Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Ancient Supercontinents and the Paleogeography of Earth
Ancient Supercontinents and the Paleogeography of Earth
Ancient Supercontinents and the Paleogeography of Earth
Ebook1,947 pages24 hours

Ancient Supercontinents and the Paleogeography of Earth

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Ancient Supercontinents and the Paleogeography of Earth offers a systematic examination of Precambrian cratons and supercontinents. Through detailed maps of drift histories and paleogeography of each continent, this book examines topics related to Earth’s tectonic evolution prior to Pangea, including plate kinematics, orogenic development, and paleoenvironments. Additionally, this book discusses the methodologies used, principally paleomagnetism and tectonostratigraphy, and addresses geophysical topics of mantle dynamics and geodynamo evolution over billions of years. Structured clearly with consistent coverage for Precambrian cratons, this book combines state-of-the-art paleomagnetic and geochronologic data to reconstruct the paleogeography of the Earth in the context of major climatic events such as global glaciations. It is an ideal, up-to-date reference for geoscientists and geographers looking for answers to questions surrounding the tectonic evolution of Earth.
  • Provides robust paleogeographies of Precambrian cratons based on high-quality paleomagnetic and geochronologic data and critically tested by global geological datasets
  • Includes links to updated databases for the Precambrian such as PALEOMAGIA and the Global Paleomagnetic Database (GPMDB)
  • Presents full-color maps of the drift histories of each continent as well as their paleogeographies
  • Discusses key questions regarding continental drift, the supercontinent cycle, and the geomagnetic dipole hypothesis and analyzes palaeography in the context of Earth’s holistic evolution
LanguageEnglish
Release dateOct 6, 2021
ISBN9780128185346
Ancient Supercontinents and the Paleogeography of Earth

Related to Ancient Supercontinents and the Paleogeography of Earth

Related ebooks

Physics For You

View More

Related articles

Reviews for Ancient Supercontinents and the Paleogeography of Earth

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Ancient Supercontinents and the Paleogeography of Earth - Lauri J. J Pesonen

    Ancient Supercontinents and the Paleogeography of Earth

    Edited by

    Lauri J. Pesonen

    Department of Physics, University of Helsinki, Helsinki, Finland

    Johanna Salminen

    Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland

    Geological Survey of Finland, Geophysical Solutions, Espoo, Finland

    Sten-Åke Elming

    Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå, Sweden

    David A.D. Evans

    Department of Earth & Planetary Sciences, Yale University, New Haven, CT, United States

    Toni Veikkolainen

    Institute of Seismology, University of Helsinki, Helsinki, Finland

    Table of Contents

    Cover image

    Title page

    Copyright

    List of contributors

    About the editors

    Preface

    Acknowledgments

    Chapter 1. Precambrian supercontinents and supercycles—an overview

    Abstract

    1.1 The history of the supercontinent research—the five milestones

    1.2 The Earth and the solar system

    1.3 Some tectonic concepts

    1.4 Precambrian supercontinents and their cyclicity—observational evidence

    1.5 How to reconstruct Precambrian terranes?

    1.6 Models of the Precambrian supercontinents—some remarks

    1.7 Precambrian paleomagnetism and paleogeography: a guideline

    1.8 Precambrian paleomagnetism applied to paleoreconstructions—an example

    1.9 Precambrian paleomagnetic databases

    1.10 Global and terrane geological maps for reconstructions

    1.11 Precambrian supercontinent cycle

    1.12 Conclusions and suggestions for future work

    1.13 How we proceed in this book

    Acknowledgments

    Appendices

    References

    Chapter 2. A mantle dynamics perspective on the drift of cratons and supercontinent formation in Earth’s history

    Abstract

    2.1 Introduction

    2.2 Methodology

    2.3 Results

    2.4 Long-term cooling of the mantle (case D)

    2.5 Discussion

    2.6 Conclusion

    Acknowledgments

    References

    Chapter 3. Precambrian geomagnetic field—an overview

    Abstract

    3.1 Introduction

    3.2 Precambrian geomagnetic field—characteristic features

    3.3 Inclination frequency analysis

    3.4 Field reversals

    3.5 Paleosecular variation

    3.6 Paleointensity

    3.7 Continental drift

    3.8 Results

    3.9 Conclusion

    Acknowledgments

    References

    Chapter 4. The Precambrian paleogeography of Laurentia

    Abstract

    4.1 Introduction and broad tectonic history

    4.2 Paleomagnetic pole compilation

    4.3 Differential motion before Laurentia amalgamation

    4.4 Paleogeography of an assembled Laurentia

    4.5 Comparing paleogeographic models to the paleomagnetic compilation

    4.6 Paleoenvironmental constraints on paleolatitude

    4.7 Evaluating Laurentia’s Proterozoic paleogeographic neighbors

    4.8 The record implies plate tectonics throughout the Proterozoic

    4.9 Conclusion

    Acknowledgments

    Notes

    Glossary

    References

    Chapter 5. The Precambrian drift history and paleogeography of Baltica

    Abstract

    5.1 Introduction

    5.2 Geological evolution of Baltica

    5.3 Material and methods

    5.4 Paleomagnetic evidence for the drift of Baltica

    5.5 Paleoproterozoic–Neoproterozoic climatic indicators for Baltica

    5.6 Drift velocities of Baltica and its subcratons with implication to tectonics

    5.7 Implications for Baltica in Superia supercraton and Nuna and Rodinia supercontinents

    5.8 Concluding remarks

    Acknowledgments

    Supplementary table

    References

    Chapter 6. The Precambrian drift history and paleogeography of Amazonia

    Abstract

    6.1 Introduction

    6.2 The Amazonian Craton

    6.3 Quality criteria of paleomagnetic poles

    6.4 Amazonian paleomagnetic data and apparent polar wander path

    6.5 Final remarks

    Acknowledgments

    References

    Chapter 7. The Precambrian drift history and paleogeography of Río de la Plata craton

    Abstract

    7.1 Introduction

    7.2 Geology of the Río de la Plata craton

    7.3 Material

    7.4 Results

    7.5 Discussion

    7.6 Conclusions

    Acknowledgements

    References

    Chapter 8. Precambrian paleogeography of Siberia

    Abstract

    8.1 Introduction

    8.2 Geology of the Siberian Craton

    8.3 Paleomagnetic data and paleolatitudes of Siberian Craton

    8.4 Possible neighbors of Siberian Craton

    8.5 Conclusion

    Acknowledgments

    References

    Chapter 9. Whence Australia: Its Precambrian drift history and paleogeography

    Abstract

    9.1 Introduction to the Precambrian geology of Australia

    9.2 Material

    9.3 Results: original and age-binned apparent polar wander paths

    9.4 Discussion

    9.5 Summary

    References

    Chapter 10. The Precambrian drift history and paleogeography of India

    Abstract

    10.1 Introduction

    10.2 Data selection

    10.3 Orogenic belts of Peninsular India

    10.4 Geomagnetic field, paleoclimate and Greater India Assembly

    10.5 India in a global context

    10.6 Conclusion

    Acknowledgments

    References

    Chapter 11. The Precambrian drift history and paleogeography of the Chinese cratons

    Abstract

    11.1 Introduction

    11.2 Precambrian geology of the north China craton

    11.3 Precambrian paleomagnetic database and apparent polar wander path of the north China craton

    11.4 Precambrian drift history of the NCC

    11.5 Precambrian drift history of the south China craton

    11.6 Precambrian drift history of the Tarim craton

    11.7 Summary

    Acknowledgments

    References

    Chapter 12. The Precambrian drift history and paleogeography of the Kalahari Craton

    Abstract

    12.1 Introduction

    12.2 Crustal architecture and geology of the Kalahari Craton

    12.3 Paleomagnetic data

    12.4 Results

    12.5 Discussion

    12.6 Summary

    Acknowledgements

    References

    Chapter 13. Constraints on the Precambrian paleogeography of West African Craton

    Abstract

    13.1 Introduction

    13.2 Geology of West African Craton

    13.3 Review of paleomagnetic data

    13.4 LIP records in West African Craton

    13.5 Paleoclimate indicators

    13.6 Precambrian paleogeography of West African Craton

    13.7 Concluding remarks

    Acknowledgments

    References

    Chapter 14. The Precambrian drift history and paleogeography of Congo−São Francisco craton

