Developments in Earth & Environmental Sciences, 8
ANTARCTIC CLIMATE EVOLUTION Edited by
FABIO FLORINDO Istituto Nazionale di Geofisica e Vulcanologia, 00143 Roma, Italy
MARTIN SIEGERT School of GeoSciences, Grant Institute, University of Edinburgh, Edinburgh EH9 3JW, UK
Amsterdam – Boston – Heidelberg – London – New York – Oxford Paris – San Diego – San Francisco – Singapore – Sydney – Tokyo
Preface Antarctic Climate Evolution is the first book dedicated to understanding the history of the world’s largest ice sheet and, in particular, how it responded to and influenced climate change during the Cenozoic. To explain the story of Antarctic ice and climate history, information on terrestrial and marine geology, sedimentology, glacier geophysics, ship-borne geophysics, and numerical ice sheet and climate modelling is presented within thirteen chapters. The book’s content largely mirrors the structure of the Antarctic Climate Evolution (ACE) program (www.ace.scar.org), an international initiative of the Scientific Committee on Antarctic Research (SCAR), affiliated with the International Polar Year 2007–2009, to investigate past changes in Antarctica by linking climate and ice sheet modelling studies with terrestrial and marine geological and geophysical evidence of past changes. The programme is designed to determine climate conditions and change in both the recent past (i.e. during the last glacial maximum, when temperatures were cooler than at present) and the more distant past (i.e. in the pre-Quaternary, when global temperature was several degrees higher than it is today). This new cross-disciplinary approach has led to a substantial improvement in the knowledge base on past Antarctic climate and to our understanding of the factors that have guided its evolution. This in turn has allowed us to build hypotheses, examinable through numerical modelling, for how the Antarctic climate is likely to respond to present and future global changes. Most of the subcommittees in ACE have been responsible for individual chapters, and in this way we have been able to cover the complete history of the Antarctic Ice Sheet and its climate evolution. The book will be of interest to research scientists from a wide range of disciplines including glaciology, palaeoclimatology, sedimentology, climate change, environmental science, oceanography and palaeoentology. It will also be valuable as a supplementary text for undergraduate courses. We are grateful to our many friends and colleagues for advice and encouragement through the gestation of the book over the last 3 years. We also acknowledge input to the ACE initiative by a number of scientists (many of them contributed to this book), including P. Barrett, A.K. Cooper, J. Francis, R. Gersonde, M.J. Hambrey, D.H. Harwood, A. Moldonado, D. Pollard,
xii
Preface
D. Sugden, G. Villa, P.-N. Webb and G.S. Wilson. We are sure that the chapter authors will join us in thanking the reviewers for their comprehensive and valuable comments and suggestions. We acknowledge their very special contributions to this book by naming them here: J. Evans, J. Francis, W. Howard, L. Krissek, A. Mackintosh, C. O’Cofaigh, G. Orombelli, A.H. Orsi, D. Pollard, C.A. Ricci, I.C. Rutt, E. Stump, C. Summerhayes and G.S. Wilson. Finally we thank Linda Versteeg-buschman, Femke Wallien and Suja Narayana of Elsevier Science for their support in the production of this book.
Fabio Florindo Martin Siegert Rome and Edinburgh, July 2008
Elsevier Radarweg 29, PO Box 211, 1000 AE Amsterdam, The Netherlands The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK First edition 2009 Copyright r 2009 Elsevier B.V. All rights reserved No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher Permissions may be sought directly from Elsevier’s Science & Technology Rights Department in Oxford, UK: phone (þ44) (0) 1865 843830; fax (þ44) (0) 1865 853333; email:
[email protected]. Alternatively you can submit your request online by visiting the Elsevier web site at http://www.elsevier.com/locate/permissions, and selecting Obtaining permission to use Elsevier material Notice No responsibility is assumed by the publisher 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. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made 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: 978-0-444-52847-6 ISSN: 1571-9197 For information on all Elsevier publications visit our website at books.elsevier.com
Printed and bound in Hungary 09 10 11 12 13 10 9 8 7 6 5 4 3 2 1
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00001-3
Chapter 1
Antarctic Climate Evolution Martin J. Siegert1, and Fabio Florindo2 1
School of GeoSciences, Grant Institute, University of Edinburgh, The King’s Buildings, West Mains Road, Edinburgh EH9 3JW, UK 2 Istituto Nazionale di Geofisica e Vulcanologia, via di Vigna Murata 605, 00143 Roma, Italy
ABSTRACT Central to the understanding of global environmental change is an appreciation of how the Antarctic Ice Sheet interacts with climate. To comprehend the processes involved one must look into the geological record for evidence of past changes. For several decades international efforts have been made to determine the glacial and climate history of Antarctica and the Southern Ocean. Much of this information derives from studies of sedimentary sequences drilled in and around the continent. In addition, there have been numerous terrestrial geological expeditions to the mountains exposed above the ice surface usually close to the margin of the ice sheet. Holistic interpretation of these data is now being made, and hypotheses on the size and timing of past changes in Antarctica are being developed. In 2004, the Scientific Committee on Antarctic Research (SCAR) commissioned a scientific research programme on Antarctic Climate Evolution (ACE) to quantify the glacial and climate history of Antarctica. This book is a result of that programme, and documents, for the first time, the state of knowledge concerning the ice and climate evolution of the Antarctic continent and its surrounding seas through the Cenozoic era.
1.1. Introduction The Antarctic Ice Sheet has existed for approximately 35 million years, but it has fluctuated considerably and has been one of the major driving forces for Corresponding author. Tel.: +44(0)131 650 7543; Fax: +44(0)131 668 3184;
E-mail:
[email protected] (M.J. Siegert).
2
M.J. Siegert et al.
Quat.
Pleistocene Pliocene 5.33 Ma Miocene 23.03 Ma Oligocene
PALAEOGENE
TERTIARY
CENOZOIC
NEOGENE
1.81 Ma
33.9 ± 0.1 Ma Eocene 55.8 ± 0.2 Ma Palaeocene 65.5 ± 0.3 Ma
Figure 1.1: Geological time periods during the Cenozoic era. Dates listed on the right hand side are taken from Gradstein et al. (2004).
changes in global sea level and climate throughout the Cenozoic (Fig. 1.1). The rates, size and frequencies of these fluctuations have been the subjects of considerable debate. Determination of the scale and rapidity of the response of large ice masses and associated sea ice to climatic forcing is of vital importance, because ice-volume variations lead to: (1) changing global sea levels on a scale of tens of metres or more, and (2) alteration to the capacity of ice sheets and sea ice as major heat sinks/insulators. It is thus important to assess the stability of the cryosphere under a warming climate (IPCC, 2007), particularly as icecore records have yielded evidence of a strong correlation between CO2 concentrations in the atmosphere and palaeotemperatures (Fig. 1.2). This concern is justified when CO2 levels are compared with those of the past. Since Antarctica is a major driver of Earth’s climate and sea level, much effort has been expended in deriving models of its behaviour. Some of these models have been successfully evaluated against modern conditions. In 2004, modelling the past record of ice-sheet behaviour in response to changes in climate (inferred from ice cores, sedimentary facies and seismic data), palaeoceanographic conditions (inferred from palaeoecology and climate proxies in ocean sediments) and palaeogeography (as recorded in landscape evolution) was seen as a critical next step, and became the focus of the ACE programme. The ACE programme facilitates research in the broad area of Antarctic climate evolution. The programme links new geophysical surveys and
Antarctic Climate Evolution
3
Figure 1.2: Variation in the Earth’s temperature during the last 65 million years, based on reconstructions from deep-marine oxygen isotope records. Future atmospheric temperature scenarios are based on IPCC (2001). Greenhouse trace gas projections are shown at top of diagram. Given the worse-case scenario, planetary temperatures could increase in 100–300 years to a level where, according to our knowledge of previous Antarctic glaciations, ice cover on Antarctica could not be sustained. The representation of palaeotemperatures is adapted from Crowley and Kim (1995). geological studies on and around the Antarctic continent with ice-sheet and climate modelling experiments. The programme determines past climate conditions and changes in both the recent past (i.e. during the Holocene, prior to anthropogenic impacts as well as at the last glacial maximum, when temperatures were cooler than at present) and the more distant past (i.e. in the pre-Quaternary, when global temperature were several degrees warmer
4
M.J. Siegert et al.
than they are today). This cross-disciplinary approach, involving climate and ice-sheet modellers, geologists and geophysicists, has led to a substantial improvement in the knowledge base on past Antarctic climate, and our understanding of the factors that have guided its evolution. This in turn allows us to build hypotheses, examinable through numerical modelling, for how the Antarctic climate is likely to respond to future global change.
1.2. Antarctic Glacial History As is discussed in Chapter 7, the East Antarctic plate formed a significant component of the Gondwanaland super-continent during the Jurassic. Since 180 Ma, this continent broke up into what are recognised today as distinct continental landmasses with the repositioning of Antarctica at southern polar latitudes in the Early Cretaceous (ca. 120 Ma). In spite of its polar position, Antarctica is thought to have remained mostly ice-free, vegetated, and with mean annual temperatures above freezing until the latter half of the Cenozoic (around 34 million years ago, Fig. 1.1), whereupon the continent became subject to repeated phases of glaciation at a variety of temporal and spatial scales. The southern continent and its surrounding ocean basins have been the target of numerous scientific expeditions and several scientific drilling project efforts that have led to significant advances in understanding of Cenozoic climate evolution, oceanography, and biota of the Antarctic continent and the Southern Ocean. The deep-ocean records document clearly the long-term cooling of climates over the past 50 million years and large-scale variability in the last 3–5 million years. They also show events that are either abrupt or brief (e.g. the Paleocene warming event with a duration of less than 1 million years; the Middle-Eocene Climatic Optimum, MECO at ca. 41 Ma), or are marked by a distinct shift in the rate at which long-term changes occur (i.e. middle-Miocene increased cooling trend). The explanation for these events include changes in atmospheric gas concentrations (e.g. carbon dioxide and methane), opening of gateways with enhanced ocean circulation, peaks in orbital forcing resulting from Croll–Milankovitch cyclicities, interactions with northern hemisphere glaciations and others. Scientific drilling on the Antarctic continental shelf and upper slope, to examine the direct record of glaciation, has been sparse and has had significant problems with recovery (o20% in diamict) using current Integrated Ocean Drilling Programme (IODP) techniques. Consequently, the linkages between Antarctic
Antarctic Climate Evolution
5
continental shelf and deep-ocean basin records are not well established and the basic problem of ice-sheet history remains unsolved. Proxy measurements (particularly oxygen isotopes) provide general details, but initiation, growth and extent of the ice sheets still are debated. 1.2.1. Late Eocene-Early Oligocene Cooling The Eocene to early Oligocene (focus of Chapter 8) represent a time of global cooling which was marked by reorganisation of global ocean circulation patterns and significant turnovers in the marine and terrestrial biota (e.g. Berggren and Prothero, 1992) that culminates in the development of the first Antarctic Ice Sheet and an important expansion of Antarctic ice volume. Global deep-sea oxygen isotope records indicate that this long-term cooling trend was not monotonic, but that it was interrupted by a series of abrupt short-term (ca. 1 million years) excursions in d18O (Zachos et al., 2001). Among these, the Oi-1 cooling event (Miller et al., 1991) at 33.55 Ma marked one of the most significant global climatic deteriorations in the Cenozoic in response to the appearance of the first continent-wide glaciation in Antarctica (e.g. Zachos et al., 1996). Coupled GCM/ice-sheet modelling has already been used to show that the formation of the East Antarctic Ice Sheet was triggered by a combination of gradual pCO2 lowering coupled with iceclimate feedbacks and orbital-forcing-induced cooling, rather than by the cooling associated with the opening of circumpolar seaways during the earliest Oligocene (e.g. Kennett et al., 1974; DeConto and Pollard, 2003; Lawver and Gahagan, 2003).
1.2.2. Oligocene–Miocene Boundary Mi-1 Glaciation The Oligocene–Miocene boundary (Chapter 9) marks a significant transition in the development of the Antarctic cryosphere, where small dynamic ice sheets of the late Oligocene rapidly expanded to continental scale in the early Miocene. The transition is recorded in benthic foraminiferal d18O records as a positive 1.0 per mil shift, representing the first of the Miocene glaciations (Mi-1). The climatic significance of this was first outlined by Zachos et al. (2001) who recognised the coincidence of the Oligocene–Miocene boundary and the Mi-1 isotope excursion with an unusual coincidence of low eccentricity and low-amplitude variability in obliquity of the Earth’s orbit. Mg/Ca reconstructions imply little or no change in temperature and that the ice-volume increase was equivalent to 90 m of sea level lowering (assuming a
6
M.J. Siegert et al.
Pleistocene calibration of 0.11m d18O per 10 m of sea level). Sediment cores recovered in the Western Ross Sea indicate orbital modulation of the ice sheet during the transition, and corroborate proxy ocean records (Naish et al., 2001). It is argued that the Mi-1 event occurred as a consequence of the aforementioned unique setup of orbital parameters during an interval of declining CO2 that led to a prolonged period of cold summer orbits, during which time a large ice sheet established. This was then followed by warmer polar summers and enhanced melt from increased eccentricity and highamplitude variability in obliquity in the early Miocene, allowing the recovery of vegetation on the Antarctic craton. 1.2.3. Middle-Late Miocene Cooling The middle-to-late Miocene period represents a time of significant ice-sheet expansion in Antarctica (Chapter 10). Interpretations of deep-sea isotope records and observations from geologic data from around the world suggest that the middle-Miocene encompassed a change from a period of warm climatic optimum, approximately 17–15 million years ago, to the beginning of major cooling between ca. 14.5–13.5 million years ago, and the formation of a quasi-permanent ice sheet on East Antarctica (Lewis et al., 2007). One outstanding question revolves around whether this cooling led to an ice sheet in East Antarctica that remained stable and in existence to the present day or underwent large-scale fluctuations. New seismic-stratigraphic data from the Ross Sea reveal at least five major intervals of ice shelf advance and retreat in the middle-Miocene. Much of this ice is sourced in West Antarctica, suggesting the presence of a large and dynamic ice sheet in a part of Antarctica that is conventionally thought to be of lesser importance at this time. One of the most vexing questions concerns the stability of Antarctic climate and ice during the late Miocene and it has been the subject of almost continuous debate for more than 20 years. A variety of indicators from the McMurdo Dry Valleys suggest the maintenance of stable, hyper-arid, colddesert conditions since 13 Ma. However, microfossil studies (mainly diatoms) in the Transantarctic Mountains, and sedimentological work within Antarctic fjords are suggestive of significant climatic dynamism extending from the late Miocene through the Plio-Pleistocene. A degree of heterogeneity in climate response is expected considering the size and diverse landscapes of Antarctica. Yet the existing state of knowledge is sufficiently contradictory that the community has evolved into two camps when it comes to describing the ice sheet in pre-Quaternary time: the ‘stabilists’ and ‘dynamicists’ (Miller and Mabin, 1998).
Antarctic Climate Evolution
7
1.2.4. Pliocene Record The Pliocene Epoch (Chapter 10) is a critical time for understanding the nature of the Antarctic Ice Sheet as IPCC (2001, 2007) projections of global temperature rise suggest that we will reach Pliocene levels within the next hundred years (Fig. 1.2). Indirect evidence, such as sea level changes and ocean floor sediments, suggests that ice volumes were subject to cyclical variability. It is believed that, since Northern Hemisphere ice sheets were not fully developed, sea level changes were driven by fluctuations of the Antarctic Ice Sheet. Many scientists believe that it was the relatively unstable West Antarctic Ice Sheet that was responsible for these changes, but the role of the much larger East Antarctic Ice Sheet remains controversial. Key to this argument is the timing of the transition of the East Antarctic Ice Sheet from a polythermal, dynamic condition to a predominantly cold-base and stable state. Two opposing and vigorously defended views prevail. The long-standing view is that the East Antarctic Ice Sheet became stable in mid-Miocene time; evidence of which is primarily from the longevity of the landscape and well-dated surfaces and ash deposits in the Dry Valleys region along the western border of the Ross Sea. Another controversial view is that terrestrial glacial deposits, known as the Sirius Group, scattered in a number of locations through the Transantarctic Mountains, indicate dynamic ice-sheet conditions as recently as Pliocene time; based on diatom biostratigraphy and preserved vegetation. The latter viewpoint is supported by work on deposits known as the Pagodroma Group along the western side of the Lambert Glacier-Amery Ice Shelf drainage system. Each argument is internally consistent and the biggest challenge is to reconcile the differing views. If the East Antarctic Ice Sheet was indeed subject to major fluctuations until Pliocene time then, taking into account IPCC projections, we have cause to be concerned about the possibility of the East Antarctic Ice Sheet reacting to climate change within the next few centuries.
1.2.5. Pleistocene Glacial Cycles (Intervals of Extreme Warmth and Cold) Studies of Antarctic ice cores show that Pleistocene climate variability in the different sectors of the southern high latitudes (detailed in Chapter 11) has occurred out of phase. This raises questions about the response of the southern high latitudes to external climate drivers, such as orbital insolation, solar variability and internal amplifiers such as thermohaline circulation and
8
M.J. Siegert et al.
carbon cycle changes that operate at both Croll–Milankovitch and millennial– decadal time scales. These questions highlight a need for appropriate time series of climate variability from all sectors of the Southern Ocean. Recovery of sediment sequences with expanded Pleistocene sections, such as those from beneath the McMurdo Ice Shelf as undertaken by the ANDRILL programme (see www.andrill.org), permits the study of the structure and timing of glacial and interglacial cycles in the Southern Ocean at millennial time scales that extend well beyond the last four major climate cycles. In addition, several groups organised under the International Marine Global Change Study (IMAGES) programme have proposed to collect long piston cores for Pleistocene research from several sectors of the Southern Ocean. With new high-resolution Pleistocene time series from both the Antarctic margin and offshore sites, we can begin to determine if the abrupt climate changes that have been documented from the Atlantic and Indian sectors, and in polar ice cores, have also occurred in the Pacific basin. During the last decade, many palaeoceanographic studies focused on millennial climate variability (e.g. on the Dansgaard-Oeschger (DO) events). They show that ocean thermohaline circulation is capable of becoming unstable in response to climate change and modifications to the cryosphere. The existing palaeoceanographic record documents mainly the North Atlantic Ocean, and modelling experiments have mostly explored the variability of North Atlantic Deep Water formation forced by fresh water flux from ice surge events. However, Southern Ocean sea ice may be important during glacial periods, and the glacial ‘on/off’ modes of global circulation could also be linked to deep-water formation in the Southern Ocean. 1.2.6. Last Glacial Cycle and Deglaciation At the last glacial maximum (LGM, B21 ka), ice-sheet expansion in Antarctica was responsible for around 15 m of global sea level fall, with growth most likely taking place over continental shelves exposed as a result of sea level fall from the development of Northern Hemisphere ice sheets (Siegert, 2001). There are currently three different ideas about the onset of deglaciation: (1) changes in the water balance of the North Atlantic, the source region for much of the global thermohaline circulation, serve to propagate the deglacial signal worldwide; (2) changes in the Southern Ocean, as recorded in some ice cores, lead deglaciation as seen in Greenland ice; and (3) synchronicity in the timing of high-latitude climate change in both hemispheres, and some tropical records, suggest that tropical forcing is a key initiator of
Antarctic Climate Evolution
9
deglaciation. It may seem surprising that this controversy has not already been settled. The most important confound for establishing synchronicity, or its absence, among the available palaeoclimate records revolves around chronology development. It is notoriously difficult to date LGM ice layers and sediments to an accuracy of better than 1–2 thousand years. It is also difficult to separate local climate or geomorphic signals from large transformations that are regionally or globally important. What is needed to resolve the deglacial synchronicity issue are better records from rapidly deposited deglacial sequences across a range of longitudes and latitudes in the Southern Ocean that use sedimentary or glacial outlet indicators to directly track regional climate systems. Currently there are too few precisely dated records of the LGM from the Southern Ocean. Chapter 12 brings together several geological datasets relevant to the LGM in an attempt to constrain the ice volume at this important time of climate change.
1.3. Structure and Content of the Book Antarctic Climate Evolution presents the state of knowledge concerning the ice and climate history of the Antarctic continent and its surrounding seas throughout the Cenozoic. It begins with two chapters that provide the historical context in which palaeoclimate knowledge has been gathered. In Chapter 2, background to the International Polar Years (IPYs) is presented. These are periods in which major advances in our understanding of the Antarctic continent have occurred. Chapter 3 summarises the history of geological investigations on the continent. The next three chapters provide background information necessary in comprehending Cenozoic change in Antarctica. Chapter 4 discusses the role of the Southern Ocean in modulating and controlling ice and climate in Antarctica, Chapter 5 presents evidence of ice-sheet changes from studies of sea-floor sediments (obtained from drilling and seismic investigations) and Chapter 6 introduces the concept of numerical ice-sheet modelling as a powerful tool in quantifying former ice sheets. From this point, the book presents a series of chapters, each of which deals with a specific time period in Antarctic history, as detailed above. Our appreciation of Antarctic climate evolution is possible because of pioneering research undertaken over the past 100 or so years. It has involved thousands of academics from a variety of disciplines and, of course, nations. Major advances in our knowledge have come about as a result of large integrated programmes of activity. Organising logistics in Antarctica necessary to gather geological records often needs the finances and assistance
10
M.J. Siegert et al.
of several nations. For example, offshore drilling has been supported by the IODP, and near-shore drilling has been organised through the ANDRILL consortium. Ice coring, subglacial lake exploration, numerical modelling and palaeoceanography each have multination programmes at their core (Chapter 13). All four IPYs (1882–1883, 1932–1933, 1957–1958 and 2007–2008) have offered momentum and encouragement for international multidisciplinary activities that have lasted well beyond their official time frames. For example, the third IPY (also known as the IGY) resulted in SCAR and the Antarctic Treaty. With several new ambitious programmes being either undertaken or planned within the fourth IPY, the next decade will undoubtedly see our knowledge of Antarctic Climate evolution develop considerably. Such knowledge will be critical to assessing how the Southern Ocean and ice sheets interrelate and feedback with global climate change.
REFERENCES Berggren, W. A., & Prothero, D. R. (1992). Eocene-oligocene climatic and biotic evolution: An overview. In: D. R. Prothero, & W. A. Berggren (Eds). EoceneOligocene Climatic and Biotic Evolution. Princeton University Press, Princeton, NJ, pp. 1–28. Crowley, T. J., & Kim, K.-Y. (1995). Comparison of longterm greenhouse projections with the geologic record. Geophy. Res. Lett., 22, 933–936. DeConto, R. M., & Pollard, D. (2003). Rapid Cenozoic glaciation of Antarctica triggered by declining atmospheric CO2. Nature, 421, 245–249. Gradstein, J. G., Ogg, A. G., Smith, F. P., Agterberg, W., & 36 others. (2004). A Geologic Time Scale 2004. Cambridge University Press, 589 pp. IPCC (2001). The Scientific Basis Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change (IPCC). In: J. T. Houghton, Y. Ding, D. J. Griggs, M. Noguer, P. J. van der Linden, & D. Xiaosu (Eds). Cambridge University Press, Cambridge, UK, 944 pp. IPCC (2007). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. In: S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, & H. L. Miller (Eds). Cambridge University Press, Cambridge, UK, 996 pp. Kennett, J. P., Houtz, R. E., Andrews, P. B., Edwards, A. R., Gostin, V. A., Hajos, M., Hampton, M. A., Jenkins, D. G., Margolis, S. V., Ovenshine, A. T., & Perch-Nielsen, K. (1974). Development of the circum-Antarctic current. Science, 186, 144–147. Lawver, L. A., & Gahagan, L. M. (2003). Evolution of Cenozoic seaways in the circum-Antarctic region. Palaeogeogr., Palaeoclimatol., Palaeoecol., 198, 11–37.
Antarctic Climate Evolution
11
Lewis, A. R., Marchant, D. R., Ashworth, A. C., Hemming, S. R., & Machlus, M. L. (2007). Major middle Miocene global climate change: Evidence from East Antarctica and the Transantarctic mountains. Geol. Soc. Am. Bull., 119, 1449–1461, doi:10.1130B26134.1. Miller, K. G., Wright, J. D., & Fairbanks, R. G. (1991). Unlocking the ice house: Oligocene-Miocene oxygen isotopes, eustacy and margin erosion. J. Geophys. Res., 96, 6829–6848. Miller, M. F., & Mabin, M. C. G. (1998). Antarctic Neogene landscapes – in the refrigerator or in the deep freeze? GSA Today, 8, 1–2. Naish, T. R., Woolfe, K. J., Barrett, P. J., et al. (2001). Orbitally induced oscillations in the East Antarctic Ice Sheet at the Oligocene/Miocene Boundary. Nature, 413, 719–723. Siegert, M. J. (2001). Ice sheets and Late Quaternary environmental change. John Wiley and Sons Ltd., Chichester, UK, 231 pp. Zachos, J. C., Pagani, M., Sloan, L., Thomas, E., & Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693. Zachos, J. C., Quinn, T. M., & Salamy, K. A. (1996). High resolution (104 years) deep sea foraminiferal stable isotope records of the Eocene-Oligocene climate transitions. Paleoceanography, 11, 251–266.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00002-5
Chapter 2
The International Polar Years: A History of Developments in Antarctic Climate Evolution Fabio Florindo1,, Antonio Meloni1 and Martin Siegert2 1
Istituto Nazionale di Geofisica e Vulcanologia, via di Vigna Murata 605, 00143 Roma, Italy 2 School of GeoSciences, Grant Institute, University of Edinburgh, The King’s Buildings, West Mains Road, Edinburgh EH9 3JW, UK
ABSTRACT The first three International Polar Years (IPYs; 1882–1883, 1932–1933, 1957– 1958) were major periods of intense multidisciplinary polar research, bringing significant new insights into global processes and laying the foundation of knowledge of the polar regions for future decades. The fourth IPY (2007–2009) continues the tradition of international science years and is one of the most ambitious internationally coordinated scientific research programmes ever attempted. In contrast to the three previous IPYs, the new IPY incorporates research within social science and its interface with the natural sciences. The new IPY also includes a wide range of education and outreach activities, and a commitment to excite and train the next generation of polar researchers. We discuss briefly the history of the IPYs, and their contribution to comprehending Antarctic Climate Evolution.
2.1. Introduction The polar regions play key roles in global climate change and have profoundly affected environments during the Cenozoic, influencing sea levels, atmospheric composition and dynamics, and ocean circulation. Starting from the end of the nineteenth century, several major internationally Corresponding author. Tel.: +39 0651860 383; Fax: +39 0651860 397;
E-mail: fl
[email protected] (F. Florindo).
14
F. Florindo et al.
coordinated explorations of the polar regions have taken place, which have improved our understanding of them and how they influence the world. The most prominent periods of exploration were: the first International Polar Year (IPY) of 1882–1883 (e.g. Heathcote and Armitage, 1959; Wood and Overland, 2006), the so-called Heroic Age of polar exploration, stimulated by the International Geographical Congress of 1895, which had made Antarctica a target, the second IPY (1932–1933) (e.g. Gerson 1958) and the International Geophysical Year (IGY) of 1957–1958 (e.g. Buedeler, 1957; Korsmo, 2007), later extended to include 1959 and which started life as the third IPY. These were major initiatives that involved an intense period of multidisciplinary polar research bringing significant new insights into global processes and laying the foundation of knowledge of the polar regions for future decades. The first two IPYs and the IGY involved an increasing number of countries and scientists, and produced unprecedented levels of knowledge and understanding in many fields of research. The 12 countries of the first IPY grew to 67 in the IGY, in which some 5,000 scientists and support staff were engaged in Antarctica alone. They not only changed the way science was conducted in the polar regions, from single nation programmes to complex multinational collaborations, they also standardised measurements, made data freely available to all and initiated the system of World Data Centers. The fourth IPY, 2007–2009 (Allison et al., 2007), is one of the largest collaborative science programs ever attempted. It continues the tradition of international science years, includes multidisciplinary research operating in both polar regions, and involves some 50,000 participants from 63 countries. A wide range of scientific problems will be addressed, including issues related to society. It differs from the three previous IPYs in that it includes all natural science disciplines – not just physics and geophysics, it incorporates the social sciences, and it includes a wide range of education and outreach activities aimed at attracting the next generation of polar scientists and engaging the attention of the public and policy makers. In this chapter, we discuss briefly the history of the polar years, and their contribution to comprehending Antarctic Climate Evolution (ACE).
2.2. The First International Polar Year (1882–1883) In August 1874, Captain Karl Weyprecht (1838–1881) (Figs. 2.1 and 2.2) returned from an Arctic expedition, of which he was leader. The
International Polar Years
15
Figure 2.1: Carl Weyprecht (1838–1881) ideas initiated the first IPY (photo courtesy: Archive, Alfred Wegener Institute for Polar and Marine Research).
Austro-Hungarian North Pole Expedition (1871–1874) aimed to explore the northwest of Nowaja Semlja, in the search for the Northeast Passage. During that Arctic expedition on the ice-strengthened schooner ‘Admiral Tegetthoff’, they discovered Franz Joseph Land (890 km from the North Pole) and gathered valuable information about the drift of icebergs and about meteorological and magnetic conditions in the Arctic. Although it was a successful expedition, it occurred to Weyprecht that single nation expeditions of this nature, often having geographical discovery as their primary goal, could only advance the frontiers of scientific knowledge to a limited extent. Knowing that answers to the fundamental questions of meteorology and geophysics were most likely to be found near the Earth’s poles, Weyprecht
16
F. Florindo et al.
Figure 2.2: The cover page of the ‘Illustrierten Wiener Extrablattes’ (Viennese Illustrated Special Edition), 25 September 1874. The cover heralds the return of the leaders of the Austro-Hungarian North Pole Expeditions, Carl Weyprecht and Julius Payer. became an avid advocate of internationally coordinated exploration of the polar regions; his views were influential in the formation of the largest coordinated series of scientific expeditions taken in the polar regions during the nineteenth century, namely what is now known as the first IPY of 1882–1883.
International Polar Years
17
During the 48th Meeting of ‘German Naturalists and Physicians’ in Graz (18 September 1875), Weyprecht gave a lecture about the ‘Basic principles of Arctic research’, in which he suggested establishing a network of fixed Arctic observation stations (Baker, 1982). In 1879, during the second International Meteorological Conference in Rome, these ideas were presented together with those of Georg von Neumayer (1826–1909), first Director of the German Hydrographic Office in Hamburg. Here it was recommended to discuss the erection of a number of observatories in the Arctic and Antarctic for simultaneous hourly meteorological and magnetic observations around the poles. These ideas generated international interest and, during the first International Polar Conference at the German Hydrographical Office in Hamburg (1–5 October 1879), an organising body called the International Polar Commission, chaired by Neumayer, was established. This commission included Denmark, Norway, Russia, Sweden, Finland, Germany, AustriaHungary, the Netherlands, France, the United States and Great Britain, with the assistance of the new Dominion of Canada. During that conference, the first IPY was planned for the biennium 1881–1882. The following year, during the second International Polar Conference in Bern (7–9 August 1880) this commission agreed to postpone the IPY and declared that it would be held in 1882–1883 to coincide with a transit of Venus across the face of the Sun, on 6 December 1882. In doing so simultaneous observations from different places on the globe could be made to calculate the astronomical unit (AU=nearly 150 million kilometres; the distance between the Earth and Sun). Between 1 and 8 August 1881, during the third International Polar Conference, held in St. Petersburg, and 5 months after Weyprecht’s death on 29th March 1881 in Michelstadt, the International Polar Commission outlined the details of the first IPY, to last from 1 August 1882 until 1 September 1883 (Heathcote and Armitage, 1959). In total, 12 countries participated in the first IPY resulting in 15 coordinated expeditions to the poles (13 to the Arctic, and 2 to peri-Antarctic islands). Fourteen research stations were established (Fig. 2.3) where researchers conducted experiments and gathered data (hourly records) over the course of the year that would greatly enhance the basis of then current knowledge of the Earth’s magnetic field, surface weather conditions and astronomy. Two of these stations were in the Southern Hemisphere: Orange Bay at the southern tip of Tierra del Fuego (established by France) and Moltke-Hafen at Royal Bay, South Georgia (established by Germany). Another 34 permanent stations were located outside Polar territories (e.g. Shanghai, Rio de Janeiro, Bombay) bringing the number of stations participating in the IPY to 48. This first IPY was primarily focused on
18
F. Florindo et al.
Figure 2.3: North polar chart, showing Arctic research stations during the first IPY. During the first IPY, eleven nations established fourteen principal research stations across the polar regions. Twelve stations were in the Arctic and two stations were in the Antarctic region (map from the Scottish Geographical Magazine, Volume I, No. 12, 1885 and scanned by the University of Texas Libraries at http://www.lib.utexas.edu/).
International Polar Years
19
physics – especially meteorology, magnetism and auroral studies, rather than on interdisciplinary work, but its investigations did extend, though locally and in a limited fashion, to other fields like botany, geology and zoology. One of the most significant results of the first IPY was the mapping of the Aurora Borealis – known at the time as the Northern Lights – showing that it often occurs in an almost circular belt centred on the north magnetic pole. A huge amount of polar data for that epoch became available to scientists. In the manner of the times, the data analysis was largely confined to the primary single disciplines that were the focus for the IPY – meteorology, geomagnetism, auroral observations. Unfortunately there was no central archive for the data and no organisation to facilitate exchange of information. As a result, no major attempts were made to synthesise the information obtained, although some scientists did manage to pull together data from others to support their own individual studies of geomagnetism and auroras. Nonetheless, the first IPY is regarded as an epoch-making event in which a major step forward was made in environmental scientific knowledge of importance not just for the polar regions but for the Earth as a whole.
2.3. The Second International Polar Year (1932–1933) The success of the first IPY stimulated an expanded effort 50 years later to hold a second IPY. During the 1920s, while conducting high-altitude weather balloon observations, scientists detected extremely strong winds at heights of 10–15 km above the surface of the Earth; these are known today as the ‘jet stream’. One such scientist was Johannes Georgi, a meteorologist of the Maritime Institute of Hamburg. During a meeting of the Deutsche Seewarte (1927), Georgi proposed to investigate this phenomenon with a coordinated international research effort that would commence on the 50th anniversary of the first IPY. In June 1928, an informal organisational meeting was held in London to discuss plans for the event and a year later, in 1929, the International Meteorological Organization (IMO), the predecessor of World Meteorological Organization (WMO), endorsed the effort and formed a commission to undertake planning for the second IPY. In August 1930, a first meeting of the International Polar Commission was held in Leningrad (now St. Petersburg) under the presidency of Dr. D. La Cour of Denmark in order to organise and integrate the total effort. Delegates of 10 nations were present: Canada, Denmark, Finland, France, Germany, Japan, Norway, UK, USSR and the USA. A further 16 nations expressed an interest in this
20
F. Florindo et al.
initiative. The aims and the scientific programme of the second IPY were presented during the second meeting of the International Polar Commission held in Innsbruck, Austria, in September 1931. When the second IPY started on 1 August 1932, forty-nine nations participated and, despite considerable economic challenges, due to the worldwide economic depression that began in 1929, it heralded advances in many fields: meteorology, atmospheric science, the mapping of the jet streams, ionospheric soundings and their role in radio communications, magnetic observations and, to a lesser extent, atmospheric electricity (Fig. 2.4). The IPY was notable for the first massive deployment of the new and somewhat experimental radiosondes for upper atmosphere measurement. Ninety-four research stations were maintained in the Arctic during the second IPY and many of these are still active today. This was the time of the Great Depression, so funds were limited. Plans for the Antarctic suffered as a result, and were not pursued as originally planned. Chile established a station at Punta Arenas and Argentina in the South Orkneys. Meteorological observations were made by Norwegian whalers in the Southern Ocean. Magnetic observations were made in several locations in the Southern Hemisphere, including Christchurch (New Zealand), Watheroo (Australia) and Cape Town (South Africa).
Figure 2.4: Stuart McVeigh, member of the Canadian team at Chesterfield Inlet on Hudson Bay, north-eastern Canada, holding airborne kite with a meteograph on it (photo courtesy: Department of Physics fonds, University of Saskatchewan Archives).
International Polar Years
21
Among the many exceptional men that participated in the second IPY, one figure stands out as symbolic – Rear-Admiral Richard E. Byrd of the USA (Fig. 2.5). Byrd had intended to sail for Antarctica in 1932 in the middle of the IPY, but due to the funding problems created by the Great Depression had to delay sailing until late 1933 and did not reach the Ross Sea until January 1934, so in one sense his expedition was not strictly speaking part of the IPY. It was noted for its exploitation of major advances in technology, in particular aviation, navigation, motor transport and radio. These technologies changed the nature of polar exploration from ship-borne or land-based expeditions using dogs and sledges, to parties employing machines and aircraft. Clearly, while those early years of flying were adventurous, flying in Antarctica entailed considerable danger. Nonetheless, the potential of utilising aircraft for exploration purposes was
Figure 2.5: Admiral Richard E. Byrd, ca. 1930 (photo courtesy: The Ohio State University Libraries; http://library.osu.edu/sites/exhibits/byrdflight/).
22
F. Florindo et al.
huge, especially in Antarctica where so much remained undiscovered. Admiral Byrd knew this, having flown over the South Pole during his first Antarctic expedition in 1929, as well as over the North Pole. It is because he took full advantage of developments in aviation that he is now remembered as one of the most famous and influential polar explorers. During Byrd’s polar expeditions during the second IPY he spent five months wintering alone whilst operating a meteorological station – the first-ever research station located inland from the coastal margin (Bolling Advance Base), on the Ross Ice Shelf, at the southern end of Roosevelt Island (123 miles from the sea). Just as in the first IPY, research during the second IPY was not limited to the polar regions. Even so, although many observatories were located in equatorial regions, those responsible at the time did not change the name of the enterprise as the main emphasis was clearly on the polar regions. The global network of observations allowed, for the first time, an appreciation of geophysical phenomena at a planetary level. A vast amount of data produced during the second IPY was collected and the Commission for the Polar Year established an official repository for IPY at the Danish Meteorological Institute in Copenhagen (Denmark). The second IPY officially continued until 1 September 1933, just prior to Byrd’s departure from the USA.
2.4. The Third International Polar Year/International Geophysical Year (1957–1958) Following the success of the first two polar years it seemed reasonable to agree that these events should occur every 50 years. However, so many scientific and technological improvements were made in a short time after the second IPY that already at the beginning of the fifties many scientists believed a third coordinated IPY would allow major advances in our knowledge of Antarctica, which should not wait until the 50th anniversary of the second IPY. On 5 April 1950, in Silver Spring, Maryland, a small group of eminent physicists gathered to meet in an informal meeting in Van Allen’s home. Among these were the house owner James Van Allen (1914–2006), Lloyd Viel Berkner (1905–1967), Siegfried Frederick Singer (born in 1924) and Sidney Chapman (1888–1970) (Fig. 2.6), all of whom had been involved in research for military applications during the World War II. They realised the potential of the new technologies such as rockets, radar and numerous other geophysical techniques perfected during the war, and hoped to
International Polar Years
23
Figure 2.6: Sidney Chapman (1888–1970) was the President of CSAGI and guided much of the IGY planning. He and Lloyd Berkner were called the ‘fathers of the IGY’ (NOAA 200th Anniversary Celebration Website, http://celebrating200years.noaa.gov/). redirect these toward creating a basic knowledge of Antarctica and the world. This desire reflected the lack of focus on the Antarctic in the previous IPYs. In the months following the ‘Van Allen’s dinner’, other geophysicists joined the proponents and, during the summer of 1950, they presented their idea for a third IPY at the Conference on the Physics of the Ionosphere, at Pennsylvania State University. From this point onwards the plan moved to an international scale and a proposal was presented to the International Council of Scientific Unions (ICSU) that the time was right for a new IPY, which would fall in 1957–1958 and coincide with an expected sunspot maximum (from the standpoint of solar-terrestrial research, the period of the previous IPY was not a particularly interesting time because it was coincident with a minimum in the 11-year solar cycle!). ICSU endorsed the proposal and broadened its scope to include studies of the whole planet, rather than just polar studies. The program was renamed the IGY and ICSU established a special committee (Comite´ Spe´cial de l’Anne´e Ge´ophysique Internationale, CSAGI), headed by S. Chapman (president) (Fig. 2.6) and
24
F. Florindo et al.
Figure 2.7: The official logo of the 1957–1958 International Geophysical Year. L.V. Berkner (vice-president), to act as the governing body for all IGY activities (Fig. 2.7). Care was taken to ensure that this committee would remain non-nationalistic, apolitical and geared towards a scientific agenda. The first meeting of CSAGI took place in Brussels, Belgium, in July 1953, and was followed by a series of general assemblies and regional conferences (held for the Arctic, Antarctic, the Americas, Eastern Europe, Africa and the western Pacific). At the general assemblies the main criteria for the IGY were established. The Soviet Union was not initially among the 26 proposing countries, probably as a result of the very peculiar cold war politics of the time. Nonetheless, how could a program aiming to be named ‘International’ take place without the participation of the Soviet Union and its numerous allies? Fortunately, a short while later also the Soviet ‘block’ joined the program with a very comprehensive scientific approach. This truly international effort, against a background of cold war mistrust and weapons escalation, was a remarkable achievement. The launch of the first artificial satellite, ‘Sputnik 1’, on 4th of October 1957, gave a temporary political victory to the Soviet Union in the race for space; but the Americans were not far behind, launching Explorer 1 on 31 January 1958. These launches began the exponential growth in geophysical knowledge about the state of the planet that was to come from the use of space probes; it was the first dramatic step towards the space-based remote sensing techniques so much in use nowadays. The IGY, began on 1 July 1957 and was completed on 31 December 1958, although a one year extension – the International Geophysical Cooperation – was allowed. With more than 10,000 scientists and 67 countries involved, the IGY was a significant event in the history of science and surely the greatest international scientific enterprise in the middle of the twentieth century.
International Polar Years
25
Important discoveries were made in the fields of cosmic ray research, climatology, glaciology, oceanography, terrestrial atmosphere and the magnetic field (Fig. 2.8). Explorer 1, in particular, brought the discovery of the so-called Van Allen radiation belts. Major deep counter currents were discovered in the ocean, for example beneath the Gulf Stream. Importantly for Antarctica, the IGY also witnessed major, internationally coordinated over-ice traverses. Using seismic exploration techniques developed several years earlier during a Norwegian-British-Swedish expedition to Dronning Maud Land (1947–1952), the ice volume of the Antarctic continent was estimated with a degree of accuracy for the first time. It showed the continent to be overlain with a vast ice sheet, several kilometres thick in places, with a bed often depressed below sea level and a separation between East and West Antarctica. Several research stations were located in the continental interior, including a US base at South Pole, a Soviet base at the Magnetic South Pole, named Vostok Station, and a Soviet station near the Pole of Relative Inaccessibility, named Sovetskya. A plinth and bust of Lenin was placed at the exact site of the Pole of Relative Inaccessibility, facing towards Moscow; it remains there today and was visited for the first time in nearly 50 years in 2006. The impact of the IGY on comprehending the dimensions of the Antarctic Ice Sheet cannot be overstated. The measurements taken remain, in some of the more remote regions, the only data ever collected on ice thickness. Perhaps the most important conclusion of the IGY was an appreciation that Antarctic scientific exploration was best served by international corporation, regardless of global geopolitics. Subsequent to the IPY, nations interested in Antarctic matters gathered to declare Antarctica free from commercial exploitation, and a site for scientific collaboration and data sharing. The Scientific Committee on Antarctic Research (SCAR) was established during the IGY from discussions that took place in 1957 and led to the first meeting of SCAR held at the Hague, from 3–6 February 1958. Its objective was to develop scientific cooperation on research of continental scope, and to ensure that such research continued beyond the narrow confines of the IGY. As an apolitical independent organisation, and an integral part of ICSU, SCAR has served to integrate and facilitate research on Antarctic matters ever since. Its members represent the national academies of science in 34 countries (as of early 2008). The need for this kind of mechanism was evident to ICSU from an examination of the shortcomings of the first and second IPYs, when the organisation underpinning of the activity and its follow-up was weak or non-existent, which meant that the data were less valuable than they might otherwise have been.
26
F. Florindo et al.
Figure 2.8: ‘Poles’ was one of six posters (Earth, The Oceans, The Poles, Weather and Climate, Sun and Earth, Space) around which the National Academy of Sciences’ IGY Committee created its booklet ‘Planet Earth’. The numbers on this poster identify points discussed in the booklet (National Academy of Sciences, 1958) (courtesy: National Academy of Sciences).
For much the same reason, another innovation of the IGY was ICSU’s creation of the network of World Data Centers to be repositories for the large amounts of data collected by the participants. The emphasis on putting data into a centre so that it would be available to all was a novel departure.
International Polar Years
27
2.5. The Fourth International Polar Year (2007–2008) The fourth IPY marks the 50th anniversary of the IGY, and will be one of the most ambitious internationally coordinated scientific research programmes ever attempted. The official IPY 2007–2008 observing period will be from 1 March 2007 to 1 March 2009, in order to include two full annual seasonal cycles in the Arctic and in the Antarctic. Planning for the fourth IPY began at SCAR’s Shanghai meeting in 2002, and took off with support from ICSU in 2003 for an International Planning Group. For former IPYs, the rationale behind them lay in coordinating international activities to explore the unknown. While exploration of the geographical unknown is also an important aspect of the fourth IPY, the main emphasis now is on integrating international efforts so as to better comprehend the behaviour of the Earth system, and understand the roles of the poles in global change. This change in emphasis is readily apparent from the six IPY themes: status, change, global linkages, new frontiers, vantage point and the human dimension (see below). The fourth IPY involves over 200 projects endorsed by the ICSU/WMO Joint Committee that steers the process. The projects are either focused on the Arctic, or on the Antarctic, or are bipolar (Fig. 2.9). Just as impressive is the amount of endorsement and publicity that has surrounded the fourth IPY. For example, the Antarctic Treaty Parties, at their summit in Edinburgh on 19th June 2006, stated ‘We express our support for a successful IPY. We believe that the scientific research undertaken during the IPY will increase knowledge of the Antarctic and will yield a better understanding of the major terrestrial, ocean and atmospheric systems that control the planet. The polar regions are sensitive barometers of climate change, and we value their biodiversity. Their health is vital to the well-being of the Earth’s systems and its inhabitants’. As in the case of the third IPY relative to the second IPY, technology has advanced considerably in the interim, and we now have a network of remote sensing, navigational and communications satellites along with massive advances in information technology that together have changed the way in which science is done. Much can now be done remotely – e.g. automated weather stations reporting back to base via satellite. There is much less emphasis therefore in the fourth IPY than there was in the IGY on establishing bases. Instead much will be done by remote observation. Another new departure is that much of the data collected will form important contributions to computer-modelling investigations of polar processes, such as ice sheet changes. Given the huge quantities of data that
Figure 2.9: IPY planning chart as of 10 October 2007: the framework of hexagons provides a visual impression of how all the endorsed IPY projects are related in terms of geography and topic and how they may be linked (source: David Carlson, IPY 2007–2008, International Programme Office, Cambridge, UK).
28 F. Florindo et al.
International Polar Years
29
are hoped to be gained, and the need to inform and run numerical models, the full results of the IPY may take several years to become clear. There will be even more emphasis in the fourth IPY than in the IGY on the capture, storage and retrieval of data for the benefit of all. The polar regions are particularly important at this time because, among other things, they are much more sensitive than are other parts of the world to climate change, and they are experiencing significant environmental change, which is having a profound impact on the ecosystems and human activity. Aside from that, through long-range climate connections what happens in the polar regions influences what happens elsewhere on the Earth. As for the past IPY/IGY initiatives, the fourth IPY will provide a framework to undertake projects that normally could not be achieved by any single nation. Thousands of scientists from more than 60 countries, including those not traditionally involved in polar research, will examine a wide range of physical, biological and social research topics. A departure from the three previous IPYs is that about 60 of the 230 or so projects will focus on education and outreach objectives, aimed at attracting the next generation of polar scientists and engaging the public in global environmental issues. As mentioned above, the various scientific proposals from the scientific community led the IPY organisers to identify six science themes, listed below. Some projects will contribute to more than one of these themes. The majority aim at understanding the changing polar environment and the impact of those changes (e.g. Allison et al., 2007): 1. Status: to determine the present environmental status of the polar regions; 2. Change: to quantify and understand past and present natural environmental and social changes in the poles and to improve projections of future change; 3. Global linkages: to advance understanding on all scales of the links and interactions between polar regions and the rest of the globe, and of the processes controlling these; 4. New frontiers: to investigate the frontiers of science in these regions; 5. Vantage point: to use the unique vantage point of the polar regions to develop and enhance observatories from the interior of the Earth to the Sun and the cosmos; 6. The human dimension: to investigate the cultural, historical and social processes that shape the sustainability of circumpolar human societies and to identify their unique contributions to global cultural diversity and citizenship.
30
F. Florindo et al.
During the IPY there will be considerable emphasis on improving our understanding of the behaviour of the Antarctic Ice Sheet and the climate of the region, and on the history of the ice sheet and climate. The ACE programme will make a significant contribution to this latter goal. IPY projects will use proxy records from sediment cores, ice cores and other sources to define how the past climate and environments changed. A greater understanding of past changes in this region is crucial to forming a better knowledge of future global environmental change and to predicting the role of the Antarctic Ice Sheet in the future as Earth warms. During the IPY new ice cores will be recovered in Greenland (North Greenland Eemian Ice Drilling, NEEM project), and Antarctica (West Antarctic Ice Sheet Divide, WAIS Divide), providing new high-resolution records of glacial and interglacial changes during the Quaternary. On a longer time scale, drilling of sediment cores along the Antarctic continental margin (ANDRILL; www.andrill.org) aims to study the glacial and climate history of Antarctica and the Southern Ocean following the late Cretaceous (e.g. see Florindo et al., 2003, for a review of the recent history of circum-Antarctic drilling by the Ocean Drilling Program and the Cape Roberts Project, and see Hambrey and Barrett, 1993, for a more comprehensive review of earlier drilling in the Ross Sea region). Finally, it should be mentioned that the IPY is not the only activity conducting special activities during the 50th anniversary of the IGY. There are also the International Year of Planet Earth (http://www.esfs.org/); the Electronic Geophysical Year (EGY) (http://www.egy.org/); the International Heliophysical Year (IHY) (http://ihy.gsfc.nasa.gov/). Various International Council for Science Unions are coordinating these efforts and the International Programme Office, established by ICSU and WMO at the British Antarctic Survey in Cambridge, UK, will serve as the official contact point for these other programmes.
REFERENCES Allison, I., Be´land, M., Alverson, K., Bell, R., Carlson, D., Danell, K., Ellis-Evans, C., Fahrbach, E., Fanta, E., Fujii, Y., Glaser, G., Goldfarb, L., Hovelsrud, G., Huber, J., Kotlyakov, V., Krupnik, I., Lopez-Martinez, J., Mohr, T., Qin, D., Rachold, V., Rapley, C., Rogne, O., Sarukhanian, E., Summerhayes, C., & Xiao, C. (2007). The scope of science for the International Polar Year 2007–2008, WMO/TD, No.1364, pp. 1–79. Baker, F. (1982). The first International Polar Year. Polar Record, 21 (132), 275–285.
International Polar Years
31
Buedeler, W. (1957). The International Geophysical Year. United Nations Educational, Scientific and Cultural Organisation (UNESCO), Oberthur Rennes, Paris, France, Vol. XV, pp. 1–72. Florindo, F., Cooper, A. K., & O’Brien, P. E. (2003). Antarctic Cenozoic palaeoenvironments: Geologic record and models. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198(1–2), pp. 1–278. Gerson, N. C. (1958). From polar years to IGY. In: H. E. Landsberg, & J. Van Mieghem (Eds). Advances in Geophysics. Academic Press, New York, Vol. 5, pp. 1–52. Hambrey, M. J, Barrett, P. J. (1993). Cenozoic sedimentary and climatic record, Ross sea region, Antarctica. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Persepective on Global Change, Part 2. Antarctic Research Series. American Geophysical Union, pp. 91–124. Heathcote, N. de V., & Armitage, A. (1959). The First International Polar Year. In: Annals of the International Geophysical Year, Vol. 1. Pergamon Press: London, pp. 6–100. Korsmo, F. L. (2007). The genesis of the international geophysical year. Phys. Today, 60(7), 38–43. National Academies Archives (1958). IGY ‘‘Planet Earth’’ Poster and Booklet, http://www7.nationalacademies.org/archives/IGYPlanetEarthPosters.html, 1–44. Wood, K. R., & Overland, J. E. (2006). Climate lessons from the first International Polar Year. Bull. Am. Meteorological Soc., 86, 1685–1697.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00003-7
Chapter 3
A History of Antarctic Cenozoic Glaciation – View from the Margin Peter Barrett Antarctic Research Centre, Victoria University of Wellington, P.O. Box 600, Wellington, New Zealand
ABSTRACT The scale and antiquity of the Antarctic Ice Sheet was sensed from the time of earliest exploration a century ago. However, significant advances in scientific thinking, along with logistics and technology for gathering data from the continent itself, were required before a clear and consistent framework for ice-sheet history and behaviour could develop and this has emerged only in the last few years. The main features of the present ice sheet were established by over-snow traverses during and following the International Geophysical Year (1957–1958), but the timing and circumstances of its origin remained uncertain. Geological records of post-Jurassic time were largely buried under the ice or the sea floor around the Antarctic margin, though a few radiometric ages from the new K–Ar dating indicated Antarctic glaciation was likely older than the Northern Hemisphere ice ages of the Quaternary Period. New post-World War II techniques in offshore surveying with marine geophysics and ship-based drilling were first applied to the Antarctic margin in the early 1970s, and were immediately productive. The Antarctic continental shelf was found to be underlain by sedimentary basins with the promise of ice-sheet history, and in early 1973, cores from Leg 28 of the Deep Sea Drilling Project (DSDP) in the Ross Sea provided the first physical record of Antarctic glaciation extending back to Oligocene times. DSDP Leg 29 drilled in the Southern Ocean for deep-sea cores from the whole Cenozoic Era, yielding the first set of oxygen isotopic ratios (d18O) from benthic calcareous microfossils, and the first estimates of ice volume. These indicated a two-stage ice-sheet history (cooling and some ice
Corresponding author. Tel.: þ64 4 463 5336; Fax: þ64 4 463 6581;
E-mail:
[email protected] (P. Barrett).
