cutting 2.5– 3 m down into two different diamictite beds. Similar to the other glaciogenic formations in
southern and central Scandinavia (Moelv and Långmarkberg formations), the Lillfjället Fm. rests on a unit of peritidal, sabkha-type dolo-
mite of the Storån Fm. Poor exposure contributes to the uncertainty about the thickness of the formation. In this account, the Lillfjället
Fm. stratigraphy has been divided into three subunits: (i) a lower diamictite-dominated unit, (ii) a middle unit composed of distinctly
laminated grey sandy mudstone and (iii) an upper diamictite-dominated unit. There is no information available from the Tossåsfjället
Group concerning palaeolatitudes and chemostratigraphy. Poor isotopic evidence (40Ar/39Ar) from the Ottfjället Dolerites, which cut
the Tossåsfjället Group, indicates that the succession is older than c. 665 Ma, so a reliable correlation with global glaciation events is
not yet possible.

Glaciogenic diamictite and associated sandy mudstone occur over a distance of a few metres to zero in the basal mylonitic foli-
in the Neoproterozoic Tossåsfjället Group of the Särv Nappe of ation. The arkose and dolerite dykes were altered to banded, felsic
the Caledonian Middle Allochthon in west-central Sweden and mafic mylonite, respectively. The nappe was subjected to
(Fig. 61.1; Gee et al. 1985). This glaciogenic unit, which was biotite-grade metamorphism, which recrystallized the rocks, and
named the Lillfjället Fm. by Kumpulainen (1980, 1981), is rep- blurs the micro-textures.
resented by less than a dozen, relatively small outcrops (stars in The tectono-sedimentary development of the Tossåsfjället
Fig. 61.1), arranged in a curvilinear manner across the nappe Group has been correlated with that of the other Neoproterozoic
over a distance of c. 70 km. successions in southern and central Scandinavia (Nystuen et al.
Along the southern margin of the nappe, close to the basal mylo- 2008). One of the features used to correlate these successions is
nites (Fig. 61.1), Strömberg (1961) and Röshoff (1975) described a the ‘carbonate –tillite’ couplet, which occurs in most of the succes-
possible glaciogenic diamictite, which was given the names Grön- sions in this region. Other correlative features include the gradual
stack and Ulvberget complexes, respectively. These outcrops are facies change from proximal to distal depositional sites and
now included in the Lillfjället Fm. systematic change in thickness of the successions below the car-
The Lillfjället Fm. has been correlated with other glacio- bonate –tillite couplet.
genic diamictites and related rocks in the Caledonides of south- These Neoproterozoic sedimentary successions were involved
ern and central Scandinavia. Descriptions of these deposits are in the Caledonian orogenic shortening and the various allochthons
given by Nystuen & Lamminen (2011) and Kumpulainen & were translated towards ESE across the western Baltica margin to
Greiling (2011). A summary of other correlatives of these glacio- their present locations. Being the uppermost tectonic unit in central
genic successions further north in Scandinavia is given by Stodt and southern Scandinavia to contain an assemblage of well pre-
et al. (2011). served rocks, the Tossåsfjället Group provides the most reliable
This chapter is based essentially on the first descriptions by information from the outermost possible succession more than
Kumpulainen (1980, 1981). The descriptions derive from two 500 km WNW of its present location (Kumpulainen & Nystuen
areas, the Lillfjället area in the south and the Stor-Lövsjön area 1985; Gayer & Greiling 1989; Nystuen et al. 2008). A pre-Iapetus
in the north (Fig. 61.1). The other exposures are more deformed plate tectonic reconstruction of Baltica and adjacent crustal units
and are not included here. (Greiling & Smith 2000) places Scotland close to the western
(present coordinates) Baltica margin, suggesting a juxtaposition(?)
of the Tossåsfjället succession with the Dalradian succession
Structural relationships (including the Port Askaig Fm.) in southwestern Scotland.

The Särv Nappe is the uppermost tectonic unit in the Scandinavian


Caledonides that hosts well-preserved glaciogenic rocks of Neo- Stratigraphy
proterozoic age. The high-quality preservation is a consequence
of the intrusion of tens of thousands of dolerite dykes, the Ottjället The unfossiliferous, Neoproterozoic Tossåsfjället Group (Fig. 61.2)
dolerite (Holmqvist 1894), into the c. 4.5-km-thick Tossåsfjället was divided by Kumpulainen (1980) into the Lunndörrsfjällen
Group. This dyke swarm reinforced the sedimentary succession (alluvial), Kråkhammaren (alluvial-marine), Storån (peritidal),
and limited Caledonian deformation to only a few shear zones Lillfjället (glaciogenic), and Lövan (marine/lacustrine-alluvial)
within the nappe. Owing to this, most of the original sedimentary formations. The stratigraphic position of the Lundörrsfjällen Fm.
structures survived. The whole nappe body, that is, the sedimen- is uncertain, because it is separated from the main nappe body
tary succession and the dyke swarm together, was folded into a by a fault zone (Gilotti & Kumpulainen 1986, fig. 5) and hence
regional fold. The sedimentary bedding in the western limb of correlation between the successions is currently not possible.
this fold is gently west-dipping but is slightly overturned in the Otherwise the other formations may be traced across the nappe
Stor-Lövsjön area in the east (Kumpulainen 1980; Gilotti & Kum- from SW to NE (Fig. 61.1). Gilotti & Kumpulainen (1986) ident-
pulainen 1986). In the overturned sequence the bedding-dyke ified in the Stor-Lövsjön area one more stratigraphic unit, an
angle still remains close to 908. Approaching the basal shear ‘unnamed unit’ (UNU). On the map (Fig. 61.3), the UNU
zone, the bedding-dyke angle decreases gradually and reduces exceeds 500 m in thickness and occupies the area between the

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 629– 634. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.61
630 R. A. KUMPULAINEN

Fig. 61.1. Generalized map of the main outcrop area of the Särv nappe in
the southern Swedish Caledonides, distinguishing between pre-glacial rocks, on
the left, and post-glacial rocks, on the right, separated by the Lillfjället Fm.
(dotted line) (modified from Kumpulainen 1980). The inset map shows the
location of the study area within the Caledonian Orogen in Scandinavia.
Fig. 61.2. Stratigraphy of the Tossåsfjället Group (modified from Kumpulainen
1980). The location of the Lunndörrsfjällen Fm. in the Särv Nappe stratigraphy
is uncertain as it may compose its own tectonic subunit within the nappe.
poorly exposed Storån Fm. in the west and Lillfjället Fm. diamic- s, silt; m, medium sand; g, gravel.
tite beds in the east. Bedding in this unit conforms to that of the
regional structure (Gilotti & Kumpulainen 1986, figs 4 & 5), justi-
fying the interpretation that the UNU is part of the Särv Nappe clast content in the lowermost diamictite bed is c. 25% dolomite,
stratigraphy. This new stratigraphic unit warrants a modification 38% other sedimentary rocks and 38% igneous rocks (granite,
of the original definition of the Lillfjället Fm. The new interpret- felsic porphyry and occasional greenstone). Higher up, the content
ation of the Lillfjället Fm. stratigraphy (Fig. 61.4) places the of sedimentary clasts (including subordinate dolomite) varies
whole Lillfjället section in the base of the formation (on between c. 10% and 25%, whereas crystalline clasts vary between
Fig. 61.3) between the Storån Fm. and the UNU. However, to 75% and 90%.
confirm this hypothesis requires a drilling operation. Resting on The diamictite units are interrupted by units of horizontally
top of the UNU, the diamictite units in the Stor-Lövsjön area are laminated sandy mudstone and weakly stratified dolomitic sand-
placed in the upper part of the formation. Hence, in the present stone. In the thick mudstone unit between D4 and D5, lamination
account the UNU is tentatively interpreted as a member within disappears gradually towards the upper part of the unit, where a
the Lillfjället Fm. (Figs 61.2 & 61.4). few lonestones are found, the largest being c. 0.4  0.6 m in size.
In this chapter, the discussion concerns the development of the The lower and upper boundaries of D1 are not exposed. The first
Lillfjället Fm. and its relation to the Storån (dolomite) and Lövan mudstone unit grades over a short interval into D2. The top of D2
(sandstone and conglomerate) formations (Figs 61.2 & 61.3). is undulating and sharply overlain by the second mudstone –
sandstone. The base and the top of the thin D3 are sharp, as are
the base and the top of D4; the top of D4 is cut erosively by a thin
Glaciogenic and associated deposits conglomerate at the base of the uppermost mudstone unit. The
base of D5 is sharp.
The Lillfjället Fm. No striated stones have been observed in this section, which may
be because these rocks, including the clasts, split across, rather
On the northern slope of the Lillfjället hill (628330 N/28500 E), five than along the matrix –clast boundary.
diamictite units and four sandy mudstone and sandstone units The UNU in the Stor-Lövsjön area is more than 500 m thick and
(Figs 61.1 & 61.5) occur with a collective thickness of c. 120 m. is composed of grey, distinctly parallel-laminated, sandy mud-
The thickness of the individual diamictite units (informally D1 stone, in which neither lonestones nor cross-bedded laminae
to D5) varies from c. 0.05 m to c. 50 m, whereas the sandy mud- have been observed.
stone units are less than 2 m thick, except for the uppermost, In the same area (628580 N/138210 E), the UNU grades gradually
which is c. 16 m thick. The diamictite is massive, and stratification upwards to the lower of the two diamictite beds (Fig. 61.4). These
has been observed only at the base of the D1 and D4 units. The two diamictite beds, c. 5–8 m (lower) and c. 5 m (upper), are
matrix in the various diamictite units is sand-dominated, poorly separated by a unit of laminated mudstone. The succession here
sorted, and contains, particularly in the lowermost diamictite, a displays a stretching lineation, which to some extent makes examin-
significant proportion of sand-grade clastic dolomite debris. The ation difficult. The diamictite units are dominated by igneous clasts;
NEOPROTEROZOIC GLACIOGENIC LILLFJÄLLET FORMATION 631

Fig. 61.3. Geological map of the Stor-Lövsjön area. The Tossåsfjället Group
stratigraphy occurs in the steeply standing succession around Stor-Lövsjön
(modified from Gilotti & Kumpulainen 1986). The Storån Fm. occurs west of
Buavan as a poorly exposed narrow belt resting on top of the shallow-marine
Kråkhammaren Fm.; neither the lower nor the upper boundary of the Storån Fm.
were observed in this area. The next belt to the east is composed of laminated
sandy mudstone, the ‘unnamed unit’ (UNU). The UNU grades upwards to
diamictite in the uppermost part of the Lillfjället Fm. In this account the UNU Fig. 61.4. Stratigraphy of the Lillfjället and adjacent formations (not to scale).
is interpreted as a member of the Lillfjället Fm. Key: 1, Kråkhammaren Fm. The Kråkhammaren Fm. subjacent to the Storån Fm. and the upper part of the
(sandstone and shale); 2, Storån Fm. (dolomite, shale and sandstone); 3, UNU Lövan Fm. are not shown here. The two sections, Lillfjället section and the
(sandy mudstone); 4, Lillfjället Fm. (diamictite); 5, lower part of the Lövan Fm. Stor-Lövsjön section, respectively, indicate in this diagram what stratigraphic
(mudstone to sandstone); 6, upper part of the Lövan Fm. (gravelly sandstone); 7, intervals they represent in the Tossåsfjället Group, respectively. The UNU was
sandstone of uncertain affiliation (Lundörrsfjällen Fm.). The inferred lower probably never deposited in the Lillfjället area.
dolomite-dominated unit in the lower part of the Lillfjället Fm. is not shown
on this map.
coarse-grained sandstone bed, c. 0.1–0.2 m thick. Within this
sandstone bed and resting on the diamictite, there is a faceted,
the largest observed, which is composed of syenite, measures flat-based clast (c. 0.1  0.25 cm) with a convex upper side.
c. 1.6  0.8 m in size. These two diamictite beds are separated This diamictite unit is also overlain by a laminated and small-scale
by a unit of thinly laminated sandstone and mudstone with some cross-laminated sandy mudstone (0.6 m thick) and a soft-sediment
minor soft-sediment deformation structures. The upper boundary deformed sandstone bed (0.2 m thick). One small lonestone of
of the mudstone with the upper diamictite is not exposed. sand grade occurs in the laminated mudstone c. 0.4 m above the
The most spectacular details exposed at this locality are clastic diamictite top.
wedges cutting into the top of each of the two diamictite beds; at
least four in the lower and five in the upper have been observed.
One wedge from each unit is described in more detail Boundary relationships to the underlying and overlying
(Fig. 61.6). The bowl-shaped upper end of the wedge in the non-glacial units
lower diamictite bed (Fig. 61.6a) is c. 0.8 –0.9 m wide. Its thick-
ness decreases rapidly downwards and at c. 1 m depth it is less In the Lillfjället section, the glaciogenic Lillfjället Fm. (Kumpulai-
than 0.2 m wide. Its total depth is almost 3.5 m. The grain size nen 1980) rests on thinly bedded dolomite of the Storån Fm.
in the wedge decreases downwards from very poorly sorted (Fig. 61.5). The exact boundary is hidden in a c. 1–2 m large unex-
coarse sandy to gravely infill to fine-grained sandstone. Other posed interval between the formations. The Storån Fm. contains
bowl-shaped depressions occur on top of this particular diamictite casts after evaporitic anhydrite that was replaced by chalcedony
unit and appear to form a semi-continuous bed on top of the lower during diagenesis. The chalcedony was then recrystallized to
diamictite. Laminated mudstone and a bed of deformed coarse microcrystalline quartz by the Caledonian metamorphism (Kum-
sand succeed this diamictite. No lonestones have been observed pulainen 1980, fig. 11). The Storån Fm. also contains occasional
in this laminated unit. sandstone and mudstone beds. In this section, the Lillfjället Fm.
The other clastic wedge, which cuts the top of the upper diamic- is overlain by coarse-grained sandstones of the Lövan Fm., but
tite, is more wedge-shaped (Fig. 61.6b). Its upper end is less the boundary area is unexposed and corresponds to a stratigraphic
than 0.4 m wide and extends downwards at least 3 m. The infill of thickness of a few tens of metres.
this wedge is vertically laminated. Similar to the other wedge, the In the Stor-Lövsjön area (Fig. 61.3), a narrow belt of locally
grain size here also decreases downwards from coarse to fine sand. derived dolomite debris suggests the presence of the Storån Fm.
The upper end of the wedge is sharply cut by a poorly sorted striking parallel and next to the underlying Kråkhammaren Fm.
632 R. A. KUMPULAINEN

Fig. 61.5. (a) Profile along the northern


Lillfjället hill side, where a dolerite dyke
embraces lenses of Storån and Lillfjället
formations. (b) Logged section of the
sedimentary succession in the hill side. No
pattern in the top of the upper mudstone
indicates massive mudstone. (c) Detail of
(b). Modified from Kumpulainen (1980).

Neither the lower nor the upper boundary of the Storån Fm. has above in the new interpretation of the Lillfjället Fm. stratigraphy
been observed along this belt. The next exposed rock unit to the (Figs 61.2 & 61.4; see ‘Discussion’).
east of the Storån Fm. is the sandy mudstone unit, UNU. The The upper boundary of the Lillfjället Fm. is located on top of the
unexposed ground on the map between these two belts of rocks sandstone bed 1 m above the second diamictite (Fig. 61.6b). A
is wide enough to accommodate a hypothetical lower diamictite- c. 4-m-thick unit of laminated sandy mudstone resting sharply
dominated unit corresponding to the Lillfjället section as suggested on top of the Lillfjället Fm. is referred to the base of the Lövan
Fm. This mudstone bed grades upwards to thinly bedded sandstone
(Fig. 61.4), which coarsens gradually into trough-cross-bedded
coarse-grained sandstone and conglomerate.

Chemostratigraphy

No data are available from this succession.

Palaeolatitudes

No data are available from Tossåsfjället Group. Data from other


successions (Greiling & Smith 2000; Hartz & Torsvik 2002;
Cawood & Pisarevsky 2006) imply high southern latitudes for
this area at the time of deposition of the Lillfjället Fm.

Geochronology

Although no data are available from the Tossåsfjället sedimentary


succession itself, the group is cut by (Fig. 61.4a) the Ottjället doler-
ite dyke swarm, which has given an 40Ar/39Ar age of 665 + 10 Ma
(Claesson & Roddick 1983). These dolerites are commonly corre-
lated with extensive dyke swarms in other parts of the Scandina-
vian Caledonides, providing ages in the interval c. 610 –550 Ma
(Svenningsen 2001; Paulsson & Andréasson 2002) and also with
the Egersund dolerite dykes in southwestern Norway (c. 616 Ma;
Bingen et al. 1998). New dating of the Ottfjäll dolerites is necess-
ary to better correlate successions from one region to another in
Scandinavia.