    Abstract

    14.1 Introduction

    14.2 The Congo−São Francisco craton

    14.3 Paleomagnetic poles

    14.4 The Congo−São Francisco craton in supercontinents

    14.5 Conclusion

    Acknowledgments

    References

    Chapter 15. Neoarchean–Paleoproterozoic supercycles

    Abstract

    15.1 Introduction

    15.2 Previous models of Archean–Paleoproterozoic crustal assemblies

    15.3 Methods and material

    15.4 Testing of the proposed models with paleomagnetic data

    15.5 Concluding remarks

    Acknowledgments

    References

    Chapter 16. Paleo-Mesoproterozoic Nuna supercycle

    Abstract

    16.1 Introduction

    16.2 The previous models of Paleo- to Mesoproterozoic Nuna

    16.3 Methods

    16.4 Paleo- to Mesoproterozoic geological evolution

    16.5 Reconstructing the Nuna supercycle

    16.6 Alternative models for Nuna

    16.7 The life-cycle of Nuna—comparison of paleomagnetic poles

    16.8 Octupole field at 1.9–1.2 Ga affecting paleoreconstructions?

    16.9 Drift velocities with the implication of tectonic style

    16.10 Paleoclimatic constraints on Nuna core on the reconstructions

    16.11 Summary and remarks

    Acknowledgements

    References

    Chapter 17. Meso-Neoproterozoic Rodinia supercycle

    Abstract

    17.1 Introduction

    17.2 Laurentia

    17.3 Baltica

    17.4 Siberia

    17.5 Amazonia

    17.6 West Africa

    17.7 Kalahari

    17.8 São Francisco/Congo and Rio Plata

    17.9 Proto-Australia

    17.10 India

    17.11 Missing-link possibilities for Chinese cratons

    17.12 Smaller cratonic fragments

    17.13 Existing Rodinia models

    17.14 Paleomagnetic tests

    17.15 2020 hindsight: a synthetic Rodinia model

    17.16 Geodynamic implications

    17.17 Rodinia development into its fourth decade

    Acknowledgments

    References

    Chapter 18. Phanerozoic paleogeography and Pangea

    Abstract

    18.1 Introduction

    18.2 Main tectonic units

    18.3 Apparent polar wander paths

    18.4 True polar wander and global APWPs

    18.5 The plate reconstructions

    18.6 The Paleozoic

    18.7 Pangea assembly and geometry

    18.8 The Mesozoic

    18.9 Pangea dispersal

    18.10 The Cenozoic

    Acknowledgments

    References

    Chapter 19. An expanding list of reliable paleomagnetic poles for Precambrian tectonic reconstructions

    Abstract

    19.1 Introduction

    19.2 Methods

    19.3 Data and discussion

    Acknowledgments

    References

    Index

    Copyright

    Elsevier

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    Copyright © 2021 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    ISBN: 978-0-12-818533-9

    For Information on all Elsevier publications visit our website at https://www.elsevier.com/books-and-journals

    Publisher: Candice Janco

    Acquisitions Editor: Amy Shapiro

    Editorial Project Manager: Emerald Li

    Production Project Manager: Bharatwaj Varatharajan

    Cover Designers: Christian J. Bilbow and Lauri J. Pesonen*

    *Lauri J. Pesonen contributed to the design of the cover image.

    Typeset by MPS Limited, Chennai, India

    List of contributors

    Paul Y.J. Antonio,     Institute of Astronomy, Geophysics and Atmospheric Sciences, University of São Paulo, São Paulo, Brazil

    María Julia Arrouy

    Consejo Nacional de Investigaciones Científicas y Técnicas, Buenos Aires, Argentina

    Centro de Investigaciones Geológicas (CIG), Universidad Nacional de La Plata, Buenos Aires, Argentina

    Leda Sánchez Bettucci,     Instituto de Ciencias Geológicas, Facultad de Ciencias, Universidad de la República, Montevideo, Uruguay

    Franklin Bispo-Santos,     Institute of Astronomy, Geophysics and Atmospheric Sciences, University of São Paulo, São Paulo, Brazil

    Linxi Chang,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    L. Robin M. Cocks,     Department of Earth Sciences, The Natural History Museum, London, United Kingdom

    Michiel O. de Kock,     Department of Geology, University of Johannesburg, Johannesburg, South Africa

    Jikai Ding,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    Cedric Djeutchou,     Department of Geology, University of Johannesburg, Johannesburg, South Africa

    Mathew Domeier,     Centre for Earth Evolution and Dynamics (CEED), University of Oslo, Oslo, Norway

    Tatiana V. Donskaya,     Institute of the Earth’s Crust, Siberian Branch of the Russian Academy of Sciences, Irkutsk, Russia

    Manoel S. D’Agrella-Filho,     Institute of Astronomy, Geophysics and Atmospheric Sciences, University of São Paulo, São Paulo, Brazil

    Bruce M. Eglington,     Department of Physics, Department of Geological Sciences, University of Saskatchewan, Saskatoon, Canada

    Sten-Åke Elming,     Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå, Sweden

    David A.D. Evans,     Department of Earth and Planetary Sciences, Yale University, New Haven, CT, United States

    Pablo R. Franceschinis

    Universidad de Buenos Aires, IGEBA, Facultad de Ciencias Exactas y Naturales, Buenos Aires, Argentina

    Consejo Nacional de Investigaciones Científicas y Técnicas, Buenos Aires, Argentina

    Dmitry P. Gladkochub,     Institute of the Earth’s Crust, Siberian Branch of the Russian Academy of Sciences, Irkutsk, Russia

    Zheng Gong,     Department of Earth and Planetary Sciences, Yale University, New Haven, CT, United States

    Uwe Kirscher,     Department of Geosciences, University of Tübingen, Tübingen, Germany

    Elina Lehtonen,     Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland

    Haiyan Li,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    Zheng-Xiang Li,     School of Earth and Planetary Sciences, Curtin University, Perth, WA, Australia

    Toni Luoto,     Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland

    Casey R. Luskin,     Department of Geology, University of Johannesburg, Johannesburg, South Africa

    Phil J. McCausland,     Western Paleomagnetic & Petrophysical Laboratory, Department of Earth Sciences, University of Western Ontario, London, ON, Canada

    Joseph G. Meert,     Department of Geological Sciences, University of Florida, Gainesville, FL, United States

    Satu Mertanen,     Geological Survey of Finland, Geophysical Solutions, Espoo, Finland

    Scott R. Miller,     Department of Geological Sciences, University of Florida, Gainesville, FL, United States

    Adam Nordsvan,     Department of Earth Sciences, University of Hong Kong, Pokfulam, Hong Kong

    Manoj K. Pandit,     Department of Geology, University of Rajasthan, Jaipur, India

    Sally Pehrsson,     Geological Survey of Canada, Ottawa, Canada

    Lauri J. Pesonen,     Department of Physics, University of Helsinki, Helsinki, Finland

    Sergei A. Pisarevsky

    School of Earth and Planetary Sciences, Curtin University, Perth, WA, Australia

    Institute of the Earth’s Crust, Siberian Branch of the Russian Academy of Sciences, Irkutsk, Russia

    Earth Dynamics Research Group, The Institute for Geoscience Research (TIGeR), School of Earth and Planetary Sciences, Curtin University, Bentley, WA, Australia

    Anthony F. Pivarunas,     Department of Geological Sciences, University of Florida, Gainesville, FL, United States

    Daniel G. Poiré

    Consejo Nacional de Investigaciones Científicas y Técnicas, Buenos Aires, Argentina

    Centro de Investigaciones Geológicas (CIG), Universidad Nacional de La Plata, Buenos Aires, Argentina

    Augusto E. Rapalini,     Universidad de Buenos Aires, IGEBA, Facultad de Ciencias Exactas y Naturales, Buenos Aires, Argentina

    Tobias Rolf,     Centre for Earth Evolution and Dynamics (CEED), University of Oslo, Oslo, Norway

    Johanna Salminen

    Geological Survey of Finland, Geophysical Solutions, Espoo, Finland

    Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland

    Phillip Schmidt,     MagneticEarth, Newrybar, NSW, Australia

    Anup K. Sinha,     Dr. K.S. Krishnan Geomagnetic Research Laboratory, Allahabad, India

    Nicholas L. Swanson-Hysell,     Department of Earth and Planetary Science, University of California, Berkeley, CA, United States

    Wilson Teixeira,     Institute of Geosciences, University of São Paulo, São Paulo, Brazil