34
P. Barrett
at 34 Ma, and persistent ice from 14 Ma – time-scale of Berggren et al., 1995). About the same time, the Dry Valleys Drilling Project (DVDP) was launched to recover onshore records of Antarctic Cenozoic climate history, and exploratory drilling from fast ice offshore soon followed. Cores from the Ocean Drilling Program in the 1980s from the Prydz Bay Shelf and the Kerguelen Plateau established the timing of the first continental ice sheet at 34 Ma. In the same period, deep continuous coring from sea ice in McMurdo Sound developed from DVDP technology succeeded in providing new detail for its subsequent history. Core recovery at B98% yielded lithological evidence of ice margin and sea level fluctuations implied by deep-sea isotope records. Further drilling with improved chronology in the 1990s yielded cores confirming ice margin and sea level changes on Milankovitch frequencies and on a scale of 10–40 m of sea level equivalent. Micropalaeontological and geochemical evidence pointed to a slight cooling from a coastal cold temperate climate, with beech forests during interglacial times. Subsequent development of ice-sheet modelling has indicated that most of the cooling that initiated ice-sheet glaciation was the consequence of a fall in atmospheric CO2 levels below a critical threshold, allowing ice sheets to form that were highly sensitive to orbital forcing. This claim has been supported by recent work on CO2 proxies and indicators of a shift in carbonate compensation depth in deep-sea sediments. Since the first measurements in the 1970s, deep-sea isotopic measurements have implied a significant increase in Antarctic ice volume at around 14 Ma that persisted to the present day. However, in the mid-1980s, marine diatoms in glacial deposits in the Transantarctic Mountains suggested periods in Pliocene times when seas invaded the East Antarctic interior, implying dynamic Antarctic Ice Sheets until Quaternary times. Evidence of continued cold in the mountains over the last 14 Ma, glaciological problems with the proposed over-riding scenario, lack of a signal in the deep-sea isotope record for the loss of most Antarctic ice in Pliocene times and possible alternative atmospheric sources for diatoms has shifted the weight of evidence in favour of persistent ice in the east Antarctic interior. However, coastal outcrops in Prydz Bay and a recent deep core from beneath the McMurdo Ice Shelf have shown that in the globally warmer Pliocene, notably around 3–5 Ma, seas around the Antarctic margin were several degrees warmer. Indeed, recent drill cores suggest that the Ross Embayment and perhaps also most of the West Antarctic interior were periodically ice-free in these times. Three decades ago, the Antarctic Ice Sheet was seen as a long-standing feature of the Earth with its origins in early Cenozoic times and its permanency assured by mid-Miocene cooling. Research in the last decade from geological drilling and glaciological remote sensing, supported by ice sheet and climate modelling, indicates the ice sheet is in fact quite responsive to changes in the global climate system, whether natural or human-induced, though at different rates in different sectors. Recent developments in both science and technology outlined here provide opportunities for projecting realistic scenarios for future ice-sheet response on human time scales.
A History of Antarctic Cenozoic Glaciation
35
3.1. Introduction This chapter traverses the growth in knowledge and understanding of Antarctic Ice Sheet history through the Cenozoic Era over the last 100 years and briefly considers the future. Important external influences on progress have been the advances in biostratigraphy and geochronology through the latter part of the last century, seismic stratigraphy and deep-sea isotopic studies world-wide over the last four decades, and the marked progress in modelling ice behaviour along with other components of the Earth system over the last two decades. Key events in this account are given in Table 3.1. Interest in the climate history of the Antarctic continent first developed with the curiosity-driven scientific expeditions from the Northern Hemisphere beginning in the late nineteenth century (the ‘‘Heroic Era’’, Fogg, 1992, p. 108ff). These included prominent explorers and scientists from Belgium, Britain, France, Germany, Japan, Norway and the United States and within two decades the extent and salient features of the present ice sheet and the continent beneath had been documented. These included: (i) An ice sheet of around 5 million square miles (13 million square kilometers) in area and rising to an elevation of at least 10,000 ft (3,000 m), along with shelf and sea ice. In terms of its history, it was thought most likely to have been more extensive in warmer Pliocene times, and to have originated in Miocene times (Taylor, 1922), though Wright and Priestley (1922, Table 17) noted Antarctic glaciation ‘‘apparently began there in Eocene or Oligocene times’’; (ii) Flat-lying ‘‘cover beds’’ similar to those in the Transantarctic Mountains (Beacon Sandstone of Late Paleozoic age), with fossil plants similar to those in India, South Africa and South America, indicating a former temperate climate in those times (Seward, 1914); (iii) An East Antarctic ‘‘shield’’, with a foundation of Precambrian rocks and a faulted Ross Sea margin comparable to that of eastern Australia. West Antarctica was seen geologically as more related to southern South America and linked to it through the Scotia Arc (David, 1914). The decades that followed saw extensive inland exploration largely by the Byrd expeditions (Fogg, 1992, pp. 134–146), with detailed observations of weather, snow and ice cover and some geological observations, but there was little advance in comprehending the basic geological history of the continent. However, there were significant advances in technology and logistics through the introduction of radio communication, seismic sounding, ships and aircraft. These advances prepared the way for US Operation Highjump
36
P. Barrett
Table 3.1: Events in progress in the evolution of antarctic glacial history. Date 1910–1912
1928–1947 1957–1964
1964–1970
1972
1970–1973
1973–1975
1978 1978–1980
1982 1984
Event Scale of Antarctic Ice Sheet recognized. Simple history proposed. Origin thought likely Miocene but possibly Eocene–Oligocene Inland exploration/technology development by Byrd expeditions International Geophysical Year (IGY) supports extensive geophysical exploration but still the lack of geological record between Jurassic and late Quaternary time. Glacial features mapped in McMurdo region thought Quaternary Miocene radiometric ages indicate pre-Quaternary glaciation a reality. Oligocene ice rafting in Southern Ocean Pre-Quaternary history thought likely from Northern Hemisphere record Recognition of widespread pre-Quaternary glacial deposits (Sirius Formation) on land in high Transantarctic Mountains Early seismic surveys of continental shelf by USNS Eltanin First drilling on shelf, showing Antarctic continental glaciation as old as 25 Ma (DSDP Leg 28) Drilling in deep-sea floor showing cooling and first ice sheet at 34 Ma and present ice sheet dating from B14 Ma (DSDP Leg 29) First onshore scientific drilling (McMurdo Dry Valleys) leading to late Neogene glacial record going back at least 4 Ma in Taylor Valley. First attempt at drilling from floating ice (DVDP 15) Collapse of West Antarctic Ice Sheet projected from rising CO2 First hole drilled through the Ross Ice Shelf (J9). Sea floor cores recover early Miocene diamict with diatomite clasts and pollen Early simple Antarctic Ice Sheet model Marine diatoms in glacial deposits of high Transantarctic Mountains (Sirius Fm) suggest East Antarctic interior seas B3 Ma ago Results from first extensive multichannel seismic survey of the Antarctic margin (R/V S.P. Lee, Ross Sea and Wilkes Land)
References Wright and Priestley (1922), Taylor (1914, 1922) Fogg (1992) Pe´we´ (1962), Bull et al. (1962), Nichols (1964)
Craddock et al. (1964), Margolis and Kennett (1970), Flint (1971)
Mercer (1972), Mayewski (1975) Houtz and Meijer (1970) Hayes, Frakes et al. (1975)
Kennett, Houtz et al. (1975) Smith (1981), Torii (1981)
Mercer (1978) Clough and Hansen (1979), Webb et al. (1979) Oerlemans (1982) Webb et al. (1984), Harwood (1986)
Cooper and Davey (1985), Eittrem and Hampton (1987)
A History of Antarctic Cenozoic Glaciation
37
Table 3.1: (Continued ). Date 1986–1988
1987 1986–1996
1988–2002
1993
1997–1999
2000 2001
2003
2004
2005
2006–2007
Event New drilling results from Ross Sea (CIROS) and Prydz Bay (ODP Leg 119) show cyclicity in early history and ice-sheet antiquity Vostok ice core provides climate record through last glacial cycle Formation of the SCAR Group of Specialists on the Evolution of Cenozoic Palaeoenvironments of the Southern High Latitudes Formation of ANTOSTRAT – Antarctic Offshore Acoustic Stratigraphy to coordinate Antarctic margin studies Ice-sheet model and deep-sea isotope record inconsistent with major Pliocene deglaciation Pre-mid-Miocene age for Sirius deposits proposed on basis of ancient ash and persistent cold in high Transantarctic Mountains Cape Roberts cores yield high-resolution record of significant orbitally forced fluctuations in Antarctic Ice Sheet and sea level from 34 to 17 Ma and slight cooling over that period New drilling results from Prydz Bay (ODP Leg 188) Erice workshop reviews ANTOSTRAT and proposes new group to integrate geophysical/ geological data on Antarctic glacial history and the new generation of coupled ice– ocean–atmosphere models (ACE) Coupled ocean–atmosphere–ice-sheet model shows major role for CO2 in early ice-sheet formation EPICA yields Dome C climate record through last 8 glacial cycles Formation of ACE as a SCAR Scientific Research Program Landscape evolution model for Lambert drainage basin shows similar fluvial and glacial influence ANDRILL results provide first continuous record of late Antarctic Cenozoic history, including first proximal record of midMiocene transition, and evidence of ice-free Ross embayment 3–4 Ma ago
References Barrett (1989), Barron, Larsen et al. (1989) Barnola et al. (1987) Webb (1990)
Cooper et al. (2002, this volume) Huybrechts (1993), Kennett and Hodell (1993) Sugden et al. (1993)
Naish et al. (2001), Barrett (2007)
O’Brein et al. (2001) Cooper et al. (2002)
DeConto and Pollard (2003) EPICA (2004) Siegert et al. (2004) Jamieson et al. (2005)
Naish et al. (2008a), Harwood et al. (2003)
38
P. Barrett
immediately following World War II. Although largely a polar training exercise, it provided much of the technological foundation for the International Geophysical Year (IGY) in Antarctica (1957–1958).
3.2. Mid-Twentieth Century Advances (1956–1972) IGY saw the establishment of a network of permanent bases around the continent from which to execute ship-based surveys of the margin, along with stations in the interior (Byrd, South Pole, Charcot and Vostok), enabling the first great scientific over-snow traverses gathering geophysical and glaciological data across the continent (Fogg, 1992, pp. 168–176). The technical and scientific success of IGY led an agreement to extend the research that had been initiated and the establishment of a Special (later Scientific) Committee on Antarctic Research (SCAR) in 1958 to facilitate and help coordinate these activities. The success of IGY also led participating nations in the years immediately following to craft an Antarctic Treaty, which was ratified in 1961 (Beeby, 1972). Though geology was subordinate to the main mission of IGY, geologists enthusiastically took advantage of the extensive access to the continent made possible by their geophysical and glaciological colleagues. Geological surveys by contrast were carried out by smaller parties visiting the numerous but limited exposed patches of rock around the continent. By this time, there had been significant advances in the Earth sciences since the Heroic Era, and the application of this new knowledge to the Earth’s only polar continent led to great interest in reports from the region. These advances included a robust geological time scale through radio-isotopic dating (Holmes, 1965), and the different character of the crust beneath oceans and continents recognized from seismology, along with the concept of continental drift from geology (Du Toit, 1937; Carey, 1958), later embodied in the theories of Sea Floor Spreading and Plate Tectonics (Dietz, 1961; Hess, 1962; Wilson, 1965). By 1970, the comprehensive surveys carried out in the IGY and the decade following had resulted in a comprehensive new view of the climate, physiography, glaciology, geology and biota of the Antarctic continent, summarized in the American Geographical Society’s map folio series (Bushnell, 1964–1975). Antarctic sedimentary strata overlying basement rocks were found to record a climate history similar to that of the other Gondwana continents, with dry warm conditions in Devonian times (390–340 Ma), ice-sheet glaciation in Carboniferous–early Permian times (340–280 Ma) and a humid temperate climate in Permian–Triassic times (280–200 Ma) with evidence of rivers, lakes and coal swamps. However, this long sedimentary record of deposition ended
A History of Antarctic Cenozoic Glaciation
39
around 180 Ma ago with Jurassic igneous activity associated with the break-up of the Gondwana supercontinent, leaving a huge gap in the Antarctic geological record until the deposition of moraines around the Antarctic margin, then presumed to have formed from ice sheets in the Quaternary Period (Harrington, 1965). Early post-IGY surveys of glacial deposits in the McMurdo Sound region by Pewe (1960), Bull et al. (1962), Nichols (1964) and others described the features they found in terms of four glacial episodes within the Quaternary Period, perhaps influenced by the prevailing view prior to the IGY in the Northern Hemisphere (Flint, 1957) and a lack of suitable fossil material or radiometric techniques for a sound chronology. Indeed, Nichols (1964) in his review of the status of Antarctic glacial geology stated, ‘‘This writer believes that as yet there is no good evidence for Tertiary Antarctic glaciation’’. In the years that followed, K–Ar dating revealed a significant preQuaternary history for Antarctic Cenozoic ice sheets. In the Jones Mountains in West Antarctica, a glaciated surface overlain by basalt was dated at more than 10 Ma (Craddock et al., 1964) and in Taylor Valley west of McMurdo Sound, small basaltic cones overlain and underlain by evidence of glaciation yielded ages ranging between 2.8 and 3.6 Ma (Armstrong et al., 1968; Denton et al., 1970). Post-IGY research world-wide was reporting evidence of pre-Quaternary glaciers in the Arctic and of late Cenozoic cooling from fossil molluscs and plants in lower latitudes, leading Flint (1971, p. 441) to conclude, ‘‘Glaciation occurred in the Miocene and Pliocene as well as in the Quaternary. Cold periods were more numerous than the four periods of the classical literature’’. Cores from the floor of the Southern Ocean suggested that Antarctic ice could have been much older, with icerafted sand grains as old as Oligocene (Margolis and Kennett, 1970). But there was no way of showing whether these came from local ice caps or continental ice sheets.
3.3. First Antarctic Drilling (1972–1975) The period following World War II was marked by an expansion in both the scientific exploration of the oceans and the search for oil on the world’s continental shelves. Echo-sounding techniques developed in war time were applied to marine seismic surveys for science and industry, and ship-based drilling for oil offshore attracted the attention of scientists for drilling through the Earth’s crust. By the mid-1960s, the main features and gross chronology of most of the world’s continental shelves were known by the oil
40
P. Barrett
industry, and the US marine science community persuaded its government to support the development of an international Deep Sea Drilling Project (DSDP) to drill the world’s ocean floor, beginning in 1968. The main aims were to test and refine the new theory of Plate Tectonics as well as gain new knowledge from this unknown 70% of the Earth’s surface that might improve our knowledge of Earth history. DSDP was remarkably successful and in its first 4 years had drilled at over 200 sites around the world but not in the Antarctic region. The first venture south was Leg 28, designed to carry out the first drilling transect between Australia and Antarctica for dating the timing of continental separation from the age of sediment resting on basaltic ocean floor, and to core for climate and ocean history in the Southern Ocean and on the Ross continental shelf. The Ross Sea region had just been surveyed by the US research vessel USNS Eltanin during which several thousand kilometres of single channel seismic surveys were taken and used to define the three main shelf basins (Houtz and Meijer, 1970; Houtz and Davey, 1973). The Eastern Basin was most promising with dipping strata at the basin margin, allowing a drill ship with a subsea floor penetration limit of around 500 m, to core most of the sedimentary section in three well-chosen sites. 3.3.1. Ship-Based Drilling (1972–1973) The Glomar Challenger sailed south on DSDP Leg 28 from Fremantle, Australia, in December 1972. In the 2 months that followed, it succeeded both to establish the history of continental separation south of Australia and to provide the first physical record of Antarctic glaciation extending back to Oligocene times (Hayes, Frakes et al., 1975). The three sites drilled in the eastern Ross Sea (Figs. 3.1 and 3.2) revealed large thicknesses of poorly sorted glacial marine debris ranging in age from late Oligocene (B25 Ma) to Quaternary. The oldest glacial sediments were cored at DSDP Site 270, where they overlay a palaeosol developed on a calc-silicate gneiss basement. DSDP Leg 29, which followed in March 1973, added a new and independent body of data with a significant bearing on Antarctic glacial history. The goal of Leg 29 was to sample the floor of the deep ocean south of Australia and New Zealand for as complete a record as possible of strata representing the whole Cenozoic Era to provide data on ocean history for the new field of palaeoceanography. This venture was also successful, both in recovering long and representative stratigraphic sequences and in finding
A History of Antarctic Cenozoic Glaciation
41
Figure 3.1: Map showing the Antarctic continent and ice drainage systems (based on Drewry, 1983), with sites of the first drill holes on the Antarctic continental shelf (DSDP Leg 28). Subsequent shelf sites in the McMurdo Sound region (box – see Fig. 3.3 for detail), Prydz Bay and the Antarctic Peninsula are also shown, each reflecting the history of the ice sheet in their respective regions. DSDP Leg 29 drilled lower latitude deep-sea sites south of New Zealand for the first Cenozoic isotopic record of ice volume and temperature (see text). Modified from Barrett (1999), with permission. well-preserved calcareous microfossils for applying a new technique, stable isotope analysis, to these cores (Kennett, Houtz et al., 1975). The result was the first record of warm early Cenozoic times followed by a decline in temperature from around early mid-Eocene times (B50 Ma ago,
42
P. Barrett
Figure 3.2: R/V Glomar Challenger at Site 270 (781S) in January 1973. Seismic section shows how three holes sampled most of the sedimentary section (De Santis et al., 1999).
Shackleton and Kennett, 1975). Two features in the isotope record stand out and are still seen as significant today: (i) An abrupt increase in d18O in earliest Oligocene times (now 34 Ma), interpreted as a consequence of global cooling, with extensive sea-ice formation around Antarctica. A major global climate event at this time had already been suspected from the change in terrestrial flora around the world. (ii) Another increase in d18O at B14 Ma, and interpreted to be the result of development of a larger relatively stable Antarctic Ice Sheet like that of today.
A History of Antarctic Cenozoic Glaciation
43
Although deep ocean sediments were useful for their continuous record of past ocean chemistry, and in providing an ice volume–temperature signal from oxygen isotopes, they could not provide information on the extent of ice or regional climate in the high latitudes. This could only come out of sediment cores from the Antarctic margin, where the direct influence of ice advance and retreat (and perhaps also sea level fall and rise) could be obtained. 3.3.2. Land-Based Drilling (1973–1974) While DSDP was planning its first Antarctic and Southern Ocean cruises, the US Office of Polar Programs was also decided to address the B180 million years gap in Antarctic geological history, as well as develop a more collaborative approach to major Antarctic science problems. This decision resulted in the launch of the Dry Valley Drilling Project (DVDP), an initiative sponsored by the United States, Japan and New Zealand to explore the late Cenozoic history of the McMurdo Dry Valleys (McGinnis, 1981; Smith, 1981), using a standard ‘‘slim-hole’’ diamond drilling system widely used in mineral exploration. The rig selected was a ‘‘Longyear 44’’, designed to take continuous core of 85, 64 or 48 mm diameter to depths of around 500 m. In the first two field seasons (1973–1974 and 1974–1975), the project drilled 3 test holes at McMurdo Station, 1 over 300 m deep, and 11 holes in the McMurdo Dry Valleys (Fig. 3.3; Torii, 1981). The most successful of these recovered 320 m of interbedded glacial and interglacial fiordal sediments of late Miocene to recent age in the lower Taylor Dry Valley (Powell, 1981; Webb and Wrenn, 1982), indicating several advances and retreats of the ice in that time. Wrenn and Webb (1982) used microfossils in the deposits to provide approximate ages on several geomorphic surfaces in the lower Dry Valleys. 3.3.3. Drilling from Sea Ice (1975) The success of DSDP Leg 28 in the Ross Sea had shown the potential for offshore sediments around Antarctica providing a direct record of the glacial history on land. However, ice flow lines and lithologies of pebbles in cores from the sites in eastern Ross Sea indicated that sediments deposited there were derived from West Antarctica (Fig. 3.1; Barrett, 1975). To find a record of the much larger and possibly older East Antarctic Ice Sheet, sites were needed in the western Ross Sea adjacent to the Victoria Land coast.
Figure 3.3: Map and cross-section of McMurdo Sound (adapted from Naish et al., 2008a). (A) The map shows sites for the Dry Valley Drilling Project (1973–1975) and subsequent offshore drilling (MSSTS-1 in 1979, CIROS 2 and 1 in 1984 and 1986, respectively, CRP-1, -2 and -3 in 1997, 1998 and 1999, and ANDRILL MIS and SMS in 2006 and 2007 – details in Table 3.2). Contour interval 200 m. Transantarctic Mountains exposures in light brown, McMurdo Volcanics in dark brown. (B) The cross-section across McMurdo sound shows how a few key drill holes have sampled the entire Cenozoic record from this section of the basin. CRP sites provide an expanded and better dated record 70 km north of the Late Eocene–Early Miocene section cored by CIROS-1.
A History of Antarctic Cenozoic Glaciation
45
A seismic survey in late 1974 showed that the thick sedimentary strata of the Victoria Land Basin did indeed extend close to the East Antarctic coast (Wong and Christoffel, 1981) and provided the justification for DVDP15, the first Antarctic offshore hole to be drilled from the fast sea ice. Experience in moving equipment 70 km across the sea ice from McMurdo Station on Ross Island to the Marble Point airstrip on the Victoria Land coast had been gained from fuel train traverses, and the same logistics were employed in transporting the DVDP rig and related equipment from McMurdo Station to the site selected for DVDP-15 around 10 km off Marble Point. The rig was set up on 2 m of sea ice in 120 m of water and began drilling in early November 1975 (Fig. 3.4). Logistic delays and drilling problems reduced the time available for drilling, which was terminated at a depth of 62 m with over 52% recovery. The character of the sediments was surprising, for below 12 m of recent glacial debris lay 50 m of basaltic sand, presumed to have been erupted as volcanic ash and then wind-blown across the sea ice before settling (Barrett and Treves, 1981). Nevertheless, the experience had shown the feasibility of drilling from a sea-ice platform.
Figure 3.4: Development of sea-ice drilling systems from 1975 to 1999. (A) DVDP-15, 1975. Water depth 112 m; depth Cored 62 m bsf (below sea floor). (B) MSSTS-1, 1979. Water depth 195 m; Depth Cored 229 m bsf. (C) CIROS-1, 1986. Water depth 205 m; depth Cored 702 m bsf. (D) CRP-3, 1999. Water depth 305 m; depth Cored 942 m bsf.
46
P. Barrett
3.4. Developments in Drilling and Thinking in the Late 1970s Although DVDP had concluded its work, the NZ Antarctic Programme was persuaded that a further effort to drill in western McMurdo Sound was justified. This was named the McMurdo Sound Sediment and Tectonic Studies Project (MSSTS), and led by New Zealand but with US and Japanese scientific participation. The United States provided the ‘‘Longyear 44’’ rig from DVDP and significant logistic support. A camp for the MSSTS-1 hole was set up on the sea ice soon after the late August winter flight to the ice, allowing the drilling to begin in October 1979. Despite the cold and difficult conditions, core was recovered to a depth of 227 m below the sea floor, when operational problems terminated drilling. After passing through a few tens of metres of Plio-Pleistocene strata, and an interval of virtually no recovery, the strata below 110 m bsf cored well. They provided a record of striking facies changes between diamictites, sand and mud, along with a pollen record and a beech leaf indicating a vegetated coastline and a cold temperate continental margin in those times. Both lithofacies and biofacies were seen as reflecting cyclic changes in extent of the continental ice-sheet margin advancing and retreating across the drill site, with associated changes in sea level by tens of metres through glacioeustacy (Barrett and McKelvey, 1986; Barrett et al., 1987). Initially, the strata were thought to extend back to Paleocene times (Webb, 1983) but further study of the sparse faunal and floral assemblages and recognition of reworked older microfossils led to a late Oligocene age assignment (Webb et al., 1986). While the NZ programme had focussed on drilling in McMurdo Sound, the US programme developed a project to core through the Ross Ice Shelf 420 km from the ocean (Clough and Hansen, 1979). The Ross Ice Shelf Project (RISP) in two successive seasons drilled through 430 m of ice to measure and sample the properties of the 230 m water column, and take cores and photographs of the sea floor beneath (Webb, 1978, 1979). The sea floor cores revealed a few tens of centimetres of Late Quaternary mud overlying a metre of midMiocene glaciomarine mudstone with diatomite clasts several millimetres across indicating an interglacial period of ice- and sediment-free biogenic sedimentation at 821S (Webb et al., 1979; Scherer et al., 1988). Terrestrial palynomorphs from the clasts indicate coastal beech forests at this time also. The MSSTS-1 results also provided a glimpse of Antarctic glacial history from the Ross Sea margin, and although there were some issues to be resolved with the drilling technology, scientific advances in the wider world were supporting the case for a better physical record of past Antarctic climatic events. These included extensive records of seismic stratigraphy from continental margins, along with a new type of analysis that recognized
A History of Antarctic Cenozoic Glaciation
47
sequences developed as a consequence of coastal advance and retreat (Vail et al., 1977). These coastal movements were taken to imply variations in ice volume causing sea level changes of hundreds of metres, and mainly from the mid-Oligocene on. At the same time, the newly developed deep-sea isotope record provided a different basis for inferring ice-sheets on Antarctica, and these indicated an earliest Oligocene initiation (Kennett, 1977). Kennett (1982) incorporated these new concepts and insights from sequence and isotope stratigraphy into a comprehensive synthesis of history and knowledge of the Earth as a system, albeit with a focus on the marine realm, an approach that is now widely accepted and practised. Neither isotopes nor sequence stratigraphy could provide direct evidence of the behaviour and history of the Antarctic Ice Sheet itself, but the first coastal Oligocene cores, from MSSTS-1 in 1979, confirmed the cyclic behaviour of ice sheet and sea level (Barrett et al., 1987). Core chronology was not yet adequate to constrain their frequency, but the role of orbital forcing in driving highfrequency Quaternary glacial cycles (Hays et al., 1976) was plainly relevant to early Antarctic Ice Sheets. While significant progress was being made in documenting past ice-sheet behaviour, a key event of the 1970s was the publication of John Mercer’s hypothesis for the likely future behaviour of the Antarctic Ice Sheet as a consequence of rising CO2 emissions (Mercer, 1978). He observed that a large area of West Antarctica lay below sea level, the ice sheet being thus inherently unstable, and that much of the ice-sheet margin was buttressed by ice shelves. Observing that present-day ice shelves form only where the January summer isotherm is below 01C, he concluded that a rise of 5–101C, projected as likely in the following 50 years or so, would be sufficient to cause the disintegration of the major ice shelves buttressing the West Antarctic Ice Sheet leading to disintegration of the ice sheet itself. Although there was significant disagreement among glaciologists on both the buttressing hypothesis and the immediacy of the threat, the hypothesis provided new impetus for seeking a geological record of the history of both East and West Antarctic Ice Sheets.
3.5. Discoveries Offshore and on the Continent in the 1980s 3.5.1. Further Sea-Ice Drilling – The CIROS Project From 1977 on, the SCAR Working Group on Geology, under the chairmanship of Cam Craddock, provided a useful forum for discussion
48
P. Barrett
and planning further drilling in McMurdo Sound in an attempt to provide records from the Antarctic margin to resolve the chronology of early icesheet formation and its subsequent evolution. Meetings with its companion group on geophysics were also important for discussions on marine geophysical surveys being planned for the Antarctic margin. The survey by S.P. Lee was particularly useful for future drilling in providing the first multichannel section of the western margin of the Victoria Land Basin (Cooper and Davey, 1985; Cooper et al., 1987) and showing a thick seawarddipping sequence that extended close enough to the coast to be drilled from the fast ice. Discussions at the SCAR meeting in Queenstown (Barrett and Webb, 1981) led to a proposal for Cenozoic investigations in the western Ross Sea (CIROS) to drill four holes in two successive seasons from the sea ice off New Harbour and off Granite Harbour (Barrett, 1982), based on the seismic surveys of Wong and Christoffel (1981) and Cooper et al. (1987). The project logistics were managed by the NZ Antarctic Programme with significant US support and scientific participation, and the drill rig and camp were set up in late 1984, 12 km off Marble Point (Fig. 3.3), for the first drill hole (CIROS-1). However, concerns grew about the poor state of the sea ice and just prior to drilling, it was decided to shift the camp and rig 20 km inshore to the CIROS-2 site in Ferrar Fiord. After a number of set-backs, CIROS-2 was successfully drilled 165 m below the sea floor to a granitoid basement, recovering several cycles of Quaternary black sand and diamict and Pliocene mudstone and diamictite. The sequence recorded a number of cycles of glacial advance from the inland ice in the Pliocene followed by cycles in which a Ross Sea ice sheet flowed westward into the Fiord (Barrett and Hambrey, 1992). The core had in fact verified the findings from DVDP-11 in 1973, the drilling system had worked and valuable experience had been gained. 3.5.2. Discoveries in the Transantarctic Mountains – The Sirius Formation/Group On land, wet-based glacial deposits at high elevations in the Transantarctic Mountains were providing new evidence of past Antarctic climate. The deposits had been found in the late 1960s, and named the Sirius Formation by Mercer (1972), who saw them as the products of temperate glaciation pre-dating the present ice sheet. However, Mayewski (1975) mapped and analysed more of these deposits, and thought they more likely represented a pre-Pliocene over-riding phase of the East Antarctic Ice Sheet. Later in the
A History of Antarctic Cenozoic Glaciation
49
mid-1980s, yet another view was proposed by Webb et al. (1984) as the result of a laboratory study of samples collected earlier by Mercer, Mayewski and others and reported in detail by Harwood (1986). In this study, samples had been processed from tens of localities along the mountains. Some were barren but others yielded a variety of types and ages of microfossils with some as old as late Cretaceous and a few marine diatoms as young as Pliocene. The latter fossils were thought to provide a maximum age for the Sirius Formation, having been deposited in East Antarctic interior seas and subsequently transported by a growing ice sheet flowed through the mountains at levels more than 1,000 m higher than today’s outlet glaciers. The discovery of plant material, and later leaves and stems, in the deposits (Webb and Harwood, 1987, 1993; Francis and Hill, 1996; Hill et al., 1996) indicated a mean annual temperature almost 201C warmer than today. Mercer’s (1978) case for a collapse of the West Antarctic Ice Sheet through a doubling of atmospheric CO2 had heightened interest in ice-sheet behaviour, but the loss of the East Antarctic Ice Sheet in recent geological times seemed less credible to many, who queried the age of the Sirius Group from recycled diatoms. Indeed, some (e.g. Clapperton and Sugden, 1990) queried the age of the diatoms themselves. However, the Pliocene age of the diatoms was confirmed by dating a volcanic ash in mudstone with the same taxa and cored in CIROS-2 (Barrett et al., 1992).
3.5.3. CIROS-1 – Dynamic Oligocene Ice Sheets Offshore, the most successful drilling for the elusive early Cenozoic preglacial record took place in late 1986. The CIROS-1 drill hole cored to 702 m below the sea floor with 98% recovery. The core reached back to the late Eocene and spanned the period to the early Miocene, based on magnetostratigraphy underpinned by biostratigraphic datums largely from diatoms, but including foraminifera, palynomorphs and nannofossils (Harwood et al., 1989; Hannah et al., 1997; Wilson et al., 1998; Roberts et al., 2003). However, the record did not quite reach the warm middle Eocene, despite coring the early Oligocene–late Eocene transition in the lower half of the core. The strata were a deep-water turbidite facies with lonestones, indicating some glacial influence. In contrast, the late Oligocene– early Miocene upper half comprised sequences of diamictite, sand and mud (Hambrey et al., 1989), seen in a much smaller way in MSSTS-1, and typical of near-shore glaciomarine sedimentation with cyclic ice margin advance and retreat and corresponding changes in sea level. The depositional patterns
50
P. Barrett
clearly showed Antarctic Ice Sheets to be dynamic and reaching the Ross Sea coast from at least late Oligocene times onwards. 3.5.4. Ice-Sheet Proxies from Seismic Stratigraphy and Deep-Sea Isotopes As the results of the MSSTS-1 and CIROS-1 records were being studied, two significant papers were published. One was by Haq et al. (1987), a better documented update of the Vail et al. (1977) monograph with a more detailed sea level curve based on coastal advance and retreat histories extracted from seismic imaging of continental margins around the world. It still showed a sharp and substantial mid-Oligocene shallowing, ostensibly from Antarctic Ice Sheet formation, along with some isolated shoaling events in the Paleogene and Cretaceous. The paper emphasized the cyclic sea level changes with a frequency of 1–3 million years, third-order cycles, inferred to be glacioeustatic. This link between ice sheets and sea level was invoked to explain cyclic patterns in Cenozoic strata in Australasia by Loutit and Kennett (1981). The other paper was a compilation of deep-sea isotope data from numerous DSDP and Ocean Drilling Program (ODP) sites in the Pacific and Atlantic Oceans by Miller et al. (1987) – the smoothed curves for both oceans were a little different in detail but their main features included both the earliest Oligocene and middle Miocene positive isotope shifts first seen by Shackleton and Kennett (1975). At the time, no cyclicity comparable to that proposed by Haq et al. (1987) was recognized in the deep-sea isotope record, but the record plainly indicated continental ice build-up in the earliest Oligocene, not the mid-Oligocene indicated by the Haq et al. (1987) analysis. A comparison of the two curves is shown in Fig. 3.5. Which was correct? CIROS-1 cores had the oldest significant diamictites in the late Oligocene, but the Transantarctic Mountains, with their origins around 55 Ma (Gleadow and Fitzgerald, 1987), could well have been a barrier to earlier continental ice. 3.5.5. DSDP–ODP Drilling Dates Initiation of Antarctic Ice Sheet By this time, the ODP had approved two further Antarctic legs, one in the Weddell Sea (DSDP 113) and one in Prydz Bay (ODP 119), seeking a glacial history from other sectors of the Antarctic margin. The Weddell Sea leg cored thick continuous Paleogene deep-sea sediments between 601 and 711S, recording a relatively warm climate, cooling sufficiently for cold deep water formation and some limited ice to form on land in the early Oligocene (Barker, Kennett et al., 1988). However, ODP 119 was able to drill several
A History of Antarctic Cenozoic Glaciation
51
Figure 3.5: A comparison of the deep-sea isotope record of Miller et al. (1987), the eustatic sea level curve of Haq et al. (1987), and the record from cores from the Eastern and Victoria Land Basins and Prydz Bay (from Barrett and Davey, 1992). Both the isotope record and the eustatic sea level curve were considered by their proponents to be sound Ice volume proxies, but showed significant differences in early Cenozoic times. Reproduced from Barrett and Davey (1992), with permission from the Royal Society of New Zealand.
52
P. Barrett
shelf sites in Prydz Bay, coring offshore-dipping strata, ranging from Cretaceous terrestrial sediments near the coast to thick late Eocene–early Oligocene diamictite mid-shelf to Neogene diamictites in the outer shelf (Hambrey et al., 1991). No better age has been extracted from these fossilpoor sediments, but the result was sufficient to make it clear that continentalscale ice sheets on Antarctica were delivering ice to the shelf edge beyond the limits of the modern ice sheet at least from early Oligocene times. ODP Leg 120 that followed was able to both confirm this physically through the occurrence of ice-rafted debris at the Eocene–Oligocene boundary and provide a better deep-sea isotope record from continuous Paleogene cores from the edge of the Kerguelen Plateau to the north (Zachos et al., 1992). The timing of the first of the big Antarctic-wide ice sheets had now been established (Wise et al., 1991). 3.5.6. Planning Future Research through SCAR The 1980s also saw work on Antarctic glacial history further facilitated by SCAR, initially through a Workshop convened by Peter Webb at the SCAR meeting in San Diego (CA) in 1986 on Cenozoic Palaeoenvironments in Southern High Latitudes. From that meeting emerged a plan for a Group of Specialists to coordinate and promote research on this broad topic. This was approved by SCAR and the Group began work with a workshop on polar drilling in Columbus, OH, in 1988. This meeting reviewed a range of drilling technologies from shelf ice, sea ice and ships, as well as shallow ship-based coring. Extensive ship-based coring programmes from the 1960s using the USNS Eltanin, USCGC Glacier and more recently the R/V N.B. Palmer provided some hard won data, but typically could not penetrate beyond the hard diamictites of the Last Glacial Maximum (with rare notable exceptions, such as the Cretaceous strata cored in Ross Sea by Domack et al., 1980). Webb (1990) provides a useful review of the state of knowledge of Antarctic glacial history at this time. At the same 1986 San Diego meeting, there was also interest in the growing body of seismic data being collected around the Antarctic margin, because of both its scientific value for Antarctic glacial history and also the political sensitivity to possible misuse. This led to the formation of the ancillary Antarctic Offshore Stratigraphy (ANTOSTRAT) Project, led by Alan Cooper (see Chapter 5). By 1990, the project had five regional working groups to organize seismic data from the five regions of the Antarctic margin, and make it available through a Seismic Data Library System (Cooper and Brancolini, 1997). However, the main goal was to use these data
A History of Antarctic Cenozoic Glaciation
53
as the basis for coring proposals in key areas in each region in the next decade under the aegis of the ODP. 3.5.7. Insights from Deep Ice Cores While the Antarctic drilling community was now focussed on a record of icesheet history going back to the Eocene times, the ice coring community was providing a remarkable climate record of the last glacial cycle with measurements of the atmospheric gas composition and proxy temperature history from the Vostok core (Barnola et al., 1987). The saw-tooth pattern of the glacial cycle, developing slowly (tens of thousands of years) with irregularities from varied orbital forcing, and deglaciation taking place relatively quickly (thousands of years), showed a close association with eustatic sea level change for the same period (Fairbanks, 1989). This gave new impetus to the importance of understanding Antarctic Cenozoic glacial history in order to appreciate its likely response to rising CO2 emissions, which by then was coming to be seen by senior climate scientists as an imminent danger (e.g. Hansen et al., 1988; Schneider, 1989).
3.6. Advances in the 1990s The 1990s saw the publication of several major reviews on the state of knowledge of Antarctic climate history, notably symposium volumes by Kennett and Warnke (1992, 1993), ANTOSTRAT monographs on the Ross Sea (Cooper, this volume) and in support of ODP drilling, three chapters in Tingey’s (1991) monograph on the geology of Antarctica (Anderson, 1991; Denton et al., 1991; McKelvey, 1991) and the review of sedimentation on the Antarctic continental shelf by Hambrey et al. (1992). By now, the broad chronological framework had been established through drilling in McMurdo Sound (CIROS-1, Barrett, 1989) and in Prydz Bay (Barron, Larsen et al., 1989) on opposite sides of the continent. However, neither the earliest Oligocene onset of glaciation nor the middle Miocene transition had been sampled by drilling on the continental shelf, and progress was hampered by a lack of chronological tools comparable with those used to date deep-sea sediments (e.g. microfossil datums based on abundant rapidly evolving taxa integrated with magnetic reversal stratigraphy in continuous sedimentary sequences). ANTOSTRAT contributed to a Detailed Planning Group of the ODP at College Station (TX) in 1994 for planning proposals for all five ANTOSTRAT sectors. However, in the event, only two legs were approved
54
P. Barrett
for drilling (Legs 178 in 1998 and Leg 188 in 2000). Both had some success in filling out the glacial chronology in each region in cores from fans and drifts on the slope and rise (Barker, Camerlenghi et al., 1999; O’Brien, Cooper et al., 2001; Cooper, this volume), though poor recovery marred the success of drilling on the continental shelf. 3.6.1. Antarctic Sea-Ice Drilling Progress in further sea-ice drilling began slowly with a SCAR workshop in Bremerhaven in 1990 for discussions that led to the formation of SCAR Group of Specialists on Global Change (GLOCHANT). While Cenozoic glacial history was seen as beyond the remit of the group, the meeting was productive in that it began a discussion among US, NZ, Italian, German and UK scientists for a workshop in Wellington (Barrett and Davey, 1992) that initiated the Cape Roberts Project (CRP). Following the meeting, national programmes agreed to plan for drilling four holes in two seasons to core a 1,500 m sequence that was thought from seismic correlation to extend the CIROS-1 record back to the Cretaceous. A new drilling system was also needed (Fig. 3.6).
Figure 3.6: Diagram showing the advances made in sea riser design after the first decade of McMurdo Sound drilling for the Cape Roberts Project, and the further developments that were required for ice-shelf drilling (Modified from Harwood et al., 2003, with permission from the ANDRILL Science Management Office).
A History of Antarctic Cenozoic Glaciation
55
The Cape Roberts Project sites were drilled, after a year’s delay on account of poor ice conditions, and with three drilling seasons (1997–1998 to 1999–2000, reported in CRST (Cape Roberts Science Team) (1998, 1999, 2000) and Davey et al. (2001). The 1,500 m-thick section cored provided the first proximal high-resolution record from the Antarctic continental shelf of climate history for the period from 34 to 17 Ma. A combination of biomagnetostratigraphic dating with Ar–Ar radiometric ages at several key points provided robust (o0.5 million years) resolution for most of the interval and excellent resolution (o0.1 million years) for three cycles around 24 Ma ago (latest Oligocene) (Florindo et al., 2005, but see update by Naish et al., 2008b). These showed for the first time that the Antarctic Ice Sheet was responding to orbital forcing with 40,000 and 100,000 year frequencies in the distant past (Naish et al., 2001). The cores also recorded the influence of orbital forcing on Antarctic Ice Sheets through Oligocene and early Miocene times as well as persistent slight cooling over this period (Barrett, 2007; Dunbar et al., 2008). This contrasted with the late Oligocene warming deduced from a review of deep-sea oxygen isotope data by Zachos et al. (2001), but was consistent with the reinterpretation of oxygen isotope data sets for this period by Pekar et al. (2006). 3.6.2. Ice Sheet and Climate Modelling ANTOSTRAT in a few short years had greatly extended the significance and value of the few drill holes that could be drilled on the Antarctic margin by providing a basis for correlating events from basin to basin. However, the behaviour of the ice that eroded and deposited the sediments was not well understood, especially on a continental scale. It was assumed that on a roughly circular continent, the extent of the ice would be similar in all directions, but Antarctica has enough irregularities, mountains and basins to see that local geological histories could vary. Ice-sheet modelling provided a means of objectively experimenting on and visualizing ice-sheet behaviour, Oerlemans (1982) providing an early example. Subsequent ice-sheet modelling by Huybrechts (1993) (Fig. 3.7) was significant because it showed a clear and consistent relationship between temperature and ice volume, including a slight increase in ice mass with initial warming, thus linking past Antarctic temperatures and sea level change. It indicated the loss of West Antarctic ice with a regional warming of 101C, in the range Mercer (1978) had suggested, and provided patterns of ice growth and decay that varied around the continent with changes in temperature. The model also supported the view that the earliest ice sheets
56
P. Barrett
Figure 3.7: Modelling the Antarctic Ice Sheet for progressively higher regional temperatures (Huybrechts, 1993). If a polar amplification factor of 2 is assumed (Manabe and Stouffer, 1980), then the effective rise in average global temperature to achieve each of these states would be 1/2 of the temperature rise shown.
would have reached the coast in the Prydz Bay, whereas the Transantarctic Mountains offered a significant barrier ice reaching to the Ross Sea. 3.6.3. Discoveries in the Transantarctic Mountains Reinterpreted Huybrechts’ model also suggested that it was not possible to create conditions in which an ice sheet could grow and over-top the present Transantarctic Mountains at their present height, depositing wet-based
A History of Antarctic Cenozoic Glaciation
57
glacial debris at the margins. Some argued that the mountains were lower in the Pliocene and had risen rapidly since then (Behrendt and Cooper, 1991), but this was subsequently countered in a review of Antarctic tectonic history by Fitzgerald (2002). However, there were other substantive concerns. New high-resolution deep-sea isotope records of the last few million years gave no indication of negative anomalies to be expected from extensive ice-sheet recession implied by Pliocene marine basins in the East Antarctic interior (Denton et al., 1991; Kennett and Hodell, 1993). Marchant et al. (1993) reported volcanic ashes at high elevations in the McMurdo Dry Valleys and dating back to B11 million years, whose survival seemed inconsistent with a warm Pliocene and an ice-free Antarctic interior. Also, diatoms were discovered in South Pole snow (Kellogg and Kellogg, 1996) and surficial rock debris, and counts showed they were extremely rare in Sirius deposits (Barrett et al., 1997). These observations led to the view that the Sirius deposits predated the widely accepted mid-Miocene cooling, the few agecritical Pliocene diatoms resulting from atmospheric transport. The report by Gersonde et al. (1997) of a Pliocene meteorite impact event ejecting diatomaceous sediment into the stratosphere from the southeastern Pacific Ocean floor provided another source for possible atmospheric contamination. However, Harwood and Webb (1998) maintained that the marine diatoms and diatomite clasts found by Harwood were larger than winds could carry. In reviewing Antarctic climate through Cenozoic times, Barrett (1996, 1999) noted that McMurdo Sound drill cores had shown Oligocene and early Miocene Antarctic Ice Sheets were of similar extent to those of recent times during glacial periods, with largely ice-free forested coasts during interglacial periods, thus resembling the dynamic behaviour of Northern Hemisphere ice sheets in the Quaternary. This contrasted with the relative stability of the Antarctic Ice Sheet at this time, when there has been little change in its central elevation, although a significant volume of ice has been lost around the margin, B15 m of sea level equivalent (SLE) since the Late Glacial Maximum (Zwartz et al., 1997). While offshore mid-Miocene to early Pliocene strata had yet to be cored, seismic stratigraphy of the continental shelf showed numerous sediment packages and unconformities indicating significant glacial deposition and erosion extending to the shelf edge. These data implied significant dynamism of the ice sheet at the margins (Anderson and Bartek, 1992), but indicated little about its interior health. The evidence cited above for a persistent late Neogene ice sheet in the East Antarctic interior was not necessarily inconsistent with clear evidence of Pliocene warmth from marginal coastal marine sediments, as in the region east of Prydz Bay (Harwood et al., 2000).
58
P. Barrett
3.7. Advances in the First Decade of the Twenty-First Century By the mid-1990s with the work of the ANTOSTRAT regional groups coming to a close, and planning well advanced for further geological data gathering by ODP and sea-ice drilling (Cape Roberts Project), the value of modelling as a means of illustrating geologically based scenarios, as well as using geological information to constrain models, seemed increasingly obvious. This was the theme of ANTOSTRAT workshops in Hobart (AU) in 1997 (Cooper and Webb, 1998; Webb and Cooper, 1998), Wellington (NZ) in 1999 and Erice (IT) in 2001 (Cooper et al., 2002), with the latter providing the starting point for the current SCAR Scientific Programme on Antarctic Climate Evolution (ACE).
3.7.1. Ice–Atmosphere–Ocean Modelling Modelling the effect of atmospheric CO2 on the sensitivity of the Antarctic Ice Sheet to orbital forcing by DeConto and Pollard (2003), and how this may have initiated Antarctic Ice Sheet formation, was a turning point in understanding past ice-sheet behaviour. For three decades, it was believed that the main cooling influence came from thermal isolation with the formation of oceanic ‘‘gateways’’ (Kennett, 1977). The modelling showed, albeit simply and with significant assumptions, that B80% of the influence could have been due to the cooling of atmospheric temperature from declining atmospheric CO2 levels. This has gained support from evidence of a 1 km drop in the carbonate compensation depth in the latest Eocene (Coxall et al., 2005) and a proxy record based on d13C from alkenones indicating an irregular decline in atmospheric CO2 levels from more than 1,000 ppm to less than 400 ppm between 40 and 24 Ma (Pagani et al., 2005). Changes in ocean circulation through opening gateways have also been accommodated (Huber and Nof, 2006). As a result of this group of studies, the significance of CO2 as a greenhouse gas influencing Antarctic climate, ice sheets and ocean chemistry has been further enhanced.
3.7.2. Landscape Modelling A different application of modelling has developed in the last few years to explore the response of a particular sector of the ice sheet to changes on geological time scales. Taylor et al. (2004) studied the Lambert drainage
A History of Antarctic Cenozoic Glaciation
59
basin on account of its significant size (16% of the area of the Antarctic Ice Sheet) and the fair degree of knowledge of its both bedrock topography and geological history. They found that once the climate had cooled to the point of initiating glaciation, the erosional history of the Lambert Basin became strongly influenced by continuing graben development, with past erosive deepening limiting the future extent of the ice margin regardless of the climatic regime. Jamieson et al. (2005) used a similar but broader analytical approach to reconstruct the Lambert drainage basin landscape since the first ice sheets developed 34 Ma ago, and found the relative influence of fluvial and glacial erosion through this time to be similar, with only slight glacial modification to long established fluvial drainage patterns. Jamieson and Sugden (2008) extended this approach to the landscape evolution of the whole continent with a similar result. They acknowledge the paucity of constraining geological evidence, and the lack of account taken of changes in area or elevation of various sectors on account of tectonism. However, the potential for development is plain, and a link with climate-oriented modelling will be of crucial importance.
3.7.3. Sea- and Shelf-Ice Drilling Extracting a direct geological record of climate and ice-sheet extent from offshore drilling in McMurdo Sound has continued on the back of the Cape Roberts Project. The successor project, ANDRILL, was designed to sample an expanded Neogene section with a pair of sites, one for the latest Neogene and the other for the middle Miocene transition. There was to be an added twist – the first site would be in the centre of the Victoria Land Basin and located on the ice shelf just south of Ross Island to obtain a history of the Ross Ice Shelf/expanded West Antarctic Ice Sheet cycle through late Neogene times. This required further technological development for maintaining an access hole through an ice shelf in deeper water and with deeper coring sought (Fig. 3.6). Both holes have now been successfully completed with excellent recovery (Naish et al., 2008a; Harwood et al., in press; Table 3.2). They also overlap with each other and CRP-1 (Fig. 3.8), and provide record of numerous cyclic ice-sheet fluctuations under varying climatic regimes over the last B18 million years. Perhaps the most interesting and unexpected result has been recovery of the 90 m of Pliocene diatomite in AND-1B, indicating open water in the Ross Embayment (and most likely throughout West Antarctica) for several hundred thousand years when global temperatures were only 2–31C
1984
1997 1998 1999
CIROS
CRP
1 2 3
2
1 1
77100uS 77100uS 77100uS
77141uS
77134uS 77105uS
77126uS
McMurdo Sound area – offshore DVDP 15 1975 15
1979 1986
77150uS 77151uS 77151uS 77135uS 77135uS 77138uS
McMurdo Sound area – onshore DVDP 1973 1 1973 2 1973 3 1974 10 1974 11 1974 12
MSSTS CIROS
77126uS 77126uS 77126uS
271 272 273
Latitude
77126uS
1973
Ross Sea DSDP 28
Site
270
Year
Project
163145uE 163143uE 163143uE
163132uE
163123uE 164130uE
164123uE
166140uE 166140uE 166140uE 163131uE 163125uE 162151uE
178130uW 178130uW 178130uW
178130uW
Longitude
154 178 295
211
195 197
122
67 47 48 3 80.2 75.1
562 629 491
634
Elevation/ depth (7) (m)
148 624 939
168
230 702
62
201 179 381 182 328 185
233 439 333
423
Depth cored (m)
86 95 97
67
62 98
52
98 96 90 83 94 98
7 37 25
62
Recovery (%)
Diamict – early Miocene Mudstone – Oligocene Sandstone – Devonian
Mudstone – late Oligocene Boulder conglomerate – late Eocene Gneiss – early Paleozoic
Black sand – early Pleistocene
Basalt – Late Quaternary Basalt – Late Quaternary Basalt – Late Quaternary Diamict – late Miocene Diamict – late Miocene Migmatite – early Paleozoic
Diamict clasts – early Pliocene Diamict – early Miocene Diamict – early Miocene
Gneiss – early Paleozoic
Oldest core
Table 3.2: Antarctic coastal and continental shelf rock-drilling sites, 1973 to 2007.
Barrett and Hambrey (1992) CRST (1998) CRST (1999) CRST (2000)
Barrett and Treves (1981) Barrett (1986) Barrett (1989)
Kyle (1981) Kyle (1981) Kyle (1981) Powell (1981) Powell (1981) Powell (1981)
Hayes, Frakes et al. (1975)
References
60 P. Barrett
63151uS
63115uS 63120uS 63116uS
3
5 6 12
Note: ?-indicates age given is uncertain.