Fig. 61.6. Two clastic wedges cutting tops of two diamictite beds, lower (a) and
upper (b), respectively, on the western shore of Storavan. Note that the Discussion
framework clasts in both diamictites are parallel to the wedges, being probably
the result of the approximately east–west-trending stretching of the nappe body A striking feature of the Neoproterozoic sedimentary successions
and simultaneous rotation of the clasts. This implies also that the wedges have in the central and southern Scandinavian Caledonides is that peri-
been stretched and elongated. The original depth of the wedges is estimated at tidal, evaporitic, dolomite extends over the whole region and is
c. 2.5– 3 m. succeeded by presumably glaciogenic sediments attesting to a
NEOPROTEROZOIC GLACIOGENIC LILLFJÄLLET FORMATION 633

significant climate change in Neoproterozoic times (Nystuen & subjected twice to permafrost conditions, due to temporary emer-
Lamminen 2011; Kumpulainen & Greiling 2011). The extensive gence of the depositional site.
dolomite unit additionally supports the interpretation that the Neo- The major difference in stratigraphy between our two sub-areas
proterozoic landscape was close to peneplained and just above the is the presence of the UNU in the Stor-Lövsjön area and its appar-
sea level of that time. ent absence in the Lillfjället area. This could be for tectonic or
There are a range of criteria that have been described to support depositional reasons. The Särv nappe is composed of coherent tec-
the glaciogenic origin of a certain rock unit or a stratigraphic tonic lenses separated by shear zones. No such deformation zone is
section (e.g. Hambrey & Harland 1981; Hambrey 1994). A observed to cut the stratigraphy in the Lillfjället area, which may
number of these criteria have been described from the Lillfjället be due to limited exposure, but the inferred absence of such a
Fm., suggesting that it is of glaciogenic origin. In the Lillfjället deformation zone is corroborated by the fact that the dyke
section, the glaciogenic evidence includes (i) the presence of dia- swarm preserves its parallel trend over Lillfjället. A shear zone
mictite and (ii) the presence of lonestones, that is, probable would be identified by an offset in the dykes’ trend. Hence, tec-
ice-rafted debris, although they are few (see below). The Stor- tonic thinning or cutting is less probable here.
Lövsjön area additionally offers (iii) periglacial clastic wedges, The UNU is a more than 500-m-thick monotonous unit of grey
(iv) periglacial involution-like structures, and (v) a flat stone on parallel-laminated sandy mudstone, where neither lonestones nor
top of the upper diamictite (Fig. 61.6b) may be glacially faceted cross-laminated units have been observed. The distinct, parallel
and subsequently reworked by wind action. No striated pavements lamination suggests that this unit was deposited in standing body
and striated stones have been found to date. of water, probably a pro-glacial lake, where currents were too
Soft-sediment deformation in the stratified units in Lillfjället is weak to produce traction deposits. Also, no mass-flow deposits
very subordinate (Fig. 61.5b,c). The absence of lamination in the have been observed. The thickness of the unit requires a significant
upper part of the thick sandy mudstone unit may be due to accumulation space or a regular subsidence of the basin floor.
partial liquefaction, possibly due to rapid deposition of the next Evidence described above suggests that the Lillfjället Fm. is gla-
diamictite unit and rapid increase of the pore-water pressure. ciogenic, and its depositional development may be summarized as
The interpretation of the origin of the large lonestone in the follows. The lower part of the formation was deposited on top of a
massive upper part of this sandy mudstone is not straightforward, peritidal dolomite unit, the Storån Fm., as described from the Lillf-
because the possible deformation structures around that boulder jället section (Figs 61.4 & 61.5). The various diamictite units were
are obliterated. deposited either as sub-aqueous meltout tills or debris-flow units.
The interpretation that the Lillfjället succession was overrun by The diamictite deposition does not require repeated ice advance
five consecutive glaciers is not justified. More likely, the Lillfjället and retreat events. A paucity of lonestones (and possible drop-
succession was formed sub-aqueously, the diamictites being stones) suggests that icebergs were not able to travel these
meltout tills or redeposited tills (debris flows), rather than having waters. The present interpretation also suggests that a correspond-
been in contact with glacier ice. ing lower diamictite-dominated section was deposited on top of the
Also in the Stor-Lövsjön area, the two diamictite units are Storån Fm. in the Stor-Lövsjön area (Fig. 61.3); this section could
massive and may have been deposited sub-aqueously; the basal be thinner or thicker than that in the Lillfjället section. For a con-
contact of the lower unit is gradual and may indicate that this firmation of this hypothesis, drilling is required in this area.
diamictite unit was dumped on its laminated muddy substratum. The UNU occupies the middle part of the Lillfjället Fm. strati-
The base of the upper diamictite is not exposed, providing no graphy (Figs 61.2 & 61.4). It occurs as a thick unit in the Stor-
further evidence as to how the upper diamictite was formed. The Lövsjön area, but appears to be absent in the Lillfjället section.
tops of both of the diamictite units are sharp, and both are cut by This thickness change is interpreted here as a difference in the sedi-
sandstone wedges forming a network on both upper surfaces. mentary system, a large probable pro-glacial lake, where the Stor-
The lower top surface additionally displays bowl-shaped struc- Lövsjön area was the depocentre of the UNU and the Lillfjället
tures, possible involution structures (cf. French 1976; Miller section was located outside that depocentre. The fine-grained clastic
1996). These features support the interpretation that, although material in this unit was provided by glacial melt waters. The scar-
they may have been formed sub-aqueously, they were later city of dropstones suggests that the glacial margin at that time was
subject to sub-aerial conditions. located on land or otherwise isolated from this depositional area.
The clastic wedges in the Stor-Lövsjön area indicate permafrost The Lillfjället Fm. has been correlated previously (Kumpulai-
conditions. Wedge (a), in the lower diamictite, is a probable peri- nen 1980) with the glaciogenic Moelv Fm. in southern Norway,
glacial ice-wedge cast: a soil wedge (cf. Washburn 1980). With the Långmarkberg Fm. in central Scandinavia and the glaciogenic
no evidence of erosion, this frost-reworked land surface was sub- deposits in northern Norway. Since the group carries no fossils, the
merged before the deposition of the laminated mudstone. Ottfjället dolerite dyke swarm remains the only significant basis
Wedge (b) (Fig. 61.6b) is different in origin, although it too for a correlation of the Lillfjället Fm. with other successions in
formed sub-aerially. The vertically laminated sand inside this Scandinavia. Earlier dating methods for mafic rocks have given
wedge indicates that the wedge is a sand wedge (Péwé 1959), very variable ages, and the 40Ar/39Ar age of 665 + 10 Ma
which formed in permafrost areas with very low precipitation (Claesson & Roddick 1983) may contain a principal methodologi-
and without snow cover. It was not filled originally with ice, but cal error. Speculating that the various dyke swarms that produce
with dry sand as the frost crack opened. The vertical lamination ages of c. 600–615 Ma could be correlated with the Ottfjället
is produced after a series of openings and fillings. After the for- dolerites, then the Lillfjället Fm. must be older than that age inter-
mation of wedge (b), erosion removed some of the land surface val, meaning that this probably glaciogenic formation can be
including a sand-filled depression that is commonly located correlated with neither the Gaskiers glaciation event (c. 580 Ma)
above a sand wedge (Berg & Black 1966). A poorly sorted gravelly nor the Marinoan glaciation event (c. 635 Ma). Modern dating
sandstone was formed possibly by winnowing on the erosional techniques are required.
surface. Subsequent submergence of the land surface allowed the
deposition of a laminated mudstone and the succeeding sandstone The author wishes to thank T. Thelander and J. P. Nystuen for discussions about
in the top of the Lillfjället Fm. Permafrost-related sand wedges glacial deposits. N. Johansson provided valuable geographical information from
may penetrate several metres into the ground (Berg & Black the Stor-Lövsjön area. The manuscript was improved by the comments of
1966). In the Stor-Lövsjön area, the estimated original depth of E. Arnaud, an anonymous reviewer and G. A. Shields. The author is grateful
the wedges is c. 2.5– 3 m, which is probably too deep for a for support from Stockholm University. This represents a contribution of the
wedge to be formed in a perennially active frozen layer. This IUGS- and UNESCO-funded IGCP (International Geoscience Programme)
suggests that the glaciogenic Stor-Lövsjön succession was Project #512.
634 R. A. KUMPULAINEN

References Harland, W. B. (eds) Earth’s Pre-Pleistocene Glacial Record.


Cambridge University Press, Cambridge, 620– 623.
Bingen, B., Demaiffe, D. & van Breemen, O. 1998. The 616 Ma old Kumpulainen, R. & Nystuen, J. P. 1985. Late Proterozoic basin
Egersund basalt dike swarm, SW Norway, in the context of late- evolution and sedimentation in the westermost part of Baltoscandia.
Neoproterozoic opening of the Iapetus Ocean. Norsk Geologisk In: Gee, D. G. & Sturt, B. A. (eds) The Caledonide Orogen –
Tidsskrift, 79, 69 –86. Scandinavia and Related Areas. John Wiley & Sons, Chichester,
Berg, T. E. & Black, R. F. 1966. Preliminary measurements of the growth 213– 232.
of nonsorted polygons, Victoria Land, Antarctica. American Geophy- Kumpulainen, R. A. & Greiling, R. O. 2011. Evidence for late Neopro-
sical Union, National Academy of Sciences – National Research terozoic glaciation in the central Scandinavian Caledonides. In:
Council Publication, 1418, 61– 108. Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geo-
Cawood, P. A. & Pisarevsky, S. A. 2006. Was Baltica right-way-up or logical Record of Neoproterozoic Glaciations. Geological Society,
upside down in the Neoproterozoic? Journal of the Geological London, Memoirs, 36, 623–628.
Society London, 163, 753– 759. Miller, J. M. G. 1996. Glacial sediments. In: Reading, H. G. (ed.) Sedi-
Claesson, S. & Roddick, J. C. 1983. 40Ar/39Ar data on the age and meta- mentary Environments, Facies and Stratigraphy. Blackwell Science,
morphism of the Ottfjället Dolerites, Särv Nappe, Swedish Caledo- Oxford, 454– 484.
nides. Lithos, 16, 61– 73. Nystuen, J. P. & Lamminen, J. T. 2011. Neoproterozoic glaciation of
French, H. M. 1976. The Periglacial Environment. Longman Group, South Norway: from continental interior to rift and pericratonic
New York. basins in western Baltica. In: Arnaud, E., Halverson, G. P. &
Gayer, R. A. & Greiling, R. O. 1989. Caledonian nappe geometry in Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
north-central Sweden and basin evolution on the Baltoscandian Glaciations. Geological Society, London, Memoirs, 36, 613–622.
margin. Geological Magazine, 126, 499– 513. Nystuen, J. P., Andresen, A., Kumpulainen, R. & Siedlecka, A. 2008.
Gee, D. G., Kumpulainen, R., Roberts, D., Stephens, M. B., Thon, A. Neoproterozoic basin evolution in Fennoscandia, East Greenland and
& Zachrisson, E. 1985. Scandinavian Caledonides –Tectonostrati- Svalbard. Episodes, 31, 35 –43.
graphic Map, Scale 1:2,000,000. Sveriges geologiska undersökning Péwé, T. L. 1959. Sand-wedge polygons (tessellations) in the McMurdo
Ba, 35. Sound region, Antarctica – in progress report. American Journal of
Gilotti, J. A. & Kumpulainen, R. 1986. Strain softening induced ductile Science, 257, 545– 552.
flow in the Särv thrust sheet, Scandinavian Caledonides. Journal of Paulsson, O. & Andréasson, P. G. 2002. Attempted break-up of Rodinia
Structural Geology, 8, 441– 455. at 850 Ma; geochronological evidence from the Seve – Kalak Super-
Greiling, R. O. & Smith, A. G. 2000. The Dalradian of Scotland: missing terrane, Scandinavian Caledonides. Journal of the Geological Society
link between the Vendian of northern and southern Scandinavia? London, 159, 751–761.
Physics and Chemistry of the Earth, 25, 495– 498. Röshoff, K. 1975. A probable glaciogenic sediment in the Särv Nappe,
Hambrey, M. J. 1994. Glacial Environments. University College London central Swedish Caledonides. Geologiska Föreningens i Stockholm
Press, London. Förhandlingar, 97, 192– 195.
Hambrey, M. J. & Harland, W. B. 1981. Part I Introduction. In: Stodt, F., Rice, A. H. N., Björklund, L., Bax, G., Halverson, G. P. &
Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene Pharaoh, T. 2011. Evidence for late Neoproterozoic glaciation in the
Glacial Record. Cambridge University Press, Cambridge, 3 – 28. Caledonides of NW Scandinavia. In: Arnaud, E., Halverson, G. P.
Hartz, E. H. & Torsvik, T. H. 2002. Baltica upside down: a new plate & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
tectonic model for Rodinia and the Iapetus Ocean. Geology, 30, Glaciations. Geological Society, London, Memoirs, 36, 603–611.
255– 258. Strömberg, A. 1961. On the tectonics of the Caledonides in the south-
Holmqvist, P. J. 1894. Om diabasen på Ottfjället. Geologiska Förenin- western part of the county of Jämtland, Sweden. Bulletin of the Geo-
gens i Stockholm Förhandlingar, 16, 175– 192. logical Institutions of the University of Uppsala, 39, 1– 92.
Kumpulainen, R. 1980. The Upper Precambrian stratigraphy and deposi- Svenningsen, O. M. 2001. Onset of seafloor spreading in the
tional environments of the Tossåsfjället Group, Särv Nappe, southern Iapetus Ocean at 608 Ma; precise age of the Sarek Dyke Swarm,
Swedish Caledonides. Geologiska Föreningens i Stockholm Förhan- northern Swedish Caledonides. Precambrian Research, 110,
dlingar, 102, 531–550. 251– 254.
Kumpulainen, R. 1981. The Late Precambrian Lillfjället Formation Washburn, A. L. 1980. Cryogeology; A Survey of Periglacial Processes
in the southern Swedish Caledonides. In: Hambrey, M. J. & and Environments. Edward Arnold, London.
Chapter 62

The Port Askaig Formation, Dalradian Supergroup, Scotland

EMMANUELLE ARNAUD1* & IAN J. FAIRCHILD2


1
School of Environmental Science, University of Guelph, Guelph, Ontario, N1G 2W1, Canada
2
School of Geography, Earth and Environmental Sciences, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK
*Corresponding author (e-mail: earnaud@uoguelph.ca)

Abstract: The Port Askaig Formation (Fm.) is a thick glaciogenic succession within the Dalradian Supergroup that consists of over
700 m of variably dolomitic diamictite, conglomerate, sandstone mudstone and minor dolomite, and is bounded by mixed siliciclas-
tic –carbonate successions of the Islay (Lossit) and Bonahaven formations. These strata are exposed in the metamorphic Caledonides
of Scotland, although excellent preservation of sedimentary structures can be found at several sites. An extensional setting for this suc-
cession has been proposed based on stratigraphic and structural arguments. The available chemostratigraphic data include the Chemical
Index of Alteration, d13C and Sr-isotope values. Palaeomagnetic analyses have been shown to be subject to post-depositional Caledonian
overprinting. There is also continued debate over the regional palaeogeographical reconstructions of the Scottish promontory for this time
period. The succession is chronologically poorly constrained with U– Pb analyses of stratigraphically much higher or lower deposits. The
thick succession is thought to record glacially influenced marine sedimentation and reworking of unstable sediments in a tectonically
active setting with evidence of ice-margin fluctuations. Alternative palaeoenvironmental interpretations that focus on glacial terrestrial
processes and emphasize climatic influence instead of tectonic activity have also been proposed. The overlying carbonate is a lithologi-
cally diverse coastal complex and so does not fit the Neoproterozoic norm. Research has to date focused on the stratigraphic and sedi-
mentological aspects of this succession, as well as some of the broader palaeogeographical and structural features of the Dalradian basin.
Future efforts should focus on the chronological, structural and palaeogeographical constraints of this succession.

The Port Askaig Fm. is exposed at several sites in Scotland, includ- wedges (Eyles & Clark 1985), the Great Breccia (Arnaud & Eyles
ing Islay, the Garvellach Islands, Schichallion, Braemar, Muckle 2002; Benn & Prave 2006), the giant cross-bedded sandstone
Fergie Burn and Fordyce (Spencer & Pitcher 1968; Spencer (Arnaud 2004) and the origin of associated carbonate strata (Fair-
1971; Litherland 1980) as well as in Donegal, Mayo and Conne- child 1980a). Anderton (1982, 1985), Yardley et al. (1982), Harris
mara in Ireland (Howarth 1971; Tanner & Shackleton 1979; Max et al. (1978, 1993) and Prave (1999) provide valuable information
1981). The type section is located at Port Askaig on the island of on the depositional setting and tectonic evolution of the Dalradian
Islay, with several good along-dip outcrops over several kilometres Supergroup, while discussing ongoing controversies. Harris et al.
of shoreline and nearby moorland (Spencer 1971). The best out- (1993) also provides a useful set of lithostratigraphic columns of
crops can be seen on the Garvellach Islands, with lateral and vertical the Dalradian Supergroup at selected sites in Ireland, Scotland
exposures of hundreds of metres to several kilometres, though only and Shetland. Field-trip guidebooks are available (itineraries I, II
the lowermost three members of the formation are exposed there and III in Hambrey et al. 1991; Arnaud & Shields 2005). Sedimen-
(Fig. 62.1). These exposures are of phenomenal quality and, com- tological studies of the overlying Bonahaven Fm. (Spencer &
pared with other Neoproterozoic successions, relatively easy to Spencer 1972; Fairchild 1977, 1980a, b, 1985b) have recently
access. They provide abundant sedimentological information as been supplemented with chemostratigraphic studies of the bound-
well as an excellent sense of lateral and vertical facies variability ing formations that constrain their global context (Brasier &
and enable relatively detailed palaeoclimatic reconstructions. Shields 2000; McCay et al. 2006; Prave et al. 2009). The under-
Other sites in Scotland and Ireland tend to have limited outcrop lying carbonates have been described on the Garvellachs
exposure, complicated structural relationships, and higher degree (Spencer 1971), but are less well exposed and less studied on
of metamorphism. The Port Askaig Fm. is also known as the Port Islay, although an updated map of northern Islay has been pub-
Askaig Tillite in the literature, but this genetic term is avoided lished (British Geological Survey 1994).
here in accordance with modern stratigraphic practice.
The Port Askaig Fm. outcrops are significant in providing the
first record of glaciogenic deposits now known to be Precambrian Structural framework
(Thomson 1871). Subsequent Survey mapping and other works
carried out until the early 1930s (references in Spencer 1971) pro- The Port Askaig Fm. and associated strata are part of the thick Dal-
vided more geological context, but with little consciousness of radian Supergroup exposed within the metamorphic Caledonides
their international significance. Bailey (1916) was the first to use between the Great Glen Fault and the Highland Boundary Fault.
way-up indicators in these rocks and to demonstrate the simplicity Regionally, the Dalradian Supergroup rocks were affected by
of the structure. Later, the stratigraphy and origin of these deposits several phases of deformation and folding associated with the
became the focus of several detailed studies (Kilburn et al. 1965; Caledonian orogeny (Treagus 1987). Slides or thrust faults that
Spencer 1971; Eyles & Eyles 1983; Eyles 1988; Arnaud & Eyles lie sub-parallel to bedding and are thought to have last moved in
2006). The most comprehensive study was Spencer’s (1971) Ordovician times, and Devonian-age granitic intrusions are also
work, in which the variable sedimentology and regional stratigra- present in the region (Harris et al. 1993). The outcrops at Port
phy were documented at various sites in Scotland and Ireland. The Askaig are exposed in the relatively structurally simple NNE-
majority of other studies have focused on the excellent exposures trending Islay anticline (Bailey 1916; Fairchild 1980c). On the
on the Garvellach Islands. Garvellach Islands, the beds dip uniformly to the south and SE
Additional papers have focused on specific aspects of this thick (c. 358). Tertiary-age dolerite dykes cross-cut the strata.
succession such as provenance (Anderton 1980; Fitches et al. 1996; The Argyll Group, which includes the Port Askaig Fm. at its
Evans et al. 1998; Cawood et al. 2003), the sandstone intrusions or base (Fig. 62.2), is thought to have accumulated in an extensional