    Trond H. Torsvik

    School of Geosciences, University of Witwatersrand, Johannesburg, South Africa

    Centre for Earth Evolution and Dynamics (CEED), University of Oslo, Oslo, Norway

    Ricardo I.F. Trindade,     Institute of Astronomy, Geophysics and Atmospheric Sciences, University of São Paulo, São Paulo, Brazil

    Toni Veikkolainen

    Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland

    Institute of Seismology, University of Helsinki, Helsinki, Finland

    Hervé Wabo,     Department of Geology, University of Johannesburg, Johannesburg, South Africa

    Chong Wang,     Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland

    Huaichun Wu,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    Hanbiao Xian,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    Tianshui Yang,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    Shihong Zhang,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    Hanqing Zhao,     State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing, P.R. China

    About the editors

    Lauri J. Pesonen is an Emeritus Professor of solid earth geophysics at Physics Department of University of Helsinki. He is a graduate at the Helsinki University of Technology and obtained PhD in 1978 at the University of Toronto. He is a member of the Finnish Academy of Science and Letters and has been the president of the Geophysical Society of Finland, and the Chair of Divison III (Paleomagnetism) of IUGG/IAGA during 1993–97. The IUA nominated the asteroid 19690Y (11263) as Pesonen according to his activities in studies of supercontinents and terrestrial impact craters. He received the Knight Medal of the First Class Finnish Lion in 1995. Lauri is author or coauthor of over 150 peer-reviewed articles. He initiated the Nordic Paleomagnetism Workshops and was a key-person in the developments of the Precambrian (PALEOMAGIA) and Holocene (GEOMAGIA) paleomagnetic databases. Lauri organized the International Supercontinent Symposium in 2012 in Helsinki. He has had academic (teaching and research) positions in Canada, Estonia, Norway, Germany, India, and Colombia. His research topics spread from supercontinents to the Earth's ancient magnetic field, impact structures, meteorite petrophysics, archeomagnetism, environmental magnetism, and biomagnetism. He has built three paleomagnetism laboratories, the first one at the Geological Survey of Finland, the second at the University of Helsinki, and the third in Tarto University, Estonia. His latest interests include constructions of exhibitions of meteorites and impactites at several Finnish museums.

    Johanna Salminen is a Docent at the University of Helsinki and has been leading the research of its Solid Earth Geophysics Laboratory during 2014–21. Since August 2021, she has been the Director of the Geophysical Laboratory of the Geological Survey of Finland. She earned an MSc in Geophysics (2004) and a PhD in Solid Earth Geophysics (2009) from the University of Helsinki. Her research interests lie in continental reconstructions, Precambrian supercontinents, deep-time evolution of the Earth, Cenozoic magnetostratigraphy, environmental magnetism, and biomagnetism. Her professional recognitions include Academy of Finland Research Fellowship (2015–20) and Early Career Scientist Award by the International Union of Geodesy and Geophysics (IAGA) (2015). She is a coleader of the Deep Time Digital Earth paleomagnetism working group.

    Sten-Åke Elming is the Professor Emeritus in Geophysics, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology. He is a graduate of Uppsala University and obtained a PhD in Geophysics at Luleå University of Technology (1982). Sten-Åke is an elected member of the Royal Swedish Academy of Sciences (2006) and was a member of the Academy jury for the Crafoord prize in geosciences, 2014. He has been a part of the committee and steering group for geosciences of the Swedish Natural Science Research Council. He has been building research capacity and departments in geophysics/geosciences at universities in developing countries, including Nicaragua, Thailand, and Mozambique. Sten-Åke is the author and coauthor of more than 100 peer-reviewed articles on topics that include basic as well as applied research, with a focus on paleomagnetism and plate tectonics, rock magnetism, regional geophysics, exploration of water, and mineral resources. He established and supported the building of two paleomagnetic laboratories, one at Luleå University of Technology and another at Prince of Songkla University, Thailand.

    David A.D. Evans is a professor of Earth and Planetary Sciences at Yale University and the Director of its Paleomagnetism Laboratory. He earned a Bachelor’s Degree in Geology & Geophysics from Yale University (1992) and a PhD in Geology from the California Institute of Technology (1998). He is author of more than 100 peer-reviewed publications on paleomagnetism and evolution of Earth’s geodynamo, Precambrian supercontinent reconstructions, and deep-time paleoclimatology and evolution. His professional accolades include a Packard Fellowship in Science and Engineering, Blavatnik Award Finalist, and the George P. Woollard Award of the Geological Society of America for outstanding contributions to geology through the application of the principles and techniques of geophysics. He has led two UNESCO International Geoscience Program (IGCP) projects on Precambrian supercontinents. Since 2016, he has served as Head of Berkeley College, a residential college at Yale University.

    Toni Veikkolainen is a geophysicist at the Institute of Seismology, University of Helsinki. He completed his PhD degree in 2014. His thesis handled the geocentric axial dipole model of the Precambrian geomagnetic field. He has been the first author or coauthor in 16 peer-reviewed publications on various aspects of geophysics, from supercontinent reconstructions to theoretical aspects of paleomagnetic data, and seismic and thermal structure of the Fennoscandian lithosphere. He has been the administrator of the global paleomagnetic database PALEOMAGIA since 2014. He has served as the secretary of the Geophysical Society of Finland since 2013. He is an active member of the Finnish amateur astronomy community and has significantly contributed to the dissemination of scientific knowledge to the public. Currently, he works with the development of seismological analysis systems and databases and monitoring of seismic risk associated with commercial projects.

    Preface

    The Earth is a unique rocky planet in our solar system, consisting of a series of spheres (core, mantle, lithosphere, hydrosphere, cryosphere, biosphere, atmosphere, and magnetosphere) with numerous mutual interactions. The Earth has undergone tectonic modes from the magma ocean phase to lid tectonics during the Hadean–Paleoarchean times, followed by modern-type plate tectonics probably since the Mesoarchean times. The current view is that the ancient continents have, since their growth from cratonic nuclei to continental terranes, moved horizontally relative to each other and to the underlying mantle, and formed clans of supercratons and supercontinents. The assembly, configuration, and breakup phases of supercontinents have been research topics of wide interest particularly with respect to mantle dynamics, tectonic evolution of the continents and oceans, growth of the atmosphere and cryosphere, changes in surface environment, and questions related to the origin and extinction of life forms. A consensus has appeared among the geoscience community that after the existence of supercratons during the Neoarchean–Paleoproterozoic times, the Earth has witnessed three eras of supercontinents or huge landmasses: the Paleo-Mesoproterozoic Nuna, the Meso-Neoproterozoic Rodinia, and the Neoproterozoic-Phanerozoic Gondwana/Pangea assemblies, which together define the supercontinent cycles. Several techniques have been applied to trace the geometries and internal configurations of the ancient continental assemblies. One of them, the paleomagnetic approach, is the only globally quantitative way to reconstruct paleogeographies of the continents by providing knowledge of their drift histories in terms of paleolatitudes, orientations, and drift velocities (kinematics). However, a look at the paleomagnetic reconstructions reveals considerable variability in the proposed geometries, configurations, and timings of the assemblies and breakups. The main cause for these differences is the distinction in paleomagnetic data used for the reconstructions, that is, which poles are used and what are their reliabilities? In order to provide a new avenue for this approach, the paleomagnetic community established a workshop series that critically evaluated all published Precambrian poles, provided a quality ranking for each pole, and made an internationally accepted database of the highest quality poles on which the future reconstructions can be built. First established in 1986, series culminated in the Eighth Nordic Paleomagnetism Workshop held in Leirubakki, Iceland, 2017, where an upgraded pole compilation was finalized and internationally agreed on (see Chapter 19, An Expanding List of Reliable Paleomagnetic Poles for Precambrian Tectonic Reconstructions, and references therein). One of the motivations for this book is to make use of the new database in building the ancient supercontinents and to study the paleogeography of the Earth; other motivations include testing of the paleomagnetically based supercontinent reconstructions with upgraded geological, geochronological, and geochemical data sets and comparing the plate velocity data as derived from new paleomagnetic observations with those provided by dynamic and kinematic mantle modeling.