2006
64152uS 64157uS 66153uS 66148uS 64100uS 62117uS
1098 1099 1100 1102 1103 1
SHALDRIL
66124uS
1097
Antarctic Peninsula ODP 178 1998
2005
66155uS 67142uS
743 1166
2000
SHALDRIL
67133S
742
ODP 188
68141uS 68123uS
740 741
77146uS
77155uS
67117uS
2
1
739
1988
2007
ANDRILL
Prydz Bay ODP 119
2006
ANDRILL
52122uW 52122uW 52150uW
54139uW
64112uW 64119uW 65142uW 65151uW 65128uW 58145uW
70145uW
74142uE 74147uE
75124uE
76143uE 76123uE
75105uE
165117uE
167101uE
506 532 442
340
1010 1400 459 431 494 488
563
989 475
416
808 551
412
380
840
23 21 4
20
47 108 111 15 363 108
437
97 381
316
226 128
487
1139
1285
40 6 64
32
99 102 5 6 12 87
14
22 19
53
32 26
34
98
98
Naish et al. (2008a) Harwood et al. (in press)
Mudstone – late Eocene/early Oligocene Muddy sand – mid-Miocene Muddy sand – early Pliocene Mudstone – Oligocene
Mud – Holocene Mud – Holocene Diamict – Pleistocene Diamict – Pleistocene? Diamict – late Miocene Mud – Late Pleistocene
Diamict – early Pliocene
http://shaldril. rice.edu/ Anderson et al. (2007)
Barker, Camerlenghi et al. (1999)
Diamict – late Eocene–early Barron, Larsen Oligocene et al. (1989) Red beds – ?Triassic Sandstone, siltstone – ?early Cretaceous Mudstone, diamict – ?Eocene– Oligocene Diamict – Pleistocene Claystone – late Cretaceous O’Brien, Cooper et al. (2001)
Diamict – late Oligocene
Basalt – early Miocene
A History of Antarctic Cenozoic Glaciation 61
62
P. Barrett
Figure 3.8: Cenozoic time–temperature graph showing the time intervals now covered in moderate detail by McMurdo Sound Drilling. ODP Legs 119 and 188 covered the same interval but with significant time gaps and lower core recovery. Intervals covered by much higher resolution deep ice cores and the continuous record from multiple sites in deep-sea sediment cores are shown for comparison. The temperature curve from Crowley and Kim (1995), which is based on the isotope summary of Miller et al. (1987), is modified to show the effect of the methane discharge at 55 Ma (Zachos et al., 2003).
higher and CO2 levels little more than those of today (Van Der Burgh et al., 1993; Ravelo et al., 2004). The middle Miocene climatic optimimum has also been cored at the Antarctic margin for the first time (AND-2, Harwood et al., in press). Much remains to be done for a detailed analysis of sedimentary features, fossil assemblages, physical, chemical and magnetic properties and geochronology for comparing this record and that of AND-1B with those of deep-sea and coastal records in lower latitudes. There is also the prospect for the first time of correlating the last million years of the AND-1B core with the
A History of Antarctic Cenozoic Glaciation
63
temperature and atmospheric CO2 record of the last eight glacial cycles in the ice core from Dome C (EPICA, 2004).
3.7.4. New Discoveries in the Transantarctic Mountains Another remarkable record of the mid-Miocene climate transition has recently been reported from on land in the Transantarctic Mountains west of the McMurdo Dry Valleys (Lewis et al., 2007). Polar conditions since the midMiocene have preserved wet-based ice marginal sediments overlain by thin dry-based glacial deposits between 1,200 and 1,500 m above sea level. Both lithofacies include volcanic ash layers dating the transition at between 13.94 and 13.62 Ma. A different but related study involved dating similar volcanic ash layers in sediments interpreted as flood deposits in the Labyrinth, a set of spectacular anastomosing channels formed within a dolerite sill in the floor of upper Wright Valley nearby (Lewis et al., 2006). The ages constrain the hypothesized megaflood event(s) to within the period 12.4 and 14.4 Ma, and also support the concept developed by Denton and Sugden (2005) that megafloods may have had a significant role in the past behaviour and erosional effectiveness of the Antarctic Ice Sheet in earlier times. Miocene megadebris flow deposits up to 400 m thick and over 200 km in extent on the slope and rise off Wilkes Land (Donda et al., 2007) might also be related. Recent reports (Wingham et al., 2006) suggest that significant subglacial flows are still taking place beneath the Antarctic Ice Sheet, though perhaps not on that scale.
3.7.5. Subglacial Lake Sediments Sought Finally, advances in the search for evidence relating to ACE in this decade have included glaciological and geophysical surveys investigating the many lakes beneath the Antarctic Ice Sheet (Siegert et al., 2005) with a view to using sediments deposited in them as recorders of the basal history of the ice sheet over the period of their existence (Mayer et al., 2007).
3.8. Future Prospects for Improving Knowledge of the History of the Antarctic Ice Sheet The activities described above indicate continuing gains in significant new knowledge and a more broadly consistent understanding of Antarctic
64
P. Barrett
ice-sheet history through improved dating resolution, more careful field observations on land and more advanced geochemical techniques and drilling technology offshore. Understanding present-day ice-sheet behaviour is rapidly advancing through modelling studies integrated with ground truth for surface velocities and mass balance from remote sensing, ice core climatology data from ITASE (2007) and ongoing climate monitoring (AGCS, 2008). However, high-resolution data on geological time scales are still needed from key sectors of the Antarctic Ice Sheet to understand its past behaviour as a basis for credible projections for the future. 3.8.1. Future Sea- and Shelf-Ice Drilling The history of drilling on the Antarctic continental shelf since 1973 is summarized in Fig. 3.9. Both ship-based and floating ice-based operations typically take 5–15 years from conception to execution and require the involvement of a committed group of scientists to execute successfully. The excellent core quality and recovery (95–98%) has come about because of the use of minerals industry technology, which is designed for optimal core recovery, and a sea riser that allows the circulation of a fluid of the right density and viscosity to remove cuttings and protect hole and core. This technique has now been used to drill over 1,200 m beneath the sea floor and in water almost 1,000 m deep. The ability to re-enter a pre-existing hole, which would be potentially useful for drilling beneath rapidly moving ice shelves, has yet to be developed, but considered feasible (A. Pyne, personal communication, 2008). This may well be useful for ANDRILL’s proposed site on the ice shelf east of Ross Island to sample through early Cenozoic strata and capture the transition from the ‘‘greenhouse’’ world on the Antarctic margin (Decesari et al., 2004). 3.8.2. Future Ship-Based Drilling Core recovery from ship-based drilling on the Antarctic continental shelf is typically poor (a few to 30%) in unconsolidated diamicton, but can be high (W95%) in calm seas and either fine-grained mud lacking lonestones or lithified sediment (including diamictite). Sites in deeper water (W400 m) are easier for operational reasons, but depth penetration has rarely been much more than 400 m in practice, on account of drill bit failure. However, hole re-entry and drill bit replacement is possible. Ship-based drilling is normally subject to more ice and weather constraints than floating sea-ice or ice-shelf
A History of Antarctic Cenozoic Glaciation
65
Figure 3.9: Cumulative metres of sedimentary strata drilled on the Antarctic continental shelf since drilling began in 1973, based on data from Table 3.2. Note that only the ODP holes drilled on the shelf are shown here – significant holes were drilled on the slope and rise also. The graph shows the episodic nature of progress in recovering core, reflecting the time and effort required from a relatively small community in both developing proposals and projects and then seeing them through to publication. The better recovery from ‘‘Slimhole’’ drilling is a consequence of the ability to recirculate drilling fluid, a smaller hole diameter and a stable drilling platform.
drilling. The upcoming Integrated Ocean Drilling Program (IODP) Expedition 323, which plans to occupy both a Neogene and a Holocene site on the continental shelf off Wilkes Land in early 2009, will be a useful test of technology and experience gained from previous legs (IODP, 2007). It will also provide the first long time scale geological history for this sector of the Antarctic margin, of interest because it will provide a test for the histories developed separately from drilling in the McMurdo region of the Ross Sea and in Prydz Bay. Despite the quality and length of core it yields, the floating ice technology can be used only in coastal or inland locations, leaving open huge areas of the
66
P. Barrett
shelf that are now extensively covered with seismic surveys. Considerable effort in the past three decades has gone into ship-based coring programmes to penetrate the layer of compact diamictite deposited when the continental shelf was largely covered with grounded ice during the Last Glacial Maximum, but with little success. SHALDRIL, a US Antarctic initiative (http://www.arf.fsu. edu/projectsShaldril.cfm), has been a recent development to address this need and help correlate the extensive seismic stratigraphy between the various sectors of the Antarctic continental shelf with other regions and other continental margins. The record thus far (Table 3.2) indicates that further technological development is needed for successful sampling.
3.8.3. Time Frame and Context for Future Ice-Sheet Investigations The need for soundly based projections for the future behaviour of the Antarctic Ice Sheet is now critical. While ice-sheet modelling and remotesensing measurements from satellites and aircraft are likely to lead this field, palaeoclimate records from ancient ice and sediment are also crucial for documenting ice-sheet response and constraining models for a warmer world. The ANDRILL Pliocene palaeoclimate record with its indications of much reduced ice on West Antarctica on many occasions alerts us to the risks of a world in which atmospheric CO2 levels are o500 ppm (Royer, 2006). However, more high-quality palaeorecords are needed close to all major sectors of the ice sheet. Future sites might be considered not only for their utility in adding to the jigsaw of Antarctic glacial history, but also how the results might be used to test or constrain future modelling and to inform the IPCC assessment process in its effort to integrate results from the whole Earth climate system. The last year has seen increased concern over the large uncertainties in estimates of the loss of Antarctic ice mass from climate change in the last decade, and the issue is complex with estimates made by different methods and over different time frames (review in AGCS, 2008). The broad pattern is of net accumulation over the dome of central East Antarctica with loss of ice around the Antarctic margin, especially the Pacific Coast of West Antarctica, though how much results from long-term trends compared with recent rises in temperature is unclear. An estimate made for Meehl et al. (2007) was around 0.270.35 mm/year of sea level equivalent (SLE) (Lemke et al., 2007), but a more recent report estimates an annual mass loss by 2006 equivalent to 0.470.2 mm SLE (Rignot et al., 2008), almost twice the value estimated for 1996.
A History of Antarctic Cenozoic Glaciation
67
Figure 3.10: History of Antarctic Climate Evolution in the context of changes in global average temperature from 1900 to the present and projected to 2100 (Meehl et al., 2007). The coloured lines beyond 2000 represent possible future global temperature paths from different greenhouse gas emission scenarios. Dashed lines show ‘‘pre-industrial’’ and þ21C temperature levels (European Commission, 2007). Drilling key Antarctic locations for high-quality geological records of past ice-sheet behaviour will help constrain models and reduce uncertainties. Research results will need to be published within the next 4 years to contribute to IPCC AR5 and within the next decade (red dashed box) to contribute to IPCC AR6. Figure 3.10 sets the progress in Antarctic geological drilling against the IPCC review process and the projected rise in global average temperature for various future scenarios. The diagram suggests that only one or two major ANDRILL-type projects in the next decade will yield results in time to have a significant influence on world’s climate community and the public at large while there is still a window of opportunity to mitigate the worst effects of climate change on polar ice sheets (Hansen et al., 2005, 2008). Parallel planning and increased resourcing for site surveys will be especially important for prospective sites beneath ice shelves, which are virtually unexplored and techniques are new, slow and laborious. Glaciological research over the last decade has shown that the Antarctic Ice Sheet can be expected to respond quite differently in different sectors and on different time frames (Vaughan, 2005). Areas of potentially significant ice
68
P. Barrett
Figure 3.11: Velocity map of Antarctic Ice Sheet showing the pattern of present-day ice flow and the main ice drainage systems. The areas for which ice-sheet history are best known – McMurdo Sound (MS) and Prydz Bay (PB) – are circled in brown. Drilling off Wilkes Land (WL) in the dashed brown circle should provide a parallel history from this sector in 2009. The red circles in Pine Island Bay (PIB), the Filchner–Ronne Ice Shelf (FRIS), the Siple Coast (SC) and the Totten Glacier (TG) represent some areas of potential interest for future drilling to seek records of ice-sheet response at the margin to climate change in warmer times in the geological past. Reproduced from Bamber et al. (2000), with permission. loss and whose past history are poorly known are shown in Fig. 3.11. The Pacific coast of West Antarctica, especially Pine Island Bay, has long been recognized as vulnerable to ice loss (Hughes, 1981). Reports of active subglacial hydrological systems beneath the Recovery Glacier, which feeds the Filchner Ice Shelf (Bell et al., 2007), the Totten Glacier in Wilkes Land (Rignot and Jacobs, 2002) and West Antarctica’s Siple Coast (Fricker et al., 2007) indicate the importance of assessing ice-sheet response to climate change in these sectors also. These are also areas of relatively rapid ice flow, though not beyond the ability of a drill system on floating ice to handle. High-resolution palaeoclimate records dating back at least through Pliocene times in each sector would be helpful. Wilkes Land is already covered with the scheduled IODP Wilkes Leg. Potential drill sites in Pine Island Bay, on the Filchner Ice Shelf and on the Siple coast would all provide challenges
A History of Antarctic Cenozoic Glaciation
69
both for site surveys, logistics and drilling, but the results could well more than justify the effort and cost. The selection of targets for geological drilling for Antarctic climate history, and the assembly of science teams to carry out the work, has until recently been relatively straightforward – so little was known and the pressures derived from our curiosity. Today, the wider community is expecting a much greater understanding of the Antarctic ice–atmosphere– ocean system, so that the threat to future ice-sheet stability can be reliably assessed. One of the key tasks will be identifying, planning, surveying and coring a few critical sites for high-quality records of past ice-sheet behaviour in the most vulnerable sectors of the Antarctic as a guide to modelling icesheet response to future greenhouse gas levels.
ACKNOWLEDGEMENTS I wish to thank a number of people for their help and companionship through the last decades of this history. First and foremost is Peter Webb, who with Barrie McKelvey was the first Antarctic expedition from Victoria University of Wellington. I first met Peter in 1968, when he visited Colin Bull, leader of the second VUW Antarctic expedition, but then Director of the Institute of Polar Studies, Ohio State University, when I was a graduate student there in 1968. Peter and I subsequently spent 10 weeks in late 1972 and early 1973 together on DSDP Leg 28, drilling in the Southern Ocean and Ross Sea, and from that experience and Peter’s involvement in the Dry Valley Drilling Project, the concept of drilling for Antarctic glacial history from the sea ice was born. Peter led the SCAR Cenozoic Group of Specialists (1986–1996) and the US component of McMurdo Sound drilling, and his contribution to Antarctic glacial history continues to this day. I am also grateful to Alex Pyne for his long-standing appreciation of the scientific goals of Ross Sea drilling and his technical skills, insight and commitment to the development of drilling technology and logistic support. This began with his observations as a graduate student and drill site science manager for the MSSTS-1 drill hole in 1979 and continued through the CIROS and Cape Roberts Projects to his development of the remarkable ANDRILL system, including its hot water drilling component for ice shelf operations. Key figures from the early days include Bob Clark, Professor of Geology at VUW from 1954 to 1985, who in 1957 initiated the University’s annual expeditions and kept them going until my arrival there in 1970, Sam Treves, University of Nebraska, lead PI for DVDP-15 in 1975, the first attempt at
70
P. Barrett
offshore drilling from the sea ice, and Trevor Hatherton and Bob Thomson, who as head of Geophysics Division and Antarctic Division, DSIR, respectively, provided practical support and approval for the next three attempts at sea-ice drilling, finally leading to the success of CIROS-1 in 1986. From the late 1980s on, Alan Cooper and the ANTOSTRAT group broadened my awareness of marine geophysics, and about the same time my appointment to the SCAR Group of Specialists on Environmental Affairs and Conservation, led initially by Nigel Bonner and then David Walton, was significant in helping me deal with the environmental implications of offshore drilling from both practical and scientific perspectives. In the 1990s, the Cape Roberts Project represented the culmination of the previous two decades of sea-ice drilling experience, brought together through the planning efforts of its International Steering Committee, variously chaired by Maria Bianca Cita (Italy), Fred Davey (NZ), Franz Tessensohn (Germany), Mike Thomson (UK), Jaap van der Meer (Holland), Peter Webb (USA) and Ken Woolfe (Australia), and the Operations Management Group chaired by Gillian Wratt (NZ). Finally, I wish to acknowledge the work of the ACE community, led by Martin Siegert and Rob Dunbar, and within that the ANDRILL Steering Committee (the next generation) – Dave Harwood and Ross Powell (USA), Tim Naish and Gary Wilson (NZ), Fabio Florindo and Franco Talarico (Italy) and Frank Niessen and Gerhard Kuhn (Germany) – in filling in key gaps in Antarctic climate history in the McMurdo Sound area, and hopefully beyond in the future. Reviews by Peter Webb, Colin Summerhayes and Jane Francis on an earlier draft were very helpful. Preparation of this chapter was supported by NZ Foundation for Research, Science and Technology, Grant No. C05X0410 ANDRILL.
REFERENCES AGCS (Mayewski, P. A., Meredith, M. P., Summerhayes, C. P., Turner, J., Worby, A., Barrett, P. J., Casassa, G., Bertler, N. A. N., Bracegirdle, T., Naveira Garabato, A. C., Bromwich, D., Campbell, H., Hamilton, G. H., Lyons, W. B., Maasch, K. A., Aoki, S., & Xiao, C.) (2008). State of the Antarctic and Southern Ocean climate system (SASOCS). Rev. Geophys., in press. Anderson, J. B. (1991). The Antarctic continental shelf: Results from marine geological and geophysical investigations. In: R. J. Tingey (Ed.). The Geology of Antarctica. Clarendon Press, Oxford, pp. 285–324.
A History of Antarctic Cenozoic Glaciation
71
Anderson, J. B., & Bartek, L. R. (1992). Cenozoic glacial history of the Ross Sea revealed by intermediate resolution seismic reflection data combined with drill site information. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change, Antarctic Research Series. AGU, Washington, DC, Vol. 56, pp. 231–263. Anderson, J. B., Wellner, J., Wise, S., Bohaty, S., Manley, P., Smith, T., Weaver, F., & Kulhanek, D. (2007). Seismic and chronostratigraphic results from SHALDRIL II, Northwestern Weddell Sea. In: A. K. Cooper, C. R. Raymond et al. (Eds). Antarctica: A Keystone in a Changing World-Online Proceedings for the Tenth International Symposium on Antarctic Earth Sciences, USGS Open-File Report. 2007-1047. Short Research Paper 094, doi:10.3133/ of2007-1047.srp094. Armstrong, R. L., Hamilton, W., & Denton, G. H. (1968). Glaciation in Taylor Valley, Antarctica, older than 2.7 million years. Science, 159, 187–189. Bamber, J. L., Vaughan, D. G., & Joughin, I. (2000). Widespread complex flow in the interior of the Antarctic Ice Sheet. Science, 287, 1248–1250. Barker, P. F., Camerlenghi, A., Acton, G. D., et al. (1999). Proceedings of the ODP, Initial Reports, Vol. 178 [Online], http://www.odp.tamu.edu/publications/ 178_IR/178TOC.HTM Barker, P. E., Kennett, J. P., et al. (1988). Proceedings of the ODP, Initial Reports, Vol. 113. College Station, TX (Ocean Drilling Program), doi:10.2973/odp.proc. ir.113.1988. Barnola, J. M., Raynaud, D., Korotkevich, Y. S., & Lorius, C. (1987). Vostok ice core provides 160,000-year record of atmospheric CO2. Nature, 329, 408–414. Barrett, P. J. (1975). Characteristics of pebbles from Cenozoic marine glacial sediments in the Ross Sea (DSDP sites 270–274) and the South Indian Ocean (Site 268). In: D. E. Hayes, L. A. Frakes, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, pp. 769–784. Barrett, P. J. (1982). Proposal for Cenozoic investigations in the western Ross Sea (CIROS). N. Z. Antarctic Rec., 2, 32–39. Barrett, P. J. (Ed.) (1986). Antarctic Cenozoic history from the MSSTS-1 drillhole, McMurdo Sound, Antarctica. NZ DSIR Bulletin, Vol. 237, 174 pp. Barrett, P. J. (Ed.) (1989). Antarctic Cenozoic history from the CIROS-1 drillhole, McMurdo Sound, Antarctica. NZ DSIR Bulletin, Vol. 245, 254 pp. Barrett, P. J. (1996). Antarctic palaeoenvironment through Cenzoic times – A review. Terra Antartica, 3, 103–119. Barrett, P. J. (1999). Antarctic climate history over the last 100 million years. Terra Antartica Rep., 3, 53–72. Barrett, P. J. (2007). Cenozoic climate and sea level history from glacimarine strata off the Victoria Land Coast, Cape Roberts Project, Antarctica, In: M. J. Hambrey, P. Christoffersen, N. F. Glasser, & B. Hubbart (Eds). Glacial Processes and Products, International Association of Sedimentologists Special Publication, Vol. 39, pp. 259–287.
72
P. Barrett
Barrett, P. J., & Hambrey, M. J. (1992). Plio-Pleistocene sedimentation in Ferrar Fiord, Antarctica. Sedimentology, 39, 109–123. Barrett, P. J., & Harwood, D. M. (1992). Geological background and rationale for drilling. In: P. J. Barrett, & F. J. Davey (Eds). Antarctic Stratigraphic Drilling, Cape Roberts Project, Workshop Report. R.l Soc. N. Z. Miscellaneous Ser., 23, 4–10. Barrett, P. J., Adams, C. J., Mclntosh, W. C., Swisher, C. C., III, & Wilson, G. S. (1992). Geochronological evidence supporting Antarctic deglaciation three million years ago. Nature, 359, 816–818. Barrett, P. J., Bleakley, N. L., Dickinson, W. W., Hannah, M. J., & Harper, M. A. (1997). Distribution of siliceous microfossils on Mount Feather, Antarctica, and the age of the Sirius Group. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 763–770. Barrett, P. J., & Davey, F. J. (1992). Cape Roberts Project Workshop Report, Royal Society of New Zealand, Miscellaneous Series, Vol. 23, 38 pp. Barrett, P. J., Elston, D. P., Harwood, D. M., McKelvey, B. C., & Webb, P.-N. (1987). Mid-Cenozoic record of glaciation and sea-level change on the margin of the Victoria Land Basin, Antarctica. Geology, 15, 634–637. Barrett, P. J., & McKelvey, B. C. (1986). Stratigraphy. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the MSSTS-1 Drillhole, McMurdo Sound, Antarctica, NZ DSIR Miscellaneous Bulletin. Vol. 237, pp. 9–51. Barrett, P. J., & Treves, S. B. (1981). Sedimentology and petrology of DVDP-15, McMurdo Sound. In: L. D. McGinnis (Ed.). Dry Valley Drilling Project, Antarctic Research Series. AGU, Washington, DC, Vol. 81, pp. 1–5. Barrett, P. J., & Webb, P.-N. (1981). Future Cenozoic research in Antarctica. SCAR Bull., 69, 595–596. Barron, J., Larson, B., & Shipboard Scientific Party (1989). Proceedings of the ODP, Initial Reports. Vol. 119, College Station, TX (Ocean Drilling Program). Beeby, C. (1972). The Antarctic Treaty. New Zealand Institute of International Affairs, Wellington, p. 26. Behrendt, J. C., & Cooper, A. K. (1991). Evidence of rapid Cenozoic uplift of the shoulder escarpment of the West Antarctic Rift System and a speculation on possible climate forcing. Geology, 19, 315–319. Bell, R. E., Studinger, M., Shuman, C. A., Fahnestock, M., & Joughin, I. (2007). Large subglacial lakes in East Antarctica at the onset of fast-flowing ice streams. Nature, 445, 904–907. Berggren, W. A., Kent, D. V., Swisher, C. C., III, & Aubry, M.-P. (1995). A revised Cenozoic geochronology and chronostratigraphy. In: W. A. Berggren, D. V. Kent, M.-P. Aubry, & J. Hardenbol (Eds). Geochronology, Time Scales and Stratigraphic Correlation. Framework for an historical geology: SEPM (Society for Sedimentary Geology) Special Publication, Vol. 54, pp. 129–212. Bull, C., McKelvey, B. C., & Webb, P.-N. (1962). Quaternary glaciations in southern Victoria Land, Antarctica. J. Glaciol., 4, 63–78.
A History of Antarctic Cenozoic Glaciation
73
Bushnell, V. C. (Ed.) (1964–1975). Antarctic Map Folio Series (19 Folios). American Geographical Society, Washington, DC. Cape Roberts Science Team (1998). Initial Report on CRP-1. Terra Antartica, 5, 1–187. Cape Roberts Science Team (1999). Studies from the Cape Roberts Project, Ross, Sea, Antarctica – Initial Report on CRP-2/2A. Terra Antartica, 6, 1–173, with supplement. Cape Roberts Science Team (2000). Studies from the Cape Roberts Project, Ross, Sea, Antarctica, Initial Report on CRP-3. Terra Antartica, 7, 1–209 , with supplement. Carey, S. W. (Ed.) (1958). Continental Drift – A Symposium. Geology Department, University of Tasmania, Hobart, p. 363. Clapperton, C. E., & Sugden, D. E. (1990). Late Cenozoic History of the Ross Embayment, Antarctica. Quaternary Sci. Rev., 9, 253–272. Clough, J. W., & Hansen, B. L. (1979). The Ross Ice Shelf Project. Science, 433–434. Cooper, A., Barrett, P., & Florindo, F. (2002). New inferences on Antarctic Ice Sheets and Cenozoic paleoclimates, Eos Trans. AGU, Vol. 83, pp. 35–36, doi:10.1029/2002EO000027. Cooper, A. K., & Brancolini, G. (1997). Antarctic seismic data library system for cooperative research (SDLS), SCAR Report, 16, Appendix 7, http://www.scar.org/ publications/reports/16/appendix7/sdls.html Cooper, A. K., & Davey, F. J. (1985). Episodic rifting of the Phanerozoic rocks of the Victoria Land Basin, Western Ross Sea, Antarctica. Science, 229, 1085–1087. Cooper, A. K., Davey, F. J., & Behrendt, J. C. (1987). Seismic stratigraphy and structure of the Victoria Land Basin, Western Ross Sea, Antarctica. In: A. K. Cooper, & F. J. Davey (Eds). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea, Earth Science Series. Circum-Pacific Council Energy Mineral Resources, Houston, TX, Vol. 5B, pp. 27–77. Cooper, A. K., & Webb, P.-N. (1998). Antarctica’s role in global change research examined. Eos Trans. AGU, Vol. 79, p. 3, doi:10.1029/98EO00005. Coxall, H., Wilson, P. A., Palike, H., Lear, C. H., & Backman, J. (2005). Rapid stepwise onset of Antarctic glaciation and deeper calcite compensation in the Pacific Ocean. Nature, 433, 53–57. Craddock, C., Bastien, T. W., & Rutford, R. H. (1964). Geology of the Jones Mountains area. In: R. J. Adie (Ed.). Antarctic Geology. North-Holland, Amsterdam, pp. 171–187. Crowley, T. J., & Kim, K. (1995). Comparison of longterm greenhouse projections with the geologic record. Geophys. Res. Lett., 22, 933–936. Davey, F. J., Barrett, P. J., Cita, M. B., van der Meer, J. J. M., Tessensohn, F., Thomson, M. R. A., Webb, P.-N., & Woolfe, K. J. (2001). Drilling for Antarctic Cenozoic climate and tectonic history at Cape Roberts, Southwestern Ross Sea, Eos Trans. AGU, Vol. 82, p. 585, doi:10.1029/01EO00339. David, T. W. E. (1914). Antarctica and some of its problems. Geogr. J., 43, 605–627.
74
P. Barrett
De Santis, L., Prato, S., Brancolini, G., Lovo, M., & Torelli, L. (1999). The Eastern Ross Sea continental shelf during the Cenozoic: Implications for the West Antarctic Ice Sheet development. Global Planetary Change, 23, 173–196. Decesari, R. C., Luyendyk, B. P., Bartek, L. R., Sorlien, C. C., Wilson, D. S., & Diebold, J. B. (2004). Ice shelf drill sites proposed to study pre-late Oligocene climate and tectonic history, Coulman High, Southwestern Ross Sea, Antarctica, Eos Trans. AGU, Vol. 85, Fall Meeting Supplement, Abstract T11A-1244. DeConto, R. M., & Pollard, D. (2003). Rapid Cenozoic glaciation of Antarctica induced by declining atmospheric CO2. Nature, 421, 245–249. Denton, G. H., Armstrong, R. L., & Stuiver, M. (1970). Late Cenozoic glaciation in Antarctica. The record in the McMurdo Sound region. Antarctic J. United States, 5, 15–21. Denton, G. H., Prentice, M. L., & Burckle, L. H. (1991). Cainozoic history of the Antarctic Ice Sheet. In: R. J. Tingey (Ed.). The Geology of Antarctica. Clarendon Press, Oxford, pp. 365–433. Denton, G. H., & Sugden, D. E. (2005). Meltwater features that suggest Miocene ice-sheet overriding of the transantarctic mountains in Victoria Land, Antarctica. Geogr. Ann., 87, 1–19. Dietz, R. S. (1961). Continent and ocean basin evolution by spreading of the sea floor. Nature, 190, 854–857. Domack, E. W., Fairchild, W. W., & Anderson, J. B. (1980). Lower cretaceous sediment from the East Antarctic continental shelf. Nature, 287, 625–626. Donda, F., O’Brien, P. E., De Santis, L., Rebesco, M., & Brancolini, G. (2007). Mega Debris Flow Deposits on the Western Wilkes land margin, East Antarctica. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES, USGS Open File Report 2007-1047, Short Research Paper 040, 4 pp., doi:10.3133/of20071047.srp040. Drewry, D. J. (1983). Antarctica: Glaciological and Geophysical Folio. Scott Polar Research Institute, Cambridge, UK9 sheets. Dunbar, G. B., Naish, T. R., Barrett, P. J., Fielding, C. R., & Powell, R. D. (2008). Constraining the Amplitude of Late Oligocene Bathymetric Changes in Western Ross Sea during Orbitally-Induced Oscillations in the East Antarctic Ice Sheet: (1) Implications for Glacimarine Sequence Stratigraphic Models. Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 50–65. Du Toit, A. (1937). Our Wandering Continents: An Hypothesis of Continental Drifting. Oliver Boyd, London. Eittrem, S. L., & Hampton, M. A. (Eds). (1987). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land Circum-Pacific Council for Energy and Mineral Resources Earth Sciences Series, 5A, Houston, 221 pp. EPICA (2004). Eight glacial cycles from an Antarctic ice core. Nature, 429, 623–628. European Commission (2007). Communication from the Commission to the Council, the European Parliament, the European Economic and Social Committee and the Committee of the Regions – Limiting global climate change
A History of Antarctic Cenozoic Glaciation
75
to 2 degrees Celsius – The way ahead for 2020 and beyond. http://ec.europa.eu/ environment/climat/future_action.htm Fairbanks, R. G. (1989). A 17,000-year glacio-eustatic sea level record; influence of glacial melting rates on the younger dryas event and deep-ocean circulation. Nature, 342, 637–642. Fitzgerald, P. F. (2002). Tectonics and landscape evolution of the Antarctic plate since the breakup of Gondwana, with an emphasis on the West Antarctic Rift System and the Transantarctic Mountains. In: J. A. Gamble, D. N. B. Skinner, & S. A. Henrys (Eds). Antarctica at the Close of the Millennium. R. Soc. N. Z. Bull., 35, 453–469. Flint, R. F. (1957). Glacial and Pleistocene geology. Wiley, New York, 553 pp. Flint, R. F. (1971). Glacial and Quaternary Geology. Wiley, New York, 892 pp. Florindo, F., Wilson, G. S., Roberts, A. P., Sagnotti, L., & Verosub, K. L. (2005). Magnetostratigraphic chronology of a late Eocene to early Miocene glacimarine succession from the Victoria Land Basin, Ross Sea, Antarctica. Global Planetary Change, 208, 207–236. Fogg, G. E. (1992). A History of Antarctic Science. Cambridge University Press, Cambridge, UK, 505 pp. Francis, J. F., & Hill, R. S. (1996). Fossil plants from the Pliocene Sirius Group, transantarctic mountains: Evidence for climate from growth rings and fossil leaves. Palaios, 11, 389–396. Fricker, H. A., Scambos, T., Bindschadler, R. & Padman, L. (2007). An active subglacial water system in West Antarctica mapped from space. Science, 315, 1544–1548. Gersonde, R., Kyte, F. T., Bleil, U., Diekmann, B., Flores, J. A., Gohl, K., Grahl, G., Hagen, R., Kuhn, G., Sierro, F. J., Vo¨lker, D., Abelmann, A., & Bostwick, J. A. (1997). Geological record and reconstruction of the late Pliocene impact of the Eltanin Asteroid in the Southern Ocean. Nature, 390, 357–363. Gleadow, A. J. W., & Fitzgerald, P. G. (1987). Uplift history and structure of the transantarctic mountains: New evidence from fission track dating of basement apatites in the dry valleys area, Southern Victoria Land. Earth Planetary Sci. Lett., 82, 1–14. Hambrey, M. J., Barrett, P. J., Ehrmann, W. U., & Larsen, B. (1992). Cenozoic sedimentary processes on the Antarctic continental shelf: The record from deep drilling. Z. Geomorphologie, 86(Suppl.), 73–99. Hambrey, M. J., Barrett, P. J., Hall, K. J., & Robinson, P. H. (1989). Stratigraphy. In: P. J. Barrett (Ed.). Antarctic Cenozoic history from the CIROS-1 drillhole, McMurdo Sound, Antarctica, DSIR Bulletin. Wellington, New Zealand, Vol. 245, pp. 23–48. Hambrey, M. J., Larsen, B., & Ehrmann, W. U. (1991). The glacial record from the Prydz Bay continental shelf, East Antarctica. In: J. Barron, B. Larsen, et al. (Eds). Ocean Drilling Program, 119, Scientific Results. College Station, Texas, pp. 77–132.
76
P. Barrett
Hannah, M. J., Cita, M. B., Coccioni, R., & Monechi, S. (1997). The Eocene– Oligocene boundary at 701S, McMurdo Sound, Antarctica. Terra Antartica, 4, 79–87. Hansen, J., Fung, I., Lacis, A., Rind, D., Lebedeff, S., Ruedy, R., Russell, G., & Stone, P. (1988). Global climate changes as forecast by Goddard Institute for Space Studies three-dimensional model. J. Geophys. Res., 93, 9341–9364. Hansen, J., Nazarenko, L., Ruedy, R., Sato, M., Willis, J., Genio, A., Koch, D., Lacis, A., Lo, K., Menon, S., Novakov, T., Perlwitz, G., Russell, G., Schmidt, G. A., & Tausnev, N. (2005). Earth’s energy imbalance: Confirmation and implications. Science, 308, 1431–1435. Hansen, J., Sato, M., Kharecha, P., Beerling, D., Masson-Delmotte, V., Pagani, M., Raymo, M., Royer, D. & Zachos, J. C. (2008). Target Atmospheric CO2: Where should Humanity Aim? http://www.columbia.edu/Bjeh1/2008/TargetCO2_ 20080407.pdf Haq, B. U., Hardenbol, H., & Vail, P. R. (1987). The chronology of fluctuating sea level since the Triassic. Science, 235, 1156–1167. Harrington, H. J. (1965). Geology and morphology of Antarctica. In: P. van Oye, & J. Van Mieghem (Eds). Biogeography and Ecology of Antarctica. Monographiae Biologicae. Junk, The Hague, Vol. 15, pp. 1–71. Harwood, D. M. (1986). Diatom Biostratigraphy and Paleoecology with a Cenozoic History of Antarctic Ice Sheets. Ph.D. Dissertation. The Ohio State University, Columbus, OH, 592 pp. Harwood, D. M., Florindo, F., Talarico, F., & Levy, R. (in press). Studies from the ANDRILL, Southern McMurdo Sound Project, Antarctica – Initial Science Report on AND-2A. Terra Antarctica. Harwood, D. M., Barrett, P. J., Edwards, A. R., Reick, H., & Webb, P.-N. (1989). Chronology. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound, Antarctica, DSIR Bulletin. Wellington, New Zealand, Vol. 245, pp. 231–239. Harwood, D., Lacy, L., & Levy, R. (Eds) (2003). Future Antarctic Margin Drilling: Developing a Science Program Plan for McMurdo Sound, Report of a Workshop, Oxford, UK, April 5–7, 2001. ANDRILL Science Management Office, University of Nebraska, Lincoln, 302 pp. Harwood, D. M., McMinn, A., & Quilty, P. G. (2000). Diatom biostratigraphy and age of the Pliocene Sørsdal formation, Vestfold Hills, East Antarctica. Antarctic Sci., 12, 443–462. Harwood, D. M., & Webb, P.-N. (1998). Glacial transport of diatoms in the Antarctic Sirius Group: Pliocene refrigerator. GSA Today, 8(1), 4–8. Hayes, D. E., Frakes, L. A., et al. (1975). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, 1012 pp. Hays, J. D., Imbrie, J., & Shackleton, N. J. (1976). Variations in the earth’s orbit: Pacemaker of the ages. Science, 194, 1121–1132.
A History of Antarctic Cenozoic Glaciation
77
Hess, H. H. (1962). History of ocean basins. In: A. E. J. Engel, et al. (Eds). Petrologic Studies: A Symposium in Honor of A.F. Buddington. Geological Society of America, Boulder, pp. 599–620. Hill, R. S., Harwood, D. M., & Webb, P.-N. (1996). Nothofagus beardmorensis (Nothofagaceae), a new species based on leaves from the Pliocene Sirius Group, Transantarctic Mountains, Antarctica. Rev. Paleobot. Palynol., 94, 11–24. Holmes, A. (1965). Principles of Physical Geology (2nd Edn.). Nelson, London, 1285 pp. Houtz, R. E., & Davey, F. J. (1973). Seismic profiler and Sonobuoy measurements in the Ross Sea, Antarctica. J. Geophys. Res., 78, 3448–3468. Houtz, R. E., & Meijer, R. (1970). Structure of the Ross Sea shelf from profiler data. J. Geophys. Res., 75, 6592–6597. Huber, M., & Nof, D. (2006). The ocean circulation in the southern hemisphere and its climatic impacts in the Eocene. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 9–28. Hughes, T. J. (1981). The weak underbelly of the West Antarctic Ice Sheet. J. Glaciol., 27, 518–525. Huybrechts, P. (1993). Glaciological modelling of the late Cenozoic East Antarctic Ice Sheet: Stability or dynamism? Geografiska Annaler, 75A, 221–238. IODP (2007). Expedition 323 – Wilkes Land, http://iodp.tamu.edu/scienceops/ expeditions/wilkes_land.html IPCC (2007). In: S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, & H. L. Miller (Eds). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, 996 pp. ITASE (2007). http://www2.umaine.edu/itase/ Jamieson, S. S. R., Hulton, N. R. J., Sugden, D. E., Payne, A. J., & Taylor, J. (2005). Cenozoic landscape evolution of the Lambert Basin, East Antarctica: The relative role of rivers and ice sheets. Global Planetary Change, 45, 35–49. Jamieson, S. S. R., & Sugden, D. E. (2008). Landscape evolution of Antarctica. In: A. K. Cooper, P. J. Barrett, H. Stagg, B. Storey, E. Stump, W. Wise, & the 10th ISAES editorial team (Eds). Antarctica: A Keystone in a Changing World. Proceedings of the 10th International Symposium on Antarctic Earth Sciences. The National Academies Press, Washington, DC, pp. 39–54. Kellogg, D. E., & Kellogg, T. B. (1996). Diatoms in South Pole ice: Implications for Eolian contamination of Sirius Group deposits. Geology, 24, 115–118. Kennett, J. P. (1977). Cenozoic evolution of Antarctic glaciation, the circumAntarctic Ocean and their impact on global oceanography. J. Geophys. Res., 82, 3843–3860. Kennett, J. P. (1982). Marine Geology. Prentice Hall, Englewood Cliffs, 813 pp. Kennett, J. P., & Hodell, D. A. (1993). Evidence for relative climatic stability of Antarctica during the early Pliocene: A marine perspective. Geografiska Annaler, 75A, 205–220.
78
P. Barrett
Kennett, J. P., Houtz, R. E., et al. (1975). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 29, 1197 pp. Kennett, J. P., & Warnke, D. A. (Eds). (1992). The Antarctic Paleoenvironment: A Perspective on Global Change, Part One, Antarctic Research Series. AGU, Washington, DC, Vol. 56, p. 400. Kennett, J. P., & Warnke, D. A. (Eds). (1993). The Antarctic Paleoenvironment: A Perspective on Global Change, Part Two, Antarctic Research Series. AGU, Washington, DC, Vol. 60, p. 273. Kyle, P. R. (1981). Geologic history of Hut Point Peninsula as inferred from DVDP 1, 2 and 3 drillcores and surface mapping. In: L. D. McGinnis (Ed.). Dry Valley Drilling Project, Antarctic Research Series. AGU, Washington, DC, Vol. 81, pp. 427–445. Lemke, P., Ren, J., Alley, R. B., Allison, I., Carrasco, J., Flato, G., Fujii, Y., Kaser, G., Mote, P., Thomas, R. H., & Zhang, T. (2007). Observations: Changes in snow, ice and frozen ground. In: S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, & H. L. Miller (Eds). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK. Lewis, A. R., Marchant, D. R., Ashworth, A. C., Hemming, S. R., & Machlus, M. L. (2007). Major middle Miocene global climate change: Evidence from East Antarctica and the transantarctic mountains. Geol. Soc. Am. Bull., 119, 1449–1461. Lewis, A. R., Marchant, D. R., Kowalewski, D. E., Baldwin, S. L., & Webb, L. E (2006). The age and origin of the labyrinth, Western Dry Valleys, Antarctica: Evidence for extensive middle Miocene subglacial floods and freshwater discharge to the Southern Ocean. Geology, 34, 315–316. Loutit, T. S., & Kennett, J. P. (1981). New Zealand and Australian Cenozoic sedimentary cycles and global sea-level changes. AAPG Bull., 65, 1586–1601. Manabe, S., & Stouffer, R. J. (1980). Sensitivity of a global climate model to an increase of CO2 concentration in the atmosphere. J. Geophys. Res., 85, 5529–5554. Marchant, D. R., Denton, G. H., & Swisher, C. C. (1993). Miocene-PliocenePleistocene glacial history of Arena Valley, Quartermain Mountains, Antarctica. Geografiska Annaler, 75A, 269–302. Margolis, S. V., & Kennett, J. P. (1970). Antarctic glaciation during the tertiary recorded in sub-Antarctic deep-sea cores. Science, 170, 1085–1087. Mayer, C., Christoffersen, P., DeConto, R., Li, Y., Powell, R., & Tulaczyk, S. (2007). SALE Glaciohydrogeology and Paleoclimate, http://scarsale.tamu.edu/ key-programs/sale-glaciohydrogeology-and-paleoclimate.html Mayewski, P. A. (1975). Glacial geology and late Cenozoic history of the transantarctic mountains, Antarctica. The Ohio State University Institute of Polar Studies Report No. 56, 168 pp.
A History of Antarctic Cenozoic Glaciation
79
McGinnis, L. D. (Ed.) (1981). Dry Valley Drilling Project. Antarctic Research Series. AGU, Washington, DC, Vol. 81. McKelvey, B. C. (1991). The Cainozoic record in South Victoria Land: A geological evaluation of the McMurdo Sound drilling projects. In: R. J. Tingey (Ed.). The Geology of Antarctica. Clarendon Press, Oxford, pp. 434–454. Meehl, G. A., Stocker, T. F., Collins, W. D., Friedlingstein, P., Gaye, A. T., Gregory, J. M., Kitoh, A., Knutti, R., Murphy, J. M., Noda, A., Raper, S. C. B., Watterson, I. G., Weaver, A. J., & Zhao, Z.-C. (2007). Chapter 10 – Global climate projections. In: S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, & H. L. Miller (Eds). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, pp. 747–845. Mercer, J. H. (1972). Some observations on the glacial geology of the Beardmore glacier area. In: R. J. Adie (Ed.). Antarctic Geology and Geophysics. Universitetsforlaget, Oslo, pp. 427–433. Mercer, J. H. (1978). West Antarctic Ice Sheet and CO2 greenhouse effect: A threat of disaster. Nature, 271, 321–325. Miller, K. G., Fairbanks, R. G., & Mountain, G. S. (1987). Tertiary oxygen isotope synthesis, sea level history, and continental margin erosion. Paleoceanography, 2, 1–19. Naish, T. R., Powell, R. D., Barrett, P. J., Levy, R. H., Henrys, S., Wilson, G. S., Krissek, L. A., Niessen, F., Pompilio, M., Ross, J., Scherer, R., Talarico, F., Pyne, A., & the ANDRILL-MIS Science team (2008a). Late Cenozoic climate history of the Ross Embayment from the AND-1B drill hole: Culmination of three decades of Antarctic margin drilling. In: A. K. Cooper, P. J. Barrett, H. Stagg, B. Storey, E. Stump, W. Wise, & the 10th ISAES editorial team (Eds). Antarctica: A Keystone in a Changing World. Proceedings of the 10th International Symposium on Antarctic Earth Sciences. The National Academies Press, Washington, DC, pp. 71–82. Naish, T. R., Wilson, G. S., Dunbar, G. B., & Barrett, P. J. (2008b). Constraining the amplitude of Late Oligocene bathymetric changes in Western Ross Sea during orbitally-influenced oscillations in the East Antarctic Ice Sheet: (2) Implications for global glacio-ustasy. Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 66–76. Naish, T. R., Woolfe, K. J., Barrett, P. J., Wilson, G. S., Atkins, C., Bohaty, S. M., Bu¨cher, C. J., Claps, M., Davey, F. J., Dunbar, G. B., Dunn, A. G., Fielding, C. R., Florindo, F., Hannah, M. J., Harwood, D. M., Henrys, S. A., Krissek, L. A., Lavelle, M. A., van der Meer, J., McIntosh, W. C., Niessen, F., Passchier, S., Powell, R. D., Roberts, A. P., Sagnotti, L., Scherer, R. P., Strong, C. P., Talarico, F., Verosub, K. L., Villa, G., Watkins, D. K., Webb, P. N., & Wonik, T. (2001). Orbitally induced oscillations in the East Antarctic Ice Sheet at the Oligocene/Miocene boundary, Nature, 413, 719–723. Nichols, R. L. (1964). Present status of Antarctic glacial geology. In: R. J. Adie (Ed.). Antarctic Geology. North-Holland, Amsterdam, pp. 123–137.
80
P. Barrett
O’Brien, P. E., Cooper, A. K., Richter, C., et al. (2001). Proceedings of the ODP, Initial Reports. Vol. 188 [Online], http://www.odp.tamu.edu/publications/ 188_IR/188ir.htm Oerlemans, J. (1982). A model of the Antarctic Ice Sheet. Nature, 297, 550–553. Pagani, M., Zachos, J. C., Freeman, K. H., Tipple, B., & Bohaty, S. (2005). Marked decline in atmospheric carbon dioxide concentrations during the Paleogene. Science, 309, 600–603. Pekar, S. F., DeConto, R. M., & Harwood, D. M. (2006). Resolving a late Oligocene conundrum: Deep-sea warming and Antarctic glaciation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 29–40. Pewe, T. (1960). Multiple glaciation in the McMurdo Sound Region, Antarctica. J. Geol., 68, 498–515. Pe´we´, T. L. (1962). Age of moraines in Victoria Land, Antarctica. J. Glaciol., 4, 93–100. Powell, R. D. (1981). Sedimentation conditions in Taylor Valley inferred from textural analyses of DVDP cores. In: L. D. McGinnis (Ed.). Dry Valley Drilling Project, Antarctic Research Series. AGU, Washington, DC, Vol. 81, pp. 331–350. Ravelo, C. R., Andreasen, D. H., Lyle, M., Lyle, A. O., & Wara, M. W. (2004). Regional climate shifts caused by gradual global cooling in the Pliocene epoch. Nature, 429, 263–267. Rignot, E., Bamber, J. L., Van Den Broeke, M. R., Davis, C., Yonghong, L., Van De Berg, W. J., & Van Meijgaard, E. (2008). Recent Antarctic ice mass loss from radar interferometry and regional climate modelling. Nature, doi:10.1038/ngeo102. Rignot, E., & Jacobs, S. S. (2002). Rapid bottom melting widespread near Antarctic Ice Sheet grounding lines. Science, 296, 2020–2023. Roberts, A. P., Wilson, G. S., Harwood, D. M., & Verosub, K. L. (2003). Glaciation across the Oligocene–Miocene boundary in southern McMurdo Sound, Antarctica: New chronology from the CIROS-1 drill hole. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 113–130. Royer, D. L. (2006). CO2-forced climate thresholds during the Phanerozoic. Geochim. Cosmochim. Acta, 70, 5665–5675. Scherer, R. P., Harwood, D. M., Ishman, S. E., & Webb, P.-N. (1988). Micropaleontological analysis of sediments from the Crary Ice Rise, Ross Ice Shelf. Antarctic J. U.S., 23, 34–36. Siegert, M. J., et al. (2004). ACE Proposal to SCAR, www.scar-ace.org Siegert, M. J., Carter, S., Tabacco, I., Popov, S., & Blankenship, D. D. (2005). A revised inventory of Antarctic subglacial lakes. Antarctic Sci., 17, 453–460. Schneider, S. (1989). Global warming: Are we entering the greenhouse century? Sierra Club, San Francisco, 346 pp. Seward, A. C. (1914). Antarctic fossil plants. British (Terra Nova) Antarctic Expedition, 1910(1), 1–49. Shackleton, N. J., & Kennett, J. P. (1975). Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: Oxygen and carbon isotope analysis in
A History of Antarctic Cenozoic Glaciation
81
DSDP Sites 277, 279 and 281. In: J. P. Kennett, R. E. Houtz, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 29, pp. 743–755. Smith, P. M. (1981). The role of the Dry Valley Drilling Project in Antarctic and international science policy. In: L. D. McGinnis (Ed.). Dry Valley Drilling Project, Antarctic Research Series. AGU, Washington, DC, Vol. 81, pp. 1–5. Sugden, D., Marchant, D. M., & Denton, G. H. (1993). The case for a stable Antarctic Ice Sheet. Geografisker Annaler, 75A, 1–273. Taylor, G. (1914). Physiography and glacial geology of East Antarctica. Geograph. J., 44, 365–382. Taylor, G. (1922). The Physiography of McMurdo Sound and Granite Harbour Region. British Antarctic (Terra Nova) Expedition, 1910–13, Harrison, London, 246 pp. Taylor, J., Siegert, M. J., Payne, A. J., Hambrey, M. J., O’Brien, P. E., Cooper, A. K., & Leitchenkov, G. (2004). Topographic controls on post-Oligocene changes in ice-sheet dynamics, Prydz Bay region, East Antarctica. Geology, 32, 197–200. Tingey, R. J. (1991). Geology of Antarctica. Oxford University Press, Oxford, 680 pp. Torii, T. (1981). A review of the Dry Valley Drilling Project 1971–76. Polar Rec., 20, 533–541. Vail, P. R., Mitchum, R. M., Jr., Thompson, S., III, Todd, R. G., Sangree, J. B., Widmier, J. M., Bubb, J. N., & Hatelid, W. G. (1977). Seismic stratigraphy and global sea level changes, 11-part contribution to stratigraphic interpretation of seismic data. In: C. E. Payton (Ed.). AAPG Memoir. Vol. 26, pp. 49–212. Van Der Burgh, J., Visscher, H., Dilcher, D. L., & Ku¨rschner, V. M. (1993). Paleoatmospheric signatures in Neogene fossil leaves. Science, 260, 1788–1790. Vaughan, D. (2005). How does the Antarctic Ice Sheet affect sea level rise? Science, 308, 1877–1878. Webb, P.-N. (1978). RISP Technical. Report, 78-1, University of Nebraska Library, 46 pp. Webb, P.-N. (1979). RISP Technical Report, 79-1, University of Nebraska Library, 127 pp. Webb, P. N. (1983). Climatic, palaeoceanographic and tectonic interpretation of the Palaeogene–Neogene biostratigraphy of the MSSTS-1 drillhole, McMurdo Sound, Antarctica. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Australian Academy of Science, Canberra, p. 560. Webb, P.-N. (1990). The Cenozoic history of Antarctica and its global impact. Antarctic Sci., 2, 3–21. Webb, P.-N., & Cooper, A. K. (1998). Antarctic Offshore Stratigraphy Project (ANTOSTRAT) Report of a Workshop on Antarctic Late Phanerozoic Earth System Science. Hobart, Australia, 6–11 July 1997, SCAR Report 16, 122 pp.