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 635– 642. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.62
636 E. ARNAUD & I. J. FAIRCHILD

Fig. 62.1. Map showing location of Port


Askaig Fm. outcrops on the Garvellach
Islands (modified from Spencer 1971).
Detailed map of outcrops on Islay can be
found in Spencer (1971) and Hambrey et al.
(1991). Note the lithology ‘sandstone’
includes from 0 to 90þ% dolomite at
different horizons.

basin experiencing increasing tectonic activity and development of Garvellach Islands, sedimentary structures are well preserved,
fault-bounded sub-basins prior to the opening of the Iapetus ocean although noticeably deformed in cleaved lithologies.
(Anderton 1982; Harris et al. 1978, 1993). Evidence that basin
extension started during accumulation of Port Askaig Fm. sedi-
ments includes rapid lateral thickness changes over short distances Stratigraphy
(Anderton 1982, 1985). For example, the Port Askaig Fm. appears
to thin rapidly westwards on Islay to the limit of its outcrop (British The Port Askaig Fm. is found within the predominantly marine
Geological Survey 1994) and the overlying Bonahaven Fm. thins succession of the Argyll Group in the Dalradian Supergroup
significantly west across the Bolsa fault in this region (Fairchild (Fig. 62.2), being underlain by the Islay (Lossit) Limestone
1980c; Anderton 1985). There is also a coarse conglomerate (mixed dolomitic-siliciclastic at its top) and overlain by the
facies of reworked limestone in the Lossit Limestone in the mixed dolomitic –siliciclastic Bonahaven Fm.
western part of its outcrop (British Geological Survey 1994). It The Port Askaig Fm. is over 700 m thick. Spencer (1971)
is within this context that the Ordovician-age Grampian slides defined five members based on predominant facies and clast lithol-
were proposed to be Proterozoic synsedimentary listric faults ogy (Fig. 62.3). Members I to III are best exposed on the Garvel-
that were later re-activated, based on their association with rapid lach Islands, whereas Members IV and V are only exposed on
thickness changes at various stratigraphic levels within the Dalra- the island of Islay. The lowermost member (Member I, Beannan
dian Supergroup (Soper & Anderton 1984; Anderton 1985, 1988). Buidhe) consists primarily of stacked carbonate-rich diamictite
Most recently, some of the diamictite units as well as repeated with commonly discontinuous sandstone and conglomerate inter-
horizons of soft-sediment deformation structures have been inter- beds. Member I also contains a thick diamictite bed called the
preted as additional indicators of early extensional tectonic activity Great Breccia, with very large clasts in a muddy sandstone
in Argyll times (Arnaud & Eyles 2002, 2006). matrix and a series of folded and boudinaged interbedded units
The rocks have undergone deformation and metamorphism as a of variable lithologies (mudstone, sandstone, dolomite and diamic-
result of the Caledonian orogeny, with open folding and upright tite) commonly referred to as the ‘Disrupted Beds’. Member II (An
cleavage developed in the least affected areas of Islay, Garvellachs Tamhanachd) consists of diamictite associated with sandstone and
and Northern Donegal (Spencer 1971; Fairchild 1980b). Locally mudstone interbeds with a base defined by the appearance of extra-
there are high-strain areas (Borradaile 1979; Fairchild 1980c) par- basinal clasts. Member III (Creagan Loisgte) consists of thick
ticularly nearing bounding thrusts, re-activated faults and some packages of laterally continuous sandstone interbedded with dia-
major lithological boundaries (Anderton 1985; Fairchild 1985b; mictite. Member IV (Ruahd Phort Beag) is again dominated by
Treagus & Treagus 2002). Pressure-solution effects are wide- diamictite, whereas Member V (Con Tom) is dominated by sand-
spread more generally, with limestone and diamictite much more stone with minor diamictite.
strongly deformed than dolostone or quartz sandstone (Fairchild Significant lateral changes to the west on Islay were mentioned
1985b). In the most-studied locations at Port Askaig and in the in the previous section. Within the Port Askaig Fm., the succession
on the island of Islay is thin at the base with the unique facies of the
overlying Disrupted Beds being the first unequivocal lithostrati-
graphic tie to the Garvellachs, but which is separated from the
underlying carbonates by ,6 m of sandstone and dolomitic con-
glomerate correlated with the Great Breccia by Spencer (1971).
The remainder of Members I, II and III are broadly similar
between Islay and the Garvellachs.

Glaciogenic deposits and associated strata

The Port Askaig Fm.

The following is a summary of a relatively thick, complex and


superbly exposed succession based on work by Kilburn et al.
(1965), Spencer (1971), Eyles & Clark (1985), Eyles (1988),
Arnaud (2002), Arnaud & Eyles (2002, 2006), Arnaud (2004)
and Benn & Prave (2006). The Port Askaig Fm. consists of a
thick succession of diamictite (44 distinct units), interbedded
with sandstone and minor units of mudstone and conglomerate.
Diamictite are massive to stratified with clasts up to several
Fig. 62.2. Generalized stratigraphy and tectonic setting of the Port Askaig Fm. metres in diameter floating in a siltstone to silty sandstone
and associated strata. P, Port Askaig Fm.; I, Islay Limestone; B, Bonahaven matrix. Some diamictite beds exhibit coarse-tail inverse grading,
Fm.; J, Jura Quartzite; S, Scarba conglomerate; *earthquake-induced and many contain sandstone stringers or inclusions of variable
liquefaction features (modified from Arnaud & Eyles 2006). lithologies (Arnaud & Eyles 2006). Some of the diamictite units
PORT ASKAIG FORMATION, SCOTLAND 637

Spencer 1971; Arnaud & Eyles 2002; Benn & Prave 2006 for
detailed descriptions).
Sandstone in the Port Askaig Fm. exhibits a variety of character-
istics including a wide range of textures (fine to very coarse), sorting
(poorly to well sorted), lithologies (dolomitic to quartzitic) and struc-
tures (massive, horizontally bedded, cross-bedded and deformed).
Thick packages of sandstone within Member III are particularly
notable as they contain thick to very thick sets of cross-bedded sand-
stone (individual set thickness averages 3 m, maximum is 11 m;
Arnaud 2004). The lowermost package of these giant cross-bedded
sandstone exhibits a preferred southerly palaeocurrent direction,
whereas cross-bedded sandstone in the rest of the succession is
more variable (Spencer 1971; Arnaud 2004). Other interbeds within
the Port Askaig Fm. include finely laminated mudstone (rhythmites),
(clastic) dolomite and massive to stratified conglomerate. Conglom-
erate often overlies and loads into diamictite, although some also
occurs interbedded with sandstone and mudstone. Outsized clasts
in laminated mudstones or diamictites occur at various horizons,
but only limited examples displaying clear deflection of underlying
laminae occur, especially in the Disrupted Beds (Spencer 1971;
Hambrey et al.1991). Petrographic study of the carbonate in the
Port Askaig Fm. indicates that it appears to be largely detrital in
origin, with that in the rhythmite facies being secondary (Fairchild
1985a; Hambrey et al. 1991, p. 27, 34).
Deformation structures have been documented at multiple dis-
tinct stratigraphic horizons. Several of these, namely sandstone
wedges, sandstone dykes and sandstone downfold structures, have
been examined in detail (Spencer 1971; Eyles & Clark 1985). The
sandstone wedges are predominantly ,10 cm wide, have sharp
irregular outer geometries and commonly penetrate up to several
metres downward into diamictite units or into carbonate units
within the underlying Islay Limestone. Some exhibit branching or
polygonal patterns on bedding plane surfaces. Sandstone dykes are
more tabular in form and intrude diamictite and siltstone. Sand-
stone downfold structures, refered to as load casts or ball and
pillow structures by Arnaud & Eyles (2006), are also common on
top of diamictite units or within interbedded sandstone and mud-
stone. These features, together with horizons of convolute and con-
torted bedding and pseudonodules, were shown to be found most
commonly within Member II sediments on the Garvellach
Islands (Arnaud & Eyles 2006).

The Lossit Limestone

Strata underlying the Port Askaig Fm. have been known as the
Islay Limestone and on the Garvellachs a succession of 72 m of
predominantly mixed dolomitic –siliciclastic sediments with
some stromatolitic and other limestone is present (Spencer
1971). There are some pure dolomicrite and dolomitic stromato-
Fig. 62.3. Stratigraphic log of the Port Askaig Fm. based on outcrops on the
lites, but for the most part the facies contain both terrigenous
Garvellach Islands (Members I to III) and on the island of Islay (Members IV
debris and reworked intrabasinal carbonate (Hambrey et al.
and V). GB, Great Breccia; DB, Disrupted Beds; XB, Giant cross-bedded 1991). The most interesting structures are crystal pseudomorphs
sandstone (modified from Arnaud & Eyles 2002). Note the dolomite interbeds originally illustrated by Spencer (1971). Similar structures from
are considered detrital in origin (see text for discussion). the Irish Dalradian were interpreted as glendonites, which are
pseudomorphs after ikaite (Johnston 1995).
On Islay, the top 300 m of the Islay Limestone have been rede-
signated as the Lossit Limestone Fm. (British Geological Survey
contain lenses of bedded siltstone sandstone and/or conglomerate
1994). Its topmost (Persabus) member (c. 80 m thick) consists of
(Spencer 1971). Clast and matrix lithology change up-section from
interbedded dolostone (locally stromatolitic or intraclastic), quart-
predominantly intrabasinal and dolomitic at the base to extrabas-
zite, slate and mixed lithologies and overlies a c. 70 m pure (Kiells)
inal and siliciclastic at the top. Pink granitoid extrabasinal clasts
limestone member, locally oolitic, which in turn overlies several
appear to have come from Palaeoproterozoic plutonic rocks of
thick slate and limestone units.
the Svecofennian-Makkovik-Ketilidian province (Evans et al.
1998). Beds are typically several metres to 10 m in thickness
with sharp conformable to erosional basal contacts. Some basal The Bonahaven Fm.
contacts are gradational with associated laminated mudstone
(Arnaud 2002; Arnaud & Eyles 2006). The Great Breccia is a par- The lithostratigraphy and geological context of the overlying
ticularly thick diamictite unit (up to 50 m), with megaclasts mixed carbonate –siliciclastic Bonahaven Fm. has been documen-
ranging from several metres to over 100 m in diameter (see ted by Spencer & Spencer (1972), who divided it into four
638 E. ARNAUD & I. J. FAIRCHILD

members (Fairchild 1977, 1980a, b, 1985b; Hambrey et al. 1991). relatively unweathered basement rocks (Panahi & Young 1997). In
On the east coast of Islay, the sub-arkosic arenite at the top of the addition, analysis of the trace element geochemistry suggests
Port Askaig Fm. is succeeded by c. 65 m of sandy facies with mud- erosion of shale developed on a post-Archaean crystalline basement,
stone and local dolomite at the top (Member 1), 25 m of quartzite although a specific source area could not be identified.
(Member 2), c. 200 m of dolomitic sandstone, oolitic facies and Brasier & Shields (2000) provided the first isotopic chemostra-
mudrock (Member 3) and 55 m of heterolithic siliclastic rocks tigraphic constraints, with reliable Sr-isotope values as low as
with a central 12 m of pure dolostone (Member 4). In turn this is 0.7067 being obtainable from the Sr-rich facies of pure Lossit
succeeded by several kilometres of pure quartzite (Jura Quartzite). (Islay) Limestone underlying the Port Askaig Fm. These
The structures at the base of the Bonahaven Fm., taken by pre- Sr-isotope values are comparable with facies underlying the
vious workers to be desiccation cracks, have been reinterpreted as earliest evidence of Neoproterozoic glaciation in other regions.
interstratal dewatering features by Tanner (1998). Primary carbon- Thomas et al. (2004) verified this result at a slightly lower
ate is first encountered near the top of Member 1, where displacive horizon in the Islay limestone (Storakaig limestone of Ballygrant)
dolomite fabrics in terrigenous mudstone with genuine desiccation (0.706651 –0.706902) using a slightly more careful sample prep-
cracks (Fairchild 1977, 1980b) are suggestive of dolocrete. The aration protocol. Sawaki et al. (2010) carried out a careful chemo-
main carbonate unit (Member 3) consists predominantly of three stratigraphic study on both Islay and the Garvellachs. The samples
dolomite facies (stromatolitic, bimodally cross-stratified intraclas- closest to the onset of glaciation, which also passed a stringent test
tic and/or oolitic sandstone, and mudstone with thin graded and/or for preservation (Mn/Sr , 0.2) were found 40 m below the base of
wave-rippled sand) (Fairchild 1980a). the Port Askaig Fm. on Garbh Eilach with values as low as
0.70640. These are close to the lowest values found in East Green-
land immediately prior to glaciation (Fairchild et al. 2000) and
Boundary relations with overlying and underlying suggest that the Garbh Eilach section is similarly complete.
non-glacial units McCay et al. (2006) described evidence for a third Dalradian
glacial from Ireland and also built on earlier d13C results of
In eastern Islay, basal sandstone, clastic dolostone and diamictite Brasier & Shields (2000). Values of þ5‰PDB in a lower (Bally-
of the Port Askaig Fm. variably rest erosionally on the underlying grant) limestone were followed by a decline to weakly positive to
dolomite (at times stromatolitic) and thin-bedded quartzite and negative values in the Lossit Limestone (Brasier & Shields 2000).
shale, which in turn rest on pure limestone (Spencer 1971; McCay et al. (2006) and Prave et al. (2009) show that in the Garvel-
British Geological Survey 1994). Over 90 m of Member I lachs section, both dolomitic and limestone facies show a change
exposed on the Garvellachs is missing here on Islay. Further upwards from negative values (–4 to –6) in the strata in which
west, Port Askaig Fm. diamictite units are less poorly exposed, Sawaki et al. (2010) measured the lightest Sr-isotope signatures,
but commonly are limited by sedimentary limestone breccias to weakly positive values. In terms of the carbonate rocks overlying
(British Geological Survey 1994). On the Garvellach Islands, the Port Askaig Fm., all these publications show that the main
diamictite sharply and erosionally overlies the Islay Limestone ferroan dolomitic part (Member 3) of the Bonahaven Fm. displays
in places (Spencer 1971). In other areas of the Garvellach negative values, and Brasier & Shields (2000) document values in
Islands (Dun Chonnuill & Garbh Eileach), the contact is grada- the Member 4 dolomite horizon exceeding þ10‰.
tional as shown by the appearance of rare clasts and the presence
of siltstone (typical of the siltstone matrix of the overlying diamic-
tite) interbeds in the upper part of the Islay limestone (Spencer Other characteristics (e.g. economic deposits, biomarkers)
1971; Arnaud 2002; Arnaud & Eyles 2006).
The upper contact of the Port Askaig Fm. is seen on Islay, where Mining of zones of epigenetic Pb–Zn mineralization formerly
it is conformable with the Bonahaven Fm., but it is difficult to occurred on Islay in the Ballygrant and Lossit limestones
define where the last evidence of glacial phenomena occur as the (British Geological Survey 1994). The Port Askaig Fm. is
dominant facies near the top is quartzite, with only rare thin dia- locally rich in detrital magnetite and in the Disrupted Beds there
mictite horizons. The highest occurrence of chess-board albite, are occasional massive layers of magnetite.
which is diagnostic of the granitic pebbles of the Port Askaig Spencer (1971) reported possible organic traces in Member 1
Fm. (Spencer 1971), is within a channelled conglomerate (Spencer & Spencer 1972) of the Bonahaven Fm. A detailed
horizon, located c. 25 m above the base of the Bonahaven Fm. description of sole structures interpreted as representing chains
(Hambrey et al. 1991, p. 39, 40). The first occurrence of dolomite of faecal pellets was made by Brasier & McIlroy (1998), although
is found near the top of member 1 of the Bonahaven Fm. where it Brasier and Shields (2000) conceded that they could well be of
occurs in tidal flat facies (Fairchild 1977, 1980c) and is succeeded inorganic origin.
by more extensive dolomite facies in Member 3, with local undo- Fairchild (1977) described clear 0.1-mm-sized mica spheres
lomitized oolitic and micritic limestone (Fairchild 1980a). within a 2 m stratigraphic interval of carbonaceous mudstones near
the top of Member 1 of the Bonahaven Fm. Petrographic evidence
indicates that they were delicate enough to collapse when mud desic-
Chemostratigraphy cated, yet were mineralized (perhaps by glauconite) when eroded as
intraclasts. They were interpreted as an unusual form of preservation
Panahi & Young (1997) carried out a study of major and trace of acritarch fossils. However, their great similarity with Triassic mica
element geochemistry, specifically utilizing the Chemical Index of spheres from SW England was noted and the latter were sub-
Alteration (CIA). Most of the 21 samples were taken from the sequently interpreted as tektites by Walkden et al. (2002); this
matrix of diamictite exposed on Garbh Eileah, Garvellach Islands. seems a more likely explanation for the Islay occurrence. Only one
Analysis showed a decrease in CIA values from the base of the other example of impact-related phenomena has been reported in
Port Askaig Fm. (values range from 68 to 77) to the upper part of the Neoproterozoic of the British Isles (Amor et al. 2008).
Member III (values range from 60 to 68). The CIA values are an indi-
cation of the extent of weathering based on the relative proportions of
alkali and alkaline earth elements and thus the decrease was inter- Palaeolatitude and palaeogeography
preted as indicating a change in sediment source. The high CIA
values reflect erosion of underlying sedimentary rocks, which had Earlier palaeomagnetic work by Tarling (1974) and Urrutia-
already experienced weathering, whereas the lower CIA values at Fucugauchi & Tarling (1983) based on large sample suites on
the top of the section indicate incorporation of sediments from the Garvellachs appeared to be consistent with low palaeolatitudes,
PORT ASKAIG FORMATION, SCOTLAND 639