    The scopes of the book are:

    1. to define the drift histories of the major Precambrian continents, their cratonic nuclei and their building blocks focusing on the last 3 billion years;

    2. to test the credibilities of previously proposed Precambrian supercontinent models, to propose alternative ones, and to study their assembly, tenure, and breakup phases in order to define the possible supercontinent cyclicity;

    3. to test the new supercontinent assemblies with upgraded geological, tectonic, geophysical, geochemical, and geochronological data;

    4. to investigate if the proposed supercontinent models and their cyclicity match various temporal features (peaks, troughs, pulses, excursions) and cycles as derived from time series of secular evolution proxies, such as plate tectonic, kinematic, mantle depletion, biogeological, atmospheric, and environmental proxies;

    5. to compare the paleomagnetically constructed supercontinent models, their phases, and their cyclicity, with predictions provided by mantle dynamic modeling;

    6. to investigate if certain properties of the Earth’s magnetic field (such as the ancient field intensity) provided by geodynamo models compare with observed paleointensity data, and if so, to shed further light on the idea that the heat-transfer phenomena occurring at the core–mantle boundary can link to observations at the surface of the Earth;

    7. and finally, the book raises challenging questions, for example, why do we have supercontinents, what forces are driving the continental blocks to amalgamate or to breakup, how long is the lifetime of a supercontinent, and many others.

    Chapter 1, Precambrian supercontinents and supercycles—an overview, provides an updated overview of the Precambrian supercontinents and their cyclicity and serves as a bridge to the subsequent chapters. The chapter also offers a short overview of the geological history of the Earth and tectonic processes taking place before the onset of modern-type plate tectonics. Since paleomagnetism is the only quantitative tool to reconstruct global paleogeographies, a few examples are shown on how it works in practice in supercontinent research. This chapter also presents secular evolution trends of the Earth as seen by isotopic, geological, geochemical, and geophysical records, and their correlations with the four phases of the supercontinent cycle—amalgamation, collision, tenure, and breakup.

    Chapter 2, A mantle dynamics perspective on the drift of cratons and supercontinent formation in Earth’s history, provides an insight of mantle dynamic modeling of the drift histories of cratonic blocks, their amalgamation into supercontinents, and their subsequent breakups. Theoretical models of lithospheric plate velocities presented in this chapter can be compared with those observed in other chapters.

    The basic assumption behind paleomagnetism, as applied to paleogeography, is that the time-averaged geomagnetic field had a Geocentric Axial Dipole (GAD) geometry. Chapter 3, Precambrian geomagnetic field—an overview, provides an updated overview of the Precambrian geomagnetic field and its time variations with strong support for the prevalence of the GAD field during the last three billion years.

    The major part of the book (Chapters 4–14) presents new and updated drift histories of the Earth’s ancient terranes and their building blocks, based on paleomagnetism and supported by geological, geophysical, isotopic, and geochemical data. The updated Precambrian supercraton and supercontinent reconstructions are presented in Chapters 15–17, providing also an insight to the cyclicity of supercontinents. Chapter 18, Phanerozoic paleogeography and pangea, bridges the Precambrian supercontinent themes and results into the modern Earth by presenting a review of Phanerozoic paleogeography and also by outlining the topics where current research on paleogeography is ongoing. The database of key paleomagnetic poles used in Chapters 4–17 is presented in Chapter 19, An expanding list of reliable paleomagnetic poles for Precambrian tectonic reconstructions, with a brief accompanying discussion.

    Acknowledgments

    This book would not be possible without the valuable help of editorial officers of the Elsevier's EMSS-team: Marisa La Fleur, Emerald Li, Mohan Narmatha, Essaki Pandyan, Amy Shapiro, Nalini Thangavelu and Bharatwaj Varatharajan. Sincere thanks to Pathamawan Sangchan for her help and patience in preparing many of the figures of Chapter 1, Precambrian supercontinents and supercycles—an overview, and Mathew Domeier for his editorial help. The guest editors of this volume are much indebted to the excellent cooperation with the leading authors and the teams of the 19 chapters. We express our gratitude to the following persons for their help during various phases of writing this book: Elina Lehtonen, Raimo Lahtinen, Petri Peltonen, Jussi Heinonen, Shihong Zhang, and Sergei Pisarevsky. Sincere thanks to Suomen Tietokirjailijat r.y. for the financial support to Lauri J. Pesonen.

    Lauri J. Pesonen¹, Johanna Salminen², ³, Sten-Åke Elming⁴, David A.D. Evans⁵ and Toni Veikkolainen⁶, ¹Department of Physics, University of Helsinki, Helsinki, Finland, ²Department of Geosciences and Geography, University of Helsinki, Helsinki, Finland, ³Geological Survey of Finland, Geophysical solutions, Espoo, Finland, ⁴Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå, Sweden, ⁵Department of Earth & Planetary Sciences, Yale University, New Haven, CT, United States, ⁶Institute of Seismology, University of Helsinki, Helsinki, Finland

    Helsinki, September 1, 2021

    Chapter 1

    Precambrian supercontinents and supercycles—an overview

    Lauri J. Pesonen¹, David A.D. Evans², Toni Veikkolainen³, Johanna Salminen³ and Sten-Åke Elming⁴,    ¹Department of Physics, University of Helsinki, Helsinki, Finland,    ²Department of Earth and Planetary Sciences, Yale University, New Haven, CT, United States,    ³Institute of Seismology, University of Helsinki, Helsinki, Finland,    ⁴Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå, Sweden

    Abstract

    There is ample evidence that supercontinent cycles on Earth have been operating since the Late Paleoproterozoic. Evidence for the supercontinent cyclicity arises from multidisciplinary observations from geology, geochronology, geophysics (e.g., paleomagnetism, seismology, heat flow), isotope geology, and geochemistry. This overview summarizes current views of Precambrian supercontinent episodicity or cyclicity. In addition, paleogeographic reconstructions based on global key paleomagnetic poles and kinematic models of Paleo-Mesoproterozoic Nuna supercycle, Meso-Neoproterozoic Rodinia supercycle, and the Phanerozoic Gondwana/Pangea supercycle are explored. The lifecycle of supercontinents is tested by geological, geophysical, and geochemical data coupled with secular evolution trends of Earth. Results suggest that (1) supercontinent cyclicity has a characteristic (quasi-) period of ~700–500 million years, supported by planetary secular evolutionary trends, but other periods are also present; (2) supercontinents Nuna, Rodinia, and Gondwana/Pangea have different configurations and secular evolutionary trends possibly due to different tectonic styles of assembly; (3) globally averaged plate velocity during the Precambrian reveals a wave-like pattern with peaks and lows corresponding with features in several secular evolution indices including the distribution of U−Pb ages, passive margins, metamorphic events, tectonic proxies, and magmatic activity; (4) the data suggest three tectonomagmatic lulls during the Proterozoic, but the proposed Mesoproterozoic quiescent period, coined as boring billion years of Earth history (1.8–0.8 Ga) appears to be seen mainly by atmospheric and biospheric data rather than tectonomagmatic activity; and (5) tectonic processes driving supercontinent cyclicity are interactive, with feedbacks from all six spheres of the Earth—the geosphere, cryosphere, hydrosphere, biosphere, atmosphere, and magnetosphere.

    Keywords

    Precambrian; paleogeography; paleomagnetism; supercontinent cycle; Kenorland; Nuna; Rodinia; Gondwana; Pangea; secular trends

    1.1 The history of the supercontinent research—the five milestones

    The research of supercontinents has a remarkable history (e.g., Nance et al., 1986; 2014; Cawood et al., 2018). Here we summarize the story in five milestones. Alfred Wegener’s iconic book of continental drift (Wegener, 1915) can be regarded as the cornerstone of scientific research on global tectonic mobilism. Another landmark was the revolutionary discovery of plate tectonics during the late 1960s, since it not only verified the horizontal movements of continental and oceanic plates, but also offered dynamic and kinematic frames for mobilism in spherical coordinates (e.g., Morgan, 1968). The third scientific landmark came from the pioneering work by Wilson (1966) and imminently subsequent researchers (e.g., Dewey and Burke, 1973; Hoffman et al., 1974; Dewey and Spall, 1975), who extended the plate-tectonic paradigm into deep time, thus outlining the concept of supercontinent cyclicity (Worsley et al., 1984), that is, the Wilson cycle. We now accept that supercontinents are products of plate tectonics and manifestations of mantle dynamic processes taking place on Earth since c.2.0 Ga (Santosh et al., 2009; Mitchell et al., 2021), although older onset times, such as >3.1 Ga by Cawood et al. (2006), >3 Ga by Condie and Kröner (2008), or younger ages such as 1.0–0.85 Ga by Stern (2005) and Hamilton (2011) have been proposed. The recognition of punctuated supercontinental history has shed further light on geodynamic feedback mechanisms between the lithosphere and the deep Earth; the understanding of these interactions can be regarded as the fourth milestone in supercontinent research (e.g., Anderson, 1982; Hoffman, 1989a; Santosh, 2010; Condie, 2020).