82
P. Barrett
Webb, P.-N., & Harwood, D. M. (1987). Terrestrial flora of the sirius formation: Its significance for late Cenozoic glacial history. Antarctic J. U.S., 22, 7–11. Webb, P.-N., & Harwood, D. M. (1993). Pliocene fossil Nothofagus (Southern Beech) from Antarctica: Phytogeography, dispersal strategies, and survival in high latitude glacial–deglacial environments. In: J. Alden, et al. (Eds). Forest Development in Cold Climates. Plenum, New York, pp. 135–165. Webb, P.-N., Harwood, D. M., McKelvey, B. C., Mercer, J. H., & Stott, L. D. (1984). Cenozoic marine sedimentation and ice-volume variation on the East Antarctic craton. Geology, 12, 287–291. Webb, P.-N., Leckie, M. R., & Ward, B. L. (1986). Foraminifera (Late Oligocene). In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the MSSTS-1 Drillhole, McMurdo Sound, Antarctica, NZ DSIR Miscellaneous Bulletin. Science Information Publishing Centre, Wellington, Vol. 237, pp. 115–126. Webb, P.-N., Ronan, T. E., Lipps, J. H., & Delaca, T. E. (1979). Miocene glaciomarine sediments from beneath the Southern Ross Ice Shelf, Antarctica. Science, 203, 435–437. Webb, P.-N., & Wrenn, J. (1982). Upper Cenozoic biostratigraphy and micropaleontology of Taylor Valley, Antarctica. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, pp. 117–1122. Wilson, G. S., Roberts, A. P., Verosub, K. L., Florindo, F., & Sagnotti, L. (1998). Magnetobiostratigraphic chronology of the Eocene–Oligocene transition in the CIROS-1 core, Victoria Land Margin, Antarctica: Implications for Antarctic glacial history. Geol. Soc. Am. Bull., 110, 35–47. Wilson, J. T. (1965). A new class of faults and their bearing on continental drift. Nature, 207, 343–347. Wingham, D. J., Siegert, M. J., Shepherd, A., & Muir, A. (2006). Rapid discharge connects Antarctic subglacial lakes. Nature, 440, 1033–1036. Wise, S. W., Jr., Breza, J. R., Harwood, D. M., & Wei, W. (1991). Paleogene glacial history of Antarctica. In: J. A. McKenzie, D. W. Muller, & H. Weissert (Eds). Controversies in Modern Geology. Cambridge University Press, Cambridge, pp. 133–171. Wong, H. K., & Christoffel, D. A. (1981). A reconnaissance seismic survey of McMurdo Sound and Terra Nova Bay, Ross Sea. In: L. D. McGinnis (Ed.). Dry Valley Drilling Project, Antarctic Research Series. AGU, Washington, DC, Vol. 81, pp. 37–62. Wrenn, J. H., & Webb, P.-N. (1982). Physiographic analysis and interpretation of the Ferrar Glacier–Victoria Valley area, Antarctica. In: C. Craddock (Ed.). Antarctic Geoscience. The University of Wisconsin Press, Madison, pp. 1091–1099. Wright, C. S., & Priestley, R. E. (1922). British (Terra Nova) Antarctic Expedition, Glaciology. Harrison and Sons, London, 518 pp. Zachos, J. C., Breza, J., & Wise, S. W. (1992). Early Oligocene ice sheet expansion on Antarctica, sedimentological and isotopic evidence from the Kergulen Plateau. Geology, 20, 569–573.
A History of Antarctic Cenozoic Glaciation
83
Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693. Zachos, J. C., Wara, M. W., Bohaty, S., Delaney, M. L., Petrizzo, M. R., Brill, A., Bralower, T. J., & Premoli-Silva, I. (2003). A transient rise in tropical sea surface temperature during the Paleocene–Eocene thermal maximum. Science, 302, 1551–1554. Zwartz, D., Lambeck, K., Bird, M., & Stone, J. (1997). Constraints on the former Antarctic Ice Sheet from sea level observations and geodynamic modelling. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 821–828.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00004-9
Chapter 4
Circulation and Water Masses of the Southern Ocean: A Review Lionel Carter1,, I. N. McCave2 and Michael J. M. Williams3 1
Antarctic Research Centre, Victoria University, P.O. Box 600, Wellington, New Zealand 2 Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge CB2 3EQ, UK 3 National Institute of Water and Atmospheric Research, P.O. Box 14901, Wellington, New Zealand
ABSTRACT The Southern Ocean is a major component of Earth’s ocean and climate. Its circulation is complex, with a zonal Antarctic Circumpolar Current (ACC) interacting with a meridional thermohaline circulation. The ACC is a highly variable, deep-reaching eastward flow driven mainly by the westerly winds. It is the longest (24,000 km), largest (transport 137–147.106 m3 s1) and only current to connect the major oceans. The Ekman component of the westerly winds also drives surface waters north. Near the ACC’s northern limit, these waters sink to form Subantarctic Mode and Antarctic Intermediate waters, which continue north at depths oB1,400 m. Interacting with the ACC is the density-forced thermohaline circulation. Super cooling and increased salinity of shelf waters off the Weddell, Wilkes Land and Ross coasts cause these waters to sink and flow equatorwards. The densest component, Antarctic Bottom Water, is captured in deep basins around Antarctica. Less dense water is entrained by the ACC and mixed with deep water moving south from the Atlantic, Indian and Pacific oceans. The resultant Lower Circumpolar Deep Water is tapped off by deep western boundary currents that enter the three oceans at depths WB2,000 m. These northward inflows, with a total volume transport of B55.106 m3 s1, disperse Antarctic and
Corresponding author. Tel.: þ64 4 463 6475; Fax: þ64 4 463 5186;
E-mail:
[email protected] (L. Carter).
86
L. Carter et al.
northern-sourced waters throughout the world ocean. Other circulation elements are the deep-reaching, cyclonic Weddell, Ross and unnamed gyres located south of the ACC. Further south again are the westward Antarctic Slope and Coastal currents that pass along the Antarctic continental margin under easterly polar winds.
4.1. Introduction The Southern Ocean has a profound influence on the world’s ocean and climate. Cold, dense water sinks to abyssal depths around the margins of Antarctica and migrates northwards into the Atlantic, Indian and Pacific oceans via deep western boundary currents (Fig. 4.1; Stommel, 1958; Warren, 1981). As succinctly noted by Warren (1971), ‘y water from the Antarctic is largely responsible for keeping the rest of the deep sea cold’. Through a process of slow upwelling, these deep cold waters rise to the upper ocean. There, they contribute to the warm surface circulation that extends west from the Pacific and Indian Oceans into the Atlantic where the warm, saline water moves north. Approaching high northern latitudes, the water cools and sinks to form North Atlantic Deep Water (NADW), which migrates south, sandwiched between northbound Antarctic Intermediate Water (AAIW) above and Antarctic Bottom Water (AABW)/Lower Circumpolar Deep Water (LCDW) below (Fig. 4.2). En route, NADW mixes with other waters and eventually rises at the Antarctic continental margin. Thus, one cycle of the global thermohaline circulation (THC) – a major regulator of Earth’s ocean and climate – is completed and another cycle begins (e.g. Broecker, 1991; Schmitz, 1995; Rahmstorf, 2002). This powerful and far-reaching influence of Antarctica and the surrounding Southern Ocean largely reflects; (i) the strong buoyancy-driven and meteorologically forced circulations, and (ii) their direct access to the major ocean basins via the Antarctic Circumpolar Current (ACC) and its offshoots, the deep western boundary currents (Fig. 4.1; Moore et al., 1999; Orsi et al., 1999; Rintoul et al., 2001). In this brief synopsis we can only provide a flavour of over 70 years of oceanographic research in the Southern Ocean. Thus, we refer the reader to the reference list for a more detailed insight into the workings of this region. We present the basic elements under two sections: (1) Section 4.2 examines the main water masses, focusing on their properties and the mechanisms that control their distribution, and (2) Section 4.3 reviews the structure and dynamics of the world’s largest ocean current, the ACC, together with that of the subpolar gyres and
Circulation and Water Masses of Southern Ocean
87
180° 150°W
150°E New Zealand Chatham Rise Campbell Plateau S. Tasman Rise
120°W
120°E Pacific Antarctic R.
cific
E. Pa
Rise
90°E
90°W
Dr Pa ake ss ag e
Un -n
am ed
G.
Ross G.
70˚S 60°S
50°S
F. Pl.
G.
d
We
40°S 60°E
ll de
60°W
SB
SW Indian R.
Mid
SAF
Atl .
ic R
ant
Legend
30°E
Depths <3500 m 30°W ACC
0°
Main DWBC inflow
Figure 4.1: The main Oceanographic elements of the Southern Ocean including: (i) the ACC contained by the Subantarctic Front (SAF) and southern limit of UCDW or southern boundary (SB); (ii) the Ross, Weddell and unnamed subpolar gyres; and (iii) the main exit points of deep western boundary currents from the Southern Ocean (blue arrows). The general path for the ACC is from Orsi et al. (1995) with modifications based on Heath (1985) and Morris et al. (2001). Bathymetric elevations are Annotated as R., ridge; K. Pl., Kerguelen Plateau; and F. Pl., Falkland Plateau. The base chart is Modified from Orsi and Whitworth (2005). currents residing south of the ACC, and the deep THC. The chapter ends with a discussion in Section 4.4 of the present debate regarding the Southern Ocean’s response to a rapidly warming climate (e.g. Gille, 2002; Curry et al., 2003; Jacobs, 2004; IPCC, 2007).
88
L. Carter et al.
ACC 0
ASF SB
SF
STF
SAF
PF
SAMW
AAIW
1000 UCDW
DEPTH(m)
2000 3000 4000
Antarctic Margin
NADW
AABW Mid-Ocean Ridge LCD
W
5000
70°S
60°S
50°S
40°S
LATITUDE
Figure 4.2: Schematic section of the main water masses and their meridional transport as compiled from Whitworth (1988); Orsi et al. (1995); Speer et al. (2000) and Rintoul et al. (2001). Water masses are SAMW, Subantarctic Mode Water; AAIW, Antarctic Intermediate Water; UCDW, Upper Circumpolar Deep Water; LCDW, Lower Circumpolar Deep Water; NADW, North Atlantic Deep Water; AABW, ‘true’ Antarctic Bottom Water (gnW28.27 kg m3). Frontal systems are ASF, Antarctic Slope Front; SB, Southern Boundary of the ACC; SF, Southern Front; PF, Polar Front (formerly the Antarctic Convergence), SAF, Subantarctic Front; STF, Subtropical Front (formerly Subtropical Convergence). The flow of ACC is directed towards the reader.
4.2. Water Mass Formation and Dispersal 4.2.1. Surface Ocean A series of ocean fronts – narrow, variable bands defined by abrupt changes in water properties, in particular, temperature and salinity – divide the surface waters of the Southern Ocean into several zones (e.g. Gordon, 1975; Deacon, 1982; Whitworth, 1988). Early studies identified (from south to north) the Polar, Subantarctic and Subtropical fronts (Fig. 4.2). More recent hydrographic transects, especially those carried out during the World Ocean Circulation Experiment (WOCE), have revealed additional boundaries located south of the Polar Front, and termed the ‘southern’ and ‘southern
Circulation and Water Masses of Southern Ocean
89
boundary’ fronts (Figs. 4.2 and 4.3; Orsi et al., 1995; Orsi and Whitworth, 2005). Furthermore, these detailed and sometimes repeated transects, along with satellite-borne observations of ocean height, temperature and drifter tracks, reveal the complex and dynamic character of the frontal systems 180° 150°W
150°E 40°S
Tasman Basin
50°S
STF SW Pacific Basin
120°W
SAF
120°E
60°S
Ba si n
P14
SF
c E. Pa
ise ific R
lia n-
An t
ar ct ic
70°S
Au st ra
SE Pacific Basin
90°E
PF
len ue rg a u e K ate Pl Enderby Basin
ll de ed W s in Ba
Crozet Basin
60°E
90°W
60°W
SB Madg. Basin Mozb. Basin
Argentine Basin
S2
Legend 30°E
Cape Basin
30°W
Depths <3500 m
0°
Figure 4.3: Location of the principal ocean frontal systems in the Southern Ocean (based on Orsi et al., 1995, but modified for the New Zealand region according to Carter et al., 1998 and Morris et al., 2001). Repeated hydrographic transects, satellite observations and drifting floats reveal the frontal systems as dynamic features with marked temporal and spatial variability but generally within the constraints imposed by the ocean floor topography (Moore et al., 1999). P14 and S2 are locations of WOCE hydrographic transects portrayed in Figs. 4.4 and 4.5. Madg., Madagascar Basin; Mozb., Mozambique Basin. Names of fronts are given in Fig. 4.2. The base chart is modified from Orsi and Whitworth (2005).
90
L. Carter et al.
(Hofmann, 1985; Davis, 1998; Moore et al., 1999; Rintoul et al., 2001; Kostianoy et al., 2004; Sokolov and Rintoul, 2007). While cognizant of these complexities, the main fronts can still be used to define the distribution of three major surface waters characterised mainly by their potential temperature (y), salinity (S) and oxygen content (see hydrographic charts in Orsi and Whitworth, 2005). (1) Near-freezing and relatively fresh Antarctic Surface Water (AASW) forms a layer about 100 m thick that extends from the Antarctic continental shelf to the Polar Front, commonly defined as the northernmost extent of the subsurface temperature minimum (Belkin and Gordon, 1996; Figs. 4.4 and 4.5). AASW temperatures are typically o01C, but may rise to 2.51C near the Front or where warm, deep water approaches the surface (Gordon, 1975; Deacon, 1982). Salinity (S) varies regionally with highest values of SW34.3 psu found in the Ross and Weddell seas, whereas AASW elsewhere around Antarctica commonly has So34.0 psu. (2) Between the Polar and Subantarctic fronts resides surface water that is transitional between AASW and Subantarctic Surface Water (SASW). The structure is complex in response to mixing and interleaving of AASW as it sinks near the Polar Front (e.g. Gordon, 1975; Rintoul et al., 2001). Thus, properties are variable, but generally S is B34.0–34.4 psu and y is 3–81C (Fig. 4.5). (3) SASW occurs north of the Subantarctic Front and encompasses water that usually warms northwards from B61C to 121C. Salinity is typically W34.3 psu except in the SE Pacific and Drake Passage where values decline to o34.16 psu. Like its more southern counterpart, SASW may be affected by vertical mixing as surface waters subduct and mix (e.g. Morris et al., 2001). The northern limit of subantarctic waters is the Subtropical Front where temperatures increase sharply by 4–51C and salinity by 0.5 psu (Fig. 4.3; Deacon, 1982). Subtropical surface water prevails north of the Subtropical Front. 4.2.2. Subantarctic Mode Water and Antarctic Intermediate Water Isopycnals – surfaces of constant density – of near-surface to deep waters in the Southern Ocean rise up in a step-like profile towards Antarctica. Close to ocean fronts, isopycnals may outcrop indicating either the rise of deep waters to the ocean surface or the descent of surface waters (Figs. 4.2, 4.4 and 4.5). In the case of the latter, winter cooling and mixing of the surface waters on the northern side of the Subantarctic Front forms Subantarctic Mode Water (SAMW) (McCartney, 1977; Morris et al., 2001; Rintoul et al., 2001). This well mixed, ventilated water descends northwards to B500 m depth along much of the front (Fig. 4.2). AAIW also descends from the surface, passing
Figure 4.4: Sections of potential temperature (A), salinity (B) and neutral density (C) from WOCE Line P14 across a major constriction in the ACC between the Ross Sea and New Zealand (see Fig. 4.3 for location). Isolines at the Antarctic margin indicate descent of dense shelf waters, which may be mixed with NADW-influenced, LCDW as suggested by the salinity field. The resultant AABW (cf. Fig. 4.5) is contained within the SW Pacific Basin. At the surface, north of the polar front, low salinity AAIW descends northwards (Fig. 4.4B). Hydrographic profiles were derived from the WOCE Southern Ocean Atlas at http://woceatlas.tamu.edu/. Names of fronts are given in Fig. 4.2.
92
L. Carter et al.
Depth (m)
[C]
°
°
°
°
AABW
LC DW
NEUTRAL DENSITY
Distance (km)
Figure 4.4: (Continued). under SAMW to reach a maximum depth of B1,400 m (Figs. 4.2 and 4.4). AAIW is identified by a salinity minimum (34.3–34.5 psu) and temperatures of B3–71C. However, the processes driving AAIW formation are unclear. Formation may be related to wind-forced or density-driven sinking of cold AASW and indeed there appears to be continuity between the AASW and AAIW salinity fields (Figs. 4.4 and 4.5). However, McCartney (1977) suggested AAIW may evolve at least in part from dense SAMW. Whatever the origin, compared to the widespread formation of SAMW, new AAIW presently appears to form mainly in the SW Atlantic and SE Pacific. From these sites AAIW circulates the oceans in anticyclonic subtropical gyres that extend north towards and locally beyond the equator before returning south as ‘old’ AAIW, which is transported within western boundary currents (Rintoul et al., 2001; Ridgway and Dunn, 2007). 4.2.3. Circumpolar Deep Water The most voluminous water mass in the Southern Ocean is Circumpolar Deep Water (CDW). It extends from B1,400 m to W3,500 m depth, but it
34.5
34.4 34.7
–
34.2
AASW
AA
34.68
IW
34.3 34.4
34.5 34.3
1000
34.4
34.6 34.66 34.71 34.74 34.76 34.78
NA DW -LC DW
34.66
AABW
3000
34.71
34.68
34.8
34.72
34.73 34.74 34.78 34.76
4000
LCDW 34.73
5000
Cape Basin We ddell Basin
African Margin
2000
Depth (m)
— – – – – –
40.0°S
– – – – – – – – — – – – – – – – – – — – – – – – – –
0
–
50.0°S
60.0°S
STF
SAF
PF
[A]
SALINITY
34.68
6000
0
4000 40.0°S
–
– – – – – – – – — – – – – – – – – – — – – – – – – –
[B]
3000
50.0°S
— – – – – –
2000
60.0°S
–
1000
0
28 27
28.12 27.4
28.2
27.2
28.2
1000
27.6 27.8 28.26 28
2000
Depth (m)
28.32
AABW 3000
28.12
28.36
28.2
4000
5000 28.4 28.4
NEUTRAL DENSITY
28.26
6000 0
1000
2000
3000
4000
Distance (km)
Figure 4.5: Hydrographic sections of the salinity (A) and neutral density (B) fields from WOCE Line S2 in the Atlantic sector of the Southern Ocean. Of note is (i) the step-wise ascent of saline, NADW-rich LCDW towards Antarctica where it is capped by fresh, cold AASW in the south and subducting AAIW in the north and (ii) the containment of a large volume of classic Antarctic Bottom Water (AABW; gnW28.27 kg m3) within the Weddell Basin (cf. Fig. 4.4). Hydrographic profiles are derived from the WOCE Southern Ocean Atlas at http://woceatlas.tamu.edu/. Names of fronts are given in Fig. 4.2.
94
L. Carter et al.
rises to meet AASW or even outcrop along the Antarctic continental margin (Figs. 4.2, 4.4 and 4.5). CDW has two basic types: (1) Upper Circumpolar Deep Water (UCDW) is identified by the oxygen minimum and elevated nutrient concentrations, and has an open ocean depth range of B1,400– 2,500 m, and (2) Lower Circumpolar Deep Water (LCDW) whose signature is the salinity maximum (34.70–34.75 psu) (Gordon, 1975; Orsi et al., 1995). This maximum reflects the input of NADW that has migrated into the Southern Ocean (Orsi et al., 1995; Rintoul et al., 2001). Upon meeting the ACC, NADW is entrained and transported east around the Antarctic continent, all the while mixing with waters from the Indian and Pacific oceans plus dense waters from Antarctica to form LCDW. Despite the vigorous mixing, the high-salinity signature of NADW is retained (Reid and Lynn, 1971; Gordon, 1975). At several locations around Antarctica, LCDW rises at the continental slope where mixing with super-cold shelf water renews not only the deep circulation of LCDW but also generates Antarctic Bottom Water (AABW), the deepest water mass in the Southern Ocean (Figs. 4.2, 4.4 and 4.5; Foster and Carmack, 1976; Jacobs et al., 1985; Orsi et al., 1999). LCDW is carried equatorwards in all three major oceans by deep western boundary currents (see Section 4.3.3 and Schmitz, 1995; Hogg, 2001). According to Rintoul et al. (2001) mixing with fresher waters, together with the biological depletion of oxygen, slowly modify LCDW into a less dense, lower oxygenated variant that returns south as UCDW. In some WOCE sections, such as P15 in the Indian Ocean, oxygen depletion of UCDW may be influenced by the direct injection of nutrient-enriched deep waters from the North Indian and North Pacific oceans. 4.2.4. Antarctic Bottom Water In their analysis of abyssal water masses, Mantyla and Reid (1983) drew attention to the often inappropriate use of the term Antarctic Bottom Water, which has tended to be used generically for any southern-sourced bottom water. They demonstrated that true AABW did not extend far from Antarctica before mixing with other waters (see also Orsi et al., 1999). This is particularly true for the deep western boundary currents in which AABW is mixed with CDW derived from the ACC. Thus, at 301N in the NW Atlantic, Amos et al. (1971) recorded o20% AABW near the seabed. To better characterise AABW and thereby improve assessments of its dispersal and contribution to bottom waters worldwide, Orsi et al. (1999) defined AABW by its neutral density (gn) whereby gnW28.27 kg m3. Such dense waters are confined mainly to the deep (down to B6,000 m),
Circulation and Water Masses of Southern Ocean
95
circum-Antarctic Basins that include the Argentine and Brazil basins (SW Atlantic), Mozambique, Crozet and Australian-Antarctic basins (Indian) and the SE Pacific Basin (Figs. 4.1, 4.4–4.6). Less dense, southern bottom water with 28.18ogno28.27 kg m3, is not confined to the circum-Antarctic basins, but instead spreads out from the deep levels of the northern edge of the ACC into all major oceans (Orsi et al., 1999; also see Section 4.3.3). AABW density is determined by a combination of different sources that produce regionally distinct waters (Mantyla and Reid, 1983; Orsi et al., 1999; Jacobs, 2004). The freshest and coldest (Sr34.64 psu; yr11C) bottom water occurs in the Weddell Sea, whereas the SE Pacific Basin has the most saline and least cold (SZ34.72 psu; 0.6oyo0.31C) bottom water. The Australian-Antarctic Basin contains water with properties intermediate between the two end members. Traditionally, the Weddell Sea was regarded as the prime source of AABW, but recently two other sources have come to the fore. The Weddell Sea’s contribution is now regarded as B50%, with the Wilkes Land margin including the Ade´lie coast (Rintoul, 1998) contributing B30% mainly to the Indian Ocean, and the Ross Sea producing B20% AABW that is destined primarily for the SE Pacific Basin (Jacobs, 2004). Equally varied are the modes of bottom water formation (Jacobs, 2004). Foster and Carmack (1976) invoked the formation of highly saline shelf water (HSSW) by brine rejection from sea ice. However, salt-driven increases in density may also be influenced by intrusions of NADW-bearing LCDW onto the upper slope and shelf (Toggweiler and Samuels, 1995). Super cold, Ice Shelf Water (ISW), formed by freezing and melting below ice shelves, can mix with HSSW and reach the outer shelf before flowing down slope (Baines and Condie, 1998). Alternatively, the simple mixing of cold AASW and highsalinity LCDW, with or without ISW, may produce negatively buoyant waters at the shelf edge (e.g. Jacobs, 2004). Finally, dense waters may sink via convection chimneys and polynyas such as the well-documented but short-lived Weddell Sea polynya (Gordon, 1982). Rates of AABW formation, as estimated from hydrographic data, usually fall within a range of 5–15 Sv (1 Sv ¼ 106 m3 s1). Anthropogenic tracers, in particular chlorofluorocarbons, record a flux of 8.1 Sv for AABW descending at the 2,500 m isobath off Antarctica (Orsi et al., 2001). This compares to 7.6 Sv of lower NADW flowing out of the Nordic and Labrador seas at the N Atlantic source. Because AABW is colder than its northern counterpart, Antarctic overturning probably plays the dominant role in cooling the deep ocean. The time at which deep water circulates through the ocean is often quoted to be a millennium or more, for example 1,000 years between the N Atlantic and Southern Oceans and a further 1,000 years from the Southern
96
L. Carter et al.
Figure 4.6: Extent of dense Antarctic Bottom Water as identified by the neutral density field at 3500 m (Orsi et al., 1999; Orsi and Whitworth, 2005). With the exception of the SE Atlantic, where AABW extends well north via the deep Argentine and Brazil basins, the remaining AABW is captured in circum-Antarctic basins where further northwards dispersal is inhibited by oceanic ridge systems shown in white. Basins annotated as Arg. B., Argentine Basin, which extends north into the Brazil basin (not shown on chart); W.B., Weddell Basin; E.B., Enderby Basin; C.B., Crozet Basin; Aust. A.B., Australian-Antarctic Basin, and SE P.B., SE Pacific Basin. Chart is generated from the WOCE Southern Ocean Atlas at http://woceatlas.tamu.edu/.
Circulation and Water Masses of Southern Ocean
97
Ocean to the North Pacific. However, in a re-analysis of radiocarbon dated ocean waters, Matsumoto (2007) indicates much shorter circulation ages thereby supporting but refining earlier radiocarbon-based studies (Stuiver et al., 1983). Thus, the circulation age for the Southern Ocean below 1,500 m is B300 14C years with a similar age for the Atlantic. For the Pacific, the basin circulation age is B900 14C years.
4.3. Ocean Circulation 4.3.1. Antarctic Circumpolar Current The Southern Ocean circulation is dominated by the ACC, a current system that is rightly described by superlatives. It is the only current to connect the major ocean basins and hence plays a prominent role in the global distribution of heat, salt and gases (Fig. 4.1). It is the longest current with an estimated pathway of B24,000 km (Whitworth, 1988). Finally, the ACC is the largest major current in terms of volume transport with a mean of 136.777.8 Sv as measured across Drake Passage (Cunningham et al., 2003). The ACC was originally termed the West Wind Drift (Deacon, 1937) in recognition of its forcing by middle-to-high latitude westerly winds (Orsi et al., 1995; Whitworth et al., 1998). However, use of the term wind drift masks the role played by the buoyancy-driven component of the circulation (see Rintoul et al., 2001). Complexities aside, the net result is an eastward current system that extends from the ocean surface to bottom, its path guided by submarine topography (Fig. 4.1; Gordon et al., 1978; Orsi et al., 1995). For much of its passage, the ACC flows along the flanks of midoceanic ridges except within major gaps in the Pacific and Indian ridge systems where the current shifts poleward (Fig. 4.1). Large submarine plateaux also exert an influence. The ACC widens to the north and south as it passes around the Kerguelen Plateau, whereas the Campbell and Falkland plateaux form constrictions (Fig. 4.1; Whitworth and Peterson, 1985; Morris et al., 2001). Interestingly, this interaction with the ocean floor was inferred as early as the 1950s. Estimates for a purely wind-driven ACC yielded transports that were excessive compared to observations. Thus, it was concluded that the wind stress was partly balanced by bottom stress (see Whitworth, 1988; Rintoul et al., 2001). The passage of westerly winds over the ACC also induces an Ekman drift to the north – a process that probably plays a role in the subduction and transport of mode and intermediate waters. This equatorward flow is compensated at depth by the southward
98
L. Carter et al.
transport and eventual upwelling of CDW thus contributing to the THC (Wyrtki, 1961; Toggweiler and Samuels, 1995). Rather than a uniform flow, the ACC is a system of deep-reaching zonal jets that separates zones of relatively quiet water. The jets are marked by the circumpolar fronts (see Section 4.2 Surface Ocean) with the northern and southern boundaries of the ACC defined, respectively, by the Subantarctic and Southern Boundary fronts (Figs. 4.1–4.3). Both eddy-resolving models and satellite observations highlight the complex flow of the frontal jets (Morrow et al., 1992; Gille, 1994; Carter and Wilkin, 1999). Meanders, eddies and intricate branches are well shown especially where the flow is constricted as off Campbell Plateau and within Drake Passage (Nowlin and Klinck, 1986; Morris et al., 2001; Cunningham et al., 2003). Most ACC transport takes place within the fronts, but their complex flow patterns and different criteria for estimating transport have led to a wide range of values for the entire ACC. Nevertheless, closely spaced and longterm monitoring sites have improved estimates of transport. In Drake Passage the mean transport is 136.777.8 Sv (Cunningham et al., 2003) compared to 147710 Sv between Australia and Antarctica (Rintoul and Sokolov, 2001). Much of the transport in the Australasian reach of the ACC occurs within the Subantarctic Front, which has a mean of 10577 Sv off Australia (Rintoul and Sokolov, 2001) and B90 Sv off southern New Zealand (Morris et al., 2001). However, ACC transport can be highly variable. Time series from Drake Passage record variations at several time scales (Whitworth and Peterson, 1985). Short-term fluctuations, related to 14-day solar and lunar tides, are superimposed on longer-period fluctuations of B1 year and longer that can lead to changes in transport of B30–40 Sv within weeks. The interaction of the ACC with the topography and southerly extensions of western boundary currents generates eddies that play important roles in the transfer of heat and momentum (Bryden and Heath, 1985; Morrow et al., 1992; Rintoul et al., 2001). Off SE New Zealand, for example interception of the ACC by the South Tasman Rise and Macquarie Ridge spawns bottomreaching eddies that migrate NE along the steep margin of Campbell Plateau (Boyer and Guala, 1972; Gordon, 1972). Both cyclonic and anticyclonic features have been observed from current meter and satellite data, which suggest a frequency of occurrence of B9 eddies annually (Stanton and Morris, 2004). Modelled eddy kinetic energy, verified by current meter data and ocean floor sedimentary evidence, attest to the power of these perturbations, which are likely to be the cause of abyssal benthic storms (Hollister and McCave, 1984; Carter and Wilkin, 1999). Seabed topography also encourages intense mixing within the ACC. The Scotia Sea and
Circulation and Water Masses of Southern Ocean
99
potentially other areas of marked seabed relief below the ACC, are zones of the highest turbulent mixing in the ocean and result in rapid upwelling that may locally short-circuit the classic meridional overturning as portrayed in Fig. 4.2 (Garabato et al., 2004, 2007).
4.3.2. Subpolar Circulation 4.3.2.1. Major gyres A key element of the circulation, south of the ACC, is three large, deepreaching cyclonic gyres that extend from the ACC to the Antarctic continental margin (Fig. 4.1). The better known are the Weddell and Ross gyres that occupy ridge-bounded sectors of the Weddell and Ross seas (e.g. Orsi et al., 1993; Jacobs et al., 2002). A third and as yet unnamed gyre was suggested to occur south of Kerguelen Plateau (Bindoff et al., 2000), and has now been confirmed by McCartney and Donohue (2007). As documented for the Weddell Gyre (Orsi et al., 1993), it is likely that all three gyres transport salt and heat from the ACC to the Antarctic continental margin where deep and bottom waters are produced (e.g. Jacobs, 2004). Furthermore, McCartney and Donohue (2007) suggest a strong connectivity between the three gyres with a westward flow along their southern limbs and an eastward flow joining their northern limbs. This latter flow is just south of another eastward flow, this time associated with an anticyclonic supergyre covering most of the S Pacific and S Indian oceans (Ridgway and Dunn, 2007). This eastward limb of the supergyre appears to reside between the Subtropical and Subantarctic fronts. As a result, it may entrain SAMW, formed in the vicinity of the Subantarctic Front and help distribute it through the ocean basins as suggested by Rintoul et al. (2001). Of these cyclonic circulations, the Weddell Gyre is the largest, extending from B501W to between 201–301E (Fig. 4.1; Orsi et al., 1993). At the surface it has the form of a NE–SW aligned, elongated gyre whereas at depth it comprises two cyclonic cells located to the east and west of 151W. Basically, the Weddell Gyre occupies the W4,000 m deep re-entrant formed by Antarctica and the ridge systems that extend eastward from near the tip of the Antarctic Peninsula (Fig. 4.1). While such a location implies containment, the northern limb of the gyre overlaps the ridges to interact with the ACC. CDW entrained from the ACC is moved within the gyre and can eventually mix with cold shelf waters to form Weddell Sea Bottom Water, the local variety of AABW and the densest water in the Southern Ocean (Foster and Carmack, 1976).
100
L. Carter et al.
The Ross Gyre extends from 1601E to 1401W, and is largely confined to the W4,000 m deep western reach of the SE Pacific Basin (Fig. 4.1). Like its Weddell Sea counterpart, the Ross Gyre is a deep-reaching feature that entrains CDW to make it available for mixing with shelf and slope waters. Situated to the south and southwest of Australia, the unnamed gyre appears to favour the southern part of the Australian-Antarctic Basin, most of which is W4,000 m deep (Fig. 4.1). Indeed, the volume transport field is compressed against the Antarctic continental margin and the associated westward slope current (Bindoff et al., 2000; McCartney and Donohue, 2007). Transport along the eastward-flowing northern limb of the gyre is estimated at 35 Sv. However, the amount of transport along the westmoving, southern limb is unclear due to merging with the slope current, the combined flows reaching 76 Sv (McCartney and Donohue, 2007).
4.3.2.2. Antarctic slope and coastal currents As summarised by Heywood et al. (2004) much of the Antarctic margin is bathed by two westward currents. One is associated with the Antarctic Slope Front that constitutes the boundary between fresh, cold Antarctic shelf waters and less cold, saline CDW (Jacobs, 1991; Whitworth et al., 1998). Deacon (1937) regarded the frontal flow as a consequence of the prevailing easterly winds and coined the name, East Wind Drift. On the basis of classical theory, he reasoned that polar easterly winds produced an onshore Ekman transport with the resultant generation of a westward geostrophic current below the wind-mixed layer (Whitworth et al., 1998; Bindoff et al., 2000). The second significant feature is the Antarctic Coastal Current, which forms a narrow, rapid flow across broad sections of the continental shelf, for example, in the SW Weddell Sea. However, where the shelf is narrow the coastal current is difficult to distinguish from flows associated with the Antarctic Slope Front and the southern limbs of the subpolar gyres where they approach the continental margin. 4.3.3. Thermohaline Circulation Wunsch (2002) drew attention to the imprecise meaning of the term, THC. Following Jacobs (2004), we prefer the broad definition of Schmitz (1995) whereby the THC is the ‘y buoyancy-driven flow field associated with water cooled (or heated) by contact with cold (or warm) air, or modified by sources and sinks of cold water. May also include flows whose characteristics are
Circulation and Water Masses of Southern Ocean
101
significantly altered by upwelling and/or mixing. Water sinking at high latitudes tends to return equatorwards in relatively strong, narrow currents called DWBC’. The high latitude sinks noted are primarily the North Atlantic and Antarctica (Stommel, 1958; Warren, 1981). As described earlier, dense water formed mainly over shelf areas of the Weddell and Ross seas, and along the Wilkes Land coast, sinks and descends down the continental slope. Descent may take place in several ways according to Baines and Condie (1998). For prominent deep-water formation areas, dense water may descend as broad sheets or plumes. For weaker sources, the resultant outflows may geostrophically descend the slope to a level where the layer thins and viscous drainage prevails. Descending flows may be disrupted by eddies into discrete parcels that overall move down slope. Further complexities are introduced by the slope topography, for example, submarine canyons and channels can contain and steer flows whose density may increase through entrainment of sediment. Whatever the mode of descent, the initial dispersal of AABW is westward as outlined by the distribution of chlorofluorocarbons (Orsi et al., 1999). Those tracers show that the densest AABW usually follows a cyclonic path presumably under the influence of the subpolar gyres and basin topography (Fig. 4.1). However, the Weddell– Enderby Basin experiences limited outflow at its northern rim where least dense AABW (gnB28.28 kg m3) passes north into adjoining basins of the Atlantic and Indian oceans (Fig. 4.6; Orsi et al., 1999). In contrast, the main inflows to the adjoining oceans are via deep western boundary currents, which carry mainly LCDW from the northern boundaries of the ACC (Mantyla and Reid, 1983; Schmitz, 1995).
4.3.3.1. Atlantic DWBC inflow Weddell Sea Bottom Water, a local variety of AABW, together with Weddell Sea Deep Water and CDW from the Drake Passage, are carried by a northbound DWBC into the Atlantic Ocean below southward-moving NADW (Warren, 1981; Schmitz, 1995). The DWBC is typically found in water depths exceeding 3,500–4,000 m along the western boundary presented by the continental margin off South America. However, the current pathway is interrupted by a succession of deep basins including (from south to north) Georgia, Argentine and Brazil basins. As noted earlier, densest AABW is captured within the deep basins leaving less dense waters to move north via gaps and channels through the inter-basin ridges. Even so, not all the deep water escapes and some recirculates within the basins themselves (Hogg and Johns, 1995; Hogg, 2001). For example, around 6.9 Sv flows from the
102
L. Carter et al.
Argentine Basin into the Brazil Basin through Vema and Hunter channels, but only 3.2 Sv leaves the Brazil Basin, leaving 3.7 Sv to recirculate.
4.3.3.2. Indian DWBC inflow The inflow of AABW into the western Indian Ocean is via Crozet and Mozambique basins (the latter being a dead end), which largely constrain AABW to south of 301S and 341S, respectively (Orsi et al., 1999). LCDW from the Crozet Basin leaks northwards through fractures zones in the SW Indian Ridge (Warren, 1974; Johnson et al., 1991; Mantyla and Reid, 1995). Deep transport is via a northbound DWBC, which like its Atlantic counterpart, has a complex pathway dictated by multiple basins and ridges (Warren, 1981; McCave et al., 2005). Basins encourage recirculation of deep waters and together with mixing, account for a northward dissipation of the DWBC. For instance, LCDW below B3,800 m has a northwards transport of 3.8 Sv through Madagascar Basin, but only 1.7 Sv exits north into the Somali Basin (Johnson et al., 1998). Crozet Basin, in the western Indian Ocean, is not the only gateway for DWBCs into the Indian Ocean. The eastern Indian Ocean is also connected to the Southern Ocean thereby allowing the northward intrusion of two other boundary currents; one along the eastern side of SE Indian Ridge and the other along the eastern flank of Ninetyeast Ridge (Warren, 1981; Toole and Warren 1993; Reid, 2005). Again, the deep water carried north is LCDW, with the denser AABW retained in Australian-Antarctic Basin (Orsi et al., 1999). When extended to the 2,000 dbar reference level, the combined northward transport of the three DWBCs into the Indian Ocean is B27 Sv (Toole and Warren, 1993).
4.3.3.3. Pacific DWBC inflow The general pathway of the DWBC into the Pacific was first outlined in the classic model of the global abyssal circulation by Stommel (1958) and Stommel and Arons (1960), and was later confirmed by the hydrographic sections of Warren (1971, 1973). Because the Tasman Basin is essentially closed at its northern end, the main inflow is off southernmost New Zealand where the DWBC enters in concert with the ACC (Fig. 4.7; Carter and McCave, 1997; Carter and Wilkin, 1999). Initially, the combined inflow intercepts Macquarie Ridge to form meanders and eddies although some current filaments pass through narrow gaps in the ridge (Boyer and Guala,
Circulation and Water Masses of Southern Ocean
103
SINKING
Thermohaline Circulation Cold deep current Cool shallow current Warm shallow current
ELLING UPW
U
G LIN EL
ING ELL PW
UP W
A.C. C.
C. A. C .
SINKING SINKING
SINKING
Figure 4.7: A schematic portrayal of the THC, which is a series of loosely linked, recirculation systems that transport, heat, salt, nutrients and ventilating gases through the world ocean. Only the main elements of the THC are shown, for example the Indian Ocean has three, deep northward inflows that include from W to E, the margin off eastern Africa (shown), and the eastern sides of the SE Indian Ridge and Ninetyeast Ridge (not shown) (image modified from Manighetti, 2001).
1972; Gordon, 1972). This perturbed combined flow continues northeast along the 3,000–3,500 m high flanks of Campbell Plateau to around 491S where the ACC veers east leaving the DWBC to continue northwards into the Pacific Ocean (Fig. 4.7). It eventually departs the Southern Ocean off Chatham Rise between 441S and 421S (McCave and Carter, 1997). There, Warren (1973) observed a volume transport of B20 Sv, the largest for a single DWBC (Schmitz, 1995).
4.4. Oceanographic Variability and Change The present phase of climate change has drawn considerable attention to the behaviour of Antarctica and the Southern Ocean (e.g. Gille, 2002; Jacobs et al., 2002; Walther et al., 2002; Cook et al., 2005). However, confident identification of any effect, especially in the oceans, has been
104
L. Carter et al.
hindered by; (i) sampling bias, (ii) the short history of observations, which typically only encompass the last 50–60 years, (iii) a multiplicity of forcing mechanisms, some with and some without clear connections to a warming climate and (iv) marked variability at a range of temporal and spatial scales (e.g. Jacobs and Giulivi, 1998; Orsi et al., 2001; Vaughan et al., 2001). Of special note is the Southern Annular Mode (SAM), which appears to dominate inter-annual to centennial variability in the Southern Ocean (Hall and Visbeck, 2002; Lovenduski and Gruber, 2005; Lovenduski et al., 2007). SAM is essentially a zone of climate variability that encircles the South Pole and strongly influences zonal winds, sea ice formation and oceanic circulation. When in a positive phase, SAM is typically associated with enhanced westerly winds over the ACC and weakened winds further north. This favours a strengthening of the ACC, a reinforcement of the northward Ekman drift and subduction of surface waters, and a northward expansion of sea ice. To compensate, there is enhanced poleward transport and rising of deep water at the Antarctic continental margin. Under a negative SAM, windiness and storminess appear to migrate to mid-latitudes thus weakening both zonal and meridional ocean transport. Despite uncertainties associated with the aforementioned limitations, it is nonetheless important to examine and evaluate recent changes in Antarctica and the Southern Ocean in light of their actual or potential influence on the global ocean and climate. Climatic trends for the past 50 years have been identified by Turner et al. (2005) using records from 19 Antarctic meteorological stations. The results emphasise the marked geographic variability of the continental climate (e.g. Vaughan et al., 2001) as well as its temporal variability at interdecadal scales. The Antarctic Peninsula has warmed at a statistically significant rate of þ0.561C/decade from 1951 to 2000. The next largest warming trend outside of the Antarctic Peninsula is in the western Ross Sea (Scott Base) with a rise of þ0.291C/decade, although the rise is not statistically significant. Elsewhere, significant trends are unclear. Annual temperature trends for coastal and interior sites on the East Antarctic Ice Sheet suggest a slight cooling. However, all but one site exhibited warmer winters. Notwithstanding its temporal and spatial variability, the upper Southern Ocean has warmed between 1955 and 2003 in concert with the world ocean (Levitus et al., 2005). Off the Antarctic Peninsula, the pronounced atmospheric warming has been accompanied by an equally marked warming of the surface ocean with summer temperatures increasing by 1.21C over the latter half of the twentieth century (Meredith and King, 2005). At water depths of 700–1,100 m, Gille (2002) reported an average 0.171C rise since 1950, which accounts for about two-thirds of the total increase in heat
Circulation and Water Masses of Southern Ocean
105
content in the ocean from 0 to 3,000 m depth (IPCC, 2007). Much of the warming is concentrated within the Subantarctic Front of the ACC. This raises the possibility that the warming at depth may result from the sinking of atmospherically warmed SAMW. Gille (2002) further suggests that warming may also be related to a general southward displacement of the ACC. This would be consistent with a suite of model simulations that suggest warming of the Southern Ocean will be accompanied by a southward shift of zonal westerly winds together with a narrowing and intensification of the ACC (IPCC, 2007 and references therein). Coherent historical trends are also evident for salinity. Subpolar regions have generally become fresher between 1955 and 1998 in contrast to subtropical and tropical regions, which display increased salinity with the exception of the central Pacific Ocean (Curry et al., 2003; Boyer et al., 2005). Such freshening of the upper Southern Ocean may be responsible for a reduction in the salinity of AAIW (Wong et al., 1999; Curry et al., 2003). Reduced salinity implies increased freshwater input that may result from one or more of the following causes: (i) greater net precipitation; (ii) changes in the extent of sea ice – a major contributor to winter salinity through brine rejection; (iii) increased melting of ice shelves, ice sheets and glaciers (e.g. Cook et al., 2005), and (iv) changes in the oceanography, especially any reduced upwelling of saline LCDW (Wong et al., 1999; Jacobs et al., 2002; Curry et al., 2003). Nearer Antarctica, hydrographic records spanning over 40 years show local variability in salinity trends. The upper 50 m of the ocean, west of the Antarctic Peninsula, has become more saline although underlying waters have freshened slightly in line with the regional trend. While the more saline surface conditions appear to be out-of-step with the strong glacier retreat on the Peninsula (e.g. Cook et al., 2005), Meredith and King (2005) suggest more saline conditions are consistent with the reduced sea ice cover and the timing of the hydrographic measurements. With less sea ice production there is less freshening of the ocean in the summer when most of the hydrographic measurements are made. On the opposite side of the continent, at Law Dome, ice core records spanning a century and longer, identify a 20% loss of sea ice since 1950 although this trend is strongly overprinted with cyclical variations with an B11 year frequency (Curran et al., 2003). Reduced sea ice along with increased precipitation and melt water from the West Antarctic Ice Sheet have been cited as contributing to the observed freshening of surface waters associated with the Ross Sea Gyre (Jacobs et al., 2002). But like Law Dome, the trend is obscured by cycles, this time by 5–6- and 9-year oscillations in HSSW formation (Assmann and Timmerman, 2005).
106
L. Carter et al.
Salinity and temperature changes have the potential to affect bottom water production and the THC. Consequently, these changes have received considerable attention (IPCC, 2007 and references therein). At glacialinterglacial time scales, palaeoceanographic evidence reveals marked variations in the position and degree of convective overturning of the N Atlantic sector of the THC (Rahmstorf, 2002 and references therein). During interglacial periods, overturning is most active and reaches its most northerly extent. In contrast, glacial periods are likely to witness a southward shift and possible slow-down in overturning. Any slow-down may be compensated by a greater production of bottom water from Antarctica as suggested by grain size (Hall et al., 2001), magnetic properties (Venuti et al., 2007) and diatom proxies (Stickley et al., 2001). However, such conclusions are sometimes at odds with geochemical tracers such as d13C (e.g. Moy et al., 2006) that point to little change in the passage of NADW through the Indian and Pacific sectors at least over recent glacial-interglacial cycles. At millennial time scales, abrupt changes such as those associated with Heinrich events, may stop N Atlantic overturning altogether as the density of the surface ocean is reduced by rapid influxes of melt water (Rahmstorf, 2002). Again, cessation of N Atlantic production may be compensated by enhanced Antarctic production. However, responses to the latest phase of climate/ocean warming are unclear. In the N Atlantic, which is the best observed deepwater source, long-term trends are equivocal due to decadal variability, a paucity of long-term observations and other factors (IPCC, 2007). A similar situation applies to Antarctica where estimations of bottom-water production are inconsistent in response to: (i) natural cycles; (ii) differences in the definitions and techniques to estimate production rates, and (iii) a bias towards summer observations (Jacobs, 2004). On the basis of chlorofluorocarbons and 14C data, which allow water masses to be traced at decadal to century scales respectively, Orsi et al. (2001) revealed no decline in bottom water production over the twentieth century as indicated earlier by Broecker et al. (1999). Nevertheless, the changes recorded in recent historical times cannot be ignored. The freshening of the Ross Gyre over the last 50 years (Jacobs et al., 2002) and an accompanying downstream freshening of AABW in the adjacent Australia-Antarctic Basin (Aoki et al., 2005) are consistent with increased freshwater input. On a larger scale, the historical salinity data of Curry et al. (2003) also reveals a freshening of deep and bottom water at Antarctic and N Atlantic sources. Simulations by 19 model runs under IPCC greenhouse gas scenario, A1B (rapid economic growth, world population peaks mid-century, new and efficient energy technologies with reliance on a range of sources) point to an average 25% reduction in N Atlantic overturning by the year 2100 (IPCC, 2007).
Circulation and Water Masses of Southern Ocean
107
None of the runs point to a shut-down; rather they favour reductions in overturning of up to 50%. While the Southern Ocean sector has received less attention from modellers, simulations based on a warmer or fresher ocean may enhance or stabilise N Atlantic overturning (Saenko et al., 2003; Weaver et al., 2003). To further emphasise the complexity of north–south relationships, the projected strengthening of the Southern Hemisphere westerly winds will increase the northward Ekman transport of upper ocean. To compensate, the poleward flow of deep water below 2,000–2,500 m depth, could be expected to strengthen and possibly stimulate the southward flow of NADW (e.g. Toggweiler and Samuels, 1995; Toggweiler et al., 2006). Again, such a trend is overprinted with marked inter-annual variability. Because of the importance, size and complexity of the Southern Ocean, the incompleteness of observations, and its high variability at a range of temporal and spatial scales, it is critical to improve our knowledge of this ocean/climate system. To re-emphasise the introduction to this chapter, the Southern Ocean has a profound influence of the distribution of salt, heat and ventilating gases throughout global seas. At the same time it is also undergoing some of the most rapid environmental changes on Earth highlighted by the warming, glacial retreat and ice shelf collapse around the Antarctic Peninsula. Thus, to address the inevitable questions relating to impacts of a rapidly changing climate on the Southern Ocean we require a strong modelling effort, supported by multi-seasonal oceanographic and remotely sensed observations and high-resolution palaeoceanographic records of past warm extremes. While this may seem to be a well-worn message, it is still appropriate at a time of certain change and uncertain consequences.
ACKNOWLEDGEMENTS We are indebted to the World Ocean Circulation Experiment (WOCE) for permission to use their data for Figures 4.1, 4.3 to 4.6, which are attributed to Orsi, A. H., T. Whitworth III, Hydrographic Atlas of the World Ocean Circulation Experiment (WOCE). Volume 1: Southern Ocean (eds. M. Sparrow, P. Chapman and J. Gould), International WOCE Project Office, Southampton, U.K., ISBN 0-904175-49-9, 2005. The chapter benefited from the critiques by the external reviewers, Will Howard and Alejandro Orsi and their input is appreciated. Funding for L. Carter was provided by the Foundation for Research Science and Technology, contracts CO50410 and VICX0704.
108
L. Carter et al.
REFERENCES Amos, A. F., Gordon, A. L., & Schneider, E. D. (1971). Water masses and circulation patterns in the region of the Blake-Bahamas Outer Ridge. Deep Sea Res., 18, 145–165. Aoki, S., Rintoul, S. R., Ushio, S., & Watanabe, S. (2005). Freshening of the Adelie Land bottom water near 1401E. Geophys. Res. Lett., 32, L23601, doi:10.1029/ 2005GL024246. Assmann, K. M., & Timmerman, R. (2005). Variability of dense water formation in the Ross Sea. Ocean Dyn., 55, 68–87. Baines, P. G., & Condie, S. (1998). Observations and modelling of Antarctic downslope flows: A review. In: S. S. Jacobs, & R. Weiss (Eds). Ocean, Ice, and Atmosphere: Interactions at the Antarctic Continental Margin. Antarctic Research Series. Vol. 75, pp. 29–49. Belkin, I. M., & Gordon, A. L. (1996). Southern ocean fronts from the Greenwich Meridian to Tasmania. J. Geophys. Res., 101, 3675–3696. Bindoff, N. L., Rosenberg, M. A., & Warner, M. J. (2000). On the circulation and water masses over the Antarctic continental slope and rise between 80 and 1501E. Deep Sea Res., 47, 2299–2326. Boyer, D. L., & Guala, J. R. (1972). Model of the Antarctic Circumpolar Current in the vicinity of the Macquarie Ridge. In: D. E. Hayes (Ed.). Antarctic Oceanology II – The Australian-New Zealand Sector. Antarctic Research Series. Vol. 19, pp. 79–94. Boyer, T. P., Levitus, S., Antonov, J. I., Locarnini, R. A., & Garcia, H. E. (2005). Linear trends in salinity for the World Ocean, 1955–1998. Geophys. Res. Lett., 32, L01604, doi:10.1029/2004GL021791. Broecker, W. S. (1991). The great ocean conveyor. Oceanography, 4, 79–89. Broecker, W. S., Sutherland, S., & Peng, T.-H. (1999). A possible 20th-century slowdown of the Southern Ocean deep water formation. Science, 286, 1132–1135. Bryden, H. L., & Heath, R. A. (1985). Energetic eddies at the northern edge of the Antarctic Circumpolar Current in the Southwest Pacific. Prog. Oceanogr., 14, 65–87. Carter, L., Garlick, R. D., Sutton, P., Chiswell, S. M, Oien, N.A., & Stanton, B.R. (1998). Ocean Circulation New Zealand. NIWA Chart Miscellaneous Series, 76. Carter, L., & McCave, I. N. (1997). The sedimentary regime beneath the deep western boundary current inflow to the Southwest Pacific Ocean. J. Sedimentary Res., 67, 1005–1017. Carter, L., & Wilkin, J. (1999). Abyssal circulation around New Zealand – A comparison between observations and a global circulation model. Mar. Geol., 159, 221–239. Cook, A. J., Fox, A. J., Vaughan, D. G., & Ferrigno, J. G. (2005). Retreating glacier fronts on the Antarctic Peninsula over the past half century. Science, 308, 541–544.