although difficulties were encountered due to the unknown age of Lithological comparisons with other diamictite-bearing succes-
the deposits and the inability to eliminate the possibility of Caledo- sions in Scotland and Ireland and throughout the North Atlantic, as
nian overprinting. Stupavsky et al. (1982) specifically addressed well as chemostratigraphic studies, have led to various regional
the overprinting issues using clasts (79 specimens from 36 cores) and global stratigraphic correlation schemes (Spencer 1975;
and matrix (2 to 3 specimens from 20 cores) of the diamictite Hambrey 1983; Prave 1999; Brasier & Shields 2000; Halverson
units as well as siltstone (2 to 3 specimens from 6 cores) from et al. 2005; McCay et al. 2006), with the most recent work
Garbh Eileach, Garvellach Islands. Remanent magnetization suggesting that the Port Askaig Fm. likely represents the oldest
measurements were used to calculate remanence angular standard of several Neoproterozoic glacial periods within the Dalradian
deviation in order to evaluate within-specimen homogeneity and basin and the North Atlantic region. A particularly distinctive
reliability of specimens that were sampled. Specimens were also characteristic is the low Sr-isotope ratios of the underlying
subjected to alternating field and thermal demagnetization. All limestone (Brasier & Shields 2000; Thomas et al. 2004; Sawaki
samples yielded similar results, thus failing the conglomerate et al. 2010). However, this has recently been called into question
test, and suggesting that these deposits have been remagnetized based on new Re-Os dates that provide a maximum age of
by Ordovician-age overprinting (Stupavsky et al. 1982). 659.6 + 9.6 Ma for the Port Askaig Fm. (Rooney et al. 2011).
The Dalradian Supergroup is generally associated with the
Proto-Iapetus Ocean, although palaeogeographical reconstructions
are uncertain, in part because of an incomplete understanding of Discussion
Dalradian basin development (Soper 1994; Tanner & Bluck
1999; Prave 1999; Dalziel & Soper 2001; Demspter et al. 2002; Although there is broad agreement that the Port Askaig Fm.
Hutton & Alsop 2004; Tanner et al. 2005). Some have suggested records environmental conditions during one of the Neoprotero-
the Dalradian experienced orogenesis prior to rifting and zoic glacial periods, several palaeoenvironmental models have
opening up of the Iapetus Ocean in Argyll-Southern Highland. been proposed with differences hinging largely on the interpret-
Such an orogenic event would suggest an affinity with Gondwana ation of diamictite units as directly deposited by ice (Kilburn
and a palaeogeographical location either off NW Gondwana et al. 1965; Spencer 1971; Benn & Prave 2006) or as primarily
(Bluck & Dempster 1991) or off Amazonia (Dalziel 1994). Others deposited in a marine setting influenced by ice-rafting and tectonic
have suggested the Dalradian experienced prolonged rifting and instability (Eyles 1988; Arnaud & Eyles 2006).
extension throughout its history consistent with a palaeogeogra- In the models that emphasize a subglacial or ice marginal origin,
phical location on the margin of Laurentia (Dalziel & Soper diamictite units are interpreted as tills recording multiple grounded
2001). Others still have suggested the Dalradian was an extension ice advances based on the lateral extent of diamictite, the discon-
of Baltica (Greiling & Smith 2000) based on the similarity between tinuous nature of some of the interbeds in the lower members, the
granitic clasts in the Port Askaig Fm. and Scandinavian intrusions. presence of large extrabasinal clasts and faceted clasts, the sand-
Recent work by Cawood et al. (2003) demonstrates the similar stone wedges that are interpreted as periglacial, and the siltstone
characteristics of the detrital zircon populations throughout with outsized clasts that are interpreted as varves with ice-rafted
the Dalradian and their close match to Laurentian sources, with debris. In these models, the Great Breccia is thought to record
the Port Askaig Fm. clasts having closest matches in the North glaciotectonic deformation of sediments based on its similarity
Atlantic Borderlands (Makkovik, Ketilidian and Svecofennian to deformed chalk rafts in tills of Norfolk (UK) and its correlative
provinces). In this regard, the distinctive provenance represents erosional unconformity on Islay. The sandstone wedges are
along-basin rather than cross-basin transport. thought to be periglacial because of their similarity to those
formed by repeated freeze and thaw. The associated interbeds
are thought to record either terrestrial or marine depositional
Geochronological constraints conditions preserved between successive ice advances.
Benn & Prave (2006) took this argument further by suggesting
Radiometrically, the Port Askaig Fm. is relatively poorly con- that the Great Breccia and associated Disrupted Beds recorded
strained. A maximum age of c. 806 Ma comes from the underlying proglacial and subglacial phases of a single glaciotectonic defor-
Grampian Shear Zone located at the base of the Dalradian Super- mation cycle. Although they acknowledged that the sedimentary
group (Noble et al. 1996). This finding is based on three dates characteristics of the Great Breccia were consistent with either a
(806 + 3, 808 þ 11/–9 and 804 þ 13/–12 Ma) obtained from glaciotectonic or non-glacial sediment gravity flow origin, they
U –Pb isotope analyses of primary monazite in pegmatite and preferred a glacial origin based on their interpretation of several
neocrystalline monazite associated with mylonitic host rocks. features within the associated Disrupted Beds as indicative of sub-
Samples were collected from the Grampian Shear Zone at Lochin- glacial deformation; namely laminae that resemble glaciotectonic
dorb and A’Bhuidheanaich, East of the Great Glen Fault and laminae, deformation indicative of shear, an increase up-section in
Inverness, Scotland. the number of extrabasinal clasts, and evidence of increasing
A minimum age for the Port Askaig Fm. is provided by two upwards cumulative strain in diamictite.
dates from the Tayvallich volcanic rocks, which are stratigraphi- In the tectonically influenced glaciomarine models, many dia-
cally 8 km above the Port Askaig Fm. at the top of the Argyll mictite units are thought to record sediment instability associated
Group in SW Scotland (Fig. 62.2; Prave 1999). Halliday et al. with basin development, whereas diamictite units in the uppermost
(1989) present various data from U –Pb, 207Pb/206Pb and part of the succession are thought to record reworking of sediments
Sm –Nd isotopic analyses of zircons from a keratophyre sampled and rainout of fine-grained sediments and ice-rafted debris in a gla-
from a small laccolithic body on the Tayvallich Peninsula. They cially influenced basin (Arnaud & Eyles 2006). This interpretation
conclude that the most likely age of the keratophyre is is based on the presence of coarse-tail inverse grading, gradational
595 þ /24 based on their Pb/Pb analyses. Field relationships basal contacts and the close association with other sediment
between the small laccolithic body and the Tayvallich volcanic gravity flow deposits (Boulton 1972; Nardin et al. 1979; Mulder
rocks are somewhat unclear, and Dempster et al. (2002) have & Alexander 2001; Arnaud & Eyles 2006). The Great Breccia is
suggested that this age may be younger than the Tayvallich volca- interpreted as a catastrophic subaqueous landslide associated
nic rocks. In an attempt to refine this age, Dempster et al. (2002) with local tectonic activity (Arnaud & Eyles 2002) based on
analysed 14 zircons from a felsic tuff collected at Port a’ Bhuailteir mapping that revealed (i) the Great Breccia to be a ‘composite
on the Tayvallich Peninsula. Concordia diagrams show a mean graded sequence’, (ii) an intimate association with undeformed
206
Pb – 238U age of 601.4 þ /23.7 Ma (n ¼ 13/14; 2s; subaqueous sediment gravity flow and traction current deposits
MSWD ¼ 0.82). and (iii) a similarity to published studies of allochtonous carbonate
640 E. ARNAUD & I. J. FAIRCHILD

megabreccia (Arnaud & Eyles 2002). The stratigraphic horizons values over vertical distances of 10 –20 m (Fairchild 1991). There
of deformation structures, including the overlying Disrupted are limestone facies in the Bonahaven Fm. that could be studied to
Beds, are interpreted as seismites indicative of local tectonic help resolve this issue. The extraordinarily high d13C signature in
instability based on their form, geographical extent, and tectonic the pure dolostone of Member 4 is an enigma; a similar strong
setting in which they are found (Arnaud & Eyles 2006). Strati- signal appears rather soon after glaciation in deposits of NW
graphic analysis of the sedimentary facies and indicators of Canada, known as the Keele peak (McCay et al. 2006). Whereas
glacial and tectonic activity suggest that repeated ice-margin fluc- Halverson et al. (2005) chose to trust the pre-Port Askaig d13C
tuation and tectonically quiet conditions occurred during depo- negative anomaly to correlate with late Cryogenian glaciations
sition of Member III sediments (Arnaud & Eyles 2006). elsewhere, it is now clear that such an anomaly is present below
The sandstone, mudstone and conglomerate interbeds record both early and late Cryogenian glacial deposits (Prave et al.
shallow marine conditions affected by sediment gravity flows 2009). Most authors believe that the low Sr-isotope ratio is a
and traction currents (Arnaud & Eyles 2006). The giant cross-beds more specific feature with which to establish an early- to mid-
are thought to result from the migration of large dunes under strong Cryogenian age for the Port Askaig Fm. The minimum value
tidal currents considering the Dalradian basin is thought to be obtained a short distance below the Port Askaig Fm. of 0.70640
narrow at this time (Eyles 1988; Arnaud 2004). In terms of the (Sawaki et al. 2010) is close to the lowest value of 0.7063 found
dolomite interbeds, there is no specific evidence of chemical depo- in East Greenland immediately prior to glaciation (Fairchild
sition and much evidence of detrital dolomite (Fairchild 1985a; et al. 2000), and suggest that the Garbh Eilach section is similarly
Hambrey et al. 1991). complete. There are difficulties in reconstructing stratigraphic pro-
The superb exposures in the Garvellach Islands have resulted in files through the entire thickness of the Dalradian succession and
numerous detailed studies, yet the resulting depositional models there is also evidence of diachroneity of widely distributed glacial
proposed for the Port Askaig Fm. have some significant differences deposits thought previously to be correlative (e.g. Kendall et al.
in their climatic and palaeoenvironmental implications. A full dis- 2006; Fanning & Link 2004), but Prave et al. (2009) were optimis-
cussion of these models is beyond the scope of this paper and the tic that there are sufficient diagnostic chemostratigraphic results
reader is referred to the original works for more details. Although to allow the global stratigraphic context of the Port Askaig Fm.
recent developments in glacial geology have allowed some of the as early- to mid Cryogenian (Sturtian in their terminology) to
earlier interpretations to be discounted, there are still instances be confirmed. Rooney et al.’s (2011) findings will ensure contin-
where the exact nature of glacial influence over these deposits is ued debate and suggest additional work is needed to constrain
debatable and difficult to establish unequivocally. the timing of glaciation.
The carbonate succession underlying the Port Askaig Fm. on
Islay is consistent with marine regression from offshore shale The authors would like to thank A. Spencer and G. Shields for helpful comments
to coastal facies. On the Garvellachs, the mixed carbonate- on earlier drafts of this paper. E. A.’s research on Neoproterozoic glaciogenic suc-
siliciclastic sediments bear local probable ikaite pseudomorphs cessions is supported by the Natural Sciences and Engineering Research Council
near the top, suggestive of cool marine conditions. Facies resem- of Canada. This represents a contribution of the IUGS- and UNESCO-funded
bling modern carbonate tropical platforms are limited to local IGCP (International Geoscience Programme) Project #512.
intraclastic dolostones in parts of the Persabus Member at the
top of the Lossit Limestone, whereas more distinctive oolitic lime-
stone facies occur below in the Kiells member. The relative strati- References
graphic positionings of these occurrences between the Garvellachs
Amor, K., Hesselbo, S. P., Porcelli, D., Thackrey, S. & Parnell, J.
and Islay cannot be resolved because of the lack of marker hor-
2008. A Precambrian proximal ejecta blanket from Scotland.
izons between the two. Geology, 36, 303–306.
The mixed carbonate-siliciclastic succession that overlies the Anderton, R. 1980. Distinctive pebbles as indicators of Dalradian prov-
Port Askaig Fm. is thought to record shallow marine sedimen- enance. Scottish Journal of Geology, 16, 143–152.
tation. No climatically distinctive facies occur in the basal part Anderton, R. 1982. Dalradian deposition and the late Precambrian –
of the Bonahaven Fm. The most prominent carbonate (Member Cambrian history of the North Atlantic region: a review of the early
3 of the Bonahaven Fm.) records a lagoonal and tidal complex evolution of the Iapetus Ocean. Journal of the Geological Society,
(Fairchild 1980a). It is generally agreed that the Bonahaven Fm. London, 139, 421–431.
is not a typical cap-carbonate sequence (Halverson et al. 2005; Anderton, R. 1985. Sedimentation and tectonics in the Scottish Dalra-
McCay et al. 2006). It is hard to pinpoint the last glacial influence dian. Scottish Journal of Geology, 21, 407–436.
(the highest characteristic conglomerate is argued to be reworked Anderton, R. 1988. Dalradian slides and basin development: a radical
and represents a tidal channel deposit, Hambrey et al. 1991) interpretation of stratigraphy and structure in the SW and Central
and the facies succession does not resemble that of carbonate Highlands of Scotland. Journal of the Geological Society, London,
successions found elsewhere. Member 3 carbonates have only 145, 669– 678.
one exact facies counterpart in the geological record – the upper Arnaud, E. 2002. Sedimentological analysis of Neoproterozoic glacio-
Canyon Fm. of East Greenland – and the latter occurs some genic successions in Norway and Scotland. Unpublished PhD thesis,
distance above the upper of two glacial deposits in that area School of Geography and Geology, McMaster University.
(Fairchild 1989). Arnaud, E. 2004. Giant cross-beds in the Neoproterozoic Port Askaig
A difference of viewpoint exists about the significance of the Formation, Scotland: implications for snowball Earth. Sedimentary
Geology, 165, 155–174.
negative C-isotope anomalies below the Port Askaig Fm. and in
Arnaud, E. & Eyles, C. H. 2002. Catastrophic mass failure of a
the overlying Bonahaven Fm. Prave et al. (2009) argue that the
Neoproterozoic glacially-influenced continental margin, the Great
C-isotope stratigraphy recorded in the Dalradian succession as a Breccia, Port Askaig Formation, Scotland. Sedimentary Geology,
whole bears close comparison with global trends and that the pub- 151, 313– 333.
lished signatures can be taken at face value as primary signatures. Arnaud, E. & Shields, G. 2005. The sedimentary record of a Neoproter-
The alternative view is that wherever impure ferroan dolomites are ozoic glaciation. (IGCP# 512 International Field workshop guide-
present, it would be expected that negative deviations from marine book), International Association of Sedimentologists Conference on
signals would be present. This applies to some (but not all) of the Glacial Sedimentary Processes and Products, University of Wales,
pre-glacial facies of the Lossit Limestone on the Garvellachs and Aberystwyth, 29 August– 3 September 2005, 60.
to the samples of Member 3 of the Bonahaven Fm. that have Arnaud, E. & Eyles, C. H. 2006. Neoproterozoic environmental change
been studied so far. The exact facies equivalents of the Bonahaven recorded in the Port Askaig Formation, Scotland: climatic and tec-
rocks in East Greenland show coherent variations in negative d13C tonic controls on sedimentation. Sedimentary Geology, 183, 99 – 124.
PORT ASKAIG FORMATION, SCOTLAND 641