    Research during the last 10 years has been able to link supercratons and supercontinents to broader Earth-system processes in an integrated fashion by coupling the supercontinent phases (assembly, tenure, breakup) with the secular evolutionary trends provided by geology, geophysics, geochronology, and geochemistry, that occur across the six spheres of our planet from the outer core to magnetosphere (Bradley, 2011). Integration of the Earth’s paleogeographic and secular evolutionary changes has led to a concept of truly global-scale supercontinent cycles with typical periodicity of c.700–500 million years (Fischer, 1984; Worsley et al., 1984; Nance et al., 1986; Reddy and Evans, 2009; Bradley, 2011; Vérard et al., 2015; Li et al., 2019; Mitchell et al., 2021). We regard this integrated approach as the fifth milestone in supercontinent research.

    Fig. 1.1 shows the Precambrian time scale and the progression from the protocontinental embryos to small cratonic nuclei, to supercratons followed by two full-fledged Precambrian supercontinents (Nuna, Rodinia) and ending with Neoproterozoic-Phanerozoic Gondwana assembly on the way toward Pangea (Bleeker, 2003; Sawada et al., 2018). Supercontinent research comprises several outstanding topics that form the scope of this book:

    1. Earth’s changing paleogeography with a focus on supercontinent cyclicity.

    2. Precambrian plate (ancient continents, cratons) motions and their velocities (kinematics), and understanding the driving forces and mantle to core dynamics, feedbacks and interactions.

    3. Secular changes of various Earth indices provided by geological, geochemical, and geophysical data, and their links and correlations with supercontinent cyclicity.

    Figure 1.1 Left: The Precambrian time scale (Ogg et al., 2016) with subdivisions into eons, eras, periods, and ages. Middle: The approximate life times of the continental embryos and protocratons (cratonic nuclei), supercratons, and supercontinents Nuna and Rodinia. The onset of Gondwana landmass during the late Neoproterozoic is also shown. Note that the lengths of eons, eras, and periods are not linear to allow all to be in the same image. Right: Number of key poles has been plotted accordingly and is based on the Leirubakki key pole list (Evans et al., 2021) (Chapter 19: An expanding list of reliable paleomagnetic poles for Precambrian tectonic reconstructions).

    After these introductory remarks, we provide a short geological history of the Earth (Section 1.2). The next Section 1.3, Section 1.4, Section 1.5, Section, 1.6, and Section, 1.7 outline supercontinent-related tectonic processes and concepts, and provide a discussion on how to reconstruct supercontinents—one of the prime topics of this book. Since most of the Precambrian continental reconstructions derive from paleomagnetic techniques, we outline the method with two examples (Section 1.8). Paleomagnetic databases play a key role in making reconstructions, and therefore we provide an overview of them with a special focus on data quality (Section 1.9). To help the reader, we provide geological maps of the Precambrian terranes and cratons of the world (Section 1.10). At the end, we summarize current views on evolution of the Precambrian supercontinents and their cycles, as monitored by geochronological, geological, geophysical, and geochemical secular evolution indices (Section 1.11). We sum up the major events archived in Precambrian rocks, signaling major changes occurring in the Earth’s spheres, and we discuss how these features link with the supercontinent cyclicity and phases (assembly, tenure, and breakup) and with deep-Earth processes, followed by conclusions and suggestions for future work (Section 1.12).

    1.2 The Earth and the solar system

    The current estimate of the age of the Universe is c.13.7 Ga. Our galaxy, the Milky Way, was formed at around 10 Ga, and the solar system c.4.57 Ga, as a result of nebular gas and dust agglomeration. From planetary studies, scientists assume that most, if not all, solar system bodies were formed within a relatively short time at about or slightly after 4.57 Ga (Fig. 1.2A). The Earth, third planet counted from the Sun, is one of the five active silicate rocky bodies in our solar system (the other ones being, in order of activity: Io, Venus, Mars, and Europa) and was formed c.4.57–4.56 Ga together with the other bodies in the solar system (Fig. 1.2A; Torsvik and Cocks, 2017; Stern, 2018; Cawood et al., 2018). During the last few years, several models have been presented for the tectonic evolution of the early Earth, for example, the ABEL (advent of bioelements) model of Maruyama et al. (2018) as modified in Fig. 1.2. Most of the proto-Earth’s tectonic products have been wiped out by meteorite bombardment, and because of the restless tectonic evolution of the Earth, the rocks originally formed close in time to the birth of the Earth may have long since been reworked by thermal processes, caused either by high internal temperatures, by impact melting or by mantle plume activity (Stern, 2018; Condie et al., 2016; O'Neill et al., 2020; King, 2020; Coltice et al., 2009). A few traces are available from detrital zircons, notably the <4.38 Ga Jack Hills zircons from Yilgarn craton of Australia (Wilde et al., 2001; Wang and Wilde, 2018) and the 3.9–3.6 Ga Isua zircons from Greenland (Nutman et al., 2009). The oldest exposed rock, the Acasta gneiss from Slave craton (later part of Laurentia) is about 4 Ga (Bowring and Williams, 1999). Most likely, the Earth faced a series of tectonic styles during its early history (4.55–3.80 Ga), such as (1) the magma ocean phase, (2) the (proto)-lid tectonic phase, followed by (3) the protoplate tectonic phase, and finally (4) the modern (i.e., subduction-type) style plate tectonic phase, which perhaps began as early as c.3.8 Ga ago and has been ongoing ever since (Jenner et al., 2009; Dhuime et al., 2015; Maruyama et al., 2018, Fig. 1.2B−E), although possibly with varying rates of globally averaged dynamic activity (Spencer et al., 2018). A debate continues on for how long in the Earth’s tectonic history the lid-tectonic styles operated: for example, only during 4.53–4.37 Ga or possibly as late as c.2.8 Ga (Bédard, 2018; Wyman, 2018) and on whether they also operated episodically, in between the time when modern plate tectonics was already in operation (Cawood et al., 2018; Hartnady and Kirkland, 2019; Tang et al., 2016). In addition to these tectonic processes occurring during 4.53–2.8 Ga, some other tectonic variants like heat pipe, slushy, sluggish, drip, and sagduction tectonics modes have been proposed but will not be discussed here (e.g., Stern, 2018; Capitanio et al., 2019; Palin et al., 2020). The onset of the geodynamo operating within the liquid outer core, generating the Earth’s magnetic field and its secular variation, was likely to have taken place at c.4.4–3.8 Ga, with specific age estimates also strongly debated (Tarduno et al., 2014; Weiss et al., 2015; Evans, 2018; Borlina et al., 2020; Veikkolainen and Pesonen, 2021, Chapter 3: Precambrian Geomagnetic Field—An Overview).

    Figure 1.2 Tectonic processes on Earth from the birth of the solar system to present. (A) The solar system 4.57 Ga; S, Sun; M, Mercury; V, Venus; E, Earth; Ma, Mars—only the first four (counted from Sun) rocky planets are shown. (B) Dry proto-Earth forms ~ 4.55 Ga ago with a magma ocean layer at the top, protomantle in the middle, and solid core in the center. In the upper mantle a two-layer convection operates whereas single cell convection occurs in the lower mantle. (C) ~ 4.53–4.37 Ga—the geospheres form: liquid core, upper, and lower mantle with tectonic boundaries or mantle transition zones at 410, 660, 1000, and 2900 km, respectively; CMB core−mantle boundary (slab stagnation boundary; Ballmer et al., 2015). In the uppermost layer stagnant lid tectonics operates. (D) The era of 4.37–4.2–3.8 Ga sees major asteroidal bombardment, which catastrophically reshapes the preexisting protolithosphere and probably also the upper mantle of the Earth. Geodynamo begins to operate. In both (C) and (D) protocrustal materials are both sinking (as blobs or slabs) and upwelling to and from the mantle. (E) From ~3.8 Ga onward Precambrian tectonism begins to resemble that of the modern plate tectonics: the basic elements of plate tectonics [subduction of the (proto)-oceanic lithosphere, accretion of island arcs onto the protolithospheric terranes, and opening of new ocean floor at MORs begins to operate]. The lithospheric terranes probably form first protocratons, followed by supercratons and later (since 2.0 Ga) supercontinents, while the rest of the globe will be occupied by superoceans. The subducted and aborted oceanic slices accumulate in the mantle graveyards, one in the transition layer between 410–660 km, the other at 1000–km layer, and the third one in the D''-layer at the top of the CMB forming the LLSVPs and PGZs. The inner core will nucleate at ~ 1.5 Ga (see fig. 3.18 in Veikkolainen and Pesonen, 2021, Chapter 3: Precambrian Geomagnetic Field—An Overview). A, anorthosites; ko, komatiites; e, eclogites; K1, KREEP basalts; Ls, lherzolite (solid); Ll, lherzolite (liquid); OL (CL), oceanic (continental) lithosphere; SCLM, Subcontinental Lithospheric Mantle. The green and red blocks denote ascending/descending upper mantle blobs signaling crust and mantle recycling. Source: Modified from the ABEL-model of Maruyama, S., Santosh, M., Azuma, S., 2018. Initiation of plate tectonics in the Hadean: eclogitization triggered by the ABEL Bombardment. Geoscience Frontiers 9, 1033–1048.