Circulation and Water Masses of Southern Ocean
109
Cunningham, S.A., Alderson, S.G., King, B.A., & Brandon, M.A. (2003). Transport and variability of the Antarctic Circumpolar Current in Drake Passage. J. Geophys. Res., 108, C5, 8084. Curran, M. A., van Ommen, T. D., Morgan, V. I., Phillips, K. L., & Palmer, A. S. (2003). Ice core evidence for Antarctic sea ice decline since the 1950s. Science, 302, 1203–1206. Curry, R., Dickson, B., & Yashayaev, I. (2003). A change in the freshwater balance of the Atlantic Ocean over the past four decades. Nature, 426, 826–829. Davis, R. (1998). Preliminary results from directly measuring mid-depth circulation in the tropical and South Pacific. J. Geophys. Res., 103, 24619–24639. Deacon, G. E. R. (1937). The hydrology of the Southern Ocean. Discov. Rep., 15, 1–123. Deacon, G. E. R. (1982). Physical and biological zonation in the Southern Ocean. Deep Sea Res., 29, 1–15. Foster, T. D., & Carmack, E. C. (1976). Frontal zone mixing and Antarctic Bottom Water formation in the southern Weddell Sea. Deep Sea Res., 23, 301–317. Garabato, A. C. N., Polzin, K. L., King, B. A., Heywood, K. J., & Visbeck, M. (2004). Widespread intense turbulent mixing in the Southern Ocean. Science, 303, 210–213. Garabato, A. C. N., Stevens, D. P., Watson, A. J., & Roether, W. (2007). Shortcircuiting of the overturning circulation in the Antarctic Circumpolar Current. Nature, 447, 194–197, doi:10.1038/nature05832. Gille, S. T. (1994). Mean sea surface height of the Antarctic Circumpolar Current from Geosat data: Method and application. J. Geophys. Res., 99, 18255–18273. Gille, S. T. (2002). Warming of the Southern Ocean since the 1950s. Science, 295, 1275–1277. Gordon, A. L. (1972). On the interaction of the Antarctic Circumpolar Current and the Macquarie Ridge. In: D. E. Hayes (Ed.). Antarctic Oceanology II – The Australian-New Zealand Sector. Antarctic Research Series. Vol. 19, pp. 71–78. Gordon, A. L. (1975). An Antarctic oceanographic section along 1701E. Deep Sea Res., 22, 357–377. Gordon, A. L. (1982). Weddell deep water variability. J. Mar. Res., 40, 199–217. Gordon, A. L., Molinelli, E., & Baker, T. (1978). Large scale relative dynamic topography of the Southern Ocean. J. Geophys. Res., 83, 3023–3032. Hall, A., & Visbeck, M. (2002). Synchronous variability in the Southern Hemisphere atmosphere, sea ice and ocean resulting from the Annular Mode. J. Clim., 15, 3043–3057. Hall, I. R., McCave, I. N., Shackleton, N. J., Weldon, G. P., & Harris, S. E. (2001). Glacial intensification of deep Pacific inflow and ventilation. Nature, 412, 809–812. Heath, R. A. (1985). A review of the physical oceanography of the seas around New Zealand – 1982. N. Z. J. Mar. Freshwater Res., 19, 79–124.
110
L. Carter et al.
Heywood, K.J., Garabato, A.C.N., Stevens, D.P., & Muench, R.D. (2004). On the fate of the Antarctic Slope Front and the origin of the Weddell Front. J. Geophys. Res., 109, C0621, 1–13. Hofmann, E. E. (1985). The large-scale horizontal structure of the Antarctic Circum-polar Current from FGGE drifters. J. Geophys. Res., 90(C4), 7087–7097. Hogg, N. G. (2001). Quantification of the deep circulation. In: G. Siedler, J. Church, & J. Gould (Eds). Ocean Circulation and Climate. Academic Press, London, pp. 259–270. Hogg, N. G., & Johns, W. E. (1995). Western boundary currents. Reviews of geophysics in US National Report to International Union of Geodesy and Geophysics 1991–1994. American Geophysical Union, Washington, D.C., pp. 1311–1334. Hollister, C. D., & McCave, I. N. (1984). Sedimentation under deep sea storms. Nature, 309, 220–225. Intergovernmental Panel on Climate Change (2007). Oceanic climate change and sea level in climate change 2007 – The physical basis. Contribution of Working Group 1 to the Fourth Assessment Report of the IPCC. In: S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, & H. L. Miller (Eds). Cambridge University Press, Cambridge, UK, pp. 385–432. Jacobs, S. S. (1991). On the nature and significance of the Antarctic Slope Front. Mar. Chem., 35, 9–24. Jacobs, S. S. (2004). Bottom water production and its links with the thermohaline circulation. Antarct. Sci., 16, 427–437. Jacobs, S. S., Fairbanks, R. G., & Horibe, Y. (1985). Origin and evolution of water 16 masses near the Antarctic continental margin: Evidence from H18 2 O/H2 O ratios in seawater. In: S. S. Jacobs (Ed.). Oceanology of the Antarctic Continental Shelf. Antarctic Research Series. Vol. 43, pp. 59–85. Jacobs, S. S., & Giulivi, C. F. (1998). Interannual Ocean and Sea Ice Variability in the Ross Sea. In: S. S. Jacobs, & R. Weiss (Eds). Ocean, Ice, and Atmosphere: Interactions at the Antarctic Continental Margin. Antarctic Research Series. Vol. 75, pp. 135–150. Jacobs, S. S., Giulivi, C. F., & Mele, P. A. (2002). Freshening of the Ross Sea during the late 20th century. Nature, 297, 386–389. Johnson, G. C., Musgrave, D. L., Warren, B. A., Ffield, A., & Olson, D. B. (1998). Flow of bottom and deep water in the Amirante passage and Mascarene Basin. J. Geophys. Res., 103, 30973–30984. Johnson, G. C., Warren, B. A., & Olson, D. B. (1991). Flow of bottom water in the Somali Basin. Deep Sea Res., 38, 637–652. Kostianoy, A. G., Ginzburg, A. I., Frankignoulle, M., & Delille, B. (2004). Fronts in the Southern Indian Ocean as inferred from satellite sea surface temperature data. J. Mar. Syst., 45, 55–73. Levitus, S., Antonov, J. I., & Boyer, T. P. (2005). Warming of the world ocean, 1955–2003. Geophys. Res. Lett., 32, L02604, doi:10.1029/2004GL021592.
Circulation and Water Masses of Southern Ocean
111
Lovenduski, N. S., & Gruber, N. (2005). The impact of the Southern Annular Mode on Southern Ocean circulation and biology. Geophys. Res. Lett., 32, L11603, doi: 10.1029/2005GL022727. Lovenduski, N. S., Gruber, N., Doney, S. C., & Lima, I. D. (2007). Enhanced CO2 outgassing in the Southern Ocean from a positive phase of the Southern Annular Mode. Global Biogeochem. Cycles, 21, GB2026, doi:10.1029/2006GB002900. Manighetti, B. (2001). Ocean circulation: The planet’s great heat engine. NIWA Water Atmos., 9, 12–14. Mantyla, A. W., & Reid, J. L. (1983). Abyssal characteristics of the world ocean waters. Deep Sea Res., 30, 805–833. Mantyla, A. W., & Reid, J. L. (1995). On the origins of deep and bottom waters of the Indian Ocean. J. Geophys. Res., 100, 2417–2439. Matsumoto, K. (2007). Radiocarbon-based circulation age of the world oceans. J. Geophys. Res., 112, C09004, doi:10.1029/2007JC004095. McCave, I. N., & Carter, L. (1997). Sedimentation beneath the deep western boundary current off northern New Zealand. Deep Sea Res., 44, 1203–1237. McCave, I. N., Kiefer, T., Thornalley, D. J. R., & Elderfield, H. (2005). Deep flow in the Madagascar-Mascarene Basin over the last 150,000 years. Philos. Trans. R. Soc. A, 363, 818–899. McCartney, M. S. (1977). Subantarctic mode water. In: M. V. Angel (Ed.). A Voyage of Discovery: George Deacon 70th Anniversary Volume, Supplement to Deep-Sea Research. Pergamon Press, Oxford, pp. 103–119. McCartney, M. S., & Donohue, K. A. (2007). A deep cyclonic gyre in the Australian–Antarctic Basin. Prog. Oceanogr., 75, 675–750. Meredith, M. P., & King, J. (2005). Rapid climate change in the ocean west of the Antarctic Peninsula during the second half of the 20th century. Geophys. Res. Lett., 32, L19604, doi:10.1029/2005GL024042. Moore, J. K., Abbott, M. R., & Richman, J. G. (1999). Location and dynamics of the Antarctic Circumpolar Front from satellite sea surface temperature data. J. Geophys. Res., 104, 3059–3073. Morris, M., Stanton, B. R., & Neil, H. L. (2001). Subantarctic oceanography around New Zealand: Preliminary results from an ongoing survey. N. Z. J. Mar. Freshwater Res., 35, 499–519. Morrow, R., Church, J., Coleman, R., Chelton, D., & White, N. (1992). Eddy momentum flux and its contribution to the Southern Ocean momentum balance. Nature, 357, 482–484. Moy, A. D., Howard, W. R., & Gagan, M. K. (2006). Late quaternary palaeoceanography of the circumpolar deep water from the South Tasman Rise. J. Quaternary Sci., 21, 763–777. Nowlin, W. D., & Klinck, J. M. (1986). The physics of the Antarctic Circumpolar Current. Rev. Geophys., 24, 469–491. Orsi, A. H., Jacobs, S. S., Gordon, A. L., & Visbeck, M. (2001). Cooling and ventilating the abyssal ocean. Geophys. Res. Lett., 28, 2923–2926.
112
L. Carter et al.
Orsi, A. H., Johnson, G. C., & Bullister, J. L. (1999). Circulation, mixing and production of Antarctic Bottom Water. Prog. Oceanogr., 43, 55–109. Orsi, A. H., Nowlin, W. D. Jr., & Whitworth III, T. (1993). On the circulation and stratification of the Weddell Gyre. Deep Sea Res., 40, 169–203. Orsi, A. H., & Whitworth III, T. (2005). Hydrographic Atlas of the World Ocean Circulation Experiment (WOCE). In: M. Sparrow, P. Chapman, & J. Gould (Eds). Southern Ocean. International WOCE Project Office, Southampton, UK, Vol. 1, ISBN 0-904175-49-9. Orsi, A. H., Whitworth III, T., & Nowlin, W. D. Jr. (1995). On the meridional extent and fronts of the Antarctic Circumpolar Current. Deep Sea Res., 42, 641–673. Rahmstorf, S. (2002). Ocean circulation and climate during the past 120,000 years. Nature, 419, 207–214. Reid, J. L. (2005). On the total geostrophic circulation of the Indian Ocean: Flow patterns, tracers and transports. Prog. Oceanogr., 56, 137–186. Reid, J. L., & Lynn, R. J. (1971). On the influence of the Norwegian-Greenland and the Weddell seas upon the bottom waters of the Indian and Pacific Oceans. Deep Sea Res., 18, 1063–1088. Ridgway, K. R., & Dunn, J. R. (2007). Observational evidence for a Southern Hemisphere oceanic supergyre. Geophys. Res. Lett., 34, L13612, doi:10.1029/ 2007GL030392. Rintoul, S., & Sokolov, S. (2001). Baroclinic transport variability of the Antarctic Circumpolar Current south of Australia (WOCE Repeat Section SR3). J. Geophys. Res., 106, C2, 2815–2832. Rintoul, S. R. (1998). On the origin and influence of Adelie Land Bottom Water. Ocean, Ice, and Atmosphere – Interactions at the Antarctic Continental Margin, Antarctic Research Series. Vol. 75, pp. 151–171. Rintoul, S. R., Hughes, C. W., & Olbers, D. (2001). The Antarctic Circumpolar Current system. In: G. Siedler, J. Church, & J. Gould (Eds). Ocean Circulation and Climate: Observing and Modelling the Global Ocean. Academic Press, London, pp. 271–302. Saenko, O. A., Weaver, A. J., & Schmittner, A. (2003). Atlantic deep circulation controlled by freshening in the Southern Ocean. Geophys. Res. Lett., 30(14), 1754, doi:10.1029/2003GL017681. Schmitz, W. J. (1995). On the interbasin-scale thermohaline circulation. Rev. Geophys., 33, 151–173. Sokolov, S., & Rintoul, S. R. (2007). On the relationship between fronts of the Antarctic Circumpolar Current and surface chlorophyll concentrations in the Southern Ocean. J. Geophys. Res., 112, C07030, doi:10.1029/2006JC004072. Speer, K., Rintoul, S. R., & Sloyan, B. (2000). The diabatic Deacon cell. J. Phys. Oceanogr., 30, 3212–3222. Stanton, B. R., & Morris, M. (2004). Direct velocity measurements in the Subantarctic Front and over Campbell Plateau, south-east of New Zealand. J. Geophys. Res., 109, C01028, doi:10.1029/2002JC001339.
Circulation and Water Masses of Southern Ocean
113
Stickley, C. E., Carter, L., McCave, I. N., & Weaver, P. P. E. (2001). Variations in LCPDW flow through the SW Pacific Gateway for the last 90 ky: Evidence from Antarctic diatoms. In: D. Seidov, M. Maslin, & B. Haupt (Eds). Ocean Past and Present, North South Connections. American Geophysical Monograph. Vol. 126, American Geophysical Union, Washington, D.C., pp. 101–116. Stommel, H. (1958). The abyssal circulation. Deep Sea Res., 5, 80–82. Stommel, H., & Arons, A. B. (1960). On the abyssal circulation of the world ocean-II: An idealized model of the circulation pattern and amplitude in oceanic basins. Deep Sea Res., 6, 217–233. Stuiver, M., Robinson, S. W., & Ostlund, H. G. (1983). Abyssal water carbon-14 distribution and the age of the world oceans. Science, 219, 849–851. Toggweiler, J. R., Russell, J. L., & Carson, S. R. (2006). Midlatitude westerlies, atmospheric CO2, and climate change. Paleoceanography, 21, PA2005, doi: 10.1029/2005PA001154. Toggweiler, J. R., & Samuels, B. (1995). Effect of sea ice on the salinity of Antarctic Bottom Waters. J. Phys. Oceanogr., 25, 1980–1997. Toole, J. M., & Warren, B. A. (1993). A hydrographic section across the subtropical Indian Ocean. Deep Sea Res. I, 40, 1973–2019. Turner, J., Colwell, S. R., Marshall, G. J., Lachlan-Cope, T. A., Carleton, A. M., Jones, P. D., Lagun, V., Reid, P. A., & Iagovkina, S. (2005). Antarctic climate change during the last 50 years. Int. J. Climatol., 25, 279–294. Vaughan, D. G., Marshall, G. J., Connolley, W. M., King, J. C., & Mulvaney, R. (2001). Devil in the detail. Science, 293, 1777–1779, doi:0.1126/science.1065116. Venuti, A., Florindo, F., Michel, E., & Hall, I. R. (2007). Magnetic proxy for the deep (Pacific) western boundary current variability across the mid-Pleistocene climate transition. Earth Planet. Sci. Lett., 259, 107–118. Walther, G., Post, E., Convey, P., Menzel, A., Parmesan, C., Beebee, T. J. C., Fromentin, J.-M., Hoegh-Guldberg, O., & Bairlein, F. (2002). Ecological responses to recent climate change. Nature, 416, 389–395, doi:10.1038/416389a. Warren, B.A. (1971). Antarctic deep water contribution to the world ocean. Research in the Antarctic. American Association for the Advancement of Science, Washington, D.C., pp. 631–643. Warren, B. A. (1973). TransPacific hydrographic sections at latitudes 431S and 281S; the SCORPIO Expedition – Deep water. Deep Sea Res., 20, 9–38. Warren, B. A. (1974). Deep flow in the Madagascar and Mascarene basins. Deep Sea Res., 21, 1–21. Warren, B. A. (1981). Deep circulation of the world ocean. In: B. A. Warren, & C. Wunsch (Eds). Evolution of Physical Oceanography. MIT Press, Cambridge, MA, pp. 6–41. Weaver, A. J., Saenko, O. A., Clark, P. U., & Mitrovica, J. X. (2003). Meltwater pulse 1A from Antarctica as a trigger of the Bølling-Allerød warm interval. Science, 299, 1709–1713. Whitworth III, T. (1988). The Antarctic Circumpolar Current. Oceanus, 32, 53–58.
114
L. Carter et al.
Whitworth III, T., Orsi, A. H., Kim, S.-J., & Nowlin, W. D. Jr. (1998). Water masses and mixing near the Antarctic Slope Front. In: S. S. Jacobs, & R. F. Weiss (Eds). Ocean, Ice and Atmosphere. Interactions at the Antarctic Continental Margin. Antarctic Research Series. Vol. 75, pp. 1–27. Whitworth, T., & Peterson, R. G. (1985). Volume transport of the Antarctic Circumpolar Current from bottom pressure measurements. J. Phys. Oceanogr., 15, 810–816. Wong, A. P. S., Bindoff, N. L., & Church, J. (1999). Large-scale freshening of intermediate waters in the Pacific and Indian oceans. Nature, 400, 440–443, doi: 10.1038/22733. Wunsch, C. (2002). What is thermohaline circulation? Science, 298, 1179–1181. Wyrtki, K. (1961). The thermohaline circulation in relation to the general circulation in the oceans. Deep Sea Res., 8, 39–64.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00005-0
Chapter 5
Cenozoic Climate History from Seismic Reflection and Drilling Studies on the Antarctic Continental Margin Alan K. Cooper1,, Giuliano Brancolini2, Carlota Escutia3, Yngve Kristoffersen4, Rob Larter5, German Leitchenkov6, Phillip O’Brien7 and Wilfried Jokat8 1
Department of Geological and Environmental Sciences, Stanford University, Stanford, CA 94306, USA 2 Istituto Nazionale di Oceanografia e di Geofisica Sperimentale, B.go Grotta Gigante 42/c, 34010 Sgonico, Trieste, Italy 3 Instituto Andaluz de Ciencias de la Tierra, CSIC-Univ. de Granada Campus de Fuentenueva s/n, 18002 Granada, Spain 4 Department of Earth Science, University of Bergen, Alle`gaten 41, N-5007 Bergen, Norway 5 British Antarctic Survey, High Cross, Madingley Road, Cambridge CB3 0ET, UK 6 Research Institute for Geology and Mineral Resources of the World Ocean, VNIIOkeangeologia, 1 Angliysky Avenue, 190 121 St. Petersburg, Russia 7 Geoscience Australia, Cnr Jerrabomberra Avenue and Hindmarsh Drive, GPO Box 378, Canberra ACT 2601, Australia 8 Alfred-Wegener Institute, Am Handelshafen 12, D-27570 Bremerhaven, Germany
ABSTRACT Seismic stratigraphic studies and scientific drilling of the Antarctic continental margin have yielded clues to the evolution of Cenozoic climates, depositional paleoenvironments and paleoceanographic conditions. This paper draws on studies of the former Antarctic Offshore Stratigraphy Project and others to review the geomorphic and lithostratigraphic offshore features that give insights into the long-duration (m.y.) and short-term (k.y.) changes that document the great variability of Cenozoic Antarctic paleoenvironments. The lithologic drilling record documents non-glacial (pre-early Eocene) to full-glacial (late Pliocene to Holocene) times, and documents times of cyclic ice-sheet fluctuations at k.y. Corresponding author. Tel.: þ01 650329 5157
E-mail:
[email protected] (A.K. Cooper).
116
A. K. Cooper et al.
scales (early Miocene to Pliocene and Holocene). Times of significant change in types and/or amounts of glaciation are also seen in the offshore lithologic record (early Oligocene, mid-Miocene, early Pliocene). Seismic data illustrate large-scale geomorphic features that point to massive sediment erosion and dispersal by ice sheets and paleoceanographic processes (e.g. cross-shelf troughs, slope-fans, risedrifts). The commonality of these features to East and West Antarctica since late Eocene time points to a continent that has been intermittently covered, partially to completely, by glaciers and ice sheets. The greatest advances in our understanding of paleoenvironments and the processes that control them have been achieved from scientific drilling, and future progress depends on a continuation of such drilling.
Citation to regional-section information: Regional-section authors, 2008, In Cooper et al., Cenozoic climate history from seismic-reflection and drilling studies on the Antarctic continental margin, In: F. Florindo and M. Siegert (Eds). Antarctic Climate Evolution, Developments in Earth and Environmental Sciences, Vol. 8, Elsevier, 537p. Citation to general summary: Cooper, A.K., G. Brancolini, C. Escutia, Y. Kristoffersen, R. Larter, G. Leitchenkov, P. O’Brien, W. Jokat, 2008, Cenozoic climate history from seismic-reflection and drilling studies on the Antarctic continental margin, In: F. Florindo and M. Siegert (Eds). Antarctic Climate Evolution, Developments in Earth and Environmental Sciences, Vol. 8, Elsevier, 537p.
5.1. Introduction The Antarctic continental margin is a tectonic collage of former rifts and subduction zones that are covered by sediments deposited when the adjacent continent was free of regional glaciers (i.e. ‘pre-ice-sheet’ times) and when glaciers extended onto the margin (i.e. glacial times). Since the 1960s, many seismic surveys and sea-floor cores and a few drill cores have been acquired on the continental margin to decipher the Cenozoic and earlier history of Antarctica’s paleoenvironments and paleoclimates – a history hidden onshore in sediments now unreachable beneath the Antarctic Ice Sheet. This chapter summarizes principally key results of seismic and drilling studies for proximal parts of the continental margin done from 1989 to 2004 by the multinational Antarctic Offshore Statigraphy project (ANTOSTRAT) to decipher Antarctic Ice Sheet history. We include some findings of the successor Antarctic Climate Evolution project (ACE) that includes the Cenozoic Antarctic Stratigraphy and Paleobathymetry project (CASP), to create a unified circum-Antarctic stratigraphy from all existing seismic and rock-core data (Davey and Cooper,
Cenozoic Climate History from Seismic Reflection and Drilling Studies
117
2007). Our summary complements istotopic and ice-rafting studies for distal parts of the margin and abyssal areas (e.g. Warnke et al., 1996; Zachos et al., 2001). We first describe work in five geographic sectors of the margin, and then summarize key results for the entire margin. Multichannel seismic (MCS) reflection data, the principal tool for deep stratigraphic studies of the continental margin, have been recorded by more than 15 nations (Fig. I-1). Many topical MCS studies with maps of select regional data exist (see citations in regional sections below), but few comprehensive data compilations are either published or openly accessible. Notable exceptions are drilling results (e.g. Deep Sea Drilling Project, Ocean Drilling Program, Cape Roberts Project, ANDRILL and other drilling projects), MCS data compilations in the Antarctic Seismic Data Library System (e.g. www.scar-sdls.org; Wardell et al., 2007), a few Antarctic and regional geosciences atlases (e.g. Hayes, 1991; Cooper et al., 1995), online
Figure I-1: Map showing locations of tracklines for multichannel seismicreflection data on the Antarctic continental margin as of late 2006 (from Wardell et al., 2007). Regions are RS, Ross Sea; WL, Wilkes Land; PB, Prydz Bay; WS, Weddell Sea; AP, Antarctic Peninsula.
118
A. K. Cooper et al.
data centres (e.g. World Data Center) and select discussions of Antarctic margin databases (e.g. Anderson, 1999). In general, Antarctica had a relatively warm climate and normal-waterdepth continental shelf (i.e. like low-latitude continental shelves today) in earliest Cenozoic and Late Cretaceous times – conditions that differ from the polar climate of the latest Cenozoic, with its thick ice sheet and an abnormally deep-water-depth and landward sloping shelf (e.g. Cooper et al., 1991b; Anderson, 1999). Ice has played an important role in continental margin evolution by eroding onshore areas (formerly with vegetation) and discharging the debris into the sea, where ocean currents distribute it to the continental shelf, slope and rise. At times the ice has strongly eroded parts of the shelf. Tectonic processes, principally variable thermal and flexural subsidence and uplift, have also modified the margin morphology and hence stratigraphy (e.g. ten Brink et al., 1995). Offshore Antarctic stratigraphic studies have thus focused on mapping geomorphology and seismic facies of characteristic features (e.g. shelf-edge fans/deltas, mound deposits, unconformities, etc.), and using the limited core and downhole data to decipher their depositional paleoenvironments and relation to nearby ice, if any. Such features help to infer and establish where and when non-glacial and glacial processes acted (e.g. Cooper et al., 1991b). Drilling is the only way to ‘ground truth’ the regional seismic surveys (i.e. via a direct tie of lithologic facies to seismic facies), and to provide the age and biostratigraphic control needed to decipher depositional and climatic paleoenvironments (e.g. Barker and Camerlenghi, 2002; Cooper and O’Brien, 2004). The following regional subchapters have been written by the regional experts listed. Their bibliographic citations are augmented in a ‘selected reference’ section that provides additional background on the prior studies done by the Antarctic geoscience community.
5.2. Ross Sea (G. Brancolini and G. Leitchenkov) The Ross Sea has four large sedimentary basins with thick Cenozoic sequences that record the proximal paleoenvironmental histories of the East and West Antarctic Ice Sheets (Cooper and Davey, 1985; Cooper et al., 1991b, c). Here, ice-sheet evolution is linked to the Cenozoic uplift histories of the Transantarctic Mountains and Marie Byrd Land. Offshore, the West Antarctic Ice Sheet (WAIS) flows across the Eastern basin, and the East Antarctic Ice Sheet (EAIS) passes over the Transantarctic Mountains and flows across the Victoria Land basin, the Northern basin and the Central trough (Fig. RS-1).
Cenozoic Climate History from Seismic Reflection and Drilling Studies
119
Figure RS-1: Map showing the structural framework of the ross sea. The limits of the basins are based on the depositional limits of the seismic Sequence RSS-1. VL, Victoria Land; VLB, Victoria Land basin; NB, Northern basin; CoH, Coulman high; CH, Central high; CT, Central trough; EB, Eastern basin; WAIS, West Antarctic Ice Sheet; EAIS, East Antarctic Ice Sheet. The heavy dashed line marks the postulated boundary between East Antarctic Ice Sheet and West Antarctic Ice Sheet drainage. Numerous seismic studies have been done in the Ross Sea region since the 1960s, with more than 45,000 km of MCS reflection data collected since 1980 (Hinz and Block, 1984; Sato et al., 1984; Cooper and Davey, 1987; Hinz and Kristoffersen, 1987; Zayatz et al., 1990; Brancolini et al., 1991) (Fig. RS-2a), to provide tectonic and deep stratigraphic control. A large number of singlechannel seismic (SCS) surveys have also been conducted for greater resolution of the shallow subsurface (Fig. RS-2b). Drilling at several sites by DSDP (Deep Sea Drilling Project, Hayes and Frakes, 1975), DVDP (Dry Valley Drilling Project, McGinnis, 1981), MSSTS (McMurdo Sound Sediment and Tectonic Study, Barrett, 1986), CIROS (Cenozoic Investigation in the Western Ross Sea, Barrett, 1989) and CRP (Cape Roberts Project,
120
A. K. Cooper et al.
Figure RS-2: (A): Multichannel seismic-reflection surveys in the Ross Sea (modified from Brancolini et al., 1995). Some of these data are available in digital format from Cooper et al. (1995) and others from the Antarctic Seismic Data Library (Childs et al., 1994; Wardell et al., 2007). (B) Singlechannel seismic surveys in the Ross Sea (modified from Barrett et al., 1999).
Cenozoic Climate History from Seismic Reflection and Drilling Studies
121
Figure RS-3: Correlation of Ross Sea area drilling and seismic stratigraphy with global oxygen isotope (Miller et al., 1987) and eustacy (Haq et al., 1987) curves (modified from Brancolini et al., 1995; Cooper et al., 1995). Regional erosional unconformities in Oligocene and younger sections are interpreted to be due in part to sub-ice erosion, especially in late Neogene time.
Cape Roberts Science Team, 1998, 1999, 2000, 2001) ANDRILL (Antarctic geological Drilling, Naish et al., 2007; Florindo et al., 2008; Harwood et al., 2008) provides geologic ground truth data (Fig. RS-3). A regional seismic stratigraphy has been derived by the ANTOSTRAT project with seismic sequences and unconformities tied to drilling data (Fig. RS-3; Cooper et al., 1995). Ross Sea seismic data are used by many to infer glacial sedimentary processes (e.g. Cooper et al., 1991b; Alonso et al., 1992; Anderson and
122
A. K. Cooper et al.
Bartek, 1992; Shipp et al., 1994; Brancolini et al., 1995; Cochrane et al., 1995; De Santis et al., 1995; Bartek et al., 1996; De Santis et al., 1999; Bart et al., 2000; Bart, 2003; Chow and Bart, 2003; Accaino et al., 2005). Characteristic features and inferred processes in Oligocene and younger strata include: 1. Landward-deepening seafloor of the continental shelf with broad (up to 100 km wide) cross-shelf troughs and banks formed by ice-stream erosion and deposition, respectively. 2. Numerous regional seismic unconformities believed to result from erosion of the continental shelf by grounded ice sheets. 3. Steep prograding sedimentary sequences (i.e. foreseest dips more than 51 and eroded topset strata) interpreted as ice-proximal till deltas from grounded ice. 4. Wedge-shape, non-reflective units interpreted as ‘till tongues’ deposited by grounded ice. 5. Shallow sediment with high velocities, considered due to overcompaction by grounded ice. The seismic stratigraphy and drilling help establish ice-sheet evolution in the Ross Sea region, and are discussed below for four key intervals.
5.2.1. Pre-Ice-Sheet (Pre-Late-Oligocene Time) This period includes seismic sequence RSS-1 between the acoustic basement and unconformity RSU-6 (Fig. RS-3, Foldouts RS-1 and RS-2) and, in the Victoria Land basin, acoustic units V4 and V5 (Cooper et al., 1987). Acoustic basement rocks have been sampled at two sites, and are Palaeozoic Beacon Formation rocks at the CRP-3 site adjacent to the coast (Cape Roberts Science Team, 2000) and are inferred palaeozoic and Cretaceous igneous and metamorphic rocks at DSDP Site 270 in the centre of the Ross Sea (Hayes and Frakes, 1975). The basins are believed to hold sedimentary rocks of Cretaceous and younger age (Hinz and Block, 1984; Cooper et al., 1991c), but Sedimentary rocks older than late Eocene have not been cored by drilling. Upper Eocene sediments have been cored in the CIROS-1 drillhole in McMurdo Sound (Coccioni and Galeotti, 1997; Fielding et al., 1997; Hannah et al., 1997; Monechi and Reale, 1997). The presence of ubiquitous lonestones (Barrett, 1989) testifies that glaciers (but not necessarily continent-size ice sheets) were calving at sea-level then. Eocene erratic rocks are found in coastal areas (Levy et al., 1995), and have flora indicating cool, but not glacial, climates in the McMurdo area (Stillwell and Feldmann, 2000). Offshore basin analysis, mainly from MCS
Cenozoic Climate History from Seismic Reflection and Drilling Studies
123
reflection data, suggests that in Cenozoic pre-ice-sheet times, the Ross Sea was dissected by high-standing subaerial ridges, now seen as the buried Coulman high and Central high (De Santis et al., 1995). Prior outcropping of these ridges is suggested by the presence of regoliths above the basement at DSDP Site 270 (Hayes and Frakes, 1975). 5.2.2. Early Glacial (Late Oligocene to Early Miocene) This period includes seismic sequences RSS-2 and -3 (Fig. RS-3 and Foldouts RS-1 and RS-2). Sedimentary rocks from this period were recovered at CIROS-1, and CRP-1, -2 and -3, and MSSTS-1 drilling sites in the McMurdo Sound area. Such rocks include compacted diamicton indicative of deposition by/under grounded ice, as well as mud and ice-rafted debris (IRD) indicative of open-water environments (Barrett, 1986; Barrett, 1989; Hannah, 1994; Cape Roberts Science Team, 2001), in lower Miocene sediments at CRP 2/2A sites. Compacted diamicton and mud layers at site CRP-1, vary with uniform cyclicity, and document systematic oscillation of the EAIS size (Naish et al., 2001). The oscillations are at orbital periodicities similar to those recorded by isotope studies in distal deep-ocean sediments. Seismic facies along the border of the Victoria Land basin suggest that tidewater glaciers all along the Transantarctic Mountains intermittently extended onto the continental shelf and carried abundant glacial sediment to the sea (Brancolini et al., 1995; Bartek et al., 1996; Henrys et al., 2001). In the eastern Ross Sea, at DSDP Site 270, Nothofagus-dominated flora in lower Miocene sediments (Kemp and Barrett, 1975) are similar to those recovered in McMurdo Sound drillcores (Hill, 1989; Mildenhall, 1989; Askin and Raine, 2000), and indicate cool-temperate climates during interglacial periods. DSDP Site 270 also recovered lower Miocene iceproximal glaciomarine sediments from the early Miocene section, but the size and character of the ice sheet that deposited these sediments is debated. Anderson and Bartek (1992) suggest, based on high-resolution singlechannel seismic data and drill cores, that by late Oligocene to early Miocene time, the continental shelf was deeply scoured and foredeepened (i.e. landward dipping) by a massive ice sheet. In contrast, Brancolini et al. (1995) and De Santis et al. (1995), utilize regional stratigraphic maps (Cooper et al., 1995) and their seismic facies analyses to postulate that, during the same period the Central high was partly exposed and partly covered by small subpolar ice caps (i.e. subpolar as defined by Anderson and Ashley, 1991). A semi-quantitative evaluation of the water depth of the Eastern basin during the early Miocene, based on the backstripping of the seismic section
124
A. K. Cooper et al.
in Foldout RS-1, indicates that the foredeepeened profile of the Eastern basin was only attained after middle Miocene time (De Santis et al., 1999). The end of this early glacial period is marked by a change in reflection geometries beneath the outer continental shelf (Foldouts RS-1 and RS-2) from principally aggrading (RSS-2 and -3) to principally prograding, (RSS-4). Cooper et al. (1991b) postulate that this change marks the start of grounded-glacier advances to the shelf edge and erosion of a normalwater-depth shelf by episodic grounded ice. 5.2.3. The Ice-Sheet Development (Mid-Early Miocene to Early Pliocene) This period includes seismic sequences RSS-4, -5 and -6 (Fig. RS-3 and Foldouts RS-1 and RS-2). Sediment from these sequences was recovered at DSDP Sites 271, 272 and 273, MSSTS-1 and DVDP 10/11. The early-middle Miocene is postulated to have been a time of major ice buildup of ice in the Ross Sea region, and the carving of the first deep troughs, similar in size to the present ice streams, across the continental shelf (Anderson and Bartek, 1992). Bart (2003) and Chow and Bart (2003) recognize at least two major WAIS expansions during the early part of the middle Miocene and five in the entire Miocene. These expansions suggest that either portions of the West Antarctic land elevation were above sea-level and/or the air and water temperatures were sufficiently cold to support a marine-based ice sheet. Drill cores from the middle Miocene have been recovered at DSDP Sites 272 and 273 (Hayes and Frakes, 1975; Savage and Ciesielsky, 1983; Leckie and Webb, 1986), and consist of diatom-bearing sediments interpreted as waterlain tills and proximal- to distal-glacimarine deposits (Hambrey and Barrett, 1993). Upper Miocene rocks are missing from all continental shelf drill cores, except in the McMurdo Sound region (MSSTS-1, DVDP-10 and -11 drill sites), where glaciomarine diamictites (tillites) and terrestrial strata are found. These deposits are interpreted as having originated from glaciers flowing out of the Transantarctic Mountains (Powell, 1981; Barrett, 1986; Ishman and Webb, 1988; McKelvey, 1991). The recent ANDRILL drilling on the Ross Ice shelf, near McMurdo Sound (MIS project; Naish et al., 2007), recovered a 1,284 m long core that records Antarctica’s history over the last 14 million years. The core indicates periods of ice-sheet growth, advancing over the drill site and then retreating again to allow the open-marine conditions to return. More than 60 of these advance-retreat cycles are present in the core. On the outer continental shelf and upper slope, well-stratified seismic sequences inferred to be of late Miocene age (i.e. RSS-6) are present in the
Cenozoic Climate History from Seismic Reflection and Drilling Studies
125
Northern and Eastern basins, but the sequences are thin or absent on the inner shelf (Cooper et al., 1995). In both basins, these sequences are characterized by steeply prograding clinoforms with relatively thin or eroded topset beds and a major seaward shift of the palaeo-shelf edge (Foldouts RS-1 and RS-2). The sequences are thought to have been deposited by intermittent grounded ice sheets carrying sediment to the continental shelf edge (Bartek et al., 1991; Cooper et al., 1991b; Anderson and Bartek, 1992). 5.2.4. The Polar Ice Sheet (Early Pliocene Through Quaternary) This period includes seismic sequences RSS-7 and -8 (Fig. RS-3) and seismic units 1–7 of Alonso et al. (1992) and Anderson and Bartek (1992) (Fig. RS-4a and b). Pliocene sediment has been recovered at drill sites in the Taylor Valley (DVDP-10 and -11) and on the continental shelf at DSDP Sites 271 and 273. DVDP drill cores contain sediments deposited by the Taylor Glacier, while the sparsely sampled Pliocene deposit at DSDP 271 and 273 contains diatomaceous glaciomarine strata (Hayes and Frakes, 1975). These rocks imply that warmer interglacial conditions than today existed at that time (Anderson and Ashley, 1991). Seismic sequences of inferred Pliocene through Quaternary age occur in the EAIS drainage in the Northern basin, where they are up to 800 m thick in the till delta fan system (Cooper et al., 1995). In the Eastern basin, Pliocene through Quaternary age strata lie within the WAIS drainage and are more than 1,000 m thick (Cooper et al., 1995). Detailed seismic stratigraphic analyses from the Eastern basin margin (Fig. RS-4a; Alonso et al., 1992; Fig. RS-4b; Anderson and Bartek, 1992), recognize a major change in the seismic character of the Pliocene deposits. Up-section, the seismic unit thicknesses decrease, the geometry of the sequences changes from principally progradational to aggradational, and numerous widespread glacial erosion surfaces are seen. These features indicate more frequent grounding events on the continental shelf and increased subglacial till deposition relative to basal transport of sediments to the grounding line. Bart et al. (2000) and Anderson et al. (1992) suggest that on at least eight occasions during Pliocene to Quaternary times, the East and West Antarctic Ice Sheets were much larger than today. The frequent and extensive grounding events on the outer continental shelf contradict the widely held view that the land-based EAIS was relatively stable and the largely marine-based WAIS was relatively dynamic (Bart and Anderson, 2000). The last glacial maximum (LGM) in the Ross Sea has been studied using seafloor cores, subbottom and swath bathymetry data (Thomas and Bentley, 1978; Kellogg et al., 1979; Denton et al., 1989; Leventer et al., 1993; Brambati et al., 1994; Hilfinger et al., 1995; Kellogg et al., 1996; Licht et al.,
126
A. K. Cooper et al.
Figure RS-4: Seismic-reflection profiles across the Eastern basin (A) and the Northern basin (B), illustrating the Neogene stratigraphic sections (from Anderson and Bartek, 1992). The shelf margin delta fan complex shown is a common feature in the seismic data from the continental margin and characterizes deposition close to the ice-sheet grounding line. See Fig. RS-1 for location. 1996; Cunningham et al., 1999; Domack et al., 1999; Licht et al., 1999; Shipp et al., 1999; Alley and Bindschadler, 2001). Radiocarbon dates from diamictons and sediment composition, indicate that ice-free conditions existed on the inner shelf at times during the period from 60 to 10 ka, and that the ice sheet was present between 26.5 and 19.5 ka. (Domack et al., 1999). The maximum ice-sheet expansion in the LGM is still debated: Kellogg et al. (1996) place the grounding line at the continental shelf edge, Domack et al. (1999) put it just north of the Coulman Island and Licht et al. (1999) interpret that it was about 100 km south of Coulman Island. The last retreat of the grounding line occurred in the western Ross Sea around 11 ka at a rate of about 100 m/year, and the grounding line reached its present position about 6 ka (Domack et al., 1999; Shipp et al., 1999).
Cenozoic Climate History from Seismic Reflection and Drilling Studies
127
5.3. Wilkes Land (C. Escutia and P. O’Brien) 5.3.1. Acoustic Stratigraphy The approximately 1,500 km long Wilkes Land segment of the continental margin (Fig. WL-1) formed during the Cretaceous separation of Australia and Antarctica (Cande and Mutter, 1982; Veevers, 1987; Sayers et al., 2001; Colwell et al., 2006; Leitchenkov et al., 2007; O’Brien and Stagg, 2007). The stratigraphy of the margin is known mainly from the seismic stratigraphic interpretation of numerous MCS surveys (Sato et al., 1984; Wanneson et al., 1985; Tsumuraya et al., 1985; Eittreim and Hampton, 1987; Ishihara et al., 1996; Tanahashi et al., 1997; Brancolini and Harris, 2000; Stagg et al., 2004a, b); complemented by surface sediment cores (Domack et al., 1980; Payne and Conolly, 1972; Domack, 1982; Tsumuraya et al., 1985; Hampton et al., 1987; Ishihara et al., 1996; Tanahashi et al., 1997; Brancolini and Harris, 2000; Leventer et al., 2001; Escutia et al., 2003; Michel et al., 2006); and limited deep geological sampling recovery at DSDP Sites 268 and 269 (Hayes and Frakes, 1975). The best-surveyed area is the eastern Wilkes Land margin (EWL) from the Ade´lie Coast to George V Land. West of this area (the western Wilkes Land margin-WWL), Japan and Russia collected widely spaced seismic lines that were then augmented during the 2001–2002 Australian Antarctic and Southern Ocean Profiling (ASSOPP) Project (Stagg et al., 2004a, b; Leitchenkov et al., 2007).
5.3.1.1. Pre-ice-sheet stratigraphy Along the Wilkes Land margin syn- and post-rift sedimentary rocks reach thicknesses of more than 7 km (Stagg et al., 2004a, b). Pre-Eocene syn-rift strata are about 3 km thick and are highly variable in seismic character, with discontinuous, faulted, and tilted strata onlapping the flanks of the acoustic basement (Eittreim and Smith, 1987; Eittreim, 1994; De Santis et al., 2003; Stagg et al., 2004a, b; Leitchenkov et al., 2007). The thickest (at least 9 km) depocentre of post-rift sedimentary rocks is located in the WWL off the Bud Coast (Close et al., 2007). In the EWL postrift strata are up to 5 km thick across the Wilkes Land continental shelf, slope and rise (Eittreim and Smith, 1987; Hampton et al., 1987; Wannesson, 1990; Tanahashi et al., 1994; De Santis et al., 2003). These strata are welllayered on the continental rise and become less stratified and more discontinuous landward (Eittreim and Smith, 1987; Eittreim, 1994;
Surveys France (IFP) 1982 US Geological Survey 1984 Japan (JNOC) 1983-1996 Italian-Australian 2000 Australian 2000-2002 Russia (RAE) 2005-2007
130°E
Terra Adélie
200 km
269
140°E
Bathymetric contour (Interval 500 m)
W21 - Seismic Line 269 - DSDP Site
Wilkes Land
135°E
W21 7 10
Mertz Glacier Tongue
W30
George V Land
P-
IF
W36
145°E 150°E 155°E 160°E
Figure WL-1: Location of multichannel seismic-reflection profiles collected on the Wilkes Land margin. Location of the seismic sections shown in Figs. 2 and 3 is indicated.
70°S
DSDP Site 268 63o 56.99'S 105o 09.34'E (off map)
Budd Coast
69°S
68°S
67°S
66°S
65°S
64°S
63°S
62°S
61°S
110°E 60°S
115°E
120°E
125°E
128 A. K. Cooper et al.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
129
Figure WL-2: Multichannel seismic line WEGA W21 and line drawings of multichannel seismic profiles IFP 107 and WEGA W35 showing the overall architecture of the Wilkes Land margin from the continental shelf to the continental rise (modified from Escutia et al., 2005). See Fig. WL-1 for location of the seismic lines. De Santis et al., 2003). A prominent regional unconformity (WL-U3) within the Cenozoic post-rift section beneath the continental shelf (Fig. WL-2) is believed to be due to erosional processes related to the first advance of grounded ice sheets onto the continental shelf (Eittreim and Smith, 1987; Tanahashi et al., 1994; Eittreim et al., 1995; Escutia et al., 1997; Escutia et al., 2005). The pre-ice-sheet strata below unconformity WL-U3, where resolvable, are flat-lying and less stratified than glacial strata above the unconformity. Pre-ice-sheet rocks have been dredged from the Wilkes Land continental shelf and slope. On the inner shelf, Mesozoic sediments have been exposed via erosion by late Cenozoic glaciers near the Mertz ice tongue. Lignite was recovered (Mawson, 1940, 1942), and lower Cretaceous brecciated, carbonaceous siltstone was cored (Domack et al., 1980). Other dredge samples in the area, acquired by Leventer et al. (2001), include sedimentary clasts of Paleogene lignites with reworked Early Cretaceous flora. On the upper continental slope off Terre Ade´lie, Sato et al. (1984) dredged samples
130
A. K. Cooper et al.
of locally derived Oligocene and Miocene limestone and undated sedimentary, metamorphic and igneous rocks of mostly ice-rafted origin.
5.3.1.2. Continental shelf glacial stratigraphy Glacial sequences on the shelf thicken seaward in prograding wedges (Fig. WL-2). The sequences are deeply eroded by broad troughs that cross the shelf. The troughs are interpreted as the erosional paths of ice streams during times of glacial maxima (Eittreim et al., 1995). Foreset strata are commonly truncated at or near the seafloor beneath the troughs (Fig. WL-2). Topset strata form the banks adjacent to the troughs. Ice is inferred to have moved slowly over bank areas and rapidly in the troughs. Geometry of strata in buried troughs on the shelf suggests to some (Eittreim et al., 1995; Escutia et al., 2000) that the locations of ice streams and their erosional troughs and banks have shifted during consecutive glacial advances. Regional glacial seismic sequences and unconformities defined by different workers (Table WL-1) in the EWL were renamed and in some cases redefined by De Santis et al. (2003). On the shelf, sequences are truncated by two Table WL-1: Summary of the terminology assigned in previous publications to the inferred wilkes Land glacial sequences and their bounding unconformities (updated from Escutia et al., 2005). Unconformities (tied with lines) and sequences (in between these lines) are listed from younger at the top to older.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
131
regional unconformities, WL-U3 and WL-U8 (Wannesson et al., 1985; Eittreim and Smith, 1987; Hampton et al., 1987; De Santis et al., 2003), and the erosion is thought to result from grounded ice sheets moving across the continental shelf (Tanahashi et al., 1994; Eittreim et al., 1995; Escutia et al., 1997; Escutia et al., 2005). Eittreim et al. (1995) calculated erosion of 300 to 600 m of strata below WL-U3. Sequences below WL-U8 are dominantly aggradational and sequences above are principally progradational. For unconformity WL-U8, Eittreim et al. (1995) estimated erosional truncation of 350 to 700 m of sediment. Unconformity WL-U8 marks changes in the geometry of the outer shelf progradating wedges, from shallower dips below WL-U8 to steeper dips above (foreset slopes up to about 101). During the open-marine Holocene, thick laminated diatom mud and oozes were deposited in deep (W1,000 m) inner shelf basins, such as for example the Ade´lie Drift (Costa et al., 2007). Based on AMS radiocarbon dates, this drift has accumulation rates on the order of 20–21 m/k.y. Opal, Ti and Ba timeseries show decadal to century variance suggestive of solar forcing and El Nin˜o Southern Oscillation (ENSO) forcing (Costa et al., 2007).
5.3.1.3. Continental slope glacial stratigraphy Although partly obscured by seafloor multiples, the stratigraphy of the continental slope consists of seaward-dipping reflectors (Eittreim and Smith, 1987; Hampton et al., 1987; Eittreim et al., 1995). Prograding strata above the WL-U8 unconformity downlap and pinch out at the base of the continental slope, but deeper units (i.e. between WL-U8 and WL-U3) continue across the margin (Hampton et al., 1987; Eittreim et al., 1995; Escutia et al., 1997; De Santis et al., 2003) (Fig. WL-2). Sediments forming prograding foresets were delivered directly to the outer shelf and upper slope as deforming tills at the base of ice streams at times of glacial maxima (Eittreim et al., 1995). Ice-stream delivery of a large volume of unconsolidated sediment to the steep slope resulted in sediment failures that led to the development of large chaotic deposits at the base of the paleoslope foresets (Eittreim et al., 1995; Escutia et al., 2000; De Santis et al., 2003; Escutia et al., 2007). More-recent slope strata are dissected by erosional submarine gullies (Eittreim et al., 1995) and slope canyons (Escutia et al., 2000). Sea-floor sediment cores from the continental slope contain debris-flow units and numerous hiatuses. The oldest sediment has been dated as late Miocene in age, indicating that gravity flows have been a dominant slope process since at least this time (Escutia et al., 2003).
132
A. K. Cooper et al.
5.3.1.4. Continental rise glacial stratigraphy On the EWL continental rise, strata above the WL-U3 unconformity include six glacial-related seismic units, WL-S4 to WL-S9 (De Santis et al., 2003; Donda et al., 2003) (Table WL-1, Fig. WL-3). The two deepest units, WL-S4 and WL-S5, consist of stratified and continuous reflectors that onlap at the base of the slope (Escutia et al., 1997; Donda et al., 2003). Acoustic signatures of isolated channel-levee complexes that characterize turbidite deposition are first observed up-section within unit WL-S5 (Escutia et al., 1997; Escutia et al., 2000; Escutia et al., 2002; Donda et al., 2003). During deposition of units WL-S6 and WL-S7, channel-levee complexes became widespread and turbidity flows were the dominant process building the sedimentary ridges on the rise. Wavy reflectors that are characteristic of bottom contour-current deposition occur on the lower rise in unit WL-S6 and on the upper rise in WL-S7. Within Unit WL-S8, there is evidence for bottom contour-current and turbidite flows, but WL-S8 mostly infills previous depressions (Escutia et al., 1997; Escutia et al., 2000; Escutia et al., 2002; Donda et al., 2003). Unit WL-S9 is a discontinuous unit on the rise, and, where present, comprises channel and levee complexes and layered reflectors (Donda et al., 2003). Recent studies on the WWL margin glacial strata show a similar evolution of the glacial sedimentary sequences (i.e. increased proximal turbidite facies up-section and influence of bottomcontour-current deposition) above unconformities ‘eoc’ (Close et al., 2007) and WL3 (Leitchenkov et al., 2007), which correlate with WL-U3 on the 4.0
600 WEST
800
1000
1200
1600
1400
W35
1800
2200
2000
2400
WEGA CHANNEL
JUSSIEU CHANNEL
2800
3000
3200
TWT(s)
WL-S6 WL-U5 WL-S5
6.0
WL-U4
WL-S4
Basement
WL-U3
PROFILE W30
7.0 4.0
3400 EAST WL-U6
WL-S8
WL-S7
5.0
2600 WL-U7
600 WEST
800
1000
1200
1400
1600
1800
2000
2200
2400 10 km
2600
2800
3000
3200
3400 EAST
TWT(s)
5.0
6.0
7.0
Figure WL-3: High-resolution multichannel line WEGA W30 across the distal continental rise. Also shown is a line drawing interpretation of this profile after Donda et al. (2003). Location of seismic line is shown in Fig. WL-1.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
133
EWL. Between 1101 and 1301 large debris-flow deposits are also reported forming throughout the Miocene (Donda et al., 2007a, b). 5.3.2. Drilling on the Wilkes Land Margin DSDP Leg 28 drilled Sites 268 and 269 on the continental rise and abyssal plain, respectively, to determine the geologic and climate history of Antarctica and the Southern Ocean (Hayes and Frakes, 1975). The drill cores document that extensive Antarctic glaciation began at least by Oligocene to early Miocene time, and that water temperatures were cooltemperate in the late Oligocene and early Miocene and cooled during the Neogene, presumably as glaciation intensified. DSDP Site 268 was drilled to a subbottom depth of 474.5 m in 3,544 m water depth with total core recovery of 14% (Hayes and Frakes, 1975). Three units were described, based on lithologies and amounts of diatoms, nannofossil ooze and ice-rafted pebbles and granules (Hayes and Frakes, 1975). Piper and Brisco (1975) interpreted the two deeper units to be contourites, based on the character of silt laminae. The shallowest of the three units, dated as Pliocene and Quaternary, was interpreted as turbidites, based on a high content of silty clay with common silt laminae and fine-sand beds 2–20 cm thick. Hayes and Frakes (1975) infer that the deepest lower Miocene and Oligocene unit was deposited when the ice sheet first advanced onto the shelf. Water at that time was warm enough to support calcareous biogenic sedimentation, but ice-rafting and contourites provide evidence for nearby ice on East Antarctica and for bottom currents, possibly generated by cold bottom water production associated with a limited ice shelf or tongue (Hayes and Frakes, 1975). DSDP Site 269 was drilled to a subbottom depth of 958 m in a water depth of 4,285 m and with 42% recovery of Eocene to recent rocks (Hayes and Frakes, 1975). The section consists predominantly of silts and clays with variable amounts of microfossils. Diatom oozes and diatom mud dominate the upper half of the section, which is dated as Quaternary to Late Miocene in age (Hayes and Frakes, 1975). In the lower half, which is late Miocene to early Miocene and Oligocene in age, diatoms are absent but calcareous nannofossils are found in trace amounts. Similar to DSDP Site 268, there is a transition in facies at DSDP Site 269 from more distal facies in the lower part of the core to more proximal facies near the surface. Piper and Brisco (1975) interpret this facies change as resulting from substantial increased supply of sand and coarse silt and clay from the Antarctic continent, possibly in response to prograding of the continental margin.