Bailey, E. B. 1916. The Islay Anticline (Inner Hebrides). Quaterly Fairchild, I. J. 1991. Origins of carbonate in Neoproterozoic stro-
Journal of the Geological Society of London, 72, 132– 164. matolites and the identification of modern analogues. Precambrian
Benn, D. I. & Prave, A. R. 2006. Subglacial and proglacial glacitectonic Research, 53, 281– 299.
deformation in the Neoproterozoic Port Askaig Formation, Scotland. Fairchild, I. J., Spiro, B., Herrington, P. M. & Song, T. 2000. Controls
Geomorphology, 75, 266–280. on Sr and C isotope compositions of Neoproterozoic Sr-rich lime-
Bluck, B. J. & Dempster, T. J. 1991. Exotic metamorphic terranes in the stones of East Greenland and North China. In: Grotzinger, J. P &
Caledonides: tectonic history of the Dalradian block, Scotland. James, N. P. (eds) Carbonate Sedimentation and Diagenesis in the
Geology, 19, 1133– 1136. Evolving Precambrian World. SEPM Special Publication, 67,
Borradaile, G. J. 1979. Strain study of the Caledonides in the Islay 297– 313.
region, SW Scotland: implications for strain histories and defor- Fanning, C. M. & Link, P. K. 2004. U–Pb SHRIMP ages of Neoproter-
mation mechanisms in greenschists. Journal of the Geological ozoic (Sturtian) glaciogenic Pocatello Formation, southeastern Idaho.
Society, London, 136, 77 –88. Geology, 32, 881– 884.
Boulton, G. S. 1972. Modern arctic glaciers as depositional models for Fitches, W. R., Pearce, N. J. G., Evans, J. A. & Muir, R. J. 1996. Pro-
former ice sheets. Journal of the Geological Society, London, 128, venance of late Proterozoic Dalradian tillite clasts, Inner Hebrides,
361– 393. Scotland. In: Brewer, S. T. (ed.) Precambrian Crustal Evolution
Brasier, M. D. & McIlroy, D. 1998. Neonereites uniserialis from in the North Atlantic Region. Geological Society, Special Publi-
c. 600 Ma year old rocks in western Scotland and the emergence of cations, London, 112, 367– 377.
animals. Journal of the Geological Society, London, 155, 5 –12. Greiling, R. O. & Smith, A. G. 2000. The Dalradian of Scotland: Missing
Brasier, M. D. & Shields, G. 2000. Neoproterozoic chemostratigraphy link between the Vendian of Northern and Southern Scandinavia?
and correlation of the Port Askaig glaciation, Dalradian Supergroup Physics and Chemistry of the Earth, Part A, 25, 495– 498.
of Scotland. Journal of the Geological Society, London, 157, Halliday, A. N., Graham, C. M., Aftalion, M. & Dymoke, P. 1989.
909– 914. Short paper: The depositional age of the Dalradian Supergroup:
BRITISH GEOLOGICAL SURVEY. 1994. North Islay. Scotland Sheet 27. Solid U–Pb and Sm–Nd isotopic studies of the Tayvallich Volcanics,
and drift geology. 1:50000 provisional series. British Geological Scotland. Journal of the Geological Society, London, 146, 3– 6.
Survey, Keyworth. Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. &
Cawood, P. A., Nemchin, A. A., Smith, M. & Loewy, S. 2003. Source of Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-
the Dalradian Supergroup constrained by U– Pb dating of detrital isotope record. Geological Society of America Bulletin, 117,
zircon and implications for the East Laurentian margin. Journal of 1181–1207.
the Geological Society, London, 160, 231–246. Hambrey, M. J. 1983. Correlation of Late Proterozoic tillites in the North
Dalziel, I. W. D. 1994. Precambrian Scotland as a Laurentia-Gondwana Atlantic region and Europe. Geological Magazine, 120, 209–232.
link: Origin and signficance of cratonic promontories. Geology, 22, Hambrey, M. J., Fairchild, I. J., Glover, B. W., Stewart, A. D.,
589– 592. Treagus, J. E. & Winchester, J. A. 1991. The Late Precambrian
Dalziel, I. W. D. & Soper, N. J. 2001. Neoproterozoic extension on the Geology of the Scottish Highlands and Islands. Geologists Associ-
Scottish Promontory of Laurentia: paleogeographic and tectonic ation Guide No. 44. The Geologists’ Association, London, England.
implications. Journal of Geology, 109, 299–317. Harris, A. L., Baldwin, C. T., Bradbury, H. J., Johnson, H. D. &
Dempster, T. J., Rogers, G. et al. 2002. Timing of deposition, orogen- Smith, R. A. 1978. Ensialic basin sedimentation: the Dalradian
esis and glaciation within the Dalradian rocks of Scotland: constraints Supergroup. In: Bowes, D. R. & Leake, B. L. (eds) Crustal Evolution
from U– Pb zircon ages. Journal of the Geological Society, London, in Northwestern Britain and Adjacent Regions. Seel House Press,
159, 83– 94. Liverpool, 115–138.
Evans, J. A., Fitches, W. R. & Muir, R. J. 1998. Laurentian clasts in Harris, A. L., Haselock, P. J., Kennedy, M. J., Mendum, J. R., Long,
a Neoproterozoic tillite from Scotland. Journal of Geology, 106, C. B., Winchester, J. A. & Tanner, P. W. G. 1993. The Dalradian
361– 366. Supergroup in Scotland, Shetland, and Ireland. In: Gibbons, W. &
Eyles, C. H. 1988. Glacially and tidally-influenced shallow marine sedi- Harris, A. L. (eds) A Revised Correlation of Precambrian Rocks in
mentation of the Late Precambrian Port Askaig Formation, Scotland. the British Isles. Geological Society, Special Report, London, 33– 53.
Palaeogeography, Palaeoclimatology, Palaeoecology, 68, 1 –25. Howarth, R. J. 1971. The Port Askaig tillite succession (Dalradian) of
Eyles, C. H. & Eyles, N. 1983. Glaciomarine model for upper Precam- Co. Donegal. Proceedings of the Royal Irish Academy, 71, 1 –36.
brian diamictites of the Port Askaig Formation, Scotland. Geology, Hutton, D. H. W. & Alsop, G. I. 2004. Evidence for a major Neopro-
11, 692– 696. terozoic orogenic unconformity within the Dalradian Supergroup
Eyles, N. & Clark, B. M. 1985. Gravity-induced soft-sediment defor- of NW Ireland. Journal of the Geological Society, London, 161,
mation in glaciomarine sequences of the Upper Proterozoic Port 629– 640.
Askaig Formation, Scotland. Sedimentology, 32, 789–814. Johnston, J. D. 1995. Pseudomorphs after ikaite in a glaciomarine
Fairchild, I. J. 1977. Phengite spherules from the Dalradian Bonahaven sequence in the Dalradian of Donegal, Ireland. Scottish Journal of
Formation, Islay, Scotland: glauconitized microfossils? Geological Geology, 31, 3 –9.
Magazine, 114, 355– 364. Kendall, B. S., Creaser, R. A. & Selby, D. 2006. Re– Os geochrono-
Fairchild, I. J. 1980a. Sedimentation and origin of a late Precambrian logy of postglacial black shales in Australia: constraints on the
‘dolomite’ from Scotland. Journal of Sedimentary Petrology, 50, timing of the Sturtian glaciation. Geology, 34, 729–732.
423– 446. Kilburn, C., Pitcher, W. S. & Shackleton, R. M. 1965. The stratigra-
Fairchild, I. J. 1980b. Stages in Precambrian dolomitization, Scotland: phy and origin of the Port Askaig boulder bed series (Dalradian).
cementing v. replacement textures. Sedimentology, 27, 631–650. Geological Journal, 4, 343– 360.
Fairchild, I. J. 1980c. The structure of NE Islay. Scottish Journal of Litherland, M. 1980. The stratigraphy of the Dalradian rocks around
Geology, 16, 189–197. Loch Creran, Argyll. Scottish Journal of Geology, 16, 105–123.
Fairchild, I. J. 1985a. Comment on ‘Glaciomarine model for upper Max, M. D. 1981. E20: Dalradian Tillite of northwestern Ireland. In:
Precambrian diamictites of the Port Askaig Formation, Scotland’. Hambrey, M. J. & Harland, W. B. (eds) Earth’s Pre-Pleistocene
Geology, 13, 89 –90. Glacial Record. Cambridge University Press, Cambridge, 640–642.
Fairchild, I. J. 1985b. Petrography and carbonate chemistry of McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial
some Dalradian dolomitic metasediments: preservation of trinity: Neoproterozoic Earth history within the British-Irish Caledo-
diagenetic textures. Journal of the Geological Society, London, nides. Geology, 34, 909– 912.
142, 167– 185. Mulder, T. & Alexander, J. 2001. The physical character of subaqueous
Fairchild, I. J. 1989. Dolomitic stromatolite-bearing units with storm sedimentary density flows and their deposits. Sedimentology, 48,
deposits from the Vendian of East Greenland and Scotland: a case 269– 299.
of facies equivalence. In: Gayer, R. A. (ed.) The Caledonide Geology Nardin, T. R., Hein, F. J., Gorsline, D. S. & Edwards, B. D. 1979. A
of Scandinavia. Graham & Trotman, London, 275– 283. review of mass movement processes, sediment and acoustic
642 E. ARNAUD & I. J. FAIRCHILD

characteristics and contrasts in slope and base of slope systems Dalradian Supergroup of Scotland. Transactions of the Royal
v. canyon fan basin floor systems. In: Doyle, L. J. & Pilkey, O. Society of Edinburgh, Earth Sciences, 73, 59 –65.
H. (eds) Geology of Continental Slopes. Society of Economic Paleon- Tanner, P. W. G. 1998. Interstratal dewatering origin for polygonal
tologists and Mineralogists, Special Publication, 61 –73. patterns of sand-filled cracks; a case study from late Proterozoic meta-
Noble, S. R., Hyslop, E. K. & Highton, A. J. 1996. High precision U –Pb sediments of Islay, Scotland. Sedimentology, 45, 71 –89.
monazite geochronology of the c. 806 Ma Grampian Shear Zone and Tanner, P. W. G. & Shackleton, R. M. 1979. Structure and stratigraphy
the implications for the evolution of the Central Highlands of of the Dalradian rocks of the Bennabeola area; Connemara, Eire. In:
Scotland. Journal of the Geological Society, London, 153, 511– 514. Harris, A. L., Holland, C. H. & Leake, B. E. (eds) The Caledo-
Panahi, A. & Young, G. M. 1997. A geochemical investigation into the nides of the British Isles – Reviewed. Geological Society, London,
provenance of the Neoproterozoic Port Askaig tillite, Dalradian Special Publications, 8, 243 – 256.
Supergroup, western Scotland. Precambrian Research, 85, 81– 96. Tanner, P. W. G. & Bluck, B. J. 1999. Current controversies in the
Prave, A. R. 1999. The Neoproterozoic Dalradian Supergroup of Scot- Caledonides. Journal of the Geological Society, London, 156,
land: an alternative hypothesis. Geological Magazine, 136, 609– 617. 1137– 1141.
Prave, A. R., Fallick, A. E., Thomas, C. W. & Graham, C. M. 2009. A Tanner, P. W. G., Alsop, G. I. & Hutton, D. 2005. Discussion on
composite C-isotope profile for the Neoproterozoic Dalradian Super- evidence for a major Neoproterozoic orogenic unconformity within
group of Scotland and Ireland. Journal of the Geological Society, the Dalradian Supergroup of NW Ireland. Journal of the Geological
London, 166, 845–857. Society, London, 162, 221–224.
Rooney, A. D., Chew, D. M. & Selby, D. 2011. Re-Os geochronology of Tarling, D. H. 1974. A paleomagnetic study of Eocambrian tillites in
the Neoproterozoic-Cambrian Datradian Supergroup of Scotland and Scotland. Journal of the Geological Society, London, 130, 163– 177.
Ireland: implications for Neoproterozoic stratigraphy, glaciations and Thomas, C. W., Graham, C. M., Ellam, R. M. & Fallick, A. E. 2004.
87
Re-Os systematics. Precambrian Research, 185, 202– 214. Sr/86Sr chemostratigraphy of Neoproterozoic Dalradian lime-
Sawaki, Y., Kawai, T. et al. 2010. 87Sr/86Sr chemostratigraphy of stones of Scotland and Ireland: constraints on depositional ages
Neoproterozoic Dalradian carbonates below the Port Askaig glacio- and time scales. Journal of the Geological Society, London, 161,
genic Formation, Scotland. Precambrian Research, 179, 150– 164. 229– 242.
Soper, N. J. 1994. Neoproterozoic sedimentation on the northeast margin Thomson, J. 1871 (for 1870). On the occurrence of pebbles and boulders
of Laurentia and the opening of Iapetus. Geological Magazine, 131, of granite in schistose rocks in Islay, Scotland. 40th Meeting British
291– 299. Association, Liverpool, Transactions, 88.
Soper, N. J. & Anderton, R. 1984. Did the Dalradian slides originate as Treagus, J. E. 1987. The structural evolution of the Dalradian of the
extensional faults? Nature, 307, 357–360. Central Highlands of Scotland. Transactions of the Royal Society of
Spencer, A. M. 1971. Late Precambrian glaciation in Scotland. Memoirs Edinburgh, 78, 1– 15.
of the Geological Society of London, 6, 1 –100. Treagus, S. H. & Treagus, J. E. 2002. Studies of strain and rheology of
Spencer, A. M. 1975. Late Precambrian glaciation in the North Atlantic conglomerates. Journal of Structural Geology, 24, 1541–1567.
region. In: Wright, A. E. & Moseley, F. (eds) Ice Ages, Ancient Urrutia-Fucugauchi, J. & Tarling, D. H. 1983. Paleomagnetic proper-
and Modern. Seel House Press, Liverpool, 217– 240. ties of Eocambrian sediments in Northwestern Scotland: implications
Spencer, A. M. & Pitcher, W. S. 1968. Occurrence of the Port Askaig for world wide glaciation in the Late Precambrian. Palaeogeography,
Tillite in north-east Scotland. Proceedings of the Geological Palaeoclimatology, Palaeoecology, 41, 325– 344.
Society of London, 1650, 195– 198. Walkden, G., Parker, J. & Kelley, S. 2002. A late Triassic impact
Spencer, A. M. & Spencer, M. O. 1972. The late Precambrian/Lower ejecta layer in southwestern Britain. Science, 298, 2185–2188.
Cambrian Bonahaven Dolomite of Islay and its stromatolites. Scottish Yardley, B. W. D., Vine, F. J. & Baldwin, C. T. 1982. The plate tec-
Journal of Geology, 8, 269–282. tonic setting of NW Britain and Ireland in late Cambrian and early
Stupavsky, M., Symons, D. T. A. & Gravenor, C. P. 1982. Evidence Ordovician times. Journal of the Geological Society of London,
for metamorphic remagnetisation of upper Precambrian tillite in the 139, 455– 463.
Chapter 63

The Neoproterozoic glaciogenic deposits of Scotland and Ireland

A. R. PRAVE1 * & A. E. FALLICK2


1
Department of Earth Science, University of St Andrews, St Andrews, KY16 9AL, UK
2
Scottish Universities Environmental Research Centre, East Kilbride, G75 0QF, UK
*Corresponding authors (e-mail: ap13@st-andrews.ac.uk)

Abstract: Of the three major Neoproterozoic supracrustal units in the Scottish and Irish Highlands (the Torridonian, Moine and
Dalradian Supergroups), only the latter contains evidence of Neoproterozoic glaciations. The Dalradian is siliciclastic-dominated and
constitutes much of the Scottish–Irish Highlands between the Great Glen and Highland Boundary Fault Zones, and their correlatives
in Ireland. At the time of writing, three stratigraphically distinct glacial intervals in the Dalradian have been documented in the literature.
The oldest is the Port Askaig Formation (Fm.) at the base of the Argyll Group (see Arnaud & Fairchild 2011). It ranges from several tens
to many hundreds of metres in thickness and occurs in numerous localities in Scotland and the north of Ireland. A second glacial is
recorded in the middle part of the Argyll Group (Easdale Subgroup) and consists of localised sedimentary breccias as well as pelites
and schists containing dropstone/lonestone units inferred to be ice-rafted debris; these rocks are patchily preserved and typically a
few metres or less in thickness. It is sharply overlain by a variably developed carbonate unit that is marked by a 1 –7-m-thick, light-
coloured, basal dolostone or dolomitic limestone interpreted as a cap carbonate. This succession is best preserved in Donegal,
Ireland, as the Stralinchy–Reelan glacial and Cranford cap-carbonate sequence. A correlative cap carbonate, the Whiteness Limestone,
has been identified in the Shetland Islands. The third and youngest glacial is represented by locally preserved dropstone and polymict
diamictite beds ranging in thickness from several to a few tens of metres in thickness in the lower Southern Highland Group. These
include the MacDuff and Loch na Cille Boulder Beds in, respectively, NE and SW Scotland, and the Inishowen Beds in Donegal, Ireland.