    The development of supercratons probably took place sometime during 3.8–2.7 Ga but their existence and configurations are poorly known and debated (Bleeker, 2003; Yakubchuk, 2019). The first supercontinent Nuna (Columbia) started to form at c.1.8 Ga (Chapter 16: Paleo-Mesoproterozoic Nuna supercycle; Elming et al., 2021) (Fig. 1.2). Evidence includes the presence of ultrahigh pressure mineral assemblages such as granulite and eclogite facies rocks, results of high pressure metamorphism (Bradley, 2011, Holder et al., 2019; Sobolev and Brown, 2019; Ganne et al., 2012; Liou et al., 2004), and sparse occurrences of putative ophiolites (e.g., Moores, 2002; Kusky et al., 2011) and collisional sutures (Gibb and Walcott, 1971; Moores, 1981; Cawood et al., 2006; Pourteau et al., 2018). The role of impact cratering in reshaping and remelting the upper portion of the Earth’s crust and mantle is generally only given brief mention in supercontinent literature (e.g., Leach et al., 2010; Sleep and Lowe, 2014; Bleeker, 2019) but not included as an important tectonic process. However, impact cratering has played a significant role during the early stages of the solar system (Fig. 1.2D; Maruyama et al., 2018; O'Neill et al., 2020; King, 2020). An example of the significance of impacts comes from our Moon. The most popular explanation for the age of the Moon is the Giant Impact Hypothesis, in which the Moon was formed as a product of the debris of the collision of a Mars sized protoplanet Theia with Earth at c.4.55–4.51 Ga ago (Canup and Asphaug, 2001; Barboni et al., 2017; Cawood et al., 2018; Maruyama et al., 2018), although recent analysis of lunar samples reveal some conflicts with this explanation (Cano et al., 2020). After the Moon-forming event, the Earth remained hot and retained enough heat to melt a large part of the mantle for several tens of million years [the magma ocean era; Maruyama et al. (2018)]. This short interval of time of the Earth’s history possibly permitted peridotitic and komatiitic rocks to form in the upper part of the crust and mantle (Fig. 1.2C) but this is still a controversial issue. In addition to the magma ocean period, the mantle was hot and the plume activity was high, with magma rising to the Earth’s surface making the lithosphere thin. This hot geodynamic regime, facilitated by high heat production due to radioactive decay, accompanied by core differentiation and the Moon-forming event, would have limited subduction of solid slabs into the mantle and thus prevent modern-type plate tectonics to take place until after c.3.8 Ga (O’Neill et al., 2007; Stern, 2018; Mundl et al., 2017; El Dien et al., 2019; Coltice et al., 2009).

    1.3 Some tectonic concepts

    Fig. 1.2E outlines some key concepts used throughout this chapter. The example shows a (present day) cross-section of an ancient terrane consisting of two collided cratons, causing increase in lithospheric thickness reaching depths to c.180–280 km and formation of lithospheric keels (Pearson et al., 2002; Artemieva and Mooney, 2002; Yakubchuk, 2019: Pourteau et al., 2018). The collision zone is a suture zone, which may contain ophiolites, occurring as obducted slices of the ocean that previously separated the two cratons (Moores, 2002; Peltonen and Kontinen, 2004). The cartoon also envisages examples of the mantle overturn process containing subducted and downwelled crustal slabs (the top-down tectonic style; Mitchell et al., 2021; Pastor-Galán et al., 2019). They form thermochemical piles in slab graveyards at depths of 410–660 km, ~1000 km and near the core−mantle boundary (CMB) where they accumulate, perhaps as the original source of the large low shear-wave velocity (LLSVPs) anomalies (Zhao, 2009; Torsvik and Cocks, 2017; Maruyama et al., 2018). The LLSVPs are also indications of upwellings of magmas, rising as mantle plumes (the bottom-up style of mantle tectonism) and expressed as large igneous provinces (LIPs) and hotspots when reaching the lithosphere (Mitchell et al., 2021). The kimberlite pipes (intrusions) are also derived from the boundaries of LLSVP anomalies although this relationship is only rigorously quantified for pipes with ages of less than c.0.3 Ga (Torsvik et al., 2010; Evans, 2010).

    Supercontinents have been defined qualitatively to include most crustal blocks, and quantitatively to include 75% of the existing crustal area, as was the case during the Phanerozoic Pangea (Meert, 2012). More sophisticated definitions of supercontinents are given by Merdith et al. (2019) and Mitchell et al. (2021) and include (1) the perimeter/area ratio of continental fragments, (2) mantle-legacy aspects and (3) the requirement of the minimum life-time of the assembly (>100 million years). Two supercontinents are thought to have existed during the Precambrian: the Paleo-Mesoproterozoic Nuna (c.1.6–1.4 Ga), and the Neoproterozoic Rodinia (c.0.9–0.7 Ga). We exclude hypothetical Pannotia (c.0.6–0.5 Ga; Murphy et al., 2020), because its existence as a full-fledged supercontinent is not well-supported by either direct evidence or indirect proxies (reviewed by Evans, 2020). We elaborate an idea of a landmass, called semisupercontinent (Evans et al., 2016) or megacontinent (Wang et al., 2021), which amalgamates prior to formation of a next supercontinent being an integral part of it. Gondwana (aggregation of several ancient continents), and also perhaps DhaBaSi (aggregation of Southern Indian cratonic blocks) and Umkondia, serve as examples of such landmasses (Choudhary et al., 2019; Wang et al., 2021).

    In Figs. 1.2E and 1.3A, various types of orogens, tectonic belts, and linear tectonic zones of the Earth’s crust are introduced. Based on end-member types, Cawood et al. (2009) grouped orogenies into three categories: accretionary, collisional, and intracratonic. An accretionary orogen forms at sites of subduction of oceanic crust beneath a continental plate and consists of magmatic arc systems (e.g., the Svecofennian orogeny in Fennoscandia; Salminen et al., 2021a,b; Chapter 15: Neoarchean-Paleoproterozoic Supercycles). A collisional orogen forms when a continental plate is subducting beneath other continental plates (e.g., Himalaya represents a modern-type collisional orogen, and Trans-Hudson is one of many Precambrian examples; Cawood et al., 2018; Zhao et al., 2004). The remnants of the subduction zones together with the terranes possibly representing fragments of different tectonic plates are called sutures (Fig. 1.3B), which may also contain slices of obducted ocean floor known as ophiolites (Cawood et al., 2013; Furnes et al., 2014). Several examples of Precambrian sutures (Moores, 1981; Pourteau et al., 2018) and associated ophiolite occurrences, for example, the Jormua ophiolite (1.98 Ga) in Fennoscandia (Peltonen and Kontinen, 2004), provide documentation for the operation of plate tectonics during the Paleoproterozoic. The ancient collisional orogens seem also to thicken the lithospheres under the collision zones and produce lithosphere keels, which may extend to depths of ~150–280 km (Fig. 1.2E) and provide additional hints of the ancient plate collisions (Yakubchuk, 2019; Pehrsson et al., 2013). Intracratonic orogens (or belts) lie within a continent, away from an active plate margin (Cawood et al., 2009), forming within-continent tectonic belts such as relative shearing of the tectonic blocks leading to upthrustings and block tiltings (e.g., the Kapuskasing structural zone in the Superior province; Evans and Halls, 2010), block rotations (e.g., the relative rotations of the N. and S. Dharwar blocks of Peninsular India; Söderlund et al., 2019), and crustal shortening (Halls, 2015). Tectonic terms used in this chapter are summarized in Figs. 1.2 and 1.3.