134
A. K. Cooper et al.
5.3.3. The Inferred Long-Term Record of Glaciations Investigators interpret the WL-U3 unconformity as having been eroded during the first grounding of an ice sheet on the continental shelf (Tanahashi et al., 1994; Eittreim et al., 1995; Escutia et al., 1997; Escutia et al., 2005), either about 40 m.y. ago (Eittreim et al., 1995) or 33.4–30 Ma (Escutia et al., 2005) (Table WL-2). Above WL-U3, early glacial strata (e.g. likely glacial outwash deposits) were provided by fluctuating temperate glaciers, and were deposited as low-dip-angle prograding foresets. The increase in stratal dips across unconformity WL-U8 in the prograding wedge at the shelf edge is interpreted to record a glacier-regime change from intermittent fluctuating glaciers to persistent oscillatory ice sheets, either on the Late Miocene (Escutia et al., 2005) or about 3 Ma (Rebesco et al., 2006) (Table WL-2). The steep foresets above WL-U8 likely consist of ice proximal (i.e. waterlain till and debris flows) and open-water sediments deposited as grounded ice sheets Table WL-2: Continental shelf and rise stratigraphy and inferred East Antarctic Ice Sheet evolution in the Wilkes Land margin and timing of events.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
135
extended intermittently onto the outer shelf – similar to sediments recovered from ODP Site 1167 on the Prydz Trough fan (O’Brien et al., 2001). On the continental rise, the up-section increase in the energy of the depositional environment in units WL-S5 to WL-S7 (i.e. seismic facies indicative of proximal turbidites and of bottom-contour-current deposition) likely resulted from enhanced shelf progradation. Maximum rates of sediment delivery to the rise appear to have occurred during the development of units WL-S6 and WL-S7, which is inferred to have been during the Miocene (Hayes and Frakes, 1975; De Santis et al., 2003; Escutia et al., 2005). During deposition of WL-S8 and WL-S9, sediment supply to the lower continental rise decreased and depocentres shifted landward to the base of the slope and outer shelf (Escutia et al., 2002; De Santis et al., 2003; Donda et al., 2003; Escutia et al., 2005; Rebesco et al., 2006). Inferred age for Units WL-S8 and WL-S9 is Pliocene to Recent (De Santis et al., 2003). Sequence WL-S9 was deposited under a polar regime with a persistent ice sheet during the Pliocene–Pleistocene. At that time, most sediment delivered to the margin was trapped on the outer shelf and slope, forming steep prograding wedges, with some sediment bypassing the slope in channelized turbidity currents (Escutia et al., 2002; De Santis et al., 2003; Escutia et al., 2005). During the Holocene open-water interglacial thick sections of diatom mud and oozes are deposited in deep inner shelf basins (Costa et al., 2007). These sediments hold an ultra-high-resolution record of climate variability likely by solar and ENSO forcing.
5.4. Prydz Bay (P. O’Brien and G. Leitchenkov) Prydz Bay is a re-entrant in the East Antarctic margin, and overlies a rift structure that extends about 500 km into the interior of the continent. The rift has channelled drainage at least since the Early Cretaceous (Fig. PB-1; Arne, 1994) and presently controls the Amery Ice Shelf drainage system, which drains more than 16% of East Antarctica. This drainage basin includes the Gamburtsev Mountains, a subglacial range in which the Cenozoic ice sheet may have nucleated. Its long history, thick sediment and Cenozoic outcrops in the flanking Prince Charles Mountains have made Prydz Bay a likely site for preserving palaeo-climate records. Seismic surveys by Australia, Russia, Japan and the US and two ODP Legs 119 and 188 (Fig. PB-1), plus field studies and exposure dating, have provided an extensive picture of palaeo-climate evolution of the region.
136
A. K. Cooper et al.
066° E 64° S
070° E
079° E 64° S
075° E
0
A′
34 0
B′
0 3 60
0 0
1165 00
32
3 60 0
00
30
34
0
30
28
00
00
32 00
0
3
32 0
0 34
0 20
Fig.PB-2
32
00
3200
66° S
2 600
66° S
3000
1167
28 0
0
0
22 0
0
0
6 2 0 24 0 2 00 1 800 0 12 6000
40 0 0 0
A
2
0
743
2
00
739 40
0
00
742
6
200
1166
Fig. PB-3 F ig
.P
4
B-
68° S
2
741
80
00
8
0
8 00
0
0
740 0
2
6
0
60400 0
B
00
68° S
800
70° S 066° E
070° E
100 km
40
0
Amery Ice Shelf
600
70° S 079° E
075° E
Seismic line Seismic line (Illustrated)
Bathymetric contour
Seismic line (single channel)
ODP Site
Figure PB-1: Map of Prydz bay showing seismic lines and ODP drill sites. Seismic lines illustrated in this chapter are in heavy black line. Foldouts are sections for A-Au and B-Bu.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
137
5.4.1. Pre-Ice-Sheet (Pre-Late Eocene) The Lambert Graben and Prydz Bay basin formed during the Carboniferous or Permian (Arne, 1994; Lisker et al., 2005) and were depocentres in pre-icesheet times. Seismic data from the Prydz Bay shelf show pre-ice-sheet sequences of parallel, moderately continuous reflectors (Figs. PB-2 and PB-3). ODP Sites 740, 741 and 1166 penetrated pre-ice-sheet sediments that were deposited in fluvial to fluvio-deltaic environments. ODP Site 740 intersected interbedded sandstone, siltstone and mudstone with reddish coloration (Shipboard Scientific Party, 1989), interpreted as fluvial flood plain deposits (Turner, 1991). The red coloration suggests a seasonal fluctuating rainfall regime but the age of this unit remains unknown (Truswell, 1991). It could be as old as Triassic, based on the presence of Triassic sediments in the northern Prince Charles Mountains (Leitchenkov, 1991; McLoughlin and Drinnan, 1997a, b). However Leitchenkov (1991) identified a thick (up to 5 km), faulted and high-velocity (up to 5.2 km/s) unit on multichannel data underlying these red beds. This sequence predates the main phase of breakup-related crustal extension, leading him to correlate the deep unit with Permian-Triassic sediments of the northern Prince Charles Mountains (Leitchenkov, 1991). If so, then the red beds in ODP Site 740 are likely Early Cretaceous or Late Jurassic in age. ODP Sites 741 and 1166 intersected Cretaceous sediments beneath the Cenozoic section. The Cretaceous comprises interbedded dark siltstone and
Figure PB-2: Diagramatic section across Prydz Bay shelf and slope based on Russian seismic lines SAE 32002 and RAE 4005. K1 are Cretaceous sediments. Location shown in Fig. PB-1.
138
A. K. Cooper et al.
Figure PB-3: Seismic section between ODP sites 742 and 1166. Sequence PS2A2 comprises fluvio-deltaic sands of late Eocene age. Sequence PS2A1 overlies an erosion surface and comprises late Eocene marine muds with lonestones. Location shown in Fig. PB-1 (modified from Erohina et al., 2004). sandstone with minor coal of probable delta plain to lagoonal origin. The ODP Site 1166 section is Turonian-Santonian(?), whereas ODP Site 741 recovered an older section of middle Aptian sediment (Fig. PB-3). Macphail and Truswell (2004) describe the palynomorphs from ODP Site 1166, and interpret the assemblage as indicating a conifer-dominated woodland vegetation, consistent with a cool, humid climate. The continental rise seaward of Prydz Bay contains up to 5 km of post-rift sediments (Figs. PB-2, Foldouts PB-1 and PB-2; Mizukoshi et al., 1986, Stagg et al., 2004a, b). The lowermost seismic stratigraphic unit has parallel, mostly continuous reflectors typical of deep-ocean deposition that probably occurred during pre-ice-sheet times (Mizukoshi et al., 1986; Kuvaas and Leitchenkov, 1992; Kuvaas et al., 2005).
5.4.2. Early Glacial (Late Eocene) In Prydz Bay, ODP Sites 739, 742 and 1166 recovered sediments deposited immediately before major glaciation (Barron et al., 1991; Cooper and O’Brien, 2004). The lithologies vary from dark siltstones to poorly sorted sands and bedded mudstone with lonestones. Seismic sections show that the
Cenozoic Climate History from Seismic Reflection and Drilling Studies
139
sands overlie an undulating erosion surface, suggesting a period of erosion, possibly related to a relative low stand of sea-level (Erohina et al., 2004, Fig. PB-3). The sand unit fines up-section into the mudstone, which contains lonestones, marine diatoms and dynocysts (Shipboard Scientific Party, 2001a). Strand et al. (2003) interpret the sand unit as a fluvial to delta plain channel deposit. They found sand-grain surface textures that suggest erosion and breakage by glaciers, implying the presence of at least valley glaciers in the hinterland of Prydz Bay. The overlying mudstone with lonestones suggests a marine transgression, with floating ice as a feature of the resulting shallow embayment. Macphail and Truswell (2004) report palynomorphs in the fine-grained units that indicate a late Eocene age (middle Nothofagites asperus zone), representing an age range from 33.9 to 39.1 Ma. This age overlaps with the age suggested by diatoms in the transgressive mudstones (33–37 Ma, Shipboard Scientific Party, 2001a). Macphail and Truswell (2004) also propose that the palynological assemblage was derived from a flora similar to stunted Nothofagus rainforest scrub, and consisted of ground-hugging plants and canopy trees about 1 m high. Today, such floras occur outside of Antarctica at higher altitudes, where cool temperatures limit tree growth. Therefore, the Prydz Bay flora reflects a cool to cold environment at sealevel. More precise temperature estimates are not possible because the plants present were tolerant of a wide range of conditions (Macphail and Truswell, 2004). 5.4.3. Ice-Sheet Development (Oligocene–Miocene) The older glacial section of the shelf comprises tabular units that pinch out shoreward due to inner-shelf erosion, and that extend seaward into prograding slope deposits (Cooper et al., 1991a). Palaeo-shelf edges for these units are better defined up-section as foreset strata steepen seaward (Fig. PB-2, Foldout PB-1). Shelf drilling (ODP Sites 739, 740, 741, 742 and 1166) recovered probable subglacial and glacimarine diamicts, with thin interbedded diatomaceous mudstones deposited during warm episodes (Hambrey et al., 1991; Erohina et al., 2004). The drilling and seismic evidence indicates glacial advance well across the Prydz Bay shelf during cold episodes, probably reaching the shelf edge. Over-compacted horizons indicate periods of glacial erosion and ice loading during the early Oligocene, Miocene and Plio-Pleistocene (Solheim et al., 1991; Shipboard Scientific Party, 2001a). Before the late Miocene, the Prydz Bay shelf prograded uniformly across its width, with the bulk of the ice and entrained sediment
140
A. K. Cooper et al.
coming from the southern end of the bay (i.e. from the Lambert Graben). The Prydz Bay continental slope became progressively steeper from the early phase of glaciation in early Oligocene time, to reach angles of as much as 81 on the present slope (Foldouts PB-1 and PB-2). On the continental rise, a pre-ice-sheet unit is overlain by one exhibiting channel-levee geometries. The nature of the change in geometry and the tracing of reflectors to the shelf drilling suggest that this change originated from the glacial expansion and increased sediment supply in the early Oligocene (Kuvaas and Leitchenkov, 1992). Overlying the channel-levee complexes are sequences that include thick mounds and sediment waves suggestive of contourite deposition, in addition to turbidite channels and associated levees formed by intensified down-slope and along-slope currents in the early Miocene (Fig. PB-2, Foldout PB-2). ODP Site 1165 (Leg 188) drilled 999 m with 69% recovery into a thick mound of lower Miocene and younger contourite sediments with turbidites only in the upper 5 m (Cooper and O’Brien, 2004). The hole penetrated the base of the mounded sequences, which was still of early Miocene age (Handwerger et al., 2004). The drilling confirmed the seismic interpretation that deposition of the thick contourite mounds had commenced by at least early Miocene time, but sediments above and below the surface were typical contourites – fissile claystones with abundant silt laminae (Handwerger et al., 2004). Therefore, there was no obvious lithological change in the hole to suggest a reason for the change from low relief submarine fans to highly mounded deposits, previously inferred to be mixed turbidite-contourites. ODP Site 1165 intersected a surface that can be mapped along the rise, and that marks a middle Miocene (14–16 Ma) change in sedimentation from laminated contourites to hemipelagic and pelagic facies (Cooper and O’Brien, 2004). Also, minerals and fossils recycled from shelf deposits first appear, suggesting the start of intense erosion by ice and overdeepening of the shelf. At this time, sedimentation rates slow more rapidly at the drill site, falling from 100 m/m.y. in the early Miocene to 37 m/m.y. in the midMiocene to 10 m/m.y. during the Plio-Pleistocene (Shipboard Scientific Party, 2001c; Florindo et al., 2003; Fig. PB-4). On shorter time scales, Gru¨tzner et al. (2003) examine the proportions of terrigenous sediment and biogenic opal in ODP Site 1165 between 3.4 and 7.6 Ma. They find high opal content from 5.8 to 5.2 Ma, which they relate to reduced sea ice and increased productivity. They also identify terrigenous intervals with high sedimentation rates from 7.2–6.6 Ma and 5.2–4.8 Ma, which they interpret as indicating high erosion rates and a fluctuating ice sheet under the influence of obliquity forcing. Gru¨tzner et al. (2003) also report cyclic variations in sediment composition and physical properties that have
Cenozoic Climate History from Seismic Reflection and Drilling Studies
141
Figure PB-4: Age-depth model for ODP Site 1165 from Shipboard Scientific Party (2001c) showing rapid sedimentation during the early Miocene, reducing rapidly through the late Miocene to Pliocene.
142
A. K. Cooper et al.
spectral peaks at B94, 41, 31, 21, and 18 k.y. cycles. Williams and Handwerger (2005) report that geophysical log parameters detect cycles of biogenic and terrigenous input at periods of B15–23 and B135 k.y., probably representing Milankovich-scale forcing of paleoenvironmental processes. Uncertainty in the age model of the hole prevents them from exact matching of peaks. 5.4.4. The Polar Ice Sheet (Late Miocene(?)–Pleistocene) In the early Pliocene, ice flow regimes changed and ice was focused into an ice stream on the western side of the bay, cutting a cross-shelf trough, the Prydz Channel. The ice stream delivered basal debris to the shelf edge, where the debris built a trough mouth fan on the upper continental slope (Prydz Channel Fan, O’Brien and Harris, 1996; O’Brien and Leitchenkov, 1997; O’Brien et al., 2004). On the banks adjacent to Prydz Channel, vertical aggradation of subglacial debris produced tabular units while glacial erosion overdeepened the inner shelf. Two ODP holes were drilled into the continental slope. ODP Site 743 was drilled to 98 mbsf into the eastern, steep part of the slope, and recovered diamict. ODP Site 1167 was drilled to 447.5 mbsf into the Prydz Fan, and recovered muddy, pebbly sands and diamicts deposited by slumping of subglacial debris interpreted to have originated at the ice grounding line at the shelf edge (Foldout PB-2, O’Brien et al., 2001; Passchier et al., 2003). ODP Site 1167 also recovered thin mudstone units deposited during periods of reduced ice extent (Shipboard Scientific Party, 2001b, Passchier et al., 2003). More than 90% of the fan was deposited before the mid Pleistocene, and there were only three advances of the Amery Ice Shelf to the shelf edge in the late Pleistocene (O’Brien et al., 2004). Clay mineralogy, magnetic properties and clast composition at ODP Site 1167 show changes suggesting that the Pleistocene peak of erosion and ice volume in the Lambert-Amery drainage system occurred in the early Pleistocene (O’Brien et al., 2004). Oxygen isotope measurements on foraminifera from ODP Site 1167 also suggest that sedimentation was reduced after the mid-Pleistocene, with the last ice advance to the shelf edge at about Marine Isotope Stage 16 (612–698 ka; Lisiecki and Raymo, 2005). However, the stratal record is fragmentary because hiatuses are common, which leads to a tentative identification of isotope stages (Theissen et al., 2003). During the mid-to-late Pleistocene, ice advances were less extensive. During the last glacial cycle, the Amery Ice Shelf grounded only 100 km north of the present ice shelf edge and far from the continental shelf edge (Domack et al., 1998; O’Brien et al., 1999) (Fig. PB-1).
Cenozoic Climate History from Seismic Reflection and Drilling Studies
143
On the continental rise, sedimentation rates decreased through the Pliocene and Pleistocene because less detritus was eroded from the continent and because sediment was deposited on the upper slope in front of the Prydz Channel. The inferred early Pliocene base of the Prydz Channel Fan is the prominent unconformity mapped by Mizukoshi et al. (1986, Reflector A) and O’Brien et al. (2004, Reflector PP-12). Drilling and seismic evidence indicate that glaciers advanced to the edge of the Prydz Bay shelf in cold episodes during the Pliocene and early Pleistocene, yet evidence of warm episodes also exists. Sediments in the Prince Charles Mountains indicate open-water fjordal environments in the Miocene to Pliocene (Hambrey and McKelvey, 2000; Whitehead et al., 2003, 2004). Lower Pliocene marine diatomite in the Vestfold Hills, on the eastern side of Prydz Bay, contains evidence of temperatures 41C warmer than today (Whitehead et al., 2001). ODP Site 1167 includes a thin mudstone horizon at 217 mbsf with calcareous nannoplankton not presently found in Prydz Bay (Shipboard Scientific Party, 2001b), suggesting warmer conditions at about 1.1 Ma (Pospical 2004; Lavelle, personal communication, 2001). These occurrences indicate warmer episodes when the Amery Ice Shelf edge retreated several hundred kilometres inland from its present position, and warmer water intruded Prydz Bay. 5.4.5. Prydz Bay Summary Seismic interpretation and drilling data reveal that the glaciation of Prydz Bay started in the latest Eocene. At that time, Prydz Bay was occupied by a fluviodeltaic plain covered with stunted cool-temperate vegetation. Rivers flowing through the Lambert Graben were fed by glaciers in the hinterland. The sea transgressed across the plain and floating ice delivered dropstones to the shallow embayment. The embayment became glaciated in the early Oligocene, with ice-sheet-scale glaciers depositing subglacial till, and glacimarine diamicts when the ice was not grounded. The ice was probably wet-based. In the early Miocene, a large temperate to polythermal ice sheet advanced and retreated across the embayment, supplying large quantities of detritus to the continental rise, where the detritus was deposited in large mounds. The mid-Miocene was marked by the start of a cooling trend and the development of a thicker, colder and more erosive ice sheet. Shelf overdeepening began, but progressively less detritus was delivered to the continental rise. In the early Pliocene, ice flow became concentrated on the western side of the bay in an ice stream that deposited sediment in a trough mouth fan. During warm phases, open-water extended landward as far as the
144
A. K. Cooper et al.
northern Prince Charles Mountains. Ice volumes and depths of erosion reached a peak in the mid Pleistocene and the cold, polar ice sheet was established. The Amery Ice Shelf no longer grounded at the shelf edge in Prydz Channel during glacial episodes.
5.5. Weddell Sea (Y. Kristoffersen and W. Jokat) The principal features in the Weddell Sea sector relevant to resolving the Antarctic paleoclimate and paleoceanographic history are prograding wedges of glacigenic sediments along the entire margin, a major troughmouth fan (Crary Fan), and numerous sediment drifts on the slope and in the deep basin, particularly along the western and northwestern side of the Antarctic Peninsula. Ice sheet flow-line patterns suggest that the continental margin of the eastern and southern Weddell Sea east of 451W receives drainage from the EAIS, whereas the continental margin west of 451W receives drainage from the WAIS (Fig. WS-1).
Figure WS-1: Track lines of multichannel seismic data in the Weddell Sea and locations of ODP drill sites. Bathymetry after Schenke et al. (1998). Areal extent of proximal and distal deposits of the Crary Trough Mouth Fan are outlined by light brown shaded area. True extent of sediment drifts in the southwestern and western Weddell Sea is poorly defined due to lack of data coverage. CTMF, Crary Trough Mouth Fan; EE, Explora Escarpment.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
145
5.5.1. The Regional Seismic and Geologic Database Modern geophysical data in the Weddell Sea comprise about 45,000 km of MCS lines from surveys principally by German, Norwegian and Russian research institutions since 1976 (Fig. WS-1). ODP drilled four sites in the Weddell Sea during ODP Leg 113, and ODP Site 693 on the Dronning Maud Land continental slope has been the most useful for stratigraphic calibration (Figs. WS-1 and WS-2). Prior to the deep drilling, stratigraphic studies in the region were conducted by Elverhøi and Maisey (1983), Hinz and Krause (1982), Hinz and Block (1984), Haugland et al. (1985) and Hinz and Kristoffersen (1987). Correlations with ODP Site 693 were made by Miller et al. (1990), Kuvaas and Kristoffersen (1991), Moons et al. (1992), Michels et al. (2002) and most extensively by Rogenhagen et al. (2004) (Fig. WS-2).
5.5.2. Acoustic Stratigraphy of the Shelf/Slope/Rise Environment-Spatial and Temporal Characteristics The continental shelf of the Weddell Sea is characterized by a prograding wedge of glacigenic sediments more than 1 km thick below the shelf edge (Fig. WS-3). The wedge downlaps onto older units, which are characterized by rather uniform thickness in the down-slope direction (Fig. WS-3a). Wedge deposition is a first order result of massive transport of unsorted texturally immature sediments by advance of a grounded ice sheet to the shelf edge (Barker et al., 1998). The acoustic response of coarse sediment in proximal positions below the shelf and uppermost slope is one of discontinuous reflection events. Continuity and definition of acoustic stratification improve in the down-slope direction as a result of progressive sorting and increased relative abundance of finer material. The shelf edge may appear rectilinear, but the three-dimensional wedge architecture in the eastern Weddell Sea reveals an amalgation of adjacent small discrete cones of glacial sediments sourced by smaller ice streams (Kristoffersen et al., 2000). The spectrum of cones reflects broad scale expansion of the EAIS, but adjacent cones may or may not be coeval. Topsets of the prograding wedge are generally truncated at the seabed. Shelf aggradation is indicated in the southern Weddell Sea west of the Crary Trough, but the vast shelf area west of 451W has not been accessible for seismic surveys (Fig. WS-1). The maximum thickness of prograding units below the mouth of the Crary Trough and also below the shelf north of Lyddan Ice Rise is more than 3 km (Rogenhagen et al., 2004). ODP Site 693 (Fig. WS-2) provides local
146
A. K. Cooper et al.
Figure WS-2: Stratigraphic summary column. Modified from Rogenhagen et al. (2004).
Figure WS-3: (a) Seismic line AWI-90110 across the Dronning Maud Land margin showing the prograding wedge (modified from Michels et al., 2002). (b) Seismic line AWI-97051 across the Larsen Shelf and Slope, showing the prograding shelf and sediment drift on the lower continental slope (modified from Michels et al., 2001). Profile locations are in Fig. WS-1.
148
A. K. Cooper et al.
Figure WS-4: Seismic line NARE-8517 across the prograding Crary Trough Mouth Fan (modified from Kuvaas and Kristoffersen, 1991). Profile location is in Fig. WS-1. calibration of the acoustic section (Miller et al., 1990), but regional extrapolations are inhibited along-slope by numerous canyons, and are inhibited down-slope by the steep Explora Escarpment (Fig. WS-1). The Crary Fan, a regional feature at the mouth of the Crary Trough, is associated with large channel/levee complexes, which extend up to 1,000 km to the north into the basin (Figs. WS-1, WS-4 and Foldout WS-1). Initial fan evolution is correlated with the resumption of sediment deposition above an Albian-early Oligocene hiatus at ODP Site 693 (Reflector W4). Sediment drifts are common within the Neogene stratigraphic interval along the continental slope (Fig. WS-1) in the western Weddell Sea (Michels et al., 2001; Maldonado et al., 2005).
5.5.3. The Weddell Sea Pre-Ice-Sheet Depositional Environment Acoustic stratigraphic information on the shelf is limited to subbottom depths comparable to the local water depth due to severe multiple reflections (e.g. Fig. WS-3). The pre-ice-sheet Cenozoic shelf edge was more than 10 km landward of its current position along the Dronning Maud Land continental margin (Kristoffersen et al., 2000), and 70 km to the south (Fig. WS-4) in the southern Weddell Sea (Kuvaas and Kristoffersen, 1991). The shoreward
Cenozoic Climate History from Seismic Reflection and Drilling Studies
149
shift in the western Weddell Sea is unknown. The deeper strata below the continental slope (i.e. below W4) appear unstructured throughout. The older sediments are thickest (5–8 km) below the Larsen Shelf in the western Weddell Sea (Rogenhagen and Jokat, 2000), and may be up to 15 km thick along the front of the Ronne and Filchner Ice shelves (Leitchenkov and Kudryavtzev, 2000). In the central Weddell Sea Basin, the pre-Oligocene section of inferred turbidites is more than 1 km thick, and thins by basal onlap towards the margins to less than 0.5 km (Rogenhagen et al., 2004). High seasonal variations in sea-surface temperatures and a well-developed seasonal thermocline characterized the early Paleogene Weddell Sea (Kennett and Barker, 1990). On Maud Rise, siliceous biogenic facies began to replace carbonate facies during the latest Eocene-earliest Oligocene (Kennett and Barker, 1990). A possible early Cenozoic seaway between East and West Antarctica could have been up to 700 m deep, and may have persisted into the Oligocene if no WAIS was present (Lawver and Gahagan, 2003). At ODP Site 693 on the middle continental slope, middle lower Oligocene and younger glacial sediments are separated by a hiatus from Albian radiolarian diatomite and claystones (i.e. Reflector W4). The unconformity may represent non-deposition and/or mild erosion (Kennett and Barker, 1990). 5.5.4. Change from Non-Glacial to Glacial Conditions Sediment fluxes on high latitude continental margins are closely connected to climate extremes. In the Weddell Sea, environmental change is manifested by a basin-wide change in acoustic character within the sedimentary section (Reflector W4) at about 1 s TWT below the sea bed (Rogenhagen et al., 2004). Younger deposits in the basin have finely laminated continuous acoustic stratification, and geometries on the slope are in the form of channel/levee complexes over a wide range of spatial scales. The change in depositional environment is interpreted to have originated from an increased sediment flux, caused by increased erosion of the continent and increased down-slope transport. At ODP Site 693 on the middle continental slope, the acoustic change correlates stratigraphically with resumed preservation of lower Oligocene sediments. The deposits include rounded dropstones in lower Oligocene (32–33 Ma) diatom muds, a signal of the first presence of glaciers on the adjacent parts of East Antarctica (Kennett and Barker, 1990). Subsequent early Miocene sedimentation rates at this site were low (7 m/Ma). A more dramatic change in sediment flux to the margin is documented by a threefold increase in sedimentation rate (to 24 m/m.y.), when sedimentation
150
A. K. Cooper et al.
resumed following a hiatus that spanned the middle Miocene. Increased sediment input is related to expansion of ice on the East Antarctic continent. The hiatus at ODP Site 693 correlates with a regional acoustic reflection event (W5) identified below the continental slope and rise along the entire Weddell Sea margin (Rogenhagen et al., 2004). Shelf progradation accelerated dramatically along the eastern and western margins of the Weddell Sea (Fig. WS-3), with grounded ice extending to the shelf edge in the late Miocene (Michels et al., 2001; Michels et al., 2002). A range of contourite drifts formed on the slope and rise in the northwestern Weddell Sea (Michels et al., 2001; Maldonado et al., 2005). Sedimentation rates at ODP Site 693 reached 60 m/m.y. in the early Pliocene, and subsequent Quaternary sedimentation rates were reduced to 16 m/m.y. (Gersonde et al., 1990). Sediment input to the margin in the southeastern Weddell Sea was focused toward a trough mouth fan. The Crary Fan began to expand at the time of change to a glacial environment (above Reflector W4, Fig. WS-4 and Foldout WS-1), and major channel/levee complexes evolved in three phases. The last of these three phases (Reflector W5, Fig. WS-4 and Foldout WS-1) was from the late Miocene on (Kuvaas and Kristoffersen, 1991; Moons et al., 1992). 5.5.5. The Glacial/Interglacial Environment The change from a glacial to an interglacial environment was associated with major changes in sediment flux. Average sediment deposition on the eastern Weddell Sea margin (101W) during the last two climatic cycles (300 k.y.) varies from 5 cm/k.y. on the upper slope to over 1 cm/k.y. on the lower slope (Grobe and Mackensen, 1992). Sedimentation was most rapid during the beginning of interglacials, with rates on the middle slope four to five times higher than during glacials. We note, however, that the grounded EAIS only reached the mid-shelf in this area during the LGM (Kristoffersen et al., 2000). Sediment input in the southern Weddell Sea was focused in the Crary Trough Mouth Fan (Figs. WS-1, WS-4 and Foldout WS-1). The fan comprises large channel-levees on the flanks of deep-water channels, such as the Cold Water Channel and the Deutschland Channel (Foldout WS-1). Grounded ice reached the shelf edge at the trough mouth during the last glaciation (Bentley and Anderson, 1998), and deposition on the levees (in water depths of 2,000–3,000 m) ranged from 100–200 cm/k.y. during the LGM to a few cm/k.y. during the present interglacial (Weber et al., 1994). Episodic sediment transport into the basin also occurred by mass flows during interglacials, probably as partial collapse of the deposits on the upper
Cenozoic Climate History from Seismic Reflection and Drilling Studies
151
continental slope. A 90-m-thick sandy turbidite unit was deposited within 0.5 m.y. during the early Gilbert Chron (4.8 Ma) at ODP Site 694 (Fig. WS-1), and may be the distal expression of mass wasting events on the continental slope in the southwestern Weddell Sea (Shipboard Scientific Party, 1988). Also, major early Pliocene drawdown of East Antarctic ice is postulated to have triggered extensive mass flows that originated from the Crary Trough Mouth Fan (Bart et al., 1999). In the western Weddell Sea, upper Miocene and younger sediments (above Reflector W5) are mostly drift deposits that reach a thickness of more than 1 km below the middle slope, seaward of the Larsen Shelf (Rogenhagen and Jokat, 2000; Michels et al., 2001; Maldonado et al., 2005). Present and past bottom currents circulated in nearly the opposite direction to channel transport, and cross-channel flow was in the same direction as the Coriolis force acting on down-slope turbidity currents in the southern Weddell Sea. Sediments scavenged from turbid channel flow by cross-channel bottom currents sourced the benthic boundary layer and enhanced formation of sediment drifts along the western and northern Weddell Basin. The actual drift distribution was mainly controlled by the physiography of the basin and bottom current flow directions (Maldonado et al., 2005). These drifts represent a storehouse of paleoceanographic and climatic proxies not yet sampled by scientific drilling. 5.5.6. Continental Margin Sediments and Ice-Sheet History The mass balance of the EAIS, the nature of the substratum and the continental topography, particularly in the coastal region, determine sediment input to the continental margin. Enhanced input of sediments to the continental margin at ODP Site 693 in the eastern Weddell Sea and development of a prograding wedge started in the latest Miocene and peaked during the earliest Pliocene (Gersonde et al., 1990). The seismic tie between ODP Site 693 and the southern Weddell Sea is uncertain, but Kuvaas and Kristoffersen (1991) suggest that fan development started in the southern Weddell Sea by the early Oligocene (above Reflector W4, Fig. WS-4 and Foldout WS-1), and that about two-thirds of the sediment thickness at the mouth of the present Crary Trough was already in place by the late Miocene (i.e. below Reflector W5). Channel-levee complexes have migrated eastward on the Crary Trough Mouth Fan, and late Miocene and younger deposition constructed a third major channel-levee complex and deposited about 1 km of sediments below the trough mouth (Fig. WS-4 and Foldout WS-1). These age relations imply that the principal input of sediments from East
152
A. K. Cooper et al.
Antarctica to the Weddell Sea margin from the early Oligocene to the late Miocene originated from a glaciated interior of the continent via the Crary Trough, and that there was effectively no input along the Dronning Maud Land margin. At this point, the significance of a local thickness maximum of glacial sediments north of Lyddan Ice Rise (Rogenhagen et al., 2004) is unclear. The EAIS expanded to the Dronning Maud Land margin during the latest Miocene–earliest Pliocene and formed a prograding wedge below the continental shelf and slope. Sea ice cover has prevented acquisition of the seismic data from west of 451W and north of the Ronne Ice Shelf (Fig. WS-1) needed to study the depositional geometries of sediments originating from the catchment area of the WAIS. Data from this area also are needed to study the relation between eastern and western sediment source regions. Moraine complexes on the shelf in the eastern Weddell Sea suggest that the EAIS was grounded on the mid-shelf and did not reach the shelf edge during the LGM (Kristoffersen et al., 2000), except at the mouth of the Crary Trough (Bentley and Anderson, 1998).
5.6. Antarctic Peninsula (R. Larter) Cenozoic tectonic processes have diversely affected the Antarctic Peninsula region and its climate record. Hence, we separately discuss four main subregions.
5.6.1. The Eastern Margin This subregion includes the Weddell Sea margin of the Antarctic Peninsula and Larsen Basin (Fig. AP-1). Persistent sea ice covers the region (Gloersen et al., 1992), hence relatively few research cruises have been conducted here. Macdonald et al. (1988) used regional geologic and aerogeophysical data to infer that a large Mesozoic–Cenozoic sedimentary basin extends B700 km south from James Ross Island. Four main seismic stratigraphic units are identified from SCS reflection data on the shelf and upper slope (Anderson et al., 1992; Sloan et al., 1995; Fig. AP-2): Unit 4: acoustic basement interpreted as Jurassic and younger volcanic rocks; Unit 3: seaward-dipping reflections interpreted as Late Cretaceous to Oligocene marine shelf deposits, the older part of which are coeval with those on nearby Seymour Island; Unit 2: prograding sequences with truncated foresets that downlap onto Unit 3, and that are thought to have been deposited by multiple advances of grounded ice across the shelf in the Miocene and early Pliocene;
Figure AP-1: (a) Track lines of multichannel (thick lines) and single-channel (thin lines) seismic data in the Antarctic Peninsula region, DSDP and ODP drill sites (filled circles, annotated with site numbers) and SHALDRIL sites (open squares). SHALDRIL sites are only marked where either W10 m subseafloor penetration was achieved or pre-quaternary sediments were recovered. Bathymetric contours are at 1,000 m intervals down to 4,000 m, with 500 m contours included locally on the shelf. Bathymetric data are based on Smith and Sandwell (1997) east of the Peninsula and north of 621S. Bathymetry for most of the area west of the Peninsula is from Rebesco et al. (1998), the exception being the 500 m contours in the southern Bellingshausen Sea, which are from O´ Cofaigh et al. (2005). To the east of the Peninsula, 500 m contours south of James Ross Island (JRI) are based on figures in Evans et al. (2005), and the 500 m contour north of James Ross Island is based on multibeam echo sounding data maps produced from data collected on RV Nathaniel B. Palmer Cruises 0003 and 0107. Bold line is location of seismic profile in Fig. AP-2. SOM: South Orkney Microcontinent; JB: Jane Basin; PB: Powell Basin; BS: Bransfield Strait; RT: Robertson Trough. (b) Expanded map of the northern Antarctic Peninsula with SHALDRIL site numbers labelled (I/II indicates first/second cruise). Same bathymetry contours shown as in (a), except 500 m contours around JRI omitted. SSI: South Shetland Islands.
154
A. K. Cooper et al.
Unit 1: aggrading reflections interpreted as deposits from fluctuating dynamic ice sheets in the Pliocene and Pleistocene. MCS reflection data show at least 8 km of sediment at the base of the continental slope, overlying likely Early Cretaceous age basement (Barker and Lonsdale, 1991). The northern continental slope has plastered contourite drift deposits up to 900 m thick, thought to have been deposited by northflowing glacially influenced bottom currents. Pudsey (2002) suggests that drift deposition began in the early Miocene at the onset of bottom water flow, or in the latest Miocene at the onset of volumnious glacially derived sediment supply to the western Weddell Sea. Late Quaternary shelf sediments have been sampled by seafloor coring (e.g. Domack et al., 2001a, b, 2005; Pudsey and Evans, 2001; Pudsey et al., 2001; Brachfeld et al., 2003; Evans et al., 2005). These researchers infer that grounded ice converged into major ice streams and advanced to the shelf edge during the LGM, that the Prince Gustav Sound Ice Shelf collapsed and reformed in the mid-Holocene, and that the recent collapse of the Larsen B Ice Shelf is unprecedented during the Holocene. Recent drilling by the SHALDRIL project has obtained the first samples from older sequences (Shipboard Scientific Party, 2005, 2006; Anderson et al., 2006, 2007).
Figure AP-2: Single-channel seismic line across the eastern margin of the Antarctic Peninsula at 65115uS, collected on RV Nathaniel B. Palmer in 1993. Line drawing interpretation shows the internal geometries and boundaries (dashed lines) between the main seismic units. Vertical exaggeration at the seafloor is 82:1. Adapted from Sloan et al. (1995). Line location is shown in Fig. AP-1.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
155
Sequences drilled by SHALDRIL are principally shallow marine sands, sandy and silty muds, and pebbly muddy sands from neritic environments, with mollusc shells distributed throughout the cores. The cores are dated as: late Eocene to early Oligocene, Oligocene, middle Miocene, early Pliocene and Holocene. 5.6.2. The South Orkney Islands Region This subregion includes the South Orkney Microcontinent (SOM) and the adjacent deep-water Jane and Powell basins (Fig. AP-1). The SOM extends about 350 km from east to west and 250 km from north to south, and is underlain by Mesozoic metamorphic and sedimentary rocks (Thomson, 1981; Dalziel, 1984). Offshore, the SOM includes four Cenozoic sedimentary basins (King and Barker, 1988) with up to 5 km of sediment (Harrington et al., 1972; Busetti et al., 2001, 2002). Powell Basin (up to 3,600 m deep) formed as the SOM rifted and drifted away from the tip of the Antarctic Peninsula in late Eocene to late Oligocene time (King and Barker, 1988; Lawver et al., 1994; Coren et al., 1997; Eagles and Livermore, 2002). Opening of Jane Basin (up to 3,300 m deep) probably began slightly later (Lawver et al., 1991, 1994), and may have continued until the middle Miocene according to Maldonado et al. (1998). From SCS data on the SOM, King and Barker (1988) defined pre-rift, synrift, and post-rift units. The post-rift sediments are less than 1 km thick (e.g. Busetti et al., 2001, 2002) and were drilled at ODP Site 695 (1,300 m water depth) and ODP Site 696 (600 m water depth) (Fig. AP-3). They comprise Oligocene or early Miocene to Quaternary terrigenous sediments, with rare coarse-grained IRD until the late Miocene (B8.7 Ma) and common IRD thereafter. Middle Miocene to Quaternary sediments are hemipelagic and diatomaceous muds and oozes (Barker et al., 1988a, b). ODP Site 696 also sampled syn-rift Eocene sandy mudstones (Sequence 2) that have nannofossil assemblages and clay minerals suggesting a relatively warm climate, and palynoflora indicating temperate beech forests and ferns on West Antarctica. Drilling results suggest intermittent glaciation with little sea ice during most of the Miocene and a persistent ice cap to sea-level on West Antarctica since the late Miocene. Herron and Anderson (1990) place the maximum late Quaternary grounding line advance at the 300 m isobath, and consider openmarine conditions to have existed over the SOM since 6,000 y. B.P. based on SCS and seafloor core data. In Powell Basin, post-early-rift sediments are up to 3 km thick. King et al. (1997) identified two seismic units with low reflectivity below and high
156
A. K. Cooper et al.
Figure AP-3: Part of seismic line AMG845-18, showing the setting of ODP Site 696 in relation to the seismic units (S1–S3) described by King and Barker (1988). Vertical exaggeration at the seafloor is 3.3:1. Adapted from Barker et al., 1988a. ODP Site 696 location shown in Fig. AP-1. reflectivity above. They interpret the change as recording the onset of glacial–interglacial cyclicity in the supply of coarse detritus to the basin in the late Miocene. A similar upward change in reflectivity is observed in Jane Basin (Maldonado et al., 1998). The reflectivity change may also be due to silica diagenesis (e.g. Lonsdale, 1990; Volpi et al., 2003). Maldonado et al. (2006) identify five seismic units in Jane and neighbouring ocean basins, and relate changes in seismic characteristics to variations in bottom water flow since the middle Miocene. ODP Site 697 was drilled in Jane Basin (Fig. AP-1) to B323 mbsf, and recovered mainly early Pliocene and younger hemipelagic sediments with IRD throughout; however, IRD is abundant only near the base of the sequence (Barker et al., 1988a, b). Other seismic studies of Powell Basin (e.g. Kavoun and Vinnikovskaya, 1994; Coren et al., 1997; Viseras and Maldonado, 1999) focus on the post-early Oligocene rift history of the basin, but are limited in paleoclimate interpretations by the lack of drilling data.
5.6.3. The South Shetland Islands Region This subregion includes the Bransfield Strait and the continental margin around the South Shetland Islands (Fig. AP-1). Bransfield Strait is a 2,000 m deep rift basin that is actively extending at 7 mm/y. (Dietrich et al., 2004). The time of initial extension and the oldest age of basin sediments are uncertain, but may be 4 Ma (Barker and Dalziel, 1983) or B6 Ma (Larter and Barker, 1991a). Gamboˆa and Maldonado (1990) speculate that
Cenozoic Climate History from Seismic Reflection and Drilling Studies
157
Bransfield Strait may have opened earlier, during the early Miocene, and have been continuous with mid-shelf basins to the southwest. From MCS data, Gamboˆa and Maldonado (1990) identify ‘rift’ and ‘drift’ sequences in Bransfield Strait. ‘Drift’ sequences prograde the shelf and are about 1 km thick. In SCS data, Jeffers and Anderson (1990) define four glacio-eustatic sequences within the ‘drift’ sequences, and interpret all four sequences as being younger than 3 Ma. Prieto et al. (1999) use different SCS data to define eight seismic units that comprise interfingering slope and basinal deposits within the ‘drift’ sequences. They interpret the slope units as having been deposited directly from grounded ice during glacial periods, and interpret the basinal units as having been deposited by mass flow processes during deglaciations and interglacial periods. From SCS data, Banfield and Anderson (1995) identify sediment mound features that they infer to be glacial grounding moraines in up to 1,000 m water depth on the southeastern flank of Bransfield Strait. They speculate that mounds at B700 m depth mark the maximum advance during the LGM. Deep troughs with mega-scale lineations are incised into the shelf and are interpreted as the paths of palaeo-ice streams (Banfield and Anderson, 1995; Canals et al., 2002). MCS profiles across the continental slope NW of the South Shetland Islands reveal a forearc basin, with more than 1.5 km sediments, that is bounded to the NW by a small accretionary prism (Maldonado et al., 1994a, b). The prism overthrusts trench-fill sediments that may have been deposited rapidly and are up to 1 km thick (Maldonado et al., 1994a, b; Kim et al., 1995). Other seismic-reflection surveys have been done in the region by British, Polish, German, Spanish, US, Italian, Chinese and Korean research groups, but published results focus on the tectonic evolution of the region and tectonic processes (Barker, 1976; Guterch et al., 1985; GRAPE Team, 1990; Acosta et al., 1992; Henriet et al., 1992; Grad et al., 1993; Barker and Austin, 1994, 1998; Bochu et al., 1995; Gra`cia et al., 1996; Jin et al., 1996; Jin and Kim, 1998; Prieto et al., 1998; Jin et al., 2002), on gas hydrates (Lodolo et al., 1993, 2002; Tinivella et al., 1998, 2002; Jin et al., 2003), and on a large submarine slide (Imbo et al., 2003). Swath bathymetry data exist over most deep-water parts (W1,000 m) of Bransfield Strait (Lawver et al., 1996; Gra`cia et al., 1997). Along Boyd Strait, ‘bundle structures’ and mega-scale glacial lineations occur and confirm palaeo-ice stream flow during glacial periods (Canals et al., 2000; COHIMAR/SEDANO Scientific Party, 2003). MCS data along outer Boyd Strait reveal glacial progradation of the margin (Maldonado et al., 1994a, b), and deep-tow boomer profiles reveal a glacier grounding zone wedge near the shelf edge (Vanneste and Larter, 1995).
158
A. K. Cooper et al.
The only scientific drilling in the South Shetland Islands region was done at SHALDRIL-I Site 1 in Maxwell Bay (Fig. AP-1), where an expanded sequence of Holocene diatomaceous muds, B105 m thick, overlying a clayrich diamicton was sampled (Shipboard Scientific Party, 2005). 5.6.4. The Pacific Margin The Pacific margin includes the region southwest from the South Shetland Islands to 701S and 801W. This is a former active margin where ridge-crest segments were progressively subducted (Larter and Barker, 1991a; Henriet et al., 1992), followed by 1–4 m.y. of uplift and then long-term subsidence (Larter and Barker, 1989, 1991b; Anderson et al., 1990; Gamboˆa and Maldonado, 1990; Bart and Anderson, 1995, 1996, 2000; Larter et al., 1997). Seismic profiles show that outer shelf sequences are separated from NW-SE trending mid-shelf basins by the Mid-shelf High (MSH) (Kimura, 1982; Anderson et al., 1990; Gamboˆa and Maldonado, 1990; Larter et al., 1997). Several research groups have conducted seismic studies in the area (Kimura, 1982; Larter and Barker, 1989, 1991b; Anderson et al., 1990; Gamboˆa and Maldonado, 1990; Henriet et al., 1992; Bart and Anderson, 1995, 1996, 2000; McGinnis and Hayes, 1995; Rebesco et al., 1996, 1997, 2002, 2006; McGinnis et al., 1997; Larter et al., 1997; Jin et al., 2002; Jabaloy et al., 2003; Herna´ndez-Molina et al., 2006a). Cenozoic sequences have been drilled at DSDP Site 325 (Hollister et al., 1976) and at multiple sites during ODP Leg 178 (Barker et al., 1999). Evidence of Oligocene glaciation exists on King George Island (Birkenmajer, 1991; Dingle and Lavelle, 1998; Troedson and Smellie, 2002), but offshore the first ice sheets on the Pacific margin are inferred from early Miocene IRD at DSDP Site 325 (Fig. AP-1). The oldest sediments from ODP Leg 178 are drift deposits at ODP Site 1095, dated at 9.6 Ma, where all cores show glacial influence and sedimentation rates decrease steadily from the late Miocene to the Quaternary. From seismic studies, Rebesco et al. (1996, 1997, 2002) suggest that sediment drift deposition began around the middle Miocene, with most growth in the late Miocene. However, Herna´ndez-Molina et al. (2004, 2006b) describe a buried sediment drift of early Miocene age (Fig. AP-1). Uenzelmann-Neben (2006) also interprets depositional patterns of early Miocene continental rise sediments as reflecting bottom current influence, but infers a different flow direction. A regular supply of both glacially derived terrigenous sediments and
Cenozoic Climate History from Seismic Reflection and Drilling Studies
159
interglacial biogenic sediments has reached the continental rise since at least the middle Miocene. The outer continental shelf is underlain by depositional sequences that thicken seaward (Larter and Barker, 1989, 1991b; Anderson et al., 1990; Gamboˆa and Maldonado, 1990; Larter and Cunningham, 1993; Bart and Anderson, 1995; Table AP-1). Aggrading sequences (S3) without a distinct paleo-shelf edge (PSE) are unconformably overlain by prograding sequences (S2, S1) with an abrupt PSE (Foldout AP-1). The change occurs at the S3/S2 boundary of Larter and Barker (1989, 1991b) and is observed all along the Pacific margin (Anderson et al., 1990; Bart and Anderson, 1995, Larter et al., 1997; Jin et al., 2002) and west along the margin of the Bellingshausen and Amundsen Seas (Nitsche et al., 1997). Drilling at ODP Sites 1097 and 1103, although with very poor recovery, suggests that the change occurred between 8 and 6 Ma (Iwai and Winter, 2002; Bart et al., 2005). Cores from S3 are diamictons with interbedded mudstones and graded sandstones, interpreted by Eyles et al. (2001) as continental slope deposits, although seismic profiles suggest a palaeo-shelf to slope transition further offshore. Cores from S1 and the upper part of S2 show abundant evidence for having been deposited subglacially (Eyles et al., 2001). Larter and Barker (1989, 1991b) and Larter et al. (1997) interpreted the S3/S2 boundary as representing the onset of frequent advances of grounded ice to the palaeo-shelf edge. However, Bart and Anderson (1995) and Bart et al. (2005, 2007) suggest that palaeo-ice streams cut erosional troughs within S3 and hence existed earlier than the S3/S2 boundary. Although equivocal, the boundary could represent a change in the typical extent of glacial advances, in the dynamic behaviour of ice sheets that advanced onto the shelf, or in the way the ice transported sediments (Larter et al., 1997; Herna´ndez-Molina et al., 2006a). Going up-section above the S3/S2 boundary, foreset stratal dips generally increase and PSE progradation in individual sequences decreases (Foldout AP-1). The regional S2/S1 boundary within the upper sedimentary section may have been produced by ice-sheet erosion during glacial periods and lower sea levels after the Late Pliocene increase in the volume of Northern Hemisphere ice sheets (Larter and Barker, 1989, 1991b). By seismic correlation across W100 km to ODP Site 1101, Rebesco et al. (2006) estimate an age of B3 Ma for a boundary that they identify as S2/S1. However, this boundary marks a change in stratal geometry that is generally characteristic of S3/S2, and Larter (2007) suggests that it probably corresponds to this earlier unconformity.
Table AP-1: Summary of stratigraphic schemes used in previous publications to describe the late Miocene to recent depositional sequences on the Pacific margin of the Antarctic Peninsula.
160 A. K. Cooper et al.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
161
Mid-shelf basins contain up to 2 km of sediment in a broad synform that is truncated at shallow depth beneath the seafloor (Kimura, 1982; Anderson et al., 1990; Gamboˆa and Maldonado, 1990; Larter et al., 1997), making the succession accessible to shallow drilling. However, these sediments have not been sampled, except for the thin Quaternary cover. Basin sediments are probably all Tertiary and may be as young as early Miocene off Adelaide Island, and middle Miocene off Anvers Island (Larter et al., 1997). The inner shelf is mostly shallower than 200 m, but has deep troughs, such as the Palmer Deep where ODP Sites 1098 and 1099 were drilled in 1,400 m of water. Holocene successions B47 and B108 m thick were recovered at ODP Sites 1098 and 1099, respectively (Shipboard Scientific Party, 1999; Domack et al., 2001; Ishman and Sperling, 2002; Leventer et al., 2002; Shevenell and Kennett, 2002). Swath bathymetric data show seafloor features of subglacial origin on the shelf (O´ Cofaigh et al., 2002; Dowdeswell et al., 2004; Amblas et al., 2006), and confirm that a grounded ice sheet with ice streams extended to the shelf edge during the LGM (Pudsey et al., 1994; Larter and Vanneste, 1995). Seafloor core data indicate that retreat of the ice sheet from the outer and middle shelf after the LGM occurred between 18,500 and 13,000 cal. y. B.P. (Pudsey et al., 1994; Heroy and Anderson, 2005).
5.7. Other Sectors of the Antarctic Continental Margin Other sectors of the Antarctic margin, than the five ANTOSTRAT project working areas discussed above, have been studied more fully than previously during the past 5–7 years, in a time of renewed interest in MCS studies of the continental margin. These regions include similar geomorphic and stratigraphic features. Although we do not include them in the discussion below due to space considerations, we include here representative citations to some of the studies published for the Bellingshausen and Amundsen seas (Hollister et al., 1976; Tucholke and Houtz, 1976; Tucholke, 1977; Kimura, 1982; Yamaguchi et al., 1988; Cunningham et al., 1994, 2002; Gohl et al., 1997, 2007; Nitsche et al., 1997, 2000; Wellner et al., 2001, 2006; Lowe and Anderson, 2002; O´ Cofaigh et al., 2005; Dowdeswell et al., 2006; Evans et al., 2006; Scheuer et al., 2006a, b; Larter et al., 2007; Uenzelmann-Neben et al., 2007) and for offshore areas of Queen Maud Land and Enderby Land (Kuvaas et al., 2004a, b, 2005; Hinz et al., 2005; Stagg et al., 2004a, b; Leitchenkov et al., 2007; Solli et al., 2007a, b, c).
162
A. K. Cooper et al.
5.8. Discussion Our discussion focuses on integrating key observations and inferences from the Antarctic continental margin to summarize a Cenozoic glacial history from the stratigraphic record. We recognize the limitations of the seismic and drilling data sets used. Seismic data image regional and up-section changes, but only provide a relative history of inferred events and processes. Drilling data provide a defined stratigraphic history, but one that is valid only at the local drilling site. There is an extensive published literature that illustrates the large spatial and temporal variability of features and processes at various scales, depending on the resolution and extent of data analysed. We selectively discuss the widespread seismic stratigraphic variations and the long- and short-period geologic transitions in cores, all of which point to a varied history of non-glacial and glacial events on the continental margin. The discussion is necessarily condensed due to the great breadth of the topics and the limited space herein. We abbreviate offshore geographic regions as: Ross Sea (RS), Antarctic Peninsula (AP), Weddell Sea (WS), Prydz Bay (PB) and Wilkes Land (WL), and prefix these abbreviations with east (E) and west (W), such as western Ross Sea (WRS) (Fig. I-1).