The late Proterozoic geological framework of the Scottish –Irish origin; research is currently being undertaken by the authors to
Highlands consists largely of three (meta)sedimentary succes- document those that are.
sions (each many kilometres thick): the Torridonian, preserved The best-known, longest-studied ancient glacial deposit in the
NW of the Moine Thrust; the Moine, in the upper plate of the Dalradian is the Port Askaig Fm. at the base of the Argyll Group
Moine Thrust zone southeastward to the Great Glen Fault; and (Fig. 63.2). It is present in numerous localities scattered across
the Dalradian, which is sandwiched between the Great Glen the Dalradian outcrop belt and is more than 500 m thick at its
and Highland Boundary Fault Zones (Fig. 63.1). Only the Dalra- type locality on the Inner Hebridean island of Islay (Fig. 63.1).
dian contains a record of Neoproterozoic glaciations. A purported The seminal study of Spencer (1971), building upon the efforts
glacial deposit was inferred for a unit in the lower part of the of earlier workers, confirmed its glaciogenicity. Another, but
Torridonian (Davison & Hambrey 1996) but this has been younger glaciogenic interval to be well documented is the
shown to be incorrect (Stewart 1997, 2002). Along most of the 12-m-thick MacDuff Boulder Bed, exposed near the eponymous
Moine outcrop belt the rocks are too strongly deformed to town along the Banffshire coastline in NE Scotland (Figs 63.1 &
obtain sedimentological information, but even in areas of low 63.2). Sutton & Watson (1954) initially, and Stoker et al. (1999)
strain no evidence has been found for glaciogenic deposits. subsequently, showed that this unit contains ice-rafted dropstones.
This is not surprising given that both the Moine and Torridonian Condon & Prave (2000) reported on other ice-rafted-debris beds a
pre-date Cryogenian time: the Moine was deposited prior to few decimetres to several metres thick along the Inishowen Penin-
c. 840 Ma, the age of cross-cutting igneous intrusions and mag- sula of Donegal, Ireland (Figs 63.1 & 63.2). They proposed corre-
matic overgrowths on detrital zircons (e.g. Vance et al. 1998; lations of these beds to the MacDuff Boulder Bed, as well as
Millar 1999; Kirkland et al. 2008), and deposition of the Torrido- another Boulder Bed, the Loch na Cille Boulder Bed, on the
nian occurred before c. 950 Ma, as based on Rb – Sr and Pb/Pb Tayvallich Peninsula in southwestern Scotland (Fig. 63.1); all of
diagenetic ages and U –Pb ages on the youngest detrital zircons these occurrences are in the lower Southern Highland Group.
(Turnbull et al. 1996; Rainbird et al. 2001). McCay et al. (2006) documented the Stralinchy –Reelan glacial
The Dalradian is a mainly metasedimentary succession sub- and Cranford Limestone cap-carbonate sequence in the middle
divided into four major units (Harris et al. 1994): the Grampian, of the Argyll Group (near the top of the Easdale Subgroup) in
Appin, Argyll and Southern Highland Groups (Figs 63.1 & Donegal (Figs 63.1 & 63.2). This work confirmed that the Dalra-
63.2). Greenschist- to amphibolite-facies metamorphism and pen- dian contained three, stratigraphically distinct glacial intervals:
etrative tectonic fabrics associated with the Neoproterozoic Knoy- (i) the Port Askaig Fm. (base of the Argyll Group); (ii) the
dartian (c. 840 –720 Ma; e.g. Piasecki & van Breeman 1983; glacial –cap carbonate couplet of the Stralinchy –Reelan and Cran-
Vance et al. 1998; Tanner & Evans 2003) and Ordovician Gram- ford Limestone formations (Easdale Subgroup); and (iii) the
pian (c. 465 Ma; Oliver et al. 1998) tectonothermal events in MacDuff –Inishowen – Loch na Cille Beds (lower Southern High-
many places obscure original sedimentary features and render pro- land Group). Subsequently, Prave et al. (2009a) interpreted the
blematic any assessment of depositional processes. Areas of low Whiteness Limestone in the Shetland Islands, some 200 km
strain are geological rosetta stones from which to establish strati- north of the Scottish mainland, as a cap-carbonate correlative
graphic relationships and make palaeoenvironmental interpret- with the Cranford Limestone.
ations, and to extrapolate those to the more deformed portions of
the Dalradian outcrop belt. It needs to be noted that in much of
the older literature on the Dalradian, the term ‘Boulder Bed’ Structural framework
appears. This term was applied to schists and pelites containing
polymict, variably shaped clasts in matrix-supported textures The rather severe tectonothermal overprinting of the Dalradian
with poor sorting. Many of these deposits are not glaciogenic in rocks during early Palaeozoic Caledonian orogenesis resulted in

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 643– 648. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.63
644 A. R. PRAVE & A. E. FALLICK

Fig. 63.1. Highly generalized geological map of the Scottish– Irish Dalradian Highlands highlighting the outcrop belts of the Argyll and Southern Highlands Groups.
Small inset map shows simplified tectonic template of the Highlands. GGF, Great Glen Fault; HBF, Highland Boundary Fault; MT, Moine Thrust. These fault
zones demarcate the three Proterozoic Supergroups of the Highlands: the Torridonian occurs NW of the Moine Thrust, the Moine is between the Moine Thrust and Great
Glen Fault, and the Dalradian (grey-shading) is bounded by the Great Glen and Highland Boundary Fault Zones. Main map, locations discussed in the text: C, Cranford;
LC, Loch na Cille; M, MacDuff; P, Port Askaig; R, Reelan; S, Stralinchy.

structural disruption of the Dalradian outcrop belt by a series of 1984). This structural and metamorphic overprinting makes
so-called tectonic ‘slides’. Most of these are interpreted as exten- assessment of original depositional settings problematic. Never-
sional collapse structures reactivated along original reverse or theless, the Dalradian Supergroup is often interpreted as recording
thrust faults (Harris et al. 1994), but several are thought to rep- deposition in a succession of rift basins (Anderton 1982, 1985;
resent original basin-bounding faults (e.g. Soper & Anderton Robertson & Smith 1999; Dalziel & Soper 2001; Arnaud &
Eyles 2006). This view, however, has been questioned, and the
exact nature of basinal genesis remains an outstanding problem
Cambrian Leny Limestone Ma
to resolve (Prave 1999). The terminal Neoproterozoic geodynamic
542
Southern Highland Gp

evolution of the Dalradian succession is, however, relatively


P

uncontroversial and records syn-depositional extensional tecton-


U

ism and continental rifting (Anderton 1979, 1982, 1985, 1988;


O

582 Soper & Anderton 1984). This phase of extensional tectonics


MacDuff–Inishowen–Loch na Cille Boulder Beds began prior to deposition of the middle glacial –cap carbonate
R

couplet (Stralinchy –Reelan and Cranford formations) and pro-


Ediacaran

gressed to active rifting through the lower Southern Highland


G

601 Ma Tayvallich volcanics


Group. Antecedent to this, the lithostratigraphic framework of
Tayvallich Sgps
R

the underlying parts of the Dalradian is not readily attributable


Crinan &

to deposition in rift basins and any such interpretation should be


E

viewed with a healthy scepticism (Prave 1999).


P
Argyll Group

‘upper’ Ardrishaig–Craignish & Port Ellen Phyllites


U

Subgroup

Cranford–Whiteness Limestones (cap carbonates)


Easdale

IRD
635 Stratigraphy
S

‘lower’ Ardrishaig–Craignish Phy.


Stralinchy
Congl. Easdale Slates
Scarba Congl. Group and subgroup correlations across the Dalradian outcrop belt
Atholl Subgroups Subgroup

Jura – Slieve Tooey Quartzites


are achievable but, because of variable exposure quality and perva-
Islay

Bonahaven Dolostone
sive deformational fabrics, finer-resolution correlations between
N

Port Askaig Formation 713


many areas are seldom attainable. Consequently, correlations
A

Ballachulish & Blair

Islay – Glencolumbkille Limestones from one locality to the next rely largely on the recognition of
Cryogenian
Appin Gp
I

packages of associated lithologies, rather than single units or


D

stratal surfaces (such as those utilized in sequence stratigraphy).


Nevertheless, over 150 years of mapping has provided a reliable
A

geological framework from which to construct a stratigraphy for


R
Grampian Gp

the Dalradian’s glaciogenic units. Like many long-studied succes-


L

sions, the Dalradian is rife with stratigraphic nomenclature. To


A

avoid getting bogged down in a morass of names, discussion is


restricted to the stratigraphic units directly pertinent to the glacio-
D

? genic rocks and their associated strata (see Fig. 63.2).


? ? (nature of base, whether unconformable or stratigraphic, remains unresolved)
metamorphic rocks having 840–800 Ma deformational ages (U-Pb zircon & monazite)

Fig. 63.2. Simplified stratigraphic framework of the Dalradian Supergroup. Glaciogenic deposits and associated strata
The three recognized glaciogenic intervals in the Dalradian (shaded) and their
inferred correlation to geochronological ages and subdivisions of The following focuses on the two younger Neoproterozoic glacial
Neoproterozoic Earth history are shown, as are other stratigraphic units units and their associated strata. The oldest glacial, the Port
mentioned in the text. IRD, ice-rafted debris. See text for details. Askaig, is discussed by Arnaud & Fairchild (2011).
NEOPROTEROZOIC GLACIOGENIC DEPOSITS OF SCOTLAND AND IRELAND 645

The Stralinchy – Reelan formations and the Cranford and carbonate is everywhere sharp. Commonly, only the basal cap
Whiteness cap carbonates dolostone is present and in these localities the top also appears
sharp. Where the entire cap-carbonate sequence is developed
McCay et al. (2006) showed that the Stralinchy Conglomerate and (the eponymous type localities of the Cranford and Whiteness
its basinal equivalent containing isolated dropstones, the Reelan Limestones), the basal cap displays a gradational contact into over-
Fm., together with the overlying Cranford Limestone constitute a lying carbonate rhythmites and/or a thin calc-pelite unit, itself
glacial –cap carbonate succession in the middle part of the transitioning into carbonate rhythmites. Thick siliciclastic-
Argyll Group in Donegal, Ireland (Figs 63.1 & 63.2). The glacial dominated intervals overlie the cap sequences.
rocks occur as lenses (from 0 to several tens of metres in thickness)
of mono- to poly-mict, poorly sorted sedimentary breccia (Stralin-
chy Conglomerate) and laterally equivalent thin-bedded pelite and MacDuff, Inishowen and the Loch na Cille Boulder Beds
limestone containing pebble- to cobble-sized dropstones and
lonestones (Reelan Fm.). Clasts consist of quartzo-feldspathic, The MacDuff Boulder Bed occurs in the core of a syncline of the
quartzitic, schistose and carbonate lithologies. The sedimentary lower Southern Highland Group rocks along the coast adjacent to
breccias alter from clast-supported to diamictic in texture, clasts the eponymous Banffshire town (Fig. 63.1); it is c. 12 m thick and
range in size from granule to metre-sized blocks, and matrix has long been recognized as glaciogenic in origin (Sutton &
composition varies along strike from siliciclastic- to carbonate- Watson 1954; Stoker et al. 1999). It was speculated to be as
dominated. Lonestones and dropstones pierce and distort young as Ordovician in age (Hambrey 1983; Molyneux 1997),
laminae in decimetre-thick intervals of finely compositionally but this is wrong: regional mapping leaves no doubt that it is
layered pelite and limestone; on-lap and splash-up structures are much older and in the Southern Highland Group (e.g. Read
associated with some of the larger clasts. Sitting sharply on these 1923). The MacDuff Boulder Bed is part of a thick succession of
glaciogenic units is the Cranford Limestone. Its base is marked siliciclastic rocks interpreted as a deep-marine fan complex (e.g.
by a 1– 7-m-thick, laterally continuous tan-coloured dolomicrite, Trewin 1987). What distinguishes it is the presence of several
that is, a cap carbonate. At the type locality, the cap is overlain dropstone intervals in which variably shaped clasts, consisting of
by 200–300 m of limestone-rhythmite and rhythmite breccia. In quartzite, gneiss, igneous and pale-coloured carbonate rocks
basinal settings, the Cranford Limestone shales out (metamor- ranging from angular to rounded and from granule to boulder in
phosed to phyllite; Alsop & Hutton 1990) and only the basal cap size, can be seen to pierce laminae, be on-lapped by subsequent
dolostone remains. laminae and, in some places, show splash-up structures. They
As documented by Prave et al. (2009a), an inferred correlative are concentrated in discrete, decimetre-thick units indicating that
cap carbonate, the Whiteness Limestone, is present in the Shetland they originated during periods of rainout, rather than as part of
Islands (these islands lie 200 km north of mainland Scotland). some continuous background sedimentation process, and an
Similar to the Cranford Limestone (and to cap carbonates else- origin as ice-rafted debris, not tills (suggested by Hambrey &
where; e.g. Hoffman & Schrag 2002; Fairchild & Kennedy Waddams 1981), is certain (Sutton & Watson 1954; Stoker et al.
2007), the Whiteness Limestone has a basal, 2– 5-m-thick, light- 1999). Stoker et al. (1999) also detailed aspects of seven
coloured dolostone to dolomitic limestone overlain by 20– 40 m decimetre-thick diamictite beds. These contain a variety of clasts
of thin-bedded limestone-pelite rhythmites. These, in turn, pass of varying sizes and shapes in a mud matrix and were interpreted
upward into a tens-of-metres-thick pelitic interval, itself overlain as resedimented glaciomarine deposits.
by many hundreds of metres of thick-bedded calcitic marbles. Condon & Prave (2000) documented glaciogenic deposits in the
The basal dolostone sits sharply on a thick (1– 2 km) sequence Southern Highland Group along the northern Inishowen Peninsula
of siliciclastic units devoid of carbonate rocks. Within 5– 10 m (Fig. 63.2), Donegal, Ireland. These consist of five dropstone units,
of the base of the cap, the underlying schistose rocks become con- each many centimetres to several decimetres thick, spaced irregu-
spicuously finer-grained and darker-coloured. It is within this larly through a c. 500-m-thick arkosic psammite-pelite succession.
interval that rare quartzo-feldspathic lonestone clasts can be The dropstone units are characterized by granule- to pebble-sized
found. The absence of any physical or textural evidence for depo- clasts of quartzite, gneiss, schist and igneous rocks; in many places
sition of the lonestones by traction currents, sediment-gravity-flow the clasts can be observed to pierce laminae and be on-lapped by
or volcanic ejecta processes, and their stratigraphic restriction to successive laminae. These units are separated by decimetre- to
subjacent to the base of the Whiteness cap, led Prave et al. metre-thick intervals of fine-grained pelite devoid of lonestones
(2009a) to interpret the lonestones as ice-rafted debris, that is, a or gritty horizons. The arkosic nature of these strata is noteworthy
glacially influenced phase of sedimentation pre-dating deposition because, for hundreds of metres above and below this interval,
of the cap carbonate. units are quartzitic in composition. Condon & Prave (2000) and
The presence of glaciogenic rocks and cap carbonates at the Condon et al. (2002) interpreted the dropstone beds as ice-rafted
southwestern (Donegal) and northeastern (Shetland) edges of the debris and speculated that the arkosic composition of the encasing
Dalradian outcrop belt has focused research efforts on mainland strata was due to cold-climate weathering processes. Condon et al.
Scotland to document correlative glacial –cap carbonate couplets (2002) also documented an iceberg dump structure, a poorly
there, or find a reason for their absence. sorted, coarse-grained dome-shaped sedimentary feature having
a flat basal surface moulded onto the top of a pelite bed. It is
c. 1–2 m in length along its longest exposed axis and c. 10–
20 cm thick along its crest before tapering off irregularly to its
Boundary relations with overlying and underlying thinned edges.
non-glacial units A distinctive unit occurs in the lower Southern Highland Group
in SW Scotland along the Tayvallich Peninsula (Fig. 63.2), the
In many places across the Dalradian outcrop belt, Caledonian Loch na Cille Boulder Bed (Elles 1934). It occurs as discontinuous
strain compromises contacts, but in areas of low strain, and lenses, typically several metres thick, and consists of matrix-
where the glaciogenic units consist of sedimentary breccia or dia- supported polymict clasts, including felsite and mafic volcanites,
mictite (e.g. the Stralinchy type locality in Donegal), the contact is granitoid, sedimentary and metamorphic rock fragments (Alsop
sharp and erosive. Where the glaciogenic units consist of drop- et al. 2000). It was recognized and mapped separately because of
stone or lonestone intervals, defining contacts relies on finding dis- its distinctiveness from the associated hundreds of metres of volca-
persed clasts and/or the contact with the overlying cap carbonate nic and pelitic rocks. Its matrix is finer-grained, lighter-coloured
to confirm being in or near glacial strata. The base of the cap and more quartzo-feldspathic than the encasing rocks, and its
646 A. R. PRAVE & A. E. FALLICK

clast content is polymict, rather than monomict volcanic detritus. confirm that negative C-isotopic excursions are antecedent to
These observations, and compatible stratigraphic position of the two stratigraphically and temporally distinct glaciations and urge
Loch na Cille Boulder Bed with the MacDuff and Inishowen caution when attempting to use such excursions as chronostrati-
beds, indicate that it, too, is a glaciogenic deposit. graphic markers (see Halverson et al. 2005). The inferred cap
carbonates exhibit a trend commonly attributed to post-Mar-
inoan-equivalent successions (e.g. Kennedy et al. 1998), namely
Boundary relations with overlying and underlying values declining from c. –3‰ in the basal cap, reaching a nadir
non-glacial units at c. – 6‰ in overlying limestone rhythmites, and then recovering
back towards positive values up-section. It should be noted that the
The lowest observed, glacially influenced deposits in the Southern basal Whiteness Limestone on the Shetland Islands has initial
Highland Group rocks occur abruptly. They are eye-catching values between 1‰ and 2‰. A more detailed C-isotopic study
because they are compositionally distinct, both in clast content of the Dalradian is given in Prave et al. (2009b).
and matrix, from encasing strata whether those are sedimentary,
as in the case of the MacDuff and Inishowen rocks, or volcaniclas-
tic, as in the Loch na Cille Beds. The upper contacts of the Inisho- Other characteristics (e.g. economic deposits, biomarkers)
wen and Loch na Cille Beds are placed at the last occurrence of
ice-rafted debris; the MacDuff Boulder Bed has no top, it is the There are no economic deposits associated with the Dalradian’s
youngest unit in the core of an exposed syncline. The impression glacial rocks. The cap carbonates have, in many places, been quar-
one gets from all three units is that glaciogenic deposition was ried, but those operations were small and not commercially viable.
intermittent, discrete events superimposed on the overall Likewise, no depositional biomarkers have been found in any of
background sedimentation. these rocks, not surprising given the metamorphic grade.