    Figure 1.3 Some concepts and tectonic styles during the Precambrian. (A) Archean supercratons consisting of subcratons (or protocratons). Also shown: intra- and intercratonic belts. (B) Supercontinent and superocean, strange attractors, spiritual interlopers, and lonely wanderers (Meert, 2014). Also shown: megacontinent and megacraton. (C) Supercontinent cyclicity showing schematic cartoons of supercratons (at 2.4 Ga), supercontinents Nuna (at 1.5 Ga), and Rodinia (at 0.9 Ga) and also the Gondwana landmass (at 0.5 Ga). (D) Tectonic styles (introversion, extroversion, and their combination) of supercontinent cycle (see Section 1.3). Figures modified from: (A) Bleeker (2003); (B) Meert (2014), (C) this chapter, (D) Nance and Murphy (2013) and Silver and Behn (2008).

    Analyses of the proposed reconstructions of successive supercontinents reveal some curiosities. Meert (2014) noted the frequent occurrence of hypothesized Precambrian cratonic connections that approximate those of Pangea. Calling such examples strange attractors (among other imaginative monikers for styles of cratonic association; see Fig. 1.3B), he suggested the possibility that implicit bias might favor familiar reconstruction models over more exotic possibilities. Baltica and Laurentia serve as a Precambrian example of such strange attractors whose evidence for repeated conjunction seems strong enough to preclude bias. It is well-established that the unity of Laurentia and Baltica (and perhaps Siberia) form the core of supercontinent Nuna (e.g., Elming et al., 2021, Chapter 16, Paleo-Mesoproterozoic Nuna Supercycle), and paleogeographic reconstructions suggest that their relative positions have remained almost (but not exactly) the same through the following cycles from Nuna to Rodinia and from Rodinia to Gondwana/Pangea (see also Salminen et al. (2021a,b); Elming et al., 2021, Chapter 15: Neoarchean-Paleoproterozoic Supercycles and Chapter 16: Paleo-Mesoproterozoic Nuna Supercycle). Fig. 1.3B shows the grouping of the majority (>75% of the crustal area) of the hypothetical continental blocks forming (by definition) a supercontinent. The remaining part of the Earth’s surface beyond this supercontinent, including any isolated lonely wanderer cratons is called the superocean realm (Li et al., 2019). Its fragments may be preserved as scattered and tectonically dismembered blocks within accretionary collages (Sengör et al., 2014); or they may be partially or completely subducted, in which case we may see traces of them as frozen reflectors in the upper mantle in seismic images (Calvert et al., 1995). The Phanerozoic slabs may also be preserved as accumulated slab graveyards in the mantle transition zones (Van der Meer et al., 2018; Fig. 1.2E). Regarding the Precambrian era, most likely the ancient subducted slabs have, if not assimilated to the mantle due to thermal processes, reached and piled-up at the CMB forming sources of LLSVPs (Van der Meer et al., 2018). Occasionally, traces of the ancient oceanic lithosphere are seen as ophiolites, as noted previously (Furnes et al., 2009; Peltonen and Kontinen, 2004).

    Within each supercontinent cycle we can recognize three phases: assembly, tenure, and breakup (Bradley, 2011; Mitchell et al., 2021; Cawood et al., 2016; Gardiner et al., 2016). The continents collide during the assembly phase. The collisional events produce tectonic belts with distinct geological and geochemical signatures (sutures, ophiolites, metamorphic belts; Moores, 1981, 2002; Holder et al., 2019; Brown and Johnson, 2019), which are used in reconstructing the ancient assemblies. The assembly phase is followed by the stable tenure phase. The phase when the assembly breaks apart is called breakup phase, and continents should share ages of rift-related magmatism prior to passive margin development. However, there is a broad consensus of supercontinent cyclicity, that is, individual continents assembled and rifted from each supercontinent, a cyclicity that is evident in geological (e.g., passive margins), geochronological (e.g., zircon age distributions), geochemical (e.g., εHf data), and plate kinematic (e.g., plate velocity) and isotopic (e.g., ⁴He/³He) proxies (Bradley, 2011; this work, Section 1.11). A supercontinent cyclicity of ~ 700–500 million years (Evans et al., 2016) can be calculated from the breakup of one supercontinent to the next (Fig. 1.3C; Condie, 2002; Bradley, 2011; Cawood et al., 2013, Li et al., 2019; Yakubchuk, 2019; Mitchell et al., 2021). We note that this period is a rough estimate since the supercontinents do not fragment at a specific time but likely in a series of events, and the same applies to their accretions (Yoshida and Santosh, 2011; Nance et al., 2014; Cawood et al., 2016; Mitchell et al., 2021). Nevertheless, the cited cyclicity of ~ 700–500 million years is roughly the same for the entire Precambrian-Phanerozoic timespan. Two further notes are added here. First, the same cyclicity period can be observed in several Earth secular evolution proxies (e.g., in U−Pb zircon ages, passive margins), but other periods are also present in various records (Li et al., 2019; see Section 1.11.2 Secular trends). Second, time series analysis of εHf and Sr-isotopes yield evidence of twice-as-long periodic signals (~1500–1000 million years), which can be either modulations of the better-defined shorter period, or products of alternating styles of supercontinental transitions (Li et al., 2019).

    Three general tectonic models for supercontinental transition, introversion, extroversion, and their combination, are introduced in Fig. 1.3D, following Murphy and Nance (2005); see also Pastor-Galan et al. (2019). These models consider alternative possible fates of internal and external oceanic tracts relative to a fragmenting supercontinent. The present Atlantic Ocean can be considered an internal ocean, which opened during the breakup of Pangea, and the modern Pacific Ocean and its precursor Panthalassa Ocean can be considered an external ocean for the Pangea supercycle. In the schematic representation of Fig. 1.3D, three cratons (labeled A, B, C) disperse from the old supercontinent by opening the internal ocean, and the rifted borders of the fragments remain as passive margins pointing inward to that new ocean.

    In the introversion model, a subsequent supercontinent assembles by closure of the internal ocean. By unknown process, but perhaps due to subduction infection similar to advancing Caribbean subduction into the Atlantic realm (Waldron et al., 2014), the cratons’ outward motion terminates and they reverse their course. Finally, the fragments re-amalgamate into a new assemblage, and the younger ocean disappears by subduction. The configuration of the new assembly can be the same as the previous one (the idealized Wilson cycle) or slightly different as in our illustration (Fig. 1.3D). In the introversion model, the ancient rifted zones and passive margins remain in-board and face each other throughout the cycle.

    In the idealized extroversion model, the rifted cratons continue to drift apart (onto the opposite side of Earth) until the nearly complete closure of the external ocean (Fig. 1.3D, middle). In this case, the new ocean widens into a superocean, while the old (external) ocean diminishes and subducts along an ever-narrowing subduction girdle. The previously inboard facing passive margins eventually face outboard, toward the new ocean.

    The third tectonic model is a combination (or succession) of the intro- and extroversion models (Fig. 1.3D, bottom). Here, the cratons have largely gone through the extroversion style but afterward some fragments (such as B and C in Fig. 1.3D, bottom) have turned to an introversion mode. The passive margins are now mixtures from intro- and extroversion models. The orthoversion model (Evans, 2003a; Mitchell et al., 2012) shares some aspects of the combination concept, in that some domains of both the interior and exterior oceans are predicted to close during supercontinental assembly. In addition, however, the orthoversion model predicts assembly of a new supercontinent ideally 90 degrees of arc distance from its predecessor when reconstructed in an absolute mantle reference frame (Mitchell et al., 2012).