5.8.1. Regional Seismic Stratigraphic Variations: Similarities and Differences Stratigraphers have long recognized sedimentary sequences and bounding regional unconformities in seismic-reflection data across the Antarctic continental margin (e.g. Hinz and Block, 1984; Wannesson et al., 1985; Hinz and Kristoffersen, 1987; Cooper et al., 1995). These sequences, principally of inferred Cenozoic age, commonly have similar seismic geometries around Antarctica (e.g. Cooper et al., 1991b; Anderson, 1999). Some geometries are unique to polar continental margins, whereas other geometries are like those of low-latitude non-polar margins (e.g. Hinz and Block, 1984; Bartek and Anderson, 1991; Table D-1). A unified circum-Antarctic seismic stratigraphy does not exist, but an International effort to compile one via the CASP project is in progress (Davey and Cooper, 2007). Numerous separate and sometimes different seismic stratigraphies exist for various localities. Many seismic features have been cited to suggest intermittent ice on the Antarctic margin (Table D-1), but those giving the strongest evidence for ice are the regional seismic unconformities, the broad erosional troughs and depositional banks on the continental shelf, and the large-scale fans and
Cenozoic Climate History from Seismic Reflection and Drilling Studies
163
Table D1-1: Common large-scale geomorphic and seismic stratigraphic features of the Antarctic continental margin (listed by location and decreasing inferred age). Feature Continental shelf A. Deep regional seismic unconformities on the shelf (lower B1/3 of sedimentary section) B. Shallow regional seismic unconformities on the shelf (upper B2/3 of section) C. Prograded and aggraded sequences under the outer shelf D. Mound features with chaotic seismic facies E. Broad cross shelf troughs and adjacent banks F. Overdeepened and foredeepened continental shelf Continental slope G. Regional seismic unconformities H. Massive sediment fan on the slope at the outlet of a broad cross-shelf trough I. Steep slopes and migrating high-relief channels J. Variable size sediment fans on the upper slope at mouths of seafloor and buried troughs. Continental rise K. Regional seismic unconformities
Regions
Processa
Timingb
PB, RS
Eustacy
Cretaceous to late Oligocene (?)
All
Ice-sheet erosion
late Oligocene (?) and younger
All
Oligocene and younger
All
Sediment carried to the shelf by glaciers Ice-sheet deposition
All
Ice-stream erosion
All
Ice-sheet erosion
mid-Miocene and younger
All
Bottom currents
PB, WS
Sediment deposited at shelf edge by one broad ice stream Bottom currents with coarse sediment Sediment deposited at shelf edge by multiple ice streams
Cretaceous and younger early Oligocene and younger (WS); early Pliocene and younger (PB) Oligocene and younger
All
WL, RS, AP
All
Bottom currents and diagenesis
late Oligocene and younger early Miocene and younger
late Miocene and younger
Cretaceous and younger
164
A. K. Cooper et al.
Table D1-1: (Continued ). Feature
Regions
L. Large sediment drift features
All
M. An up-section landward shift of depocenters from rise to slope.
All
Processa Bottom currents and downslope sediment supply Reduction of sediment supply to rise
Timingb early Miocene and younger mid-Miocene and younger
a
The principal process is listed – others such as lithospheric loading, sediment loading, paleoceanographic processes, diagenesis, etc. are also commonly involved. b Initiation time of features varies in different regions.
prograding deposits on the continental slope at the mouths of the shelf troughs. These features are observed on all sediment-covered segments of the Antarctic margin, in East and West Antarctica, and principally occur in the upper part of the stratigraphic section. These features are increasingly common up-section, indicating more abundant glacial events more recently. The ubiquitous overdeepened and foredeepened depth profile of the continental shelf is the ultimate evidence of sustained strong glacial erosion of the entire Antarctic margin. Ten Brink et al. (1995) modelled the geometries of the seafloor and stratigraphic sections on the continental shelf and upper slope, incorporating lithospheric, glacial and eustatic processes in the models. They showed that multiple advances and retreats of grounded ice sheets across the continental shelf, coupled with redistribution of sediment from onshore and shelf areas to the continental slope, are required to match the observed geometries. Eustatic and paleoceanographic processes are important for sediment redistribution, especially on the continental slope and rise, but are not sufficient by themselves to explain the shelf erosion and prograding geometries beneath the outer continental shelf. Bartek et al. (1991) illustrated that the stratal signatures of the Oligocene and younger sections in the eastern Ross Sea (i.e. unconformities and prograding sections) are similar to those of low-latitude non-polar margins, indicating that the Neogene stratal signature results from glacio-eustatic fluctuations. Seismic data provide a relative history of increasing circum-Antarctic glacial events, as noted above, but drilling and seafloor coring provide the only absolute age control and ground truth of glacial lithologies and processes. Definitive ages are limited due to the small number of cored sites, which in turn are biased toward sampling of shallow younger sections.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
165
The lithostratigraphic record from proximal drilling and dredging on the Antarctic margin ‘establishes’ a general history (at the drill sites) of no regional glaciers in Cretaceous and earlier times on the shelf (PB, RS, WL) and on the slope (WS). Evidence of glacial episodes is first seen in the late Eocene to early Oligocene as diamicts from grounded ice on the shelf (PB, RS) and from glacial erratics on the slope (WS). Upper Oligocene glacial marine deposits are sampled on the shelf (RS) and slope (WS). Lower Miocene sections show increasing evidence of ice and deep-ocean currents, with IRD and sediment drifts on the rise (PB, WL, AP, WS?) and glacial marine sediments and diamicts on the shelf (RS). The middle Miocene has an increased glacial hemipelagic signature on the rise (PB), and glacial marine deposits on the shelf (RS). The late Miocene and early Pliocene were times of enhanced glacial activity, as recorded by: (1) shelf deposits of glacial diamicts (PB, AP) and glacial marine sediments (RS, AP); (2) glacial marine sediments on the slope (WS, WL); and (3) rise deposits of glacial hemipelagics (PB, WL, RS, AP) and turbidites (WL, WS, AP). Upper Neogene deposits are principally glacial on the shelf (PB, RS, AP), slope (PB, WS) and rise (PB, WL, RS, AP, WS). Based on recovered drill cores and Antarctic Peninsula coastal geology, the general history of events for East and West Antarctica is essentially the same since the early Oligocene. However, the middle Miocene and Oligocene history for offshore West Antarctica is based on only two Ross Sea DSDP cores (Hayes and Frakes, 1975) and two tentatively dated SHALDRIL cores from the western Weddell Sea (Anderson et al., 2006; Shipboard Scientific Party, 2006). The above general history at the sparse drill sites has been greatly expanded by many investigators who have traced seismic unconformities and seismic stratigraphic units from core sites and rare onshore sedimentary sections up to hundreds of km to infer ages and lithofacies for key stratigraphic features listed in Table D-1. The expanded seismic stratigraphic history includes inferences of the following (although uncertainties about these inferences are inherently greater than the uncertainties about ages at drillsites): A pre-ice-sheet to early-glacial period (Cretaceous to early Oligocene): On the inner and mid shelf, variable seismic facies (WRS, PB), narrow channel geometries (PB) and sea-floor dredged/cored rocks indicate the presence of pre-ice-sheet subaerial, fluvial and shallow marine environments in the Cretaceous. These are unconformably overlain (WRS, PB) by upper Eocene to lower Oligocene early glacial sediments. This transition, from pre-ice-sheet to early glacial conditions, has not been sampled elsewhere on the shelf. Beneath the slope and upper rise, thick sediment sections are observed (all areas), but are not yet well imaged and mapped in most areas.
166
A. K. Cooper et al.
Aggrading shelf period (Oligocene): On the shelf, stratal geometries in drilled Oligocene glaciomarine sections mostly aggrade the PSEs in PB and ERS, but PSEs appear to prograde where the shelf is strongly uplifted and eroded (WWS). Aggrading PSE geometries occur in unsampled areas of the outer shelf (WL, AP). On the slope, geometries show high-relief paleo-slopecanyons (PB, WL, RS, WS), and in the WS, deposition of the massive Crary Fan (WS) began. On the rise, seismic facies indicate higher energy depositional environments with paleo-channel-levee systems first developing (PB, WL, RS, AP, WS). Where drilled (PB, WL, RS), lithologies from this period have glacial components indicating onshore glaciers. Backstripping calculations indicate normal shelf water-depths (PB, RS). Uniform prograding shelf (early and middle Miocene): In many regions, seismic sequences uniformly prograde the continental shelf edge, with an up-section increase in the dips of foreset-beds (glaciomarine deposits) and variable erosion of topset strata (with diamicton) (PB, WL, RS, AP, WS). Sea-level stratigraphic control began to shift to ice-dominated stratigraphic control, with documented cyclic shelf erosion by grounded ice sheets. Initial regional erosion and overdeepening of East Antarctic shelves commenced (PB, RS, WL). Slope geometries indicate canyon shifting and infilling (PB, WL, RS) and fan growth (WS), with rise geometries showing the construction of large drift mounds (WWS, PB, WL, AP) and channel-levees (WL, WS, AP, PB). Abundant contourite deposits (PB, AP) with some turbidites (AP, WL) are documented. Glacial and interglacial sediment volumes decreased on the rise (PB, AP), but increased on the slope (WS). Local and focused prograding shelf (late Miocene to Pleistocene): A prominent regional unconformity occurs in the late Miocene to early Pliocene across all margin segments (PB (A, PP12); WL (U8); RS (U2); AP (BGMS) WS (W5)). The unconformity marks a circum-Antarctic change from areas of uniform PSE progradation to local-arcuate and broadly focused PSE progradation into small and/or overlapping upper-slope fans (e.g. QML, WL, AP) and broad trough-mouth fans (e.g. PB, WS) lying at or near the end of cross-shelf troughs. Strong regional shelf erosion, by both narrow and wide ice streams early in this period, was followed by regional deposition of topset banks composed principally of glacial diamicton (PB, AP, WL, WS?) and glacial marine (RS) sediment. Periods of pelagic sedimentation on the shelf indicate open water and sea ice. On the slope, foreset dips steepened as fans formed above the unconformity. On the rise, sedimentation rates decreased (PB, AP) and depocentres shifted progressively landward, moving from the rise to beneath the slope (PB, WL, AP,
Cenozoic Climate History from Seismic Reflection and Drilling Studies
167
WS, AP?). Stratigraphic control was dominated by episodic grounded ice, with sediment deposition by ice and sediment distribution by ocean-currents. Sediment drape (late Pleistocene and Holocene): Thin well-layered acoustic units commonly infill shelf depressions and drape across the slope and rise (All regions). These units provide a record of pelagic sedimentation from the last few interglacial and glacial periods. Deep inner-shelf basins, in particular, trap Holocene biogenic sediments with that were deposited at very high sedimentation rates; these biogenic sediments yield an ultra-highto high-resolution (i.e. decadal to millennial) record of climate variability (AP, WL, PB).
5.8.2. Long- and Short-Period Transitions in the Geologic Record The seismic stratigraphic record provides a regional framework that illustrates distinct changes in the morphology of the Antarctic margin over the last 60 m.y. Drill cores provide the direct ‘ground truth’ geologic record of both long-period (m.y.) and short-period (k.y.) transitions (Table D-2). Some of the transitions in drill cores are reflected in the seismic stratigraphic framework and others are not. All transitions, however, are important in deciphering Antarctic paleoenvironments. In this discussion, we focus on lithostratigraphic changes in drill cores, and leave the discussion of isotopic, biostratigraphic and other relevant variations to authors of other chapters of this book. Our intent is to use the proximal lithostratigraphic drilling record from the continental margin to independently evaluate paleoenvironmental history, where possible. Drill cores from two segments of the continental shelf provide a general lithostratigraphic framework for the long-period systematic transition from non-glacial (Mesozoic) to fluctuating glacial and interglacial (late Cenozoic) paleo-depositional environments. At the mid-shelf of Prydz Bay, cores document the changes from subaerial non-glacial (Cretaceous) to fluvial/ lagoonal early glacial (late Eocene) to shallow marine early glacial (early Oligocene) to subglacial deep shelf (early Pliocene) to interglacial openmarine (Holocene) conditions. A similar transition is documented in the Ross Sea (McMurdo Sound), from subaerial non-glacial (Mesozoic) to shallow marine early glacial (early Oligocene) to fluctuating subglacial and marine glacial (early Miocene to Holocene) to interglacial open-marine (Holocene) environments. Large hiatuses exist in the shelf cores. Improved resolution of the continuity and timing of changes is seen in drill cores from the continental rise, where geologic continuity and core recovery
168
A. K. Cooper et al.
Table D2-2: Lithostratigraphic transitions in the geologic records from drill cores from the antarctic margin (listed by duration and decreasing inferred age). Feature
Regions
Long-period changes (m.y.) On the shelf: up-section PB lithologic changes from alluvial to fluvial to shallow marine to marine glacial to subglacial On the rise, sedimentation PB, AP, rates of hemipelagic WS sediments decrease smoothly while on the slope, sedimentation rates increase stepwise (i.e. in distinct stratigraphic units) On the rise, up-section PB increases in IRD, diatom content; shift kaolinite and glauconite Short-period changes (k.y.) On the inner shelf: cyclic RS changes from diamict (glacial) to glacial marine (interglacial) facies (in early Miocene at Milankovitch frequencies) On the rise: cyclic changes PB, AP from terrigenous (glacial) to biogenic (interglacial) facies at Milankovitch (PB) and similar variable (AP) frequency On the slope: shift in PB glacial sediment clast type: sandstone to granite a
The principal process is listed.
Processa Subsiding graben/ shelf with increasing ice to the shelf
Timing Cretaceous to Pliocene
Likely decrease in early Miocene to onshore sediment early Pliocene supply coincident with a shift in deposition from the rise to the slope
Erosion of shelf basins by grounded ice sheets
middle Miocene (B17–14 m.y.)
Glaciers fluctuating early Oligocene to onto and off the middle Miocene shelf during glacial and interglacial times
Glaciers fluctuating early Miocene to onto and off the early Pliocene shelf during glacial and interglacial times Likely change in ice Between 1.1 Ma and source area: 780 k.y. offshore to onshore?
Cenozoic Climate History from Seismic Reflection and Drilling Studies
169
are greater than from the shelf or slope. On the rise, there is a distinct up-section change in seismic character from well layered below to channellevy development above (all areas) that is widely inferred due to a large influx of sediment when onshore ice sheets initiated in late Eocene to early Oligocene time. Here, the pre-ice to glacial transition has not yet been sampled by drilling. Yet, higher in the stratigraphic section of the rise, drill cores show a clear long-term parabolic decrease in the sedimentation rates within sediment drift deposits from the early Miocene (PB: 10-fold decrease) and the late Miocene (AP: 6-fold decrease) to the present. The large decreases occurred when the PSEs were prograding over distances of several tens of kilometres and aggrading up to several hundred metres, although the detailed timing of the prograding and aggrading is unknown. Regardless, the distinct changes in geometries of seismic sequences beneath the outer paleoshelves in PB and AP are not seen as abrupt changes in sediment deposition rates on the rise. Drilling on the slope (WS) shows large incremental increases in sedimentation rates for this same general period (i.e. early Miocene to early Pliocene), indicating that sediment coming from the shelf may not have reached the rise. The decreases in sedimentation rates on the rise may also reflect a decrease in the amount of sediment being eroded from onshore and shelf areas. A notable long-term transition occurs in middle Miocene sediments (17–14 Ma) from the rise (PB). Up-section increases in IRD, diatom content, and recycled organic matter, along with changes in the types of clay and the first appearance of glauconite, point to greater ice nearby and initial erosion of shelf sedimentary basins. Evidence for strong erosion on the shelf is also seen in truncated foreset strata beneath the outer PB paleo-shelf. RS shelf drill cores are marked by a long hiatus, from mid- to late-Miocene. The hiatus and truncated shelf reflectors point to an increase in shelf erosion and overdeepening in the RS, similar to the erosion recorded on the rise at PB. Short-period fluctuations in shelf and rise drill cores provide the strongest evidence that erosion onshore and on the shelf by fluctuating grounded ice sheets was the mechanism for sediment supply and distribution by glacial processes, as inferred from seismic-reflection data. On the RS shelf, cyclic fluctuations in glacial diamict and interglacial glaciomarine lithofacies at Milankovitch frequencies are documented for lower Miocene nearshore facies at the front of the Transantarctic Mountains, and resulted from ice advancing onto and retreating off the shelf at this time (Naish et al., 2001). On the rise (PB and AP), alternating dark- and light-coloured lithofacies with varying amounts of terrigeneous (dark) and biogenic (light) components are described throughout the lower Miocene to lower Pliocene intervals from
170
A. K. Cooper et al.
visual observation of cores, downhole logging and physical properties measurements (PB); similar compositions variations are described from upper Miocene to Pliocene intervals in the AP. The facies are inferred to be of glacial and interglacial origin, respectively, and occur at Milankovitch frequencies in PB and similar order-of-magnitude frequencies in AP. Hence, drilling on the rise in East Antarctica and in West Antarctica has provided similar geologic evidence for fluctuating ice sheets on the shelves during the period of principal shelf progradation and aggradation – from the early Miocene to the early Pliocene. The geologic transition in the late Miocene to the early Pliocene is the initiation of broad and narrow shelf troughs and widespread banks, lobes and upper-slope fans along the Antarctic margin (all areas); this transition is difficult to see in drill cores, but has been imaged seismically. On the rise (AP, PB), sedimentation rates decrease uniformly during this period and lithologies (e.g. clays, IRD, etc.) do not show systematic long-term changes. However, seismic geometries beneath the adjacent continental shelves show abrupt changes to rapidly prograding sections (S2/S3 beneath AP; PP-12 beneath PB). Elsewhere, drilling information is insufficient to explain why large geomorphic changes on the shelf, probably due to changes in glacial regime, are not reflected in the rates or types of sediment delivered to the rise. 5.8.3. Sea-Level and Ice-Volume Changes Lithostratigraphic data from Antarctic margin drill cores show clear evidence on the shelf (PB, RS) for linked sea-level and ice-volume changes. This is best shown in the Oligocene through early Miocene record of cyclic glacial and interglacial lithologies near the coast in Cape Roberts cores (WRS) (Barrett, 2007). Lithostratigraphic data from the slope (PB) and rise (PB, AP) show additional direct evidence for cyclic ice-volume changes. Seismic-reflection data provide indirect evidence across the entire margin for the linked sea-level and ice-volume fluctuations that have been noted by many investigators and have been modelled in the RS (Bartek et al., 1991) and presented conceptually for all margins (ten Brink et al., 1995). The Antarctic drill cores are too limited, however, to establish the timing, magnitude and extent of individual ice sheet advances onto the continental shelf, other than for the LGM. A comparison of the Antarctic margin proximal stratigraphic record with the global record of sea-level variations (and linked ice-volume variations) from coastal onlap, backstripping, and isotopic records since the middle
Cenozoic Climate History from Seismic Reflection and Drilling Studies
171
Figure D-1: Graph showing sea-level and isotope curves and principal stratigraphic events for the Antarctic continental margin. Curves are from Miller et al. (2005). Events are from text and Tables D-1 and D-2. Curve A is for benthic foraminifera. Curve B is derived from stratigraphic backstripping (W9 M.a.) and isotopic measurements (0–9 M.a.); Curve C is from Haq et al. (1987). Curves B and C have different sea-level change scales. For some periods, Antarctic stratigraphic events correlate with isotopic shifts (e.g. PB: early Oligocene unconformity and first ice sheets at the coast and midMiocene lithology changes and ice buildup) and with the Haq Curve (e.g. mid-to-late Miocene shelf erosion and early Pliocene and younger slope fan development with long-term sea-level lowerings).
Eocene is shown in Fig. D-1. Large differences appear between the global sea-level curves (see Miller et al. (2005) for the explanation), yet the Antarctic stratigraphic features can potentially be linked to parts of all of the curves principally because of the current uncertainty in ages of Antarctic features,
172
A. K. Cooper et al.
especially for the Paleogene. The closest links of Antarctic features with the global curves are: Late Eocene and early Oligocene lowering of sea-level (first glaciers into PB and RS), followed by sea-level rise in the early Oligocene (flooding of PB). A long period from the Oligocene into the late Miocene of cyclic sea-level, seen in cyclic coastal deposits (WRS) and as glacial/interglacial rise-drift deposits (PB, AP) Abrupt sea-level lowering in the middle Miocene, seen in lithologies of rise-drift deposits (PB), followed by further sea-level lowering (with ice buildup) and enhanced erosional deepening and prograding of continental shelves by glaciers (as seen in seismic profiles from all areas). The systematic decrease in sea-level (and increase in ice) from the midMiocene to the present that corresponds with the decrease in sedimentation rates on the rise and increase in sedimentation rates on the slope. This is most pronounced since early Pliocene time, when an extensive system of shelf erosional troughs and upper-slope fans developed. Greater resolution of the link between ice-volume variations and sea-level changes requires further drilling on the Antarctic margin.
5.9. Summary The Antarctic continental margin holds a thick Cenozoic sedimentary section that is characterized by both long-period and short-period lithostratigraphic transitions, which are seen locally in drill cores and regionally in seismic-reflection data. Age resolution is inadequate to link individual stratigraphic events, but is sufficient to make general statements about glacial history. The transitions point to the last 40 m.y. being a period of increasing glaciation and sediment distribution by glacial processes via short-period fluctuations (e.g. Milankovitch frequencies) of grounded ice sheets across the continental shelf and accompanying sea-level changes. The proximal history is generally similar to that of distal proxy records from isotopic studies in adjacent ocean basins (e.g. Zachos et al., 2001) and from stratigraphic studies on low-latitude continental margins (e.g. Miller et al., 2005, 2000). Key inferences from the Antarctic margin for the Cenozoic, based on published data and inferences from extensive seismic-reflection and limited drilling records, are: Although tectonic histories differ around the Antarctic margin, similar geomorphic features (e.g. overdeepened and foredeepened seafloor; broad
Cenozoic Climate History from Seismic Reflection and Drilling Studies
173
erosional troughs, sediment fans and drift mounds) are seen everywhere on the margin, as a result of ubiquitous fluctuating glaciers eroding and distributing sediments. East and West Antarctic margin segments may have similar glacial histories, based on similar geomorphologies and known ages offshore for glacial strata. Current differences result partly from lack of sufficient drilling into likely Paleogene offshore sections beneath the margin. Seismic geometries and facies from all segments of the continental margin show evidence for up-section increases in the dynamic movement of sediment across the margin (e.g. shelf troughs, slope sediment fans, channel-levee systems) and along the margin (e.g. rise-drift deposits), reflecting increased glacial and ocean current activity from the Oligocene to the present. Stratal geometries of the continental shelf and slope were controlled principally by eustatic changes (with ice fluctuations) from the Paleogene to about the middle Miocene, and thereafter principally by fluctuating grounded glaciers (in tandem with sea-level changes) on the shelf, leading to the overdeepening and foredeepening of the shelf. Extensive prograding and aggrading of the continental shelf from the early Miocene to the latest Neogene is the principal result of sediment dispersal by ice sheets during glacial and interglacial periods at near-Milankovitch periodicities, as documented from drilling of drift deposits on the continental rise in East Antarctica (PB) and West Antarctica (AP), and from nearcoastal sequences (RS). The principal locus of sediment deposition on the margin has shifted from the outer rise (and beyond) during the Paleogene, to the inner rise and slope in the early Miocene to early Pliocene, and to the mid slope thereafter. The depocentre shift reflects the increases in glacial activity (and increases in ocean currents) and decrease in sediment being supplied due to erosion of onshore and shelf areas. Specific circum-Antarctic glacial events in the evolution of the margin include: first glaciers at the coast and initiation of channel-levee systems on the rise and the Crary Fan (early Oligocene); fluctuating glaciers, initial rapid progradation of the continental shelf, and initial growth of drift mounds and large levees on the rise (early Miocene); onshore ice buildup and initial overdeepening of the continental shelves (middle Miocene); dynamic ice movements and initial widespread development of cross-shelf troughs and upper-slope fans (early Pliocene); widespread deposition of biogenic interglacial sediment in deep inner-shelf troughs (Holocene). Additional advances in our understanding of Antarctica’s glacial history and the varied effects of ice sheets on the paleoceanographic and
174
A. K. Cooper et al.
lithostratigraphic processes of the Antarctic continental margin can only be achieved through additional offshore deep stratigraphic drilling studies, such as the current IODP, ANDRILL and SHALDRIL projects.
ACKNOWLEDGEMENTS We thank many people who collaborated, by sharing data and ideas, on geoscience research projects under the umbrella of the highly successful ANTOSTRAT project for nearly 15 years. This synthesis, which reflects our views, would not have been possible without the efforts of these many investigators, most of whom continue their collaborative Antarctic studies, now under the successor ACE project. We appreciate the helpful review comments by Larry Krissek and Peter Barrett on this chapter. We thank John Anderson for compiling and providing the supplementary references to provide further historic perspective to community efforts partly summarized in this chapter.
REFERENCES The reference section includes many supplementary circum-Antarctic references to research studies that are directly related to the topic of this chapter. They are selected from a compilation provided by John Anderson to further assist colleagues in locating, using and citing prior collaborative research data and interpretations. Abelmann, A., Gersonde, R., & Spiess, V. (1990). Pliocene-Pleistocene paleoceanography in the Weddell Sea – Siliceous microfossil evidence. In: U. Bleil, & J. Thiede (Eds). Geological History of the Polar Oceans: Arctic versus Antarctic. Kluwer Academic Publishers, Boston, MA, pp. 729–759. Abreu, V. S., & Anderson, J. B. (1998). Glacial eustasy during the Cenozoic: Sequence stratigraphic implications. Am. Assoc. Petrol. Geol. Bull., 82, 1385–1400. Abreu, V. S., & Haddad, G. A. (1999). Glacioeustatic fluctuations: The mechanism linking stable isotope events and sequence stratigraphy from the early Oligocene to middle Miocene time. In: P. C. DeGraciansky, J. Hardenbol, T. Jacquin, P. R. Vail, & M. B. Farley (Eds). Sequence Stratigraphy of the European Basins. Society of Economic Paleontologists and Mineralogists, Tulsa, OK, Special Publication No. 60, pp. 245–259. Abreu, V. S., Savini, R., & Barrocas, S. (1992). Paleoceanography and microbiostratigraphy of the Moby Dick Group, Melville Peninsula, Northern King
Cenozoic Climate History from Seismic Reflection and Drilling Studies
175
George Island, Western Antarctica. In: Fourth International Conference on Paleo Oceanography, Siena, Italy, p. 48. Accaino, F., Boehm, G., & Brancolini, G. (2005). Analysis of Antarctic glaciations by seismic reflection and refraction tomography. Mar. Geol., 216, 145–154. Acosta, J., Herranz, P., Sanz, J. L., & Uchupi, E. (1992). Antarctic continental margin: Geological image of the Bransfield Trough, an incipient ocean basin. In: C. W. Poag, & P. C. Graciansky (Eds). Geological Evolution of the Atlantic Continental Rises. Van Nostrand Reinhold, New York, pp. 49–61. Adamson, D., & Pickard, J. (1983). Late Quaternary ice movement across the Vestfold Hills, East Antarctica. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 465–469. Adamson, D. A., & Pickard, J. (1986). Cainozoic history of the Vestfold Hills. In: J. Pickard (Ed.). Antarctic Oasis – Terrestrial Environments and History of the Vestfold Hills. Academic Press, New York, pp. 63–97. Alley, R. B. (1990). West Antarctic collapse – how likely? Episodes, 13, 231–238. Alley, R. B., & Bindschadler, R. A. (Eds). (2001). The West Antarctic Ice Sheet: Behavior and Environment. American Geophysical Union, Antarctic Research Series, Washington, DC, Vol. 77. Alley, R. B., Blankenship, D. D., Bentley, C. R., & Rooney, S. T. (1987a). Till beneath ice stream B. 3. Till deformation: Evidence and implications. J. Geophys. Res., 92, 8921–8929. Alley, R. B., Blankenship, D. D., Bentley, C. R., & Rooney, S. T. (1987b). Till beneath ice stream B. 4. A coupled ice-till flow model. J. Geophys. Res., 92, 8931–8940. Alley, R. B., Blankenship, D. D., Bentley, C. R., & Rooney, S. T. (1989). Sedimentation beneath ice shelves – the view from ice stream B. Mar. Geol., 85, 101–120. Alonso, B., Anderson, J. B., Diaz, J. I., & Bartek, L. R. (1992). Pliocene–Pleistocene Seismic Stratigraphy of the Ross Sea: Evidence for Multiple Ice Sheet Grounding Episodes, Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 57, pp. 93–103. Amblas, D., Urgeles, R., Canals, M., Calafat, A. M., Rebesco, M., Camerlenghi, A., Estrada, F., De Batist, M., & Hughes-Clarke, J. E. (2006). Relationship between continental rise development and palaeo-ice sheet dynamics, Northern Antarctic Peninsula Pacific margin. Quat. Sci. Rev., 25, 933–944. Anderson, J. B. (1971). Marine origin of sands in the Weddell Sea. Antarct. J. U.S., 6, 168–169. Anderson, J. B. (1972a). The Marine Geology of the Weddell Sea. Ph.D. Dissertation, Florida State University, Tallahassee, 222 pp. Anderson, J. B. (1972b). Nearshore glacial-marine deposition from modern sediments of the Weddell Sea. Nat. Phys. Sci., 240, 189–192. Anderson, J. B. (1975a). Ecology and distribution of foraminifera in the Weddell Sea of Antarctica. Micropaleontology, 21, 69–96. Anderson, J. B. (1975b). Factors controlling CaCO3 dissolution in the Weddell Sea from foraminiferal distribution patterns. Mar. Geol., 19, 315–332.
176
A. K. Cooper et al.
Anderson, J. B. (1983). Ancient glacial-marine deposits: Their spatial and temporal distribution. In: B. F. Molnia (Ed.). Glacial-Marine Sedimentation. Plenum Press, New York, pp. 3–92 Anderson, J. B. (1985). Antarctic Glacial Marine Sedimentation: A Core Workshop. Geological Society of America, Boulder, CO, 66 pp Anderson, J. B. (1986). Critical evaluation of some criteria used to infer Antarctica’s glacial and climatic history from deep-sea sediments. S. Afr. J. Sci., 82, 503–505. Anderson, J. B. (1990). Sediments. In: G. P. Glasby (Ed.). Antarctic Sector of the Pacific. Elsevier Oceanography Series. New York, Vol. 51, pp. 187–206. Anderson, J. B. (1991). The Antarctic continental shelf: Results from marine geological and geophysical investigations. In: R. J. Tingey (Ed.). The Geology of Antarctica. Clarendon Press, Oxford, pp. 285–334. Anderson, J. B. (1999). Antarctic Marine Geology. Cambridge University Press, Cambridge, 289 pp. Anderson, J. B., & Kurtz, D. (1985). The use of silt grain size parameters as a paleovelocity gauge: A critical review and case study. Geo Mar. Lett., 5, 55–59. Anderson, J. B., & Molnia, B. F. (1989). Glacial-Marine Sedimentation. Short Course in Geology. American Geophysical Union, Washington, DC, Vol. 9, 127 pp. Anderson, J. B., & Smith, M. J. (1989). Formation of modern sand-rich facies by marine currents on the Antarctic continental shelf. In: R. Morton, & D. Nummedal (Eds). Shelf Sedimentation, Shelf Sequences, and Related Hydrocarbon Accumulation: Gulf Coast Section of the Society of Economic Paleontologists and Mineralogists 7th Annual Research Conference Proceedings, Vol. 85, pp. 41–52. Anderson, J. B., & Ashley, G. M. (Eds). (1991). Glacial Marine Sedimentation; Paleoclimatic Significance. The Geological Society of America, Boulder, Co, Special Paper 261. Anderson, J. B., & Bartek, L. R. (1991). High resolution seismic record of Antarctic glacial history. Korean J. Polar Res., 2, 71–78. Anderson, J. B., & Thomas, M. A. (1991). Marine ice-sheet decoupling as a mechanism for rapid, episodic sea-level change: The record of such events and their influence on sedimentation. Sediment. Geol., 70, 87–104. Anderson, J. B., & Bartek, L. R. (1992). Cenozoic Glacial History of the Ross Sea Revealed by Intermediate Resolution Seismic Reflection Data Combined with Drill Site Information. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 56, pp. 231–263. Anderson, J. B., & Andrews, J. T. (1999). Radiocarbon constraints on ice sheet advance and retreat in the Weddell Sea, Antarctica. Geology, 27, 179–182. Anderson, J. B., Balshaw, K., Domack, E., Kurtz, D., Milam, R., & Wright, R. (1979a). Geologic survey of the East Antarctic continental margin aboard USCGC Glacier. Antarct. J. U.S., 14, 142–144. Anderson, J. B., Kurtz, D. D., & Weaver, F. M. (1979b). Sedimentation on the Antarctic continental slope. In: L. J. Doyle, & O. Pilkey (Eds). Geology of Continental Slopes. Society of Economic Paleontologists and Mineralogists, Tulsa, OK, Special Publication No. 27, pp. 265–283.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
177
Anderson, J. B., Domack, E. W., & Kurtz, D. D. (1980a). Observations of sedimentladen icebergs in Antarctic waters: Implications to glacial erosion and transport. J. Glaciol., 25, 387–396. Anderson, J. B., Kurtz, D. D., Domack, E. W., & Balshaw, K. M. (1980b). Glacial and glacial marine sediments of the Antarctic continental shelves. J. Geol., 88, 399–414. Anderson, J. B., Brake, C., Domack, E. W., Myers, N., & Singer, J. (1983a). Sedimentary dynamics of the Antarctic continental shelf. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 387–389. Anderson, J. B., Brake, C., Domack, E., Myers, N., & Wright, R. (1983b). Development of a polar glacial-marine sedimentation model from Antarctic Quaternary deposits and glaciological information. In: B. F. Molnia (Ed.). Glacial-Marine Sedimentation. Plenum Press, New York, pp. 233–264. Anderson, J. B., Brake, C. F., & Myers, N. C. (1984). Sedimentation on the Ross Sea continental shelf, Antarctica. Mar. Geol., 57, 295–333. Anderson, J. B., Wright, R., & Andrews, B. (1986). Weddell Fan and associated abyssal plain, Antarctica: Morphology, sediment processes, and factors influencing sediment supply. Geo Mar. Lett., 6, 121–129. Anderson, J. B., Pope, P. G., & Thomas, M. A. (1990). Evolution and hydrocarbon potential of the northern Antarctic Peninsula continental shelf. In: B. St John (Ed.). Antarctica as an Exploration Frontier – Hydrocarbon Potential, Geology and Hazards. American Association of Petroleum Geologists, Tulsa, OK, Studies in Geology, Vol. 31, pp. 1–12. Anderson, J. B., Andrews, B. A., Bartek, L., & Truswell, E. M. (1991a). Petrology and palynology of Weddell Sea glacial sediments: Implications for subglacial geology. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 231–235. Anderson, J. B., Bartek, L. R., & Thomas, M. A. (1991b). Seismic and sedimentologic record of glacial events on the Antarctic Peninsula shelf. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 687–691. Anderson, J. B., Reid, D. E., & Taviani, M. (1991). Compositional and geochemical characteristics of recent biogenic carbonates from the Ross Sea (Antarctica). In: C. A. Ricci (Ed.). Earth Science Investigations in Antarctica. Memorie della Societa Geologica Italiana, Vol. 46, pp. 561–562. Anderson, J. B., Kennedy, D. S., Smith, M. J., & Domack, E. W. (1991c). Sedimentary facies associated with Antarctica’s floating ice masses. In: J. B. Anderson, & G. M. Ashley (Eds). Glacial Marine Sedimentation; Paleoclimatic Significance. Geological Society of America, Boulder, CO, Special Paper 261, pp. 1–25. Anderson, J. B., Shipp, S. S., Bartek, L. R., & Reid, E. D. (1992a). Evidence for a Grounded Ice Sheet on the Ross Sea Continental Shelf During the Late Pleistocene
178
A. K. Cooper et al.
and Preliminary Paleodrainage Recostruction. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 57, pp. 39–62. Anderson, J. B., Shipp, S. S., & Siringan, F. P. (1992b). Preliminary seismic stratigraphy of the northwestern Weddell Sea continental shelf. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing, Tokyo, pp. 603–612. Anderson, J. B., Wellner, J. S., Bohaty, S., Manley, P. L., & Wise, S. W. (2006). Antarctic Shallow Drilling Project Provides Key Core Samples. Eos Trans. Am. Geophys. Union, 87, 402–408. Anderson, J. B., Wellner, J., Wise, S., Bohaty, S., Manley, P., Smith, T., Weaver, F., & Kulhanek, D. (2007). Seismic and chronostratigraphic results from SHALDRIL II, northwestern Weddell Sea. US Geological Survey and The National Academies, USGS OF-2007-1047, doi:10.3133/of2007-1047.srp094. Andrews, B. A. (1984). A Petrologic Study of Weddell Sea Sediments: Implications for Provenance and Glacial History. M.A. Thesis, Rice University, Houston, TX, 203 pp. Angino, E. E., & Andrews, R. S. (1968). Trace element chemistry, heavy minerals, and sediment statistics of Weddell Sea sediments. J. Sediment. Petrol., 38, 634–642. Arne, D. C. (1994). Phanerozoic exhumation history of northern Prince Charles Mountains (East Antarctica). Antarct. Sci., 6, 69–84. Ashley, G. M., Shaw, J., & Smith, N. D. (Eds). (1985). Glacial Sedimentary Environments, Short Course No. 16. Society of Paleontologists and Mineralogists, Tulsa, OK, 246 pp. Askin, R. A. (1988a). Campanian to Paleocene palynological succession of Seymour and adjacent islands, northeastern Antarctic Peninsula. In: R. M. Feldman, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America, Boulder, Co, Memoir, Vol. 169, pp. 131–153. Askin, R. A. (1988b). The palynological record across the Cretaceous/Tertiary transition on Seymour Island, Antarctica. In: R. M. Feldman, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America, Boulder, CO, Memoir, Vol. 169, pp. 155–162. Askin, R. A. (1992). Late Cretaceous–early Tertiary Antarctic outcrop evidence for past vegetation and climates. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change I. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 56, pp. 61–73. Asking, R. A., & Raine, J. L. (2000). Oligocene and early Miocene terrestrial palinology of the Cape Roberts drillhole CRP-2/21, Victoria Land Basin, Antarctica. Terra Antart., 7(4), 493–501. Baegi, M. (1985). Turbidite Deposition on the Bellingshausen Abyssal Plain: Sedimentologic Implications. M.A. Thesis, Rice University, Houston, TX, 109 pp.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
179
Balshaw, K. M. (1981). Antarctic glacial chronology reflected in the Oligocene through Pliocene sedimentary section in the Ross Sea. M.A. Thesis, Rice University, Houston, TX, 140 pp. Bandy, O. L., Casey, R. E., & Wright, R. C. (1971). Late Neogene planktonic zonation, magnetic reversals, and radiometric dates, Antarctic to the Tropics. In: J. L. Reid (Ed.). Antarctic Oceanology I. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 15, pp. 1–26. Banfield, L. A. (1994). Seismic Facies investigation of the Late Cenozoic Glacial History of Bransfield Basin, Antarctica. M.A. Thesis, Rice University, Houston, TX, 184 pp. Banfield, L. A., & Anderson, J. B. (1995). Seismic facies investigation of the late Quaternary glacial history of Bransfield Basin, Antarctica. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 123–140. Bardin, V. I. (1982). Composition of East Antarctic moraines and some problems of Cenozoic history. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 1069–1076. Barker, P. F. (1976). The tectonic framework of Cenozoic volcanism in the Scotia Sea region: A review. In: O. Gonzalez-Ferran (Ed.). Symposium on Andean and Antarctic Volcanology Problems. Proceedings of the IAVCEI, Santiago, Chile, pp. 330–346. Barker, P. F. (1982). The Cenozoic subduction history of the Pacific margin of the Antarctic Peninsula: Ridge crest-trench interactions. J. Geol. Soc. London, 139, 787–801. Barker, P. F., & Dalziel, I. W. D. (1976). The evolution of the southwest Atlantic Ocean basin: Leg 36 data. In: P. F. Barker, & I. W. D. Dalziel (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, p. 993. Barker, P. F., & Burrell, J. (1977). The opening of Drake Passage. Mar. Geol., 25, 15–34. Barker, P. F., & Dalziel, I. W. D. (1983). Progress in geodynamics in the Scotia Arc region. In: R. Cabre (Ed.). Geodynamics in the Eastern Pacific Region, Caribbean and Scotia Arcs. American Geophysical Union, Washington, DC, Geodynamics Series, Vol. 9, pp. 137–170. Barker, P. F., & Lonsdale, M. J. (1991). A multichannel seismic profile across the Weddell Sea margin of the Antarctic Peninsula: Regional tectonic implications. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, UK, pp. 237–241. Barker, D. H. N., & Austin, J. A. (1994). Crustal diapirism in Bransfield Strait, West Antarctica: Evidence for distributed extension in marginal basin formation. Geology, 22, 657–660.
180
A. K. Cooper et al.
Barker, D. H. N., & Austin, J. A. (1998). Rift propagation, detachment faulting, and associated magmatism in Bransfield Strait, Antarctic Peninsula. J. Geophys. Res., 103, 24017–24043. Barker, P. F., & Camerlenghi, A. (2002). Glacial history of the Antarctic Peninsula from Pacific margin sediments. In: P. F. Barker, A. Camerlenghi, G. D. Acton, & A. T. S. Ramsay (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 178 [Online]. Available from World Wide Web: http://wwwodp.tamu.edu/publications/178_SR/synth/synth.htm Barker, P. F., Kennett, J. P., et al. (1988a). Proceedings of the Ocean Drilling Program, Initial Reports. Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 113, 785 pp. Barker, P. F., Kennett, J. P., et al., (1988b). Weddell Sea palaeoceanography: Preliminary results of ODP Leg 113. Palaeogeogr. Palaeoclimatol. Palaeoecol., 67, 75–102. Barker, P. F., Barrett, P., Camerlenghi, A., Cooper, A., Dacey, F., Domack, E., Escutia, C., Kristoffersen, Y., & O’Brien, P. E. (1998). Ice sheet history from continental margin sediments: The ANTOSTRAT approach. Terra Antart., 5, 737–760. Barker, P. F., Camerlenghi, A., Acton, G. D., et al. (1999). Proceedings of the Ocean Drilling Program, Initial Reports, Vol. 178 [CD-ROM]. Available from Ocean Drilling Program, Texas A&M University, College Station, TX. Barnes, P. W. (1987). Morphologic studies of the Wilkes Land continental shelf, Antarctica-glacial and iceberg effects. In: S. L. Eittreim, & M. A. Hampton (Eds). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, CPCEMR Earth Science Series, Vol. 5A, pp. 175–194. Barnes, P. W., & Lien, R. (1988). Icebergs rework shelf sediments to 500 m off Antarctica. Geology, 16, 1130–1133. Baroni, C., & Orombelli, G. (1991). Holocene raised beaches at Terra Nova Bay, Victoria Land, Antarctica. Quat. Res., 36, 157–177. Barrera, E., & Huber, B. T. (1993). Eocene to Oligocene oceanography and temperatures in the Antarctic Indian Ocean. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment; A Perspective on Global Change II. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 60, pp. 49–65. Barrett, P. J. (1975). Textural characteristics of Cenozoic preglacial and glacial sediments at Site 270, Ross Sea, Antarctica. In: D. E. Hayes, & L. A. Frakes (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, pp. 757–767. Barrett, P. J. (1981). History of the Ross Sea region during the deposition of the Beacon Supergroup 400–180 million years ago. J. R. Soc. N. Z., 11, 447–459. Barrett, P. J. (Ed.) (1986). Antarctic Cenozoic History from the MSSTS-1 Drillhole, McMurdo Sound. DSIR Bulletin 237. Science Information Publishing Centre, Wellington, New Zealand.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
181
Barrett, P. J. (Ed.) (1989). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. DSIR Bullettin, N.Z., 245. Barrett, P. J. (Ed.) (1991). Antarctica and global climate change: A geological perspective. In: C. Harris, & B. Stonehouse (Eds). Antarctica and Global Climate Change. Belhaven Press, London, pp. 35–50. Barrett, P. J. (1997). Antarctic paleoenvironment through Cenozoic times: A review. Terra Antart., 3, 103–119. Barrett, P. J. (2007). Cenozoic climate and sea level history from glacimarine strata off the Victoria Land coast, Cape Roberts Project, Antarctica. In: M. J. Hambrey, P. Christoffersen, N. F. Glasser, & B. Hubbart (Eds). Glacial Processes and Products. International Association of Sedimentologists, Oxford, Special Publication, Vol. 39, pp. 259–287. Barrett, P. J., & McKelvey, B. C. (1986). Stratigraphy. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the MSSTS-1 Drillhole, McMurdo Sound. Science Information Publishing Centre, Wellington, New Zealand, DSIR Bulletin, Vol. 237, pp. 9–52. Barrett, P. J., Pyne, A. R., & Ward, B. L. (1983). Modern sedimentation in McMurdo Sound, Antarctica. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 550–554. Barrett, P. J., Elston, D. P., Harwood, D. M., McKelvey, B. C., & Webb, P. N. (1987). Mid-Cenozoic record of glaciation and sea-level change on the margin of the Victoria Land basin, Antarctica. Geology, 15, 634–637. Barrett, P. J., Hambrey, M. J., Harwood, D. M., Pyne, A. R., & Webb, P. N. (1989). Synthesis. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. DSIR Bulletin 245. DSIR Publishing, Wellington, pp. 241–251. Barrett, P. J., Adams, C. J., McIntosh, W. C., Swisher, C. C. III, & Wilson, G. S. (1992). Geochronological evidence supporting Antarctic deglaciation three million years ago. Nature, 359, 816–818. Barrett, P. J., Henrys, S. A., Bartek, L. R., Brancolini, G., Busetti, M., Davey, F. J., Hannah, M. J., & Pyne, A. R. (1995). Geology of the margin of the Victoria Land Basin off Cape Roberts, southwest Ross Sea. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 183–208. Barrett, P. J., Davey, F. J., De Santis, L., Alonso, B., Bart, P. J., Bartek, L. R., Brancolini, G., & Wise, S. W. Jr. (1999). Past Behavior of the West Antarctic Ice Sheet. IODP proposal 489/revision 3. Barrie, C. Q., & Piper, D. J. W. (1982). Late Quaternary Marine Geology of Makkovik Bay, Labrador. Geological Survey of Canada Paper 81–17, 37 pp. Barron, J., Larsen, B., & Shipboard Scientific Party (Eds) (1989). Proceedings of the Ocean Drilling Program, Initial Reports. College Station, TX, Vol. 119, 942 pp.
182
A. K. Cooper et al.
Barron, J. A. (1996). Diatom constraints on the position of the Antarctic Polar Front in the middle part of the Pliocene. Mar. Micropaleontol., 27, 195–213. Barron, J. A., Larson, B., & Baldauf, J. G. (1991). Evidence for for late Eocene to early Oligocene Antarctic glaciation and observations on late Neogene glacial history of Antarctica: Results from Leg 119. In: J. A. Barron, B. Larsen, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 119, pp. 869–894. Bart, P. J. (1993). Glacial and Tectonic Development of the Antarctic Peninsula. M.A. Thesis, Rice University, Houston, TX, 152 pp. Bart, P. J. (1998). Seismic-Stratigraphic Analysis of Shelf-Margin Delta/Slope Fans and Basin Floor Fans on High-Latitude and Middle-Latitude Margins (Ross Sea and Alabama/West Florida Shelf): Paleoclimatic and Eustatic Implications. Ph.D. Dissertation, Rice University, Houston, TX, 236 pp. Bart, P. J. (2003). Were West Antarctic Ice Sheet grounding events in the Ross a consequence of East Antarctic Ice Sheet expansion during the middle Miocene? Earth Planet. Sci. Lett., 216, 93–107. Bart, P. J., & Anderson, J. B. (1995). Seismic record of glacial events affecting the Pacific margin of the northwestern Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 74–95. Bart, P. J., & Anderson, J. B. (1996). Seismic expression of depositional sequences associated with expansion and contraction of ice sheets on the northwestern Antarctic Peninsula continental shelf. In: M. De Batist, & P. Jacobs (Eds). Geology of Siliciclastic Shelf Seas. Geological Society, London, Special Publication, Vol. 117, pp. 171–186. Bart, P. J., & Anderson, J. B. (2000). Relative temporal stability of the Antarctic Ice Sheets during the late Neogene based on the minimum frequency of outer shelf grounding events. Earth Planet. Sci. Lett., 182(3/4), 259–272. Bart, P. J., De Batist, M., & Miller, H. (1994). Neogene collapse of glaciallydeposited, shelf-edge deltas in the Weddell Sea: Relationships between deposition during glacial periods and sub-marine fan development. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin. Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea-Level Change. Terra Antart., Vol. 1, pp. 317–318. Bart, P. J., DeBatist, M., & Jokat, W. (1999). Interglacial collapse of Crary Trough Mouth Fan, Weddel Sea, Antarctica: Implications for Antarctic glacial history. J. Sed. Res., 69, 1276–1289. Bart, P. J., Anderson, J. B., Trincardi, F., & Shipp, S. S. (2000). Seismic data from the Northern basin, Ross Sea, record extreme expansion of the East Antarctic Ice Sheet during late Neogene. Mar. Geol., 166, 31–50. Bart, P. J., Egan, D., & Warny, S. A. (2005). Direct constraints on Antarctic Peninsula Ice Sheet grounding events between 5.12 and 7.94 Ma. J. Geophys. Res., 110, F04008, doi:10.1029/2004JF000254.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
183
Bart, P. J., Hillenbrand, C. D., Ehrmann, W., Iwai, M., Winter, D., & Warny, S. A. (2007). Are Antarctic Peninsula Ice Sheet grounding events manifest in sedimentary cycles on the adjacent continental rise? Mar. Geol., 236, 1–13. Bartek, L. R., & Anderson, J. B. (1991). Facies distribution resulting from sedimentation under polar interglacial climatic conditions within a high-latitude marginal basin, McMurdo Sound, Antarctica. In: J. B. Anderson, & G. M. Ashley (Eds). Glacial Marine Sedimentation; Paleoclimatic Significance. Geological Society of America, Boulder, CO, Special Paper 261, pp. 27–49. Bartek, L. R., Vail, P. R., Andersonl, J. B., Emmet, P. A., & Wu, S. (1991). Effect of Cenozoic ice sheet fluctuations in Antarctica on the stratigraphic signature of the Neogene. J. Geophys. Res., 96, 6753–6778. Bartek, L. R., Sloan, L. C., Anderson, J. B., & Ross, M. I. (1992). Evidence from the Antarctic continental margin of late Paleogene ice sheets: a manifestation of plate reorganization and synchronous changes in atmospheric circulation over the emerging Southern Ocean. In: D. R. Prothero, & W. A. Berggren (Eds). EoceneOligocene Climatic and Biotic Evolution. Princeton University Press, Princeton, NJ, pp. 131–159. Bartek, L. R., Anderson, J. B., & Henrys, S. A. (1994). Contributions of high resolution single-channel data to understanding sequence architecture of a glaciated continental margin, Ross Sea, Antarctica. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin. Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea-Level Change. Terra Antart., 1 (2), 329–332. Bartek, L. R., Henrys, S. A., Anderson, J. B., & Barrett, P. J. (1996). Seismic stratigraphy of McMurdo Sound, Antarctica: Implication for glacially influenced early Cenozoic eustatic change. Mar. Geol., 130, 79–98. Barton, C. M. (1965). The Geology of the South Shetland Islands. III. The Stratigraphy of King George Island. British Antarctic Survey Scientific Reports, No. 44, 37 pp. Behrendt, J. C., LeMasurier, W. E., Cooper, A. K., Tessensohn, F., Trehu, A., & Damaske, D. (1991). Geophysical studies of the West Antarctic Rift System. Tectonics, 10(6), 1257–1273. Belknap, D. F., & Shipp, R. C. (1991). Seismic stratigraphy of glacial marine units, Maine inner shelf. In: J. B. Anderson, & G. M. Ashley (Eds). Glacial Marine Sedimentation: Paleoclimatic Significance. Geological Society of America, Boulder, CO, Special Paper, Vol. 261, pp. 137–158. Bentley, M. J., & Anderson, J. (1998). Glacial and marine geological evidence for the ice sheet configuration in the Weddell Sea – Antarctic Peninsula region during the last glacial maximum. Antarct. Sci., 10, 309–325. Bialas, J., & GRAPE Team (1990). Seismic investigations along the Antarctic Peninsula. In: A. K. Cooper, & P. N. Webbs (Eds). International Workshop on Antarctic Offshore Seismic Stratigraphy (ANTOSTRAT). Overview and Extended Abstracts, pp. 72–78.