Chemostratigraphy
Palaeolatitude and palaeogeography
Figure 63.3 provides a summary diagram of our C-isotopic data
associated with the Dalradian’s glaciogenic intervals. Dramatic There are no reliable depositional palaeomagnetic data on the
declines, from values of þ2‰ to þ6‰ to as low as –10‰ to Dalradian rocks; this renders palaeolatitudinal placements, at
–12‰, and recoveries mark the carbonate strata predating both best, speculative. Generalized palaeogeographies envisage east
the Port Askaig and Stralinchy-Reelan formations. These data to west (current-day coordinates) transitions from shallow-marine
shelf to deeper-marine settings in rift basins along the eastern
margin of Laurentia for most of Dalradian time (Anderton 1982,
1985; Harris et al. 1994). Although several nicely detailed local
studies have been undertaken (Klein 1970; Anderton 1976; Fair-
INISHOWEN–MACDUFF–LOCH NA CILLE GLAC. child 1980), much additional work is needed to help refine recon-
structions. Complicating matters are the wildly varying estimates
Marinoan–equivalent for the magnitude of pre-Devonian displacement along the Great
cap carbonates Glen Fault Zone; restoration of the Dalradian block can vary by
Whiteness Limestone
Cranford Limestone
as much as 1000 km (e.g. Dewey & Strachan 2003). Most recon-
structions incorporate the Dalradian as part of eastern Laurentia,
the position of which within a disassembling Rodinia is often
depicted as moving along a broadly low- to mid-latitude arc
during Cryogenian time (e.g. Dalziel 1997; Pisarvesky et al. 2008).
STRALINCHY – REELAN GLACIAL
Ardrishaig Phyllite (& correlatives) Geochronological constraints
pre–glacial decline
and recovery Geochronological constraints are frustratingly sparse for the
(Trezona anomaly Dalradian:
equivalent?)
Easdale Subgroup † The only direct depositional age for the Dalradian is from
carbonate rocks an ash bed in the Tayvallich Volcanics near the base of the
Southern Highland Group. It has a U –Pb zircon age of
601 + 4 Ma (Dempster et al. 2002) and occurs more than
Bonahaven 500 m beneath the youngest Dalradian glacial beds (in this
Formation
instance, the Loch na Cille Boulder Bed). The top of the Dalra-
PORT ASKAIG GLACIAL dian, well above the youngest glacial rocks, passes stratigraphi-
cally into the Cambrian Leny Limestone (Tanner 1995).
pre–glacial decline † The nature of the base of the Dalradian, whether unconformable
and recovery Islay–Glencolumbkille–
Limestones or depositional on rocks containing U –Pb zircon and monazite
deformational ages ranging from c. 840–800 Ma (Highton
et al. 1999; Noble et al. 1996), remains unresolved. If the
base defines an unconformity, then these are maximum ages
-15 -10 -5 0 5 10 for the initiation of Dalradian sedimentation; if the base is tran-
sitional (albeit tectonized), then the lower parts of the Dalradian
δ C‰ (V-PDB)
13
could be much older. Regardless, the earliest record of glacia-
Fig. 63.3. C-isotopic trends for part of the Dalradian Supergroup with inferred tion in the Dalradian, the Port Askaig Fm. at the base of the
correlations to known Neoproterozoic chemostratigraphic events. Data are Argyll Group, occurs some 10 km above these rocks.
from McCay et al. (2006), Prave et al. (2009a) and Prave & Fallick † The youngest concordant detrital zircon U –Pb ages for the
(unpublished). See text for discussion. Argyll Group are c. 1000 Ma (Cawood et al. 2003). These are
NEOPROTEROZOIC GLACIOGENIC DEPOSITS OF SCOTLAND AND IRELAND 647

not robust constraints for depositional ages, but do confirm a References


post-early Neoproterozoic age for the initiation of Argyll
Group sedimentation. Alsop, G. I. & Hutton, D. W. 1990. A review and revision of Dalradian
† Time lines for the Dalradian have been proposed using chron- stratigraphy in central and southern Donegal, Ireland. Irish Journal of
ostratigraphic constraints established for global climatic Earth Sciences, 10, 181–198.
events and secular variations in the isotopic compositions of Alsop, G. I., Prave, A. R., Condon, D. J. & Phillips, C. A. 2000.
Neoproterozoic oceans. The proposed linkage by McCay Cleaved clasts in Dalradian conglomerates: possible evidence for
et al. (2006) of the three Dalradian glacials to Neoproterozoic Neoproterozoic compressional tectonism in Scotland and Ireland?
glaciations known from elsewhere is viable and testable. In Geological Journal, 35, 87– 98.
this framework, the Port Askaig Fm. would be chronostratigra- Anderton, R. 1976. Tidal shelf sedimentation: an example from the
Scottish Dalradian. Sedimentology, 23, 429–458.
phically tied to the mid-late Neoproterozoic Sturtian glacia-
Anderton, R. 1979. Slope, submarine fans, and syn-depositional faults:
tion(s). We recognize that the age and number of glaciations sedimentology of parts of the Middle and Upper Dalradian in the
during this climatic phase remain contentious, but Brasier & SW Highlands of Scotland. In: Harris, A. L., Holland, C. H. &
Shields (2000) have argued, based on C and Sr isotopes, that Leake, B. E. (eds) The Caledonides of the British Isles – Reviewed.
the Port Askaig is best correlated to the c. 713 Ma Gubrah Geological Society, London, Special Publications, 8, 483–488.
glacial in Oman (Bowring et al. 2007). This, then, would be Anderton, R. 1982. Dalradian deposition and the late Precambrian –
the age of (or close to) initiation of Argyll Group sedimentation. Cambrian history of the North Atlantic region: a review of the early
The Cranford –Whiteness Limestones are interpreted as evolution of the Iapetus Ocean. Journal of the Geological Society,
Marinoan-equivalent cap carbonates (McCay et al. 2006; London, 139, 421– 431.
Prave et al. 2009a), thereby making this part of the Easdale Anderton, R. 1985. Sedimentation and tectonics in the Scottish Dalra-
Subgroup 635 Ma, that is the age for the end of the Marinoan dian. Scottish Journal of Geology, 21, 407– 436.
glaciation (Hoffmann et al. 2004; Condon et al. 2005). The Anderton, R. 1988. Dalradian slides and basin development: a radical
MacDuff – Inishowen –Loch na Cille beds, which occur above reinterpretation of stratigraphy and structure in the SW and Central
the 601 Ma Tayvallich Volcanics and their correlatives, can Highlands of Scotland. Journal of the Geological Society, London,
be reasonably linked to the 582 Ma Gaskiers glaciation 145, 669–678.
(Bowring et al. 2003, 2007). Arnaud, E. & Eyles, C. H. 2006. Neoproterozoic environmental change
recorded in the Port Askaig Formation, Scotland: climatic and
The best these data permit is to indicate that the post-Easdale Sub- tectonic controls on sedimentation. Sedimentary Geology, 183,
group portion of the Dalradian belongs to the Ediacaran Period and 99– 124.
that the underlying Dalradian is Cryogenian in age (Fig. 63.2). Arnaud, E. & Fairchild, I. J. 2011. The Port Askaig Formation, Dalra-
Such broad age constraints are unsatisfying. The irony is that, dian Supergroup, Scotland. In: Arnaud, E., Halverson, G. P. &
although it is one of the most-studied and mapped regions in the Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic
world, the Scottish –Irish Highlands is one of the most poorly Glaciations. Geological Society, London, Memoirs, 36, 635– 642.
constrained, geochronologically. Bowring, S., Myrow, P., Landing, E., Ramezani, J. & Grotzinger, J.
2003. Geochronological constraints on terminal Neoproterozoic
events and the rise of metazoans. Geophysical Research Abstracts,
Discussion 5, 13219.
Bowring, S. A., Grotzinger, J. P., Condon, D. J., Ramezani, J.,
Two main data sets have enabled construction of a glacial-based Newall, M. & Allen, P. A. 2007. Geochronologic constraint of
stratigraphic framework for the Scottish –Irish Dalradian. The the chronostratigraphic framework of the Neoproterozoic Huqf
first is the presence of discrete glacially influenced deposits Supergroup, Sultanate of Oman. American Journal of Science, 307,
ranging from thick diamictites to thin, discontinuous ice-rafted- 1097–1145.
debris beds (in areas of higher metamorphic grade, these are recog- Brasier, M. D. & Shields, G. 2000. Neoproterozoic chemostratigraphy
nized as fine-grained schists with lonestones). Excepting the Port and correlation of the Port Askaig glaciation, Dalradian Supergroup
Askaig Fm., with its impressive, laterally extensive diamictite of Scotland. Journal of the Geological Society, London, 157,
beds, the record of Neoproterozoic glaciation in the Dalradian 909– 914.
is modest and mostly represented by thin ice-rafted-debris Cawood, P. A., Nemchin, A. A., Smith, M. & Loewy, S. 2003. Source of
intervals. The second data set comes from carbonate units that the Dalradian Supergroup constrained by U –Pb dating of detrital
are lithologically and C-isotopically compatible with Neoprotero- zircon and implications for the East Laurentian margin. Journal of
zoic cap carbonates. the Geological Society, London, 160, 231–246.
The Scottish – Irish Dalradian succession contains evidence for Condon, D. J. & Prave, A. R. 2000. Two from Donegal: Neoproterozoic
three distinct glaciations, from oldest to youngest, (i) the Port glacial episodes on the NE margin of Laurentia. Geology, 28,
951– 954.
Askaig Fm. at the base of the Argyll Group (note that no associated
Condon, D. J., Prave, A. R. & Benn, D. 2002. Neoproterozoic glacial –
cap carbonate has been recognized), (ii) the Stralinchy – rainout intervals: observations and implications. Geology, 30, 35– 38.
Reelan formations (and correlative ice-rafted-debris beds) and Condon, D., Zhu, M., Bowring, S., Wang, W., Yang, A. & Jin, Y. 2005.
the Cranford and Whiteness cap carbonates in the Easdale U–Pb ages from the Neoproterozoic Doushantuo Formation, China.
Subgroup, and (iii) the MacDuff –Inishowen –Loch na Cille Science, 308, 95– 98.
Boulder Beds and ice-rafted-debris units in the lower Southern Dalziel, I. W. D. 1997. Neoproterozoic-Paleozoic geography and tec-
Highland Group. Geochronological constraints are sparse for the tonics: review, hypothesis, environmental speculation. Geological
Dalradian, but those that exist, combined with C-isotope chemo- Society of America Bulletin, 109, 16– 42.
stratigraphy, enable placing the base of the Ediacaran Period in Dalziel, I. W. D. & Soper, N. J. 2001. Neoproterozoic extension on the
the Easdale Subgroup and the underlying Dalradian units in the Scottish Promontory of Laurentia: paleogeographic and tectonic
Cryogenian. implications. Journal of Geology, 109, 299–317.
Davison, S. & Hambrey, M. J. 1996. Indications of glaciation at the base
Numerous colleagues have provided insight and help over the years and we would of the Proterozoic Stoer Group (Torridonian), NW Scotland. Journal
like to particularly acknowledge I. Alsop, D. Benn, D. Condon, G. McCay, of the Geological Society, London, 153, 139–149.
G. Oliver, C. Rose, J. Soper, R. Strachan and G. Thomas. We thank The Carnegie Dempster, T. J., Rogers, G. et al. 2002. Timing of deposition, orogen-
Trust for The Universities of Scotland, The Russell Trust and The Schroder Foun- esis and glaciation within the Dalradian rocks of Scotland: Con-
dation (Islay) for financial support. This represents a contribution of the IUGS- straints from U –Pb zircon ages. Journal of the Geological Society,
and UNESCO-funded IGCP (International Geoscience Programme) Project #512. London, 159, 83 – 94.
648 A. R. PRAVE & A. E. FALLICK

Dewey, J. F. & Strachan, R. A. 2003. Changing Silurian –Devonian Highland region of the Scottish Caledonides. Transactions of the
relative plate motion in the Caledonides: sinistral transpression to Royal Society of Edinburgh: Earth Sciences, 73, 119– 134.
sinistral transtension. Journal of the Geological Society, London, Pisarvesky, S. A., Murphy, J. B., Cawood, P. A. & Collins, A. S.
160, 219– 229. 2008. eeLate Neoproterozoic and early Cambrian palaeogeography:
Elles, G. L. 1934. The Loch na Cille Boulder Bed and its place in the models and problems. Journal of the Geological Society, London,
Highland succession, Geological Society of London Quarterly 294, 9– 31.
Journal, 91, 111– 147. Prave, A. R. 1999. The Neoproterozoic Dalradian Supergroup
Fairchild, I. 1980. Sedimentation and origin of a late Precambrian ‘dolo- of Scotland: an alternative hypothesis. Geological Magazine, 136,
mite’ from Scotland. Journal of Sedimentary Petrology, 50, 423–446. 609– 617.
Fairchild, I. & Kennedy, M. J. 2007. Neoproterozoic glaciation in the Prave, A. R., Strachan, R. A. & Fallick, A. E. 2009a. Global C cycle
Earth System. Journal of the Geological Society, London, 164, perturbations recorded in marbles: a record of Neoproterozoic Earth
895– 921. history within the Shetland Islands, Scotland. Journal of the Geologi-
Hambrey, M. J. 1983. Correlation of Late Proterozoic tillites in the North cal Society, London, 166, 129–135.
Atlantic region and Europe. Geological Magazine, 120, 209– 232. Prave, A. R., Fallick, A. E., Thomas, C. W. & Graham, C. M. 2009b. A
Hambrey, M. J. & Waddams, P. 1981. Glaciogenic boulder-bearing composite C-isotopic profile for the Neoproterozoic Dalradian Super-
deposits in the Upper Dalradian MacDuff Slates, northeastern group of Scotland and Ireland. Journal of the Geological Society,
Scotland. In: Hambrey, M. J. & Harland, W. W. (eds) Earth’s London, 166, 845–857.
Pre-Pleistocene Glacial Record. Cambridge University Press, Rainbird, R. H., Hamilton, M. A. & Young, G. M. 2001. Detrital zircon
Cambridge, 571– 575. geochronology and provenance of the Torridonian, NW Scotland.
Harris, A. L., Haselock, P. J., Kennedy, M. & Mendum, J. R. 1994. The Journal of the Geological Society, London, 158, 15 –27.
Dalradian Supergroup in Scotland, Shetland and Ireland. In: Read, H. H. 1923. The geology of the country around Banff, Huntly and
Gibbons, W. & Harris, A. L. (eds) A Revised Correlation of Precam- Turriff. Memoir Geological Survey Scotland, 86 & 96, 240.
brian Rocks in British Isles. Geological Society, London, Special Robertson, S. & Smith, M. 1999. The significance of the Geal Charn–
Report, 22, 33 –53. Ossian Steep Belt in basin development and inversion in the
Highton, A. J., Hyslop, E. K. & Noble, S. R. 1999. U– Pb zircon geo- Central Scottish Highlands. Journal of the Geological Society,
chronology of migmatisation in the northern-central Highlands: evi- London, 156, 1175– 1182.
dence for pre-Caledonian (Neoproterozoic) tectonometamorphism Soper, N. J. & Anderton, R. 1984. Did the Dalradian slides originate as
in the Grampian block, Scotland. Journal of the Geological Society, extensional faults? Nature, 307, 357–360.
London, 156, 1195–1204. Spencer, A. M. 1971. Late Pre-cambrian Glaciaion in Scotland. Geo-
Hoffman, P. F. & Schrag, D. P. 2002. The snowball Earth hypothesis: logical Society, London, Memories, 6, 100.
testing the limits of global change. Terra Nova, 14, 129–155. Stewart, A. D. 1997. Discussion on: Indications of glaciation at the base
Hoffmann, K. H., Condon, D. J., Bowring, S. A. & Crowley, J. L. 2004. of the Proterozoic Stoer Group (Torridonian), NW Scotland. Journal
A U–Pb zircon date from the Neoproterozoic Ghaub Formation, of the Geological Society, London, 154, 375– 376.
Namibia: constraints on Marinoan glaciation. Geology, 32, 817–820. Stewart, A. D. 2002. The Later Proterozoic Torridonian Rocks of Scot-
Kennedy, M. J., Runnegar, B., Prave, A. R., Hoffmann, K. H. & land: their Sedimentology, Geochemistry and Origin. Geological
Arthur, M. 1998. Two or four Neoproterozoic glaciations? Society, London, Memoirs, 24, 130.
Geology, 26, 1059–1063. Stoker, M. S., Howe, J. A. & Stoker, S. J. 1999. Late Vendian? Cam-
Kirkland, C. L., Strachan, R. A. & Prave, A. R. 2008. Detrital zircon brian glacially influenced deepwater sedimentation, MacDuff Slate
signature of the Moine Supergroup, Scotland: contrasts and compari- Formation (Dalradian), NE Scotland. Journal of the Geological
sons with other Neoproterozoic successions within the circum-North Society, London, 156, 55 –61.
Atlantic region. Precambrian Research, 163, 332–350. Sutton, J. & Watson, J. V. 1954. Ice-borne boulders in the MacDuff
Klein, G. De V. 1970. Tidal origin of a Precambrian quartzite: lower fine- Group of the Dalradian of Banffshire. Geological Magazine, 91,
grained quartzite (Middle Dalradian) of Islay, Scotland. Journal of 391– 398.
Sedimentary Petrology, 40, 973– 985. Tanner, P. W. G. 1995. New evidence that the Lower Cambrian Leny
McCay, G. A., Prave, A. R., Alsop, G. I. & Fallick, A. E. 2006. Glacial Limestone at Callender, Perthshire, belongs to the Dalradian Super-
trinity: Neoproterozoic Earth history within the British– Irish Caledo- group, and a reassessment of the ‘exotic’ status of the Highland
nides. Geology, 34, 901–912. Border Complex. Geological Magazine, 132, 473– 483.
Millar, I. L. 1999. Neoproterozoic extensional basic magmatism associ- Tanner, P. W. G. & Evans, J. A. 2003. Late Precambrian U– Pb titanite
ated with the West Highland granite gneiss in the Moine Supergroup age for peak regional metamorphism and deformation (Knoydartian
of NW Scotland. Journal of the Geological Society, London, 156, orogeny) in the western Moine, Scotland. Journal of the Geological
1153– 1162. Society, London, 160, 555–564.
Molyneux, S. G. 1997. An upper Dalradian microfossil reassessed. Trewin, N. H. 1987. MacDuff, Dalradian turbidite fan and glacial depos-
Journal of the Geological Society, London, 155, 741–743. its. In: Trewin, N. H., Kneller, B. C. & Gillen, C. (eds) Geology of
Noble, S. R., Hyslop, E. K. & Highton, A. J. 1996. High precision U –Pb the Aberdeen Area. Geological Society of Aberdeen. Scottish Aca-
monazite geochronology of the c.806 Ma Grampian Shear Zone and demic Press, Edinburgh, 79 –88.
implications for the evolution of the Central Highlands of Scotland. Turnbull, M. J. M., Whitehouse, M. J. & Moorbath, S. 1996. New
Journal of the Geological Society, London, 153, 511–514. isotopic age determinations for the Torridonian, NW Scotland.
Oliver, G. J. H., Chen, F., Buchwald, R. & Hegner, E. 1998. Fast tec- Journal of the Geological Society, London, 153, 955– 964.
tonometamorphism and exhumation in the type area of the Barrovian Vance, D., Strachan, R. A. & Jones, K. A. 1998. Extensional
and Buchan zones. Geology, 28, 459– 462. versus compressional settings for metamorphism: garnet chrono-
Piasecki, M. A. J. & van Breemen, O. 1983. Field and isotope evidence metry and pressure– temperature –time histories in the Moine Super-
for a c.750 Ma tectonothermal event in Moine rocks in the Central group, NW Scotland. Geology, 26, 927– 930.
Chapter 64