    1.4 Precambrian supercontinents and their cyclicity—observational evidence

    Geological proxies of supercontinent cyclicity include number of orogens, rift basins, and passive margins (e.g., Hoffman, 1989b; Condie, 2002; Bradley, 2008, 2011; Cawood et al., 2018), LIPs (Doucet et al., 2020; Prokoph et al., 2004; Ernst et al., 2013), mineral deposits (Pehrsson et al., 2016), distributions of orogenic-diagnostic rock types (like eclogites), metamorphic events (Brown, 2007; Sobolev and Brown, 2019), and isotope-geochemical data, notably the εHf, δ¹⁸O, ⁴He/³He indices (e.g., Mitchell et al., 2021; Bradley, 2011; Silver and Behn, 2008). Present-day geophysical datasets can illustrate the legacy of orogens or ancient rifts, including gravity and aeromagnetic anomaly patterns (e.g., Finn and Pisarevsky, 2008; Gradmann and Ebbing, 2015; Golynsky and Jacobs, 2001; Ebbing et al., 2021). Seismic reflection studies yield evidence of frozen subducted slabs at the bottom of the crust and partly penetrating in the upper mantle within Precambrian crustal terrains (BABEL Working Group, 1990; Calvert et al., 1995; Korja and Heikkinen, 2005; Fig. 1.2). These frozen reflectors were generally undated, diminishing their value of true evidence of ancient subduction, but a recent report by Wan et al. (2020) verifies that they are indeed of Proterozoic age.

    Time-constrained oxygen (δ¹⁸O) and hafnium (εHf) isotopes in dated zircon crystals provide a geochemical proxy to explore time series throughout the supercontinent cycles (Valley et al., 2005; Gardiner et al., 2016; Spencer et al., 2019). The δ¹⁸O of zircon is a sensitive tracer of the orogenic cycle, since higher values (isotopically heavier) result from greater degrees of sediment recycling or evolved continental input (Valley et al., 2005). The opposite is true for Hf: isotopically heavier values indicate more juvenile sources (Hawkesworth et al., 2010). Therefore the degree of continental contribution in magmatic systems can be assessed with δ¹⁸O and Hf-isotopic measurements of zircons. Increases in δ¹⁸O values temporally coincide with the assembly phases of three supercontinents (Nuna, Rodinia, and Pangea) and indicate increased crustal reworking as expected during supercontinent amalgamation. During the breakup phase of each of the three supercontinent cycles, δ¹⁸O values decrease, trending toward more mantle-like values, which is consistent with models invoking more mantle-derived magmatism associated with either mantle plumes and/or slab rollback during supercontinent breakup (Van Kranendonk and Kirkland, 2016; Mitchell et al., 2021).

    Other trace elements (e.g., Nb, Rb, Ni) and their ratios (e.g., Nb/Th, Rb/Sr) also provide data to delineate crustal growth and provenance histories as will be discussed in Section 1.11. In addition, several isotopes (e.g., ²³⁸U) and isotope ratios (e.g., ⁸⁷Sr/⁸⁶Sr) as well as derived indices (e.g., δ¹³C) can be used to isolate Earth’s paleoclimatic events such as glaciations (δ¹³C, ²³⁸U), oxygenation events (e.g., δ⁵³Cr, δ⁸²Se, δ⁹⁸Mo), erosion events (⁸⁷Sr/⁸⁶Sr), biogeological events given by nutrient productivity indices (e.g., P, Ni, Se) (Large et al., 2018), and mantle processes (⁴He/³He, Nb/Th; Silver and Behn, 2008). See reviews of the various indices (proxies) by Kendall (2021). These are further discussed in Section 1.11.

    1.5 How to reconstruct Precambrian terranes?

    Ancient supercontinents can be reconstructed in several ways using geological (e.g., LIP barcodes), geophysical (e.g., paleomagnetism), tectonic (e.g., matching orogenic belts), geochemical (matching chemostratigraphic plots), and geochronological (matching coeval age provinces) data (e.g., Hoffman, 1991; Dalziel, 1997; Zhao et al., 2002; Bleeker, 2003; Rogers and Santosh, 2004; Johansson, 2009). Paleomagnetism is a geophysical discipline primarily using well-dated key paleomagnetic poles, and is the only quantitative technique for global paleocontinental reconstructions in Precambrian time (Buchan et al., 2001; Evans and Pisarevsky, 2008, Pesonen et al., 2012; Buchan, 2013; Evans et al., 2021, Chapter 19: An Expanding List of Reliable Paleomagnetic Poles for Precambrian Tectonic Reconstructions). The key pole technique is of vital importance since it provides a way to anchor the apparent polar wander paths (APWPs) with precisely dated poles (Chapters 4–15). The APWPs allow the paleogeography to be constrained by applying Euler techniques in placing the continental terranes onto their ancient latitudes and azimuthal orientations on the globe (see Section 1.8). Due to the axial symmetry of the geocentric axial dipole- (GAD-) field, paleolongitudes remain undeterminable without additional assumptions such as fixed mantle structures (see Mitchell et al., 2012; Tegner et al., 2019; Torsvik et al., 2021, Chapter 18: Phanerozoic Paleogeography and Pangea). Paleomagnetically derived paleogeographic assemblies of continental terranes (or possible supercontinents) are then tested with geological, tectonic, or geophysical data. Paleolatitude changes derived from APWPs also give us a direct way to calculate drift velocities for cratons (minimum estimates, due to the freedom of absolute paleolongitude), which provide kinematic aspects to the reconstructions. Paleomagnetically inferred plate velocities can be compared with the other speedometers [e.g., number of orogens, distributions of passive margins, certain isotope indices (e.g., εHf), deep mantle structure indicators (kimberlites, LIPs, and certain metal deposits like orogenic gold), Bradley, 2008; Torsvik et al., 2014]. These data also provide a tool to compare drift velocities of different cratons, which provides a way to test the reconstructions (see Elming et al., 2021, Chapter 16: Paleo-Mesoproterozoic Nuna Supercycle), and explore velocities during different phases of the supercontinent cycle (O’Neill et al., 2007; Antonio et al., 2017; Pesonen et al., 2012; Condie et al., 2015a,b; Section 1.12). One overarching caveat to the use of paleomagnetic APWPs to calculate velocities is that any given APWP segment might contain a nonnegligible component of true polar wander (TPW; Evans, 2003a); the actual plate velocity (craton moving relative to the deep mantle) equals the vectorial difference between the total APW-derived velocity and the TPW component (Torsvik and Cocks, 2017).

    Paleomagnetically derived paleolatitudes, calculated from inclinations (but reduced to reference locations) or from the APWPs, for each of the cratons in the reconstructions, can be tested against the data provided by independent paleoclimatic indicators, such as laterites, evaporites, carbonates, red sandstones, regoliths (low to moderate latitudes), and glaciogenic rocks (centered on high latitudes) (see Fig. 1.16F; Pesonen et al., 2003; Evans, 2003b; 2006; Williams and Schmidt, 2018).

    The reconstructions of ancient terranes can be made using the following techniques:

    1. Matching pairs of coeval poles, or portions of the APWPs (Evans and Pisarevsky, 2008; Fairchild et al., 2017). These methods are the only directly quantitative ways to reconstruct ancient assemblies of Precambrian continents. Additional constraints can be achieved by comparing not only the shapes of the APWPs but also their lengths, the drift velocities derived from the APWPs (see above), or the polarity patterns (superchrons, hyperactivity zones) of the poles along the APWPs (e.g., Irving and McGlynn, 1976; Pesonen and Neuvonen, 1981; Swanson-Hysell et al., 2014). Updated examples and discussions of the use of these methods in making continental reconstructions can be found in Chapters 4–14 and 18.

    2. Matching coeval orogenic belts (collisional or accretionary) (Hoffman, 1996; Zhao et al., 2002; Johansson, 2009, 2014) and rifts (passive margins formed in the ancient trailing edges) (e.g., Dalziel, 1997; Bradley, 2008). Meert (2014) provides the merits and caveats of this method with examples.

    3. Matching continuous linear or radial mafic dyke swarms (e.g., Park et al., 1995; Bleeker and Ernst, 2006; Goldberg, 2010), prominent shear zones (e.g., Onstott and Hargraves, 1981; D’Agrella-Filho et al., 1998), or breakup fractures (Brookfield, 1993) from one craton to the other.

    4. Seeking continuations of nonorogenic geological intrusions, for example, rapakivi-anorthosites (Piper, 1980a; Vigneresse, 2005; Pesonen et al., 2012) or kimberlite corridors (Kumar et al., 2007; Torsvik et al., 2010; Pesonen et al., 2005; Tappe et al., 2018).

    5. Aligning sediment provenances (e.g.,

    Enjoying the preview?
    Page 1 of 1