Neoproterozoic glacial deposits of Tasmania

CLIVE R. CALVER
Mineral Resources Tasmania, P.O. Box 56, Rosny Park, Tasmania, Australia 7018 (e-mail: ccalver@mrt.tas.gov.au)

Abstract: In Tasmania, Neoproterozoic glaciogenic deposits were laid down in one or more epicratonic basins, probably situated at the
eastern margin of the Australian– Antarctic craton. Rifting and volcanism took place in the late Cryogenian to early Ediacaran. On King
Island, north of Tasmania, the Cottons Breccia consists of 50–200 m of diamictite, conglomerate and sandstone. Limestone and dolo-
stone clasts are abundant in the diamictite, although carbonate is unknown in the underlying successions. The Cottons Breccia is overlain
by 10 m of laminated dolostone and limestone with a negative, upward-decreasing d13C profile. Rift volcanics and shallow intrusives
higher in the sequence are dated at c. 575 Ma. In NW Tasmania, two diamictite units are found in the Togari Group. The Julius
River Member, 200 m thick, contains dominantly dolostone clasts and overlies a shallow-marine dolostone unit with vase-shaped micro-
fossils and C-isotopes consistent with a mid-Cryogenian age. Some clasts in the Julius River Member contain a stromatolite (Baicalia cf.
B. burra) very similar to a form that is abundant in the middle part of the Burra Group, Adelaide rift basin. The Julius River Member is
immediately overlain by black shale and impure carbonate dated by Re–Os at 641 + 5 Ma. The younger diamictite in the Togari Group
is the Croles Hill Diamictite, 70 m thick, with predominantly volcanic clasts, underlain by a shale and mafic-volcaniclastic succession
and overlain by thin mudstone followed by thick rift tholeiites. At one locality this diamictite is underlain by a rhyodacite flow dated at
582 + 4 Ma. In southern Tasmania, diamictites are found in the Wedge River Beds and in the Cotcase Creek Formation (Fm.) (Weld
River Group). Laminated siltstone with dropstones is associated with the diamictites in the Cotcase Creek Fm. The southern
Tasmanian deposits are poorly constrained in age.
Chronometric and other evidence suggests correlation of the Julius River Member, Cottons Breccia and Croles Hill Diamictite with the
Sturt, Elatina and Gaskiers glacial phases, respectively. However, a glacial origin for the Julius River Member and Croles Hill Diamictite
remains uncertain.

Proterozoic rocks of King Island and western Tasmania comprise glacials of South Australia. Meffre et al. (2004) and Calver et al.
a continental fragment that lay at the eastern margin of the (2004) isotopically dated the overlying volcanics and shallow
Australian/east Antarctic craton in the Neoproterozoic (Burrett intrusives respectively, the latter authors specifically in the context
& Berry 2000; Li et al. 2008). This fragment probably rifted at of correlation and age constraints for the Cottons Breccia. Direen
c. 580 Ma and accreted back to the craton in a late phase & Jago (2008) contended that the Cottons Breccia is a non-
of the Cambrian, Delamerian/Ross Orogeny (Berry et al. 2008). glaciogenic mass flow deposit laid down in a developing rift.
Alternatively, King Island (and perhaps western Tasmania as
well) comprised a rifted margin that never completely separated
from the craton (Meffre et al. 2000, 2004). Cryogenian to Structural framework
Ediacaran successions unconformably overlie Mesoproterozoic
to early Neoproterozoic sediments and metasediments in many The Grassy Group unconformably overlies a thick, Mesopro-
places, but Cambrian and later tectonism, as well as significant terozoic to early Neoproterozoic siliciclastic shelf succession
lateral variability in the Cryogenian –Ediacaran successions (Black et al. 2004). The Grassy Group was deposited in an epi-
themselves, have made correlation across Tasmania problematic. cratonic rift basin that may have evolved into an east-facing
Consequently, three regions with known or probable Neoprotero- passive margin soon after deposition of the Cottons Breccia
zoic glaciogenic deposits are described separately below: King (Direen & Crawford 2003; Meffre et al. 2004; Direen & Jago
Island, NW Tasmania and southern Tasmania. 2008). The thick mafic volcanics (rift tholeiites and picrites) that
begin c. 100 m above the Cottons Breccia have been interpreted
as analogous to the mafic volcanic packages (‘seaward dipping
King Island: the Cottons Breccia reflector sequences’) at the base of Mesozoic –Cenozoic volcanic
continental margins (Direen & Crawford 2003; Meffre et al.
Introduction 2004). The distribution of these volcanics can be traced as an
offshore magnetic anomaly that extends southward to offshore
The Cottons Breccia (Jago 1974) occurs near the base of the Grassy western Tasmania, but it is not continuous with the similar-aged
Group on King Island (Fig. 64.1). It consists of up to 200 m of Spinks Creek Volcanics (next section), suggesting the Grassy
diamictite and minor conglomerate and sandstone, and crops out Group and Togari Group filled separate structural basins. King
over a strike length of about 8 km along or near the SE coast of Island was probably weakly deformed in the Cambrian, Tyennan
the island from Cottons Flat to north of Cumberland Creek Orogeny (Delamerian/Ross Orogeny correlate). Further defor-
(Fig. 64.1b). In a few places there is excellent exposure on the mation may have occurred in the Devonian, Tabberabberan
coast, although the basal part of the unit is relatively poorly Orogeny, followed by local granitoid intrusion in the Early
exposed inland. Waterhouse (1916) first described the unit as a Carboniferous. In the Cottons Flat –Cumberland Creek area, the
lithified ‘glacial till’; Carey (1947) suggested it was an Adelaidean Grassy Group dips moderately east, and a weak, subvertical clea-
tillite. Jago (1974) named and described the Cottons Breccia in vage is developed in shales. Minor cross-faulting results in dif-
some detail, but left its mode of origin unresolved. No type ferent parts of the succession being exposed along the coast.
section was nominated. Further description and discussion were Mineral assemblages in the volcanics indicate low greenschist-
provided by Jago (1981) and Waldron & Brown (1993), the facies metamorphism (Meffre et al. 2004). Near Grassy, how-
latter authors favouring a debris-flow origin. Calver & Walter ever, the sequence has been strongly contact metamorphosed and
(2000) gave evidence, including d13C stratigraphy of the overlying metasomatized by the Early Carboniferous Sandblow Granite
dolostone, for correlation of the Cottons Breccia with the Elatina (Fig. 64.1b).

From: Arnaud, E., Halverson, G. P. & Shields-Zhou, G. (eds) The Geological Record of Neoproterozoic Glaciations. Geological Society, London,
Memoirs, 36, 649– 657. 0435-4052/11/$15.00 # The Geological Society of London 2011. DOI: 10.1144/M36.64
650 C. R. CALVER

Fig. 64.1. (a) Distribution of Proterozoic rocks and major tectonic elements in Tasmania. (b) Simplified geological map of SE King Island with localities mentioned in
text. (c) Generalized stratigraphic section through Grassy Group in Bold Head–Cumberland Creek area.

Stratigraphy (c. 100 mm). The larger clasts are commonly oriented with long
axes parallel to stratification. At the Gut (Fig. 64.1b) a 15-m-
The Grassy Group (Fig. 64.1c) unconformably overlies the Fraser thick interval contains carbonate boulders up to 3 m long, but
Fm., a thick succession of siliceous siltstone and mudstone of this horizon cannot be traced any distance laterally. Clast litholo-
Mesoproterozoic or early Neoproterozoic age (Black et al. 2004; gies comprise abundant carbonate (including cream and grey
Direen & Jago 2008). The basal unit of the Grassy Group is the dolostone, dark limestone, oolitic limestone and dolostone con-
Robbins Creek Fm., consisting of laminated siltstone and shale glomerate), abundant fine-grained siliciclastics (mainly fine-
with minor lavas and a thin basal conglomerate (Calver 2008a, b). grained quartzarenite, siltstone, sandstone and mudstone), and
The overlying Cottons Breccia shows considerable lithological rare chert, red jasper, basic volcanics and dacite (Jago 1974;
variation, both laterally and vertically, and varies irregularly Waldron & Brown 1993; Calver & Walter 2000). Common striated
in thickness from c. 50 m in the north (Cumberland Creek) to clasts were noted by Solomon (1969) but have not been confirmed
c. 200 m in the south (City of Melbourne Bay) (Jago 1974; by subsequent workers (Jago 1974; Waldron & Brown 1993).
Calver 2008b). The Cottons Breccia is overlain by the Cumberland Diamictite in the lower part of the Cottons Breccia has clasts of
Creek Dolostone (c. 10 m thick), which passes up gradationally predominantly carbonate and a grey muddy carbonate matrix,
into green and black, then red, shale (the Yarra Creek Shale: while in the upper part, siliciclastic clasts become more abundant
Calver et al. 2004). The Yarra Creek Shale varies in thickness and the matrix is a red-brown mudstone (Calver & Walter 2000).
from c. 10 m to 150 m, apparently in response to movement Jago (1974) noted graded bedding (30 –300 mm) in southern out-
on penecontemporaneous faults (Direen & Jago 2008; Calver crops and a number of lenses of well-sorted sandstone and siltstone
2008a). A differentiated sill of intermediate composition (Grimes up to 300 mm thick and 10 m long, within the diamictite. Between
Intrusive Suite) intrudes the lower Yarra Creek Shale and older Cottons Flat and the Gut there is a middle unit, up to 20 m thick, of
units, and is locally amygdaloidal. The Yarra Creek Shale is con- green mafic-volcaniclastic sandstone, with shards visible in thin
formably overlain by the tholeiitic City of Melbourne Volcanics section. At Cottons Flat, diamictite in the lower part of the
consisting of peperites and volcanic breccias overlain by pillow formation includes minor intervals of well-laminated siltstone
lavas. The picritic Shower Droplet Volcanics follow, then the tho- crowded with dropstones up to 100 mm in size (Jago 1974, figs
leiitic Grahams Road Volcanics (Waldron & Brown 1993; Meffre 23, 24; P. Hoffman, pers. comm.). At the Gut, there is an upward
et al. 2004; Calver 2008b). The top of the Grassy Group is transition from diamictite through 2 m of closed-framework con-
unknown and lies offshore. Aeromagnetic mapping shows that glomerate, then 0.8 m of plane-laminated red sandstone, pebbly
the volcanics continue offshore and may be several kilometres at the base, which in turn is conformably and gradationally over-
thick in total (Direen & Crawford 2003). lain by the Cumberland Creek Dolostone.
The Cumberland Creek Dolostone is a pale grey to pale pinkish-
grey, fine-grained laminated dolostone (weathering to a pale
Glaciogenic deposits and associated strata yellow-brown), passing up into thinly interbedded dolostone and
shale, then shale with thin, pale grey, fine-grained limestone
The Cottons Breccia is dominantly massive or crudely stratified beds, c. 10 m thick in total. Thin sections of the dolostone show
diamictite with minor conglomerate, sandstone and laminated dolomicrite or turbid dolomicrosparite, with scattered rounded
pebbly siltstone (Jago 1974). Clasts in the diamictite are pre- micritic peloids 0.1–0.25 mm in diameter, and irregular micro-
dominantly sub-angular to sub-rounded, and the maximum size fenestrae (Calver & Walter 2000). Sharp-crested intrastratal anti-
in any outcrop is typically 300–500 mm, but commonly less clines, 100– 200 mm in amplitude, are found here and there in
NEOPROTEROZOIC GLACIAL DEPOSITS OF TASMANIA 651

the lower, dolomitic part of the unit. The overlying Yarra Creek NW Tasmania: Julius River Member and Croles
Shale is massive or plane-laminated, and pale yellow-brown to Hill Diamictite
red except for a middle interval with beds of black shale that
have the characteristic microfabric of benthic microbial mats Introduction
(Calver & Walter 2000).
In the Smithton Synclinorium of NW Tasmania, two diamictite
units (the Julius River Member and the Croles Hill Diamictite)
Boundary relations with overlying and underlying
are found in a Cryogenian to Early Cambrian succession known
non-glacial units as the Togari Group (Calver 1998; Everard et al. 2007). The
Julius River Member is found within the upper part of the
The base of the Cottons Breccia is not exposed, but its variable Black River Dolomite, and crops out at widespread localities
thickness suggests erosional incision into the underlying Robbins within the Smithton Synclinorium (Fig. 64.2). It was first briefly
Creek Fm. (Calver 2008a). The top of the Cottons Breccia is, at described by Longman & Matthews (1962). Griffin & Preiss
least locally, a conformable and gradational contact with the over- (1976) described outcrops now recognized as belonging to the
lying Cumberland Creek Dolostone (previous section). However, Julius River Member, and the stromatolites contained in some of
at a locality north of Conglomerate Creek, the contact with the the clasts. The Julius River Member was described under the
dolostone is abrupt and disconformable (Jago 1974). informal name ‘Trowutta Breccia’ by Jago (1981). The unit was
formally defined by Everard et al. (2007), with a type section,
c. 200 m thick, on the Arthur River (Fig. 64.2). Calver (1998)
Chemostratigraphy
reported C and Sr chemostratigraphic data for the Togari Group
and suggested correlation of the Julius River Member with the
Carbonate d13C and d18O data for six samples of the Cumberland
Sturtian glacial deposits of the Adelaide Geosyncline. Kendall
Creek Dolostone were presented by Calver & Walter (2000). The
et al. (2007) supported this correlation with a Re – Os age determi-
unit undergoes a steep upward decline in d13C, from –1.9‰ at
nation on black shale immediately overlying the Julius River
the base to –5.0‰ near the top. O-isotopic compositions are
Member.
moderately to strongly depleted ( –7.9 to –16.7‰ VPDB).
The Croles Hill Diamictite (Everard et al. 2007) is part of the
Kanunnah Subgroup, which overlies the Black River Dolomite
Other characteristics (Fig. 64.2). Up to 250 m thick, the formation is best developed
in the south of the Synclinorium, with significant localities also
At Grassy, the Grassy Group hosts economically important calcic near Forest in the NE and at Robbins Passage in the NW of the
scheelite skarns in the contact aureole of the Sandblow Granite Synclinorium (Fig. 64.2; Brown 1989, p. 25; Calver et al. 2004).
(Danielson 1975). Everard et al. (2007) formally defined the Croles Hill Diamictite,
with a type section on the Arthur River (Fig. 64.2).

Palaeolatitude and palaeogeography


Structural framework
A palaeomagnetic study by T. Raub (pers. comm.) shows a prob-
able primary magnetization in the upper Cottons Breccia and The Togari Group unconformably overlies a thick early Neoproter-
basal Cumberland Creek Dolostone, with both polarities preserved ozoic siliciclastic shelfal succession, the Rocky Cape Group. The
and directed approximately to present-day north and south. These part of the Togari Group below the Julius River Member records
units appear to have been deposited within 58 of the equator. epicratonic, shallow-marine, predominantly carbonate sedimen-
tation. In the Kanunnah Subgroup above the Julius River
Member, basalts (dominantly rift tholeiites) and associated volca-
Geochronological constraints niclastics are abundant. The NE-trending Roger River Fault was an
east-side-down growth fault at this time (Everard et al. 2007).
Calver & Walter (2000) argued for correlation between the Cottons NW –SE extension is also shown by the approximately coeval
Breccia and the Elatina (‘Marinoan’) glacials of the Adelaide intrusion of the Tayatea Dyke Swarm (588 – 600 Ma) into base-
Geosyncline on the basis of, inter alia, the strong lithologic and ment east of the Smithton Synclinorium (Brown 1989). Rifting
d13C-chemostratigraphic resemblance between the Cumberland was evidently aborted at this location as the Kanunnah Subgroup
Creek Dolostone and the Nuccaleena Fm. (Williams et al. 2011). is conformably followed by thick upper Ediacaran shallow-marine
Correlation of the Yarra Creek Shale with the lower Brachina carbonates of the Smithton Dolomite.
Fm. of the Adelaide Geosyncline is also supported by lithostrati- The succession was gently folded in the Cambrian Tyennan
graphy, including the presence of microbialite black shale beds Orogeny and again in the Devonian, Tabberabberan Orogeny.
both in the Yarra Creek Shale and low in the Brachina Fm. and The Smithton Synclinorium (a large, gently north-plunging syn-
equivalents (Logan et al. 1999; Calver 2000). There are two clinal structure) is probably largely a Devonian structure. Dips
radiometric minimum age constraints on the Cottons Breccia: (i) are mainly gentle to moderate except along the western margin,
a Nd –Sm isochron age of 579 + 16 Ma on the Shower Droplet where steep dips are associated with east-directed thrusting
Volcanics and Grahams Road Volcanics (Meffre et al. 2004), (Everard et al. 2007). The rocks are undeformed to weakly
(n ¼ 5, 2s error limits, MSWD ¼ 1.6; initial 1Nd of 4.2); (ii) a deformed except locally in the south, where pelitic lithologies
U –Pb (SHRIMP on zircon) age of 574.7 + 3.0 Ma on the are moderately cleaved. Low-grade regional metamorphism (of
Grimes Intrusive Suite (Calver et al. 2004) (n ¼ 26, MSWD ¼ prehnite-pumpellyite grade, Griffin & Preiss 1976) affects the
0.73; precision limit

You might also like