184
A. K. Cooper et al.
Bibby, J. S. (1966). The Stratigraphy of Part of North-East Graham Land and the James Ross Island Group. British Antarctic Survey Scientific Reports, No. 53, 37 pp. Birkenmajer, K. (1980). Report on geological investigations of King George Island, South Shetland Islands (West Antarctica) in 1978/79. Stud. Geol. Pol., 64, 89–105. Birkenmajer, K. (1981a). Geological relations at Lions Rump, King George Island (South Shetland Islands, Antarctica). Stud. Geol. Pol., 72, 75–87. Birkenmajer, K. (1981b). Lithostratigraphy of the Point Hennequin Group (Miocene volcanics and sediments) at King George Island (South Shetland Islands, Antarctica). Stud. Geol. Pol., 72, 59–73. Birkenmajer, K. (1984). Pre-Quaternary glaciations of West Antarctica: evidence from the South Shetland Islands. Pol. Polar Res., 5, 319–329. Birkenmajer, K. (1985). Onset of Tertiary continental glaciation in the Antarctic Peninsula sector (West Antarctica). Acta Geol. Pol., 35, 1–31. Birkenmajer, K. (1987). Oligocene-Miocene glacio-marine sequences of King George Island (South Shetland Islands), Antarctica. Palaeontol. Pol., 49, 9–40. Birkenmajer, K. (1988). Tertiary glacial and interglacial deposits, South Shetland Islands, Antarctica: Geochronology versus biostratigraphy (a progress report). Bull. Pol. Acad. Sci. Earth Sci., 36, 133–145. Birkenmajer, K. (1991). Tertiary glaciation in the South Shetland Islands, West Antarctica: Evaluation of data. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, UK, pp. 629–632. Birkenmajer, K., & Gazdzicki, A. (1986). Oligocene age of the Pecten conglomerate on King George Island, West Antarctica. Bulletin of the Polish Academy of Sciences. Earth Sciences, 34, 219–226. Birkenmajer, K., & Luczkowska, E. (1987). Early Miocene foraminiferal zonation, southeast Pacific Basin, Antarctic Peninsula sector. Bulletin of the Polish Academy of Sciences, Earth Sciences, 35, 1–10. Birkenmajer, K., Soliani, E. Jr., & Kawashita, K. (1988). Early Miocene K-Ar age of volcanic basement of the Melville Glaciation deposits, King George Island, West Antarctica. Bulletin of the Polish Academy of Sciences. Earth Sciences, 36, 25–34. Birkenmajer, K., Soliani, E. Jr., & Kawashita, K. (1989). Geochronology of Tertiary glaciations on King George Island, West Antarctica. Bulletin of the Polish Academy of Sciences. Earth Sciences, 37, 27–48. Blank, R. G., & Margolis, S. V. (1975). Pliocene climatic and glacial history of Antarctica as revealed by Southeast Indian Ocean deep-sea cores. Geol. Soc. Am. Bull., 86, 1058–1066. Blankenship, D. D., Bentley, C. R., Rooney, S. T., & Alley, R. B. (1987). Till beneath ice stream B, 1. Properties derived from seismic travel times. J. Geophys. Res., 92, 8903–8911. Bochu, Y., Guangyu, W., Bangyan, C., & Shengyuan, C. (1995). The characteristics of geophysical field and tectonic evolution in the Bransfield Strait. Antarct. Res., 6, 12–23.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
185
Bohaty, S. M., & Harwood, D. M. (1995). Pliocene Silicoflagellate Paleotemperature Variation from the Kerguelen Plateau, Abstracts of Programs. Geological Society of America, Vol. 27, No. 3, p. 41. Bornhold, B. D. (1983). Ice-rafted debris in sediments from Leg 71, southwest Atlantic Ocean. In: W. J. Ludwig, V. A. Krasheninnikov, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 71, pp. 307–316. Boulton, G. S. (1978). Boulder shapes and grain-size distributions of debris as indicators of transport paths through a glacier and till genesis. Sedimentology, 25, 773–799. Boulton, G. S. (1990). Sedimentary and sea level changes during glacial cycles and their control on glacimarine facies architecture. In: J. A. Dowdeswell, & J. D. Scourse (Eds). Glacimarine Environments: Processes and Sediments. Geological Society, London, England, Special Publication No. 53, pp. 15–52. Boulton, G. S., & Paul, M. A. (1976). The influence of genetic processes on some geotechnical properties of glacial tills. Q. J. Eng. Geol., 9, 159–194. Boulton, G. S., & Jones, A. S. (1979). Stability of temperate ice caps and ice sheets resting on beds of deformable sediments. J. Glaciol., 24, 29–43. Brachfeld, S., Domack, E., Kissel, C., Laj, C., Leventer, A., Ishman, S., Gilbert, R., Camerlenghi, A., & Eglinton, L. B. (2003). Holocene history of the Larsen-A Ice Shelf constrained by geomagnetic paleointensity dating. Geology, 31, 749–752. Brady, H. T. (1982). Late Cenozoic history of Taylor and Wright Valleys and McMurdo Sound inferred from diatoms in Dry Valley Drilling Project cores. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 1123–1131. Brake, C. F., & Anderson, J. B. (1983). The bathymetry of the North Victoria Land continental margin. Mar. Geol., 6, 139–147. Brambati, A., Ceccaroni, L., D’Onofrio, S., Fanzutti, G. P., Finocchiaro, F., Frignani, M., Langone, L., Melis, R., & Ravaioli, M. (1994). Palaeoenvironmental inferences from the Cora ANTA91-30 Drigalsky Basin, Ross Sea, Antarctica. Terra Antart., 1, 335–337. Brancolini, G., Gantar, C., Lodolo, E., & Marchetti, A. (1991). Acquisition, processing and preliminary results of the 1988/89 Antarctic geophysical survey. Mem. Soc. Geol. It., 46, 543–560. Brancolini, G., Cooper, A. K., & Coren, F. (1995). Seismic Facies and Glacial History in the Western Ross Sea (Antarctica), Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 68, pp. 209–234. Brancolini, G., & Harris, P. T. (2000). Post-cruise report AGSO Survey 217: Joint Italian/Australian marine geoscience expedition aboard the R.V. Tangaroa to the George Vth Land region during February–March 2000. Australian Geological Survey Organisation Record 2000/38. Bruno, L. A., Baur, H., Graf, T., Schluchter, C., Signer, P., & Wieler, R. (1997). Dating of Sirius Group tillites in the Antarctic Dry Vallies with cosmogenic 3He and 21Ne. Earth Planet. Sci. Lett., 147, 37–54.
186
A. K. Cooper et al.
Burckle, L. H. (1984). Diatom distribution and paleoceanographic reconstruction in the Southern Ocean; present and last glacial maximum. Mar. Micropaleontol., 9, 241–261. Burckle, L. H., & Abrams, N. (1986). Diatom species diachrony in late Neogene sediments of the Southern Ocean. S. Afr. J. Sci., 82, 505–506. Burckle, L. H., & Cirilli, J. (1987). Origin of diatom ooze belt in the Southern Ocean; implications for late Quaternary paleoceanography. Micropaleontology, 33, 82–86. Burckle, L. H., Robinson, D., & Cooke, D. (1982). Reappraisal of sea-ice distribution in Atlantic and Pacific sectors of the Southern Ocean at 18,000 yr BP. Nature, 299, 435–437. Busetti, M., & Cooper, A. K. (1994). Possible ages and origins of unconformity U6. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea Level Change. Terra Antart., 1 (2), 341–344. Busetti, M., Zanolla, C., & Marchetti, A. (2001). Geological structure of the South Orkney microcontinent. Terra Antart., 8, 71–78. Busetti, M., Marchetti, A., Zanolla, C., & De Cillia, C. (2002). Tectonic history of the South Orkney microcontinent from seismic structure and stratigraphy. In: J. A. Gamble, D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium. Royal Society of New Zealand Bulletin. Royal Society of New Zealand, Wellington, Vol. 35, pp. 507–513. Canals, M., De Batist, M., Baraza, J., Acosta, J., & the GEBRA Team. (1994). New reflection seismic data from Bransfield Strait, Preliminary Results of the GEBRA-93 Survey. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea Level Change. Terra Antart., 1 (2), 291–292. Canals, M., Urgeles, R., & Calafat, A. M. (2000). Deep sea-floor evidence of past ice streams off the Antarctic Peninsula. Geology, 28, 31–34. Canals, M., Casamor, J. L., Urgeles, R., Calafat, A. M., Domack, E. W., Baraza, J., Farran, M., & De Batist, M. (2002). Seafloor evidence of a subglacial sedimentary system off the northern Antarctic Peninsula. Geology, 30, 603–606. Cande, S. C., & Mutter, J. C. (1982). A revised identification of the oldest sea-floor spreading anomalies between Australia and Antarctica. Earth Planet. Sci. Lett., 58, 151–160. Cande, S. C., & Kent, D. V. (1995). Revised calibration of the geomagnetic polarity timescale for the Late Cretaceous and Cenozoic. J. Geophys. Res., 100, 6093–6095. Cape Roberts Scientific Team (1998). Initial report and scientific results on CRP-1, Cape Roberts Project. Terra Antart., 5 (1), 1–187, 5 (3), 255–711. Cape Roberts Scientific Team (1999). Initial report and scientific results on CRP-2, Cape Roberts Project. Terra Antart., 6(1/2), 1–173.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
187
Cape Roberts Scientific Team (2000). Initial report and scientific results on CRP-2 and CRP-3, Cape Roberts Project. Terra Antart., 7 (1/2, 3, 4), 1–654. Cape Roberts Scientific Team (2001). Initial report and scientific results on CRP-3, Cape Roberts Project. Terra Antarct., 8 (3, 4), 121–621. Carey, S. W., & Ahmad, N. (1961). Glacial marine sedimentation. Geol. Arctict., 2, 865–894. Carmack, E. C., & Foster, T. D. (1975). On the flow of water out of the Weddell Sea. Deep-Sea Res., 22, 711–724. Carrara, P. (1981). Evidence for a former large ice sheet in the Orville Coast-Ronne Ice Shelf area, Antarctica. J. Glaciol., 27, 487–491. Chase, T. E., Seekins, B. A., Young, J. D., & Eittreim, S. L. (1987). Marine topography of offshore Antarctica. In: S. L. Eittreim, & M. A. Hampton (Eds). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, CPCEMR Earth Science Series, Vol. 5A, pp. 147–150. Childs, J. R., Sliter, R. W., & Cooper, A. K. (1994). A progress report on the Antarctic Seismic Data Library System for Cooperative Research (SDLS). Terra Antart., 1(2), 243–246. Chow, J. M., & Bart, P. J. (2003). West Antarctic Ice Sheet grounding events on the Ross Sea outer continental shelf during the middle Miocene. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 169–186. Chriss, T., & Frakes, L. A. (1972). Glacial marine sedimentation in the Ross Sea. In: R. J. Adie (Ed.). Antarctic Geology and Geophysics. Universitetsforlaget, Oslo, Norway, International Union of Geological Sciences, Series B, Vol. 1, pp. 747–762. Ciesielski, P. F. (1975). Biostratigraphy and paleoecology of Neogene and Oligocene silicoflagellates from cores recovered during Antarctic Leg 28, Deep Sea Project. In: D. E. Hayes, L. A. Frakes, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, pp. 625–664. Ciesielski, P. F., & Weaver, F. M. (1974). Early Pliocene temperature changes in the Antarctic seas. Geology, 2, 511–515. Ciesielski, P. F., Dinkelman, M. G., Ledbetter, M. T., & Ellwood, B. B. (1978). New Evidence on the Timing of Late Miocene–Middle Pliocene Glacial Fluctuations from the Southwest Atlantic Sector of the Southern Ocean, Abstracts of Programs. Geological Society of America, Vol. 10, p. 380. Ciesielski, P. F., Ledbetter, M. T., & Ellwood, B. B. (1982). The development of Antarctic glaciation and the Neogene paleoenvironment of the Maurice Ewing Bank. Mar. Geol., 46, 1–51. Clapperton, C. M., & Sugden, D. E. (1982). Late Quaternary glacial history of George VI Sound area, West Antarctica. Quat. Res., 18, 243–267. Clapperton, C. M., & Sugden, D. E. (1990). Late Cenozoic glacial history of the Ross Embayment, Antarctica. Quat. Sci. Rev., 9, 253–272. Close, D. I., Stagg, H. M. J., & O’Brien, P. E. (2007). Seismic stratigraphy and sediment distribution on the Wilkes land and Terre Ade´lie margins, East Antarctica. Mar. Geol., 239, 33–57.
188
A. K. Cooper et al.
Coccioni, R., & Galeotti, S. (1997). Foraminiferal biostratigraphy and palaeoecology of the CIROS-1 Core from McMurdo Sound (Ross Sea, Antarctica). Terra Antart., 4(2), 103–117. Cochrane, G. R., De Santis, L., & Cooper, A. K. (1995). Seismic velocity expression of glacial sedimentary rocks beneath the Ross Sea from sonobuoy seismicrefraction data In: Seismic Stratigraphic Atlas of the Ross Sea. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 68, pp. 261–270. COHIMAR/SEDANO Scientific Party (2003). Uncovering the footprint of former ice streams off Antarctica. Eos Trans. Am. Geophys. Union, 84, 97–103. Colhoun, E. A. (1991). Geological evidence for changes in the East Antarctic Ice Sheet (60 degrees–120 degrees E) during the last glaciation. Polar Rec., 27, 345–355. Colhoun, E. A., Mabin, M. C. G., Adamson, D. A., & Kirk, R. M. (1992). Antarctic ice volume and contribution to sea-level fall at 20,000 yr BP from raised beaches. Nature, 358, 316–319. Colwell, J. B., Stagg, H. M. J., & Direen, N. G. (2006). Geology of deep-water margin of East Antarctica between Queen Mary and George V Lands. In: D. K. Futterer (Ed.). Terra Nostra: Proceedings of the Ninth International Symposium on Antarctic Earth Sciences IX. Springer, Amsterdam. Cooke, D. W., & Hays, J. D. (1982). Estimates of Antarctic Ocean seasonal sea-ice cover during glacial intervals. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 1017–1025. Cooper, A. K., & Davey, F. J. (1985). Episodic rifting of the Phanerozoic rocks of the Victoria Land Basin, Western Ross Sea, Antarctica. Science, 229, 1085–1087. Cooper, A. K., & Davey, F. J. (Eds). (1987). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea. Circum-Pacific Council of Energy and Mineral Resouces, Houston, TX, pp. 27–76. Cooper, A. K., & O’Brien, P. E. (2004). Leg 188 synthesis: Transitions in the glacial history of the Prydz Bay Region, East Antarctica, from ODP drilling. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. Available from Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 188 [CD-ROM], pp. 1–42. Cooper, A. K., Davey, F. J., & Behrendt, J. C. (1987). Seismic stratigraphy and structure of the Victoria Land basin, western Ross Sea, Antarctica. In: A. K. Cooper, & F. J. Davey (Eds). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, CPCEMR Earth Science Series, Vol. 5B, pp. 27–76. Cooper, A., Stagg, H., & Geist, E. (1991a). Seismic stratigraphy and structure of Prydz Bay, Antarctica: Implictions from Leg 119 drilling. In: J. Barron, B. Larsen, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, College Station, TX, Vol. 119, pp. 5–26. Cooper, A. K., Barrett, P. J., Hinz, K., Traube, V., Leitchenkov, G., & Stagg, H. M. J. (1991b). Cenozoic prograding sequences of the Antarctic continental margin: A record of glacio-eustatic and tectonic events. Mar. Geol., 102, 175–213.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
189
Cooper, A. K., Davey, F. J., & Hinz, K. (1991c). Crustal extension and origin of sedimentary basin beneath the Ross Sea and Ross Ice Shelf, Antarctica. In: M. R. A. Thompson, J. A. Crame, & J. W. Thompson (Eds). Geological Evolution of Antarctica. Cambridge University Press, pp. 285–291. Cooper, A. K., Eittreim, S., ten Brink, U., & Zayatz, I. (1993). Cenozoic glacial sequences of the Antarctic continental margin as recorders of Antarctic Ice Sheet fluctuations. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironmant: A Perspective on Global Change II. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 60, pp. 75–90. Cooper, A. K., Brancolini, G., Behrendt, J. C., Davey, F. J., Barrett, P. J., & ANTOSTRAT Ross Sea Regional Working Group. (1994). A record of Cenozoic tectonism throughout the Ross Sea and possible controls on the glacial record. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea Level Change. Terra Antart., Vol. 1, pp. 353–355. Cooper, A. K., Barker, P. F., & Brancolini, G. (Eds). (1995). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68. Coren, F., Ceccone, G., Lodolo, E., Zanolla, C., Zitellini, N., Bonazzi, C., & Centonze, J. (1997). Morphology, seismic structure and tectonic development of the Powell Basin, Antarctica. J. Geol. Soc. London, 154, 849–862. Costa, E., Dunbar, R. B., Kryc, K. A., Mucciarone, D. A., Brachfeld, S., Roark, E. B., Manley, P. L., Murray, R. W., & Leventer, A. (2007). Solar forcing and El Nin˜o-Southern-Oscillation (ENSO) influences on productivity cycles interpreted from late-Holocene high-resolution marine sediment record, Ade´lie Drift, East Antarctic Margin. In: A. K. Cooper, C. R. Raymond, et al. (Eds) Antarctica: A Keystone in a Changing World-Online Proceedings of the 10th International Symposium on Antarctic Earth Sciencies. USGS Open-File Report 2007-1047, Short Research Paper 036, 6 pp.; doi:10.3133/of2007-1047.srp036. Cunningham, A. P., Larter, R. D., & Barker, P. F. (1994). Glacially prograded sequences on the Bellingshausen Sea continental margin near 901W. Terra Antart., 1, 267–268. Cunningham, A. P., Larter, R. D., Barker, P. F., Gohl, K., & Nitsche, F.-O. (2002). Tectonic evolution of the Pacific margin of Antarctica: 2. Structure of late Cretaceous–early Tertiary plate boundaries in the Bellingshausen Sea from seismic reflection and gravity data. J. Geophys. Res., 107, 2346, doi:10.1029/ 2002JB001897. Cunningham, W. L., Leventer, A., Andrews, J. T., Jennings, A. E., & Licht, K. J. (1999). Late Pleistocene–Holocene marine conditions in the Ross Sea, Antarctica: Evidences from diatom record. Holocene, 2(9), 129–139. Dalziel, I. W. D. (1984). Tectonic evolution of a fore-arc terrane, southern Scotia Ridge, Antarctica. Geological Society of America, Boulder, CO, Special Paper, Vol. 200, 32 pp.
190
A. K. Cooper et al.
DaSilva, J. L., Anderson, J. B., & Stravers, J. (1997). Seismic facies changes along a nearly continuous 24 degree latitudinal transect: The fjords of Chile and the northern Antarctic Peninsula. Mar. Geol., 103, 103–123. Davey, F. J. (1987). Geology and structure of the Ross Sea region. In: A. K. Cooper, & F. J. Davey (Eds). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, CPCEMR Earth Science Series, Vol. 5B, pp. 1–16. Davey, F. J., & Brancolini, G. (1995). The late Mesozoic and Cenozoic structural setting of the Ross Sea region. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 167–182. Davey, F. J., Bennett, D. J., & Houtz, R. E. (1982). Sedimentary basins of the Ross Sea, Antarctica. N. Z. J. Geol. Geophys., 25, 245–255. Davey, F. J., & Cooper, A. K. (Conveners). (2007). Circum-Antarctic Stratigraphy and Paleobathymetry (CASP) project: Report on a meeting held at the 10th ISAES on September 1, 2007. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES. USGS Open-File Report 2007–1047, Extended Abstract 218, pp. 1–5. De Batist, M., Bart, P., Kuvaas, B., Moons, A., & Miller, H. (1994). Detailed seismic stratigraphy of the Crary Fan, southeastern Weddell Sea. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea-Level Change. Terra Antart., Vol. 1, pp. 321–323. Debenham, F. (1919). A new mode of transportation by ice: The raised marine muds of South Victoria Land. Q. J. Geol. Soc. London, 75, 51–76. del Valle, R. A., Elliot, D. H., & Macdonald, D. I. M. (1992). Sedimentary basins on the east flank of the Antarctic Peninsula: Proposed nomenclature. Antarct. Sci., 4, 477–478. del Valle, R. A., Fourcade, N. H., & Medina, F. A. (1982). The stratigraphy of Cape Lamb and The Naze, Vega and James Ross islands, Antarctica. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 275–280. Denton, G. H., Armstrong, R. L., & Suiver, M. (1971). The late Cenozoic glacial history of Antarctica: The late Cenozoic glacial ages. Yale University Press, New Haven, CT, p. 267–306 Denton, G. H., Bockheim, J. G., Wilson, S. C., & Stuiver, M. (1989). Late Wisconsin and early Holocene glacial history inner Ross embayment, Antarctic. Quat. Res., 31, 151–182. Denton, G. H., Prentice, M. L., & Burckle, L. H. (1991). Cainozoic history of the Antarctic Ice Sheet. In: R. J. Tingey (Ed.). The Geology of Antarctica. Clarendon Press, Oxford, pp. 365–433.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
191
Denton, G. H., Sugden, D. E., Marchant, D. R., Hall, B. L., & Wilch, T. I. (1993). East Antarctica Ice Sheet sensitivity to Pliocene climate change from a Dry Valleys perspective. Geogr. Ann., 75A, 155–204. De Santis, L., Anderson, J. B., Brancolini, G., & Zayatz, I. (1995). Seismic Record of Late Oligocene Through Miocene Glaciation on the Central and Eastern Continetal Shelf of the Ross Sea. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 68, pp. 253–260. De Santis, L., Prato, S., Brancolini, G., Lovo, M., & Torelli, L. (1999). The eastern Ross Sea continental shelf during the Cenozoic: Implications for the West Antarctic Ice Sheet development. In: F. M. Van der Wateren, & S. A. P. L. Cloetingh (Eds). Lithosphere dynamics and environmental change of the Cenozoic West Antarctic Rift System. Global Planet. Change, 23 (Special Issue), 173–196. De Santis, L., Brancolini, G., & Donda, F. (2003). Seismic-stratigraphic analysis of the Wilkes Land continental margin (East Antarctica). Influence of glaciallydriven processes on the Cenozoic deposition. Deep Sea Res. II, 50(8–9), 1563–1594. Dewar, G. J. (1970). The Geology of Adelaide Island. British Antarctic Survey Scientific Reports, No. 57, 66 pp. Diester-Haass, L., Robert, C., & Chamley, H. (1993). Paleoceanographic and paleoclimatic evolution in the Weddell Sea (Antarctica) during the middle Eocene-late Oligocene, from a coarse sediment fraction and clay mineral data (ODP Site 689). Mar. Geol., 114, 233–250. Dietrich, R., Ru¨lke, A., Ihde, J., Lindner, K., Miller, H., Niemeier, W., Schenke, H.-W., & Seeber, G. (2004). Plate kinematics and deformation status of the Antarctic Peninsula based on GPS. Global Planet. Change, 42, 313–321. Dingle, R. V., & Lavelle, M. (1998). Antarctic Peninsular cryosphere: Early Oligocene (c. 30 Ma) initiation and a revised glacial chronology. J. Geol. Soc. London, 155, 433–437. Ditchfield, P. W., Marshall, J. D., & Pirrie, D. (1994). High latitude paleotemperature variation; new data from the Tithonian to Eocene of James Ross Island, Antarctica. Palaeogeogr., Palaeoclimatol. Palaeoecol., 107, 79–102. Doake, C. S. M. (1982). State of balance of the ice sheet in the Antarctic Peninsula. Ann. Glaciol., 3, 77–82. Doake, C. S. M. (1985). Antarctic mass balance: Glaciological evidence from Antarctic Peninsula and Weddell Sea sector. In: Glaciers, Ice Sheets, and Sea Level: Effect of a CO2– induced Climatic Change. US National Research Council, Washington, DC, pp. 197–209. Doake, C. S. M., Crabtree, R. D., & Dalziel, I. W. D. (1983). Subglacial morphology between Ellsworth Mountains and Antarctic Peninsula: New data and tectonic significance. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 270–273. Doake, C. S. M., & Vaughan, D. G. (1991). Rapid disintegration of the Wordie Ice Shelf in response to atmospheric warming. Nature, 350, 328–330.
192
A. K. Cooper et al.
Doktor, M., Gazdzicki, A., Marenssi, S. A., Porebski, S. J., Santillana, S. N., & Vrba, A. V. (1988). Argentine-Polish geological investigations on Seymour (Marambio) Island, Antarctica, 1988. Pol. Polar Res., 9, 521–541. Domack, E. W. (1982). Sedimentology of glacial and glacial marine 1712 deposits on the George V – Adelie continental shelf, East 1713 Antarctica. Boreas, 11, 79–97. Domack, E. W. (1987). Preliminary stratigraphy from a portion of the Wilkes Land continental shelf: Evidence from till provenance. In: S. L. Eittreim, & M. A. Hapton (Eds). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, Vol. 5A, pp. 195–204. Domack, E. W. (1988). Biogenic facies in the Antarctic glacimarine environment: Basis for a polar glacimarine summary. Palaeogeogr. Palaeoclimatol. Palaeoecol., 63, 357–372. Domack, E. W. (1990). Laminated terrigenous sediments from the Antarctic Peninsula: The role of subglacial and marine processes. In: J. A. Dowdeswell, & J. D. Scourse (Eds). Glacimarine Environments: Processes and Sediments. Geological Society, London, England, Special Publication No. 53, pp. 91–103. Domack, E. W., & Anderson, J. B. (1983). Marine geology of the George V continental margin: Combined results of Deep Freeze 79 and the 1911–14 Australasian expedition. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 402–406. Domack, E. W., & Williams, C. R. (1990). Fine structure and suspended sediment transport in three Antarctic fjords. In: D. H. Elliot (Ed.). Contributions to Antarctic Research I. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 50, pp. 71–89. Domack, E. W., & Ishman, S. (1993). Oceanographic and physiographic controls on modern sedimentation within Antarctic fjords. Geol. Soc. Am. Bull., 105, 1175–1189. Domack, E. W., & Ishman, S. (1994). Oceanographic controls on benthic foraminifers from the Bellingshausen margin of the Antarctic Peninsula. Mar. Micropaleontol., 24, 119–155. Domack, E. W., & McClennen, C. E. (1996). Accumulation of glacial marine sediments in fjords of the Antarctic Peninsula and their use as late Holocene paleoenvironmental indicators. In: R. Ross, E. E. Hofmann, & L. B. Quentin (Eds). Foundations for Ecological Research West of the Antarctic Peninsula. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 70, pp. 135–154. Domack, E. W., Fairchild, W. W., & Anderson, J. B. (1980). Lower Cretaceous sediment from the East Antarctic continental shelf. Nature, 287, 625–626. Domack, E. W., Jull, A. J. T., Anderson, J. B., Linick, T. W., & Williams, C. R. (1989). Application of tandem accelerator mass-spectrometer dating to Late Pleistocene-Holocene sediments of the East Antarctic continental shelf. Quat. Res., 31, 277–287.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
193
Domack, E. W., Jull, A. J. T., & Donahue, D. J. (1991). Holocene chronology for the unconsolidated sediments at Hole 740A: Prydz Bay, East Antarctica. In: J. Barron, B. Larsen, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. College Station, TX, Vol. 119, pp. 747–750. Domack, E. W., Mashiotta, T. A., Burhley, L. A., & Ishman, S. E. (1993). 33-year cyclicity in organic matter preservation in Antarctic fjord sediments. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change II. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 60, pp. 265–272. Domack, E. W., Ishman, S. E., Stein, A. B., McClennen, C. E., & Jull, A. J. T. (1995). Late Holocene advance of the Mu¨ller Ice Shelf, Antarctic Peninsula: Sedimentological, geochemical and palaeontological evidence. Antarct. Sci., 7, 159–170. Domack, E., O’Brien, P. E., Harris, P. T., Taylor, F., Quilty, P. G., DeSantis, L., & Raker, B. (1998). Late Quaternary sedimentary facies in Prydz Bay, East Antarctica and their relationship to glacial advance onto the continental shelf. Antarct. Sci., 10, 227–235. Domack, E. W., Jacobson, E. A., Shipp, S., & Anderson, J. B. (1999). Late Pleistocene–Holocene retreat of the West Antarctic Ice Sheet system in the Ross Sea: Part 2: Sedimentologic and stratigraphic signature. GSA Bull., 111, 1517–1536. Domack, E., Leventer, A., Dunbar, R., Taylor, F., Brachfeld, S., & Sjunneskog, C., ODP Leg 178 Scientific Party. (2001a). Chronology of the Palmer Deep site, Antarctic Peninsula: A holocene palaeoenvironmental reference for the circumAntarctic. Holocene, 11, 1–9. Domack, E., Leventer, A., Gilbert, R., Brachfeld, S., Ishman, S., Camerlenghi, A., Gavahan, K., Carlson, D., & Barkoukis, A. (2001b). Cruise reveals history of Holocene Larsen Ice Shelf. Eos Trans. Am. Geophys. Union, 82, 13–17. Domack, E., Duran, D., Leventer, A., Ishman, S., Doane, S., McCallum, S., Amblas, D., Ring, J., Gilbert, R., & Prentice, M. (2005). Stability of the Larsen B Ice Shelf on the Antarctic Peninsula during the Holocene epoch. Nature, 436, 681–685. Donda, F., Brancolini, G., De Santis, L., & Escutia, C. (2007a). Sedimentary processes in the Wilkes Land margin: A record of the Cenozoic East Antarctic Ice Sheet evolution. J. Geol. Soc. Lond., 164, 243–256. Donda, F., O’Brien, P. E., De Santis, L., Rebesco, M., & Brancolini, G. (2007b). Maga debris flow deposits on the western Wilkes Land margin, East Antarctica. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th International Symposium on Antarctic Earth Sciences. USGS Open-File Report 2007-1047, Short Research Paper 040, 4, doi:10.3133/of2007-1047.srp040. Donda, F., Brancolini, G., De Santis, L., & Trincardi, F. (2003). Seismic facies and sedimentary processes on the continental rise off Wilkes Land (East
194
A. K. Cooper et al.
Antarctica). Evidence of bottom current activity. Deep Sea Res. II, 50(8–9), 1509–1528. Dowdeswell, J. A., Evans, J., O´ Cofaigh, C., & Anderson, J. B. (2006). Morphology and sedimentary processes on the continental slope off Pine Island Bay, Amundsen Sea, West Antarctica. Geol. Soc. Am. Bull., 118, 606–619. Dowdeswell, J. A., O´ Cofaigh, C., & Pudsey, C. J. (2004). Thickness and extent of the subglacial till layer beneath an Antarctic paleo-ice stream. Geology, 32, 13–16. Drewry, D. J. (1981). The record of late Cenozoic glacial events in East Antarctica (60 degrees–171 degrees E). In: M. J. Hambrey, & W. B. Harland (Eds). Earth’s Pre-Pleistocene Glacial Record: IGCP Project No. 38. Cambridge University Press, Cambridge, pp. 212–216. Drewry, D. J. (1983). Antarctica: Glaciological and Geophysical Folio. Scott Polar Research Institute, University of Cambridge, Cambridge, England, 9 pp Drewry, D. J., & Cooper, A. P. R. (1981). Processes and models of Antarctic glaciomarine sedimentation. Ann. Glaciol., 2, 117–122. Dunbar, R. B., Anderson, J. B., Domack, E. W., & Jacobs, S. S. (1985). Oceanographic influences on sedimentation along the Antarctic continental shelf. In: S. S. Jacobs (Ed.). Oceanology of the Antarctic Continental Shelf. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 43, pp. 291–312. Dunbar, R. B., Leventer, A. R., & Stockton, W. L. (1989). Biogenic sedimentation in McMurdo Sound, Antarctica. Mar. Geol., 85, 155–179. Eagles, G., & Livermore, R. A. (2002). Opening history of Powell Basin, Antarctic Peninsula. Mar. Geol., 185, 195–205. Edwards, B. D., Lee, H. J., Karl, H. A., Reimnitz, E., & Timothy, L. A. (1987). Geology and physical properties of Ross Sea, Antarctica, continental shelf sediment. In: A. K. Cooper, & F. J. Davey (Eds). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, CPCEMR Earth Science Series, Vol. 5B, pp. 191–216. Edwards, D. S. (1968). The Detrital Mineralogy of Surface Sediments of the Ocean Floor in the Area of the Antarctic Peninsula, Antarctica. M.A. Thesis, Florida State University, Tallahassee, FL, 143 pp. Ehrmann, W. U. (1991). Implications of sediment composition on the southern Kerguelen Plateau for paleoclimate and depositional environment. In: J. Barron, B. Larsen, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 119. College Station, TX, pp. 185–210. Ehrmann, W. V. (1995). Clay mineral assemblages in Cenozoic sediments of Ross Sea, Antarctica. In: Proceedings of the VII ISAES 1995, Siena, Italy, 118 pp. Ehrmann, W. U., Melles, M., Kuhn, G., & Grobe, H. (1992). Significance of clay mineral assemblages in the Antarctic Ocean. Mar. Geol., 107, 249–273. Eittreim, S. L. (1994). Transition from continental to oceanic crust on the WilkesAdelie margin of Antarctica. J. Geophys. Res., 99, 24189–24205. Eittreim, S. L., & Smith, G. L. (1987). Seismic sequences and their distribution on the Wilkes Land margin. In: S. L. Eittreim, & M. A. Hampton (Eds). The
Cenozoic Climate History from Seismic Reflection and Drilling Studies
195
Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, Earth Sciences Series 5A, pp. 15–43. Eittreim, S. L., & Hampton, M. A. (Eds). (1987). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, pp. 1–221. Eittreim, S. L., Cooper, A. K., & Wannesson, J. (1995). Seismic stratigraphic evidence of ice-sheet advances on the Wilkes Land margin of Antarctica. Sed. Geol., 96, 131–156. Elliot, D. H. (1965). Geology of north-west Trinity Peninsula, Graham Land. Brit. Antarct. Surv. Bull., 7, 1–24. Elliot, D. H. (1966). Geology of the Nordenskjo¨ld Coast and a comparison with north-west Trinity Peninsula, Graham Land. Brit. Antarct. Surv. Bull., 10, 1–43. Elliot, D. H. (1988). Tectonic setting and evolution of the James Ross Basin, northern Antarctic Peninsula. In: R. M. Feldman, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America, Boulder, CO, Memoir 169, pp. 541–555. Elliot, D. H., & Trautman, T. A. (1982). Lower Tertiary strata on Seymour Island, Antarctic Peninsula. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 287–297. Elliot, D. H., Hoffman, S. M., & Rieske, D. E. (1992). Provenance of Paleocene strata, Seymour Island. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing Co., Tokyo, Japan, pp. 347–355. Elverhøi, A. (1984). Glacigenic and associated marine sediments in the Weddell Sea, fjords of Spitsbergen, and the Barents Sea: A Review. Mar. Geol., 57, 53–88. Elverhøi, A., & Maisey, G. (1983). Glacial erosion and morphology of the eastern and southeastern Weddell Sea Shelf. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Australian Academy of Science, Canberra, ACT, pp. 483–487. Elverhøi, A. (1981). Evidence for a late Wisconsin glaciation of the Weddell Sea. Nature, 293, 641–642. Elverhøi, A., & Roaldset, E. (1983). Glaciomarine sediments and suspended particulate matter, Weddell Sea Shelf, Antarctica. Polar Res., 1, 1–21. Engelhardt, H., Humphrey, N., Kamb, B., & Fahnestock, M. (1990). Physical conditions at the base of a fast moving Antarctic ice stream. Science, 248, 57–59. Erohina, T., Cooper, A. K., Handwerger, D., & Dunbar, R. (2004). Seicmic stratigraphic correlations between ODP Sites 742 and 1166: Implications for depositional paleoenvironments in Prydz Bay, Antarctica. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 188 [CD-ROM]. Available from Ocean Drilling Program, Texas A&M University, College Station, TX, pp. 1–21.
196
A. K. Cooper et al.
Escutia, C., Eittreim, S. L., & Cooper, A. K. (1997). Cenozoic glaciomarine sequences on the Wilkes Land continental rise, Antarctica. In: Proceedings Volume-VII International Symposium on Antarctic Earth Sciences, pp. 791–795. Escutia, C., Eittreim, S. L., Cooper, A. K., & Nelson, C. H. (2000). Morphology and acoustic character of the Antarctic Wilkes Land turbidite systems: Ice-sheet sourced versus river-sourced fans. J. Sed. Res., 70(1), 84–93. Escutia, C., Nelson, C. H., Acton, G. D., Cooper, A. K., Eittreim, S. L., Warnke, D. A., & Jaramillo, J. (2002). Current controlled deposition on the Wilkes Land continental rise. In: D. Stow, et al. (Eds). Deep-Water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics. The Geological Society of London, Memoirs, Vol. 22, pp. 373–384. Escutia, C., Warnke, D. A., Acton, G. D., Barcena, A., Burckle, L., Canals, M., & Frazee, C. S. (2003). Sediment distribution and sedimentary processes across the Antarctic Wilkes Land margin during the Quaternary. Deep Sea Res. II, 50, 1481–1508. Escutia, C., De Santis, L., Donda, F., Dunbar, R. B., Brancolini, G., Eittreim, S. L., & Cooper, A. K. (2005). Cenozoic Ice Sheet history from east Antarctic Wilkes Land continental margin sediments. Global Planet. Change, 45, 1–3. Escutia, C., Donda, F., Lobo, F. J., & Tanahashi, M. (2007). Extensive debris flow deposits on the eastern Wilkes Land margin: a key to changing glacial regimes. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th International Symposium on Antarctic Earth Sciences. USGS Open-File Report 2007-1047, Short Research Paper 026, 4, doi:10.3133/of2007-1047.srp026. Evans, J., Pudsey, C. J., O´ Cofaigh, C., Morris, P., & Domack, E. (2005). Late Quaternary glacial history, flow dynamics and sedimentation along the eastern margin of the Antarctic Peninsula Ice Sheet. Quat. Sci. Rev., 24, 741–774. Evans, J., Dowdeswell, J. A., O´ Cofaigh, C., Anderson, J. B., & Benham, T. J. (2006). Extent and dynamics of the West Antarctic Ice Sheet on the outer continental shelf of Pine Island Bay during the last glaciation. Mar. Geol., 230, 53–72. Eyles, N., Daniels, J., Osterman, L. E., & Januszczak, N. (2001). Ocean Drilling Program Leg 178 (Antarctic Peninsula): Sedimentology of glacially influenced continental margin topsets and foresets. Mar. Geol., 178, 135–156. Fechner, N., & Jokat, W. (1995). A seismic transect across the Ronne Shelf. In: VII International Symposium on Antarctic Earth Sciences, Abstracts. Tipografia Senese, Siena, Italy, 25 pp. Fielding, C. R., Woolfe, K. J., Purdon, R. G., Lavelle, M., & Howe, J. A. (1997). Sedimentological and stratigraphical re-evaluation of the CIROS-1 Core, McMurdo Sound, Antarctica. Terra Antart., 4(2), 149–160. Fillon, R. H. (1972). Evidence from the Ross Sea for widespread submarine erosion. Nature, 238, 40–42. Fisco, M. P. P. (1982). Sedimentation on the Weddell Sea Continental Margin and Abyssal Plain, Antarctica. M.A. Thesis, Rice University, Houston, TX, 170 pp.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
197
Fitzgerald, P. G., & Stump, E. (1997). Cretaceous and Cenozoic episodic denudation of the Transantarctic Mountains, Antarctica; new constraints from apatite fission track thermochronology in the Scott Glacier region. J. Geophys. Res. B. Solid Earth Planets, 102(4), 7747–7765. Fitzsimons, S. J. (1990). Ice-marginal depositional processes in a polar maritime environment, Vestfold Hills, Antarctica. J. Glaciol., 36, 279–286. Flemming, R. F., & Barron, J. A. (1996). Evidence of Pliocene Nothofagus in Antarctica from Pliocene marine sedimentary deposits (DSDP site 274). Mar. Micropaleontol., 27, 227–236. Florindo, F., Bohaty, S. M., Erwin, P. S., Richter, C., Roberts, A. P., Whalen, P. A., & Whitehead, J. M. (2003). Magnetobiostratigraphic chronology and palaeoenvironmental history of Cenozoic sequences from ODP sites 1165 and 1166, Prydz Bay, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 69–100. Florindo, F., Harwood, D., Levy, R., and SMS Project Science Team (2008). ANDRILL’s success during the 4th International Polar Year. Scientific Drilling, 6, 29–31, doi:10.2204/iodp.sd.6.03.2008. Flower, B. P., & Kennett, J. P. (1994). The middle Miocene climatic transition: East Antarctic Ice Sheet development, deep ocean circulation and global carbon cycling. Palaeogeogr. Palaeoclimatol. Palaeoecol., 108, 537–555. Foldvik, A., & Gammelsrød, T. (1988). Notes on southern ocean hydrography, sea-ice and bottom water formation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 67, 3–17. Foster, T. D., & Carmack, E. C. (1976). Frontal zone mixing and Antarctic bottom water formation in the southern Weddell Sea. Deep Sea Res., 23, 301–317. Frakes, L. A. (1983). Problems in Antarctic marine geology: A Review. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Sciences. Cambridge University Press, New York, pp. 375–378. Frakes, L. A., & Francis, J. E. (1988). A guide to Phanerozoic cold polar climates from high-latitude ice-rafting in the Cretaceous. Nature, 333, 547–549. Frezzotti, M. (1992). Fluctuations of ice tongues and ice shelves derived from satellite images in Terra Nova Bay Area, Victoria Land, Antarctica. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing Co., Tokyo, Japan, pp. 733–739. Fu¨tterer, D. K., & Melles, M. (1990). Sediment patterns in the southern Weddell Sea: Filchner Shelf and Filchner Depression. In: U. Bleil, & J. Thiede (Eds). Geologic History of the Polar Oceans: Arctic versus Antarctic. Kluwer Academic Publishers, Boston, MA, pp. 381–401. Gamboˆa, L. A. P., & Maldonado, P. R. (1990). Geophysical investigations in the Bransfield Strait and in the Bellingshausen Sea–Antarctica. In: B. St John (Ed.). Antarctica as an Exploration Frontier–Hydrocarbon Potential, Geology and Hazards. American Association of Petroleum Geologists, Tulsa, OK, Studies in Geology, Vol. 31, pp. 127–141. Gersonde, R., Abelman, A., et al. (1990). Biostratigraphic synthesis of Neogene siliceous microfossils from the Antarctic Ocean, ODP Leg 113 (Weddell Sea).
198
A. K. Cooper et al.
In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 113, pp. 915–936. Gill, A. E. (1973). Circulation and bottom water production in the Weddell Sea. Deep Sea Res., 20, 111–140. Glasby, G. P., Barrett, P. J., McDougall, J. C., & McKnight, D. G. (1975). Localized variations in sedimentation characteristics in the Ross Sea and McMurdo Sound Regions, Antarctica. N. Z. J. Geol. Geophy., 18, 605–621. Gloersen, P., Campbell, W. J., Cavalieri, D. J., Comiso, J. C., Parkinson, C. L., & Zwally, H. J. (1992). Arctic and Antarctic Sea Ice, 1978–1987: Satellite Passive Microwave Observations and Analysis, Special Publication, Vol. 511. National Aeronautics and Space Administration, Washington, DC, 290 pp. Gohl, K., Nitsche, F., & Miller, H. (1997). Seismic and gravity data reveal Tertiary interplate subduction in the Bellingshausen Sea, southeast Pacific. Geology, 25, 371–374. Gohl, K., Teterin, D., Eagles, G., Netzeband, G., Grobys, J. W. G., Parsiegla, N., Schlu¨ter, P., Leinweber, V., Larter, R. D., Uenzelmann-Neben, G., & Udintsev, G. B. (2007). Geophysical Survey Reveals Tectonic Structures in the Amundsen Sea Embayment, West Antarctica. US Geological Survey Open File Report, OF-20071047, doi:10.3133/of2007-1047.srp047. Goodell, H. G., Watkins, N. D., Mather, T. T., & Koster, S. (1968). The Antarctic glacial history recorded in sediments of the Southern Ocean. Palaeogeogr. Palaeoclimatol. Palaeoecol., 5, 41–62. Goodell, H. G., Houtz, R., Ewing, M., Hayes, D., Naini, B., Echols, R. J., Kennett, J. P., & Donahue, J. G. (1973). Marine Sediments of the Southern Oceans. Antarctic Map Folio Series – Folio 17. American Geographical Society, New York, p. 18 Goodwin, I. D. (1993). Holocene deglaciation, sea-level change, and the emergence of the Windmill Islands, Budd Coast, Antarctica. Quat. Res., 40, 70–80. Gordon, A. L. (1971). Oceanography of Antarctic Waters. In: J. L. Reid (Ed.). Antarctic Oceanology I. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 15, pp. 169–203. Gordon, A. L., & Tchernia, P. (1972). Waters of the continental margin off Ade´lie Coast, Antarctica. In: D. E. Hayes (Ed.). Antarctic Oceanology II. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 19, pp. 59–69. Gra`cia, E., Canals, M., Farra`n, M., Prieto, M. J., & Sorribas, J.GEBRA Team. (1996). Morphostructure and evolution of the central and eastern Bransfield Basins (NW Antarctic Peninsula). Mar. Geophys. Res., 18, 429–448. Gra`cia, E., Canals, M., Farra`n, M., Sorribas, J., & Palla`s, R. (1997). Central and eastern Bransfield basins (Antarctica) from high-resolution swath-bathymetry data. Antarct. Sci., 9, 168–180. Grad, M., Guterch, A., & Janik, T. (1993). Seismic structure of the lithosphere across the zone of subducted Drake plate under the Antarctic plate, West Antarctica. Geophys. J. Int., 115, 586–600.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
199
GRAPE Team (1990). Preliminary results of seismic reflection investigations and associated geophysical studies in the area of the Antarctic Peninsula. Antarct. Sci., 2, 223–234. Griffith, T. W. (1988). A Geological and Geophysical Investigation of Sedimentation and Recent Glacial History in the Gerlache Strait Region, Graham Land, Antarctica. M.A. Thesis, Rice University, Houston, TX, 449 pp. Griffith, T. W., & Anderson, J. B. (1989). Climatic control of sedimentation in bays and fjords of the northern Antarctic Peninsula. Mar. Geol., 85, 181–204. Grinnell, D. V. (1971). Physiography of the continental margin of Antarctica from 1251E to 1501E. Antarct. J. U.S., 6, 164–165. Grobe, H., Fuetterer, D. K., & Spiess, V. (1990). Oligocene to quaternary sedimentation processes on the Antarctic continental margin, ODP Leg 113, Site 693. In: P. F. Barker, & J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 113. College Station, TX, pp. 121–131. Grobe, H., & Mackensen, A. (1992). Late Quaternary climatic cycles as recorded in sediments from the Antarctic continental margin. In: J. P. Kennett, & D. Warnke (Eds). The Antarctic Paleoenvironment: A perspective on global change. American Geophysical Union, Washington, DC, Antarctica Research Series Vol. 56, pp. 349–376. Grobe, H., Huybrechts, P., & Fu¨tterer, D. K. (1993). Late Quaternary record of sealevel changes in the Antarctic. Geol. Rundsch., 82, 263–275. Gru¨tzner, J., Rebesco, M. A., Cooper, A. K., Forsberg, C. F., Kryc, K. A., & Wefer, G. (2003). Evidence for orbitally controlled size variations of the East Antarctic Ice Sheet during the late Miocene. Geology, 31, 777–780. Guterch, A., Grad, M., Janik, T., Perchuc´, E., & Pajchel, J. (1985). Seismic studies of the crustal structure in West Antarctica 1979–1980 – Preliminary results. Tectonophysics, 114, 411–429. Haase, G. M. (1986). Glaciomarine sediments Along the Filchner/Rønne Ice Shelf, southern Weddell Sea – first results of the 1983/84 Antarktis-II/4 Expedition. Mar. Geol., 72, 241–258. Hambrey, M. J., & Harland, W. B. (Eds). (1981). Earth’s Pre-Pleistocene Glacial Record. Cambridge University Press, New York, p. 1004. Hambrey, M. J., & Barrett, P. J. (1993). Cenozoic Sedimentary and Climatic Record, Ross Sea Region, Antarctica, Antarctic Research Series, Vol. 60. American Geophysical Union, Washington, DC, pp. 91–124. Hambrey, M. J., & McKelvey, B. (2000). Neogene sedimentation of the western margin of the Lambert Glacier. Sedimentology, 47, 577–607. Hambrey, M. J., Larsen, B., & Ehrmann, W. U.ODP Leg 119 Shipboard Scientific Party. (1989). Forty million years of Antarctic glacial history yielded by Leg 119 of the Ocean Drilling Program. Polar Rec., 25, 99–106. Hambrey, M. J., Ehrmann, W. U., & Larsen, B. (1991). Cenozoic glacial record of the Prydz Bay continental shelf, East Antarctica. In: J. Barron, B. Larson, et al.
200
A. K. Cooper et al.
(Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 119, pp. 77–132. Hampton, M. A. (1972). The role of subaqueous debris flows in generating turbidity currents. J. Sediment. Petrol., 42, 775–793. Hampton, M. A., Eittreim, S. L., & Richmond, B. M. (1987). Post-breakup sedimentation on the Wilkes Land Margin, Antarctica. In: S. L. Eittreim, & M. A. Hampton (Eds). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, Earth Sciences Series 5A, pp. 75–88. Handwerger, D. A., Cooper, A. K., O’Brien, P. E., Williams, T., Barr, S. R., Dunbar, R. B., Leventer, A., & Jarrard, R. D. (2004). Synthetic seismograms linking ODP sites to seismic profiles, continental rise and shelf of Prydz Bay, Antarctica. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 188 [CDROM]. Ocean Drilling Program, Texas A&M University, College Station, TX, pp. 1–28. Hannah, M. J. (1994). Eocene dinoflagellates from CIROS-1 drill-hole, McMurdo Sound, Antarctica. Terra Antart., 1(2), 371–373. Hannah, M. J., Cita, M. B., Coccioni, R., & Monechi, S. (1997). The Eocene/Oligocene boundary at 701 South, McMurdo Sound, Antarctica. Terra Antart., 4(2), 79–87. Hansom, J. D. (1983). Ice-formed intertidal boulder pavements in the sub-Antarctic. J. Sediment. Petrol., 53, 135–145. Haq, B. U., Hardenbol, J., & Vail, P. R. (1988). Mesozoic and Cenozoic chronostratigraphy and cycles of sea-level change. In: C. K. Wilgus, B. S. Hastings, C. G. St. C. Kendall, H. W. Posamentier, C. A. Ross, & J. C. Van Wagoner (Eds). Sea Level Changes: An Integrated Approach. Society of Economic Paleontologists and Mineralogists, Tulsa, OK, Special Publication, Vol. 42, pp. 71–108. Haq, B. U., Harnderbol, J., & Vail, P. R. (1987). Chronology of fluctuating sea level since Triassic. Science, 235, 156–1167. Harrington, P. K., Barker, P. F., & Griffiths, D. H. (1972). Continental structure of the South Orkney Islands area from seismic refraction and magnetic measurements. In: R. J. Adie (Ed.). Antarctic Geology and Geophysics. Universitets Forlaget, Oslo, pp. 27–32. Harris, P. T., O’Brien, P. E., Quilty, P. G., Taylor, F., Domack, E., DeSantis, L., & Raker, B. (1997). Vincennes Bay, Prydz Bay and Mac. Robertson Shelf, PostCruise Report: AGSO Cruise 186, Anare Voyage 5, 1996/97. AGSO, Canberra City, Australia, 75 pp. Harwood, D. M. (1991). Cenozoic diatom biogeography in the southern high latitudes: Inferred biogeographic barriers and progressive endemsim. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 667–673. Harwood, D. M., Florindo, F., Talarico, F., & Levy, R. (2008). Studies from the ANDRILL, Southern McMurdo Sound Project, Antarctica – Initital Science Report on AND-2A. Terra Antartica.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
201
Haugland, K. (1982). Seismic reconnaissance survey in the Weddell Sea. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 405–413. Haugland, K., Kristoffersen, Y., & Velde, A. (1985). Seismic investigations in the Weddell Sea embayment. Tectonophys, 114, 293–313. Hayes, D. E. (Ed.) (1991). Marine Geological and Geophysical Atlas of the CircumAntarctic to 301S. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 54. Hayes, D. E., & Davey, F. J. (1975). A geophysical study of the Ross Sea, Antarctica. In: D. E. Hayes, & L. A. Frakes (Eds). Initial Reports of the Deep Sea Drilling Project Leg 28. US Government Printing Office, Washington, DC, Vol. 28, pp. 887–907. Hayes, D. E., & Frakes, L. A. (Eds). (1975). Initial Reports of the Deep Sea Drilling Project Leg 28. US Government Printing Office, Washington, DC, 28, 1017. Hays, J. D., & Opdyke, N. D. (1967). Antarctic radiolaria, magnetic reversals, and climatic change. Science, 158, 1001–1011. Heezen, B. C., Tharp, M., & Bentley, C. R. (1972). Morphology of the Earth in the Antarctic and Sub-Antarctic. Antarctic Map Folio Series – Folio 16. American Geographical Society, New York, 16 pp. Henriet, J. P., Meissner, R., & Miller, H.GRAPE Team. (1992). Active margin processes along the Antarctic Peninsula. Tectonophysics, 201, 229–253. Henrys, S. A., Bartek, L. R., Anderson, J. B., & Barrett, P. J. (1994). Seismic stratigraphy in McMurdo sound: Correlation of high resolution data sets. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea-Level Change. Terra Antart., 1, 373–374. Henrys, S. A., Bu¨cker, C. J., Niessen, F., & Bartek, L. R. (2001). Correlation of seismic reflectors with the CRP-3 Drillhole, Victoria Land Basin, Antarctica. Terra Antart., 8(3), 127–136. Herna´ndez-Molina, F. J., Larter, R. D., Rebesco, M., & Maldonado, A. (2004). Miocene changes in bottom current regime recorded in continental rise sediments on the Pacific margin of the Antarctic Peninsula. Geophys. Res. Lett., 31, L22606, doi: 10.1029/2004GL020298. Herna´ndez-Molina, F. J., Larter, R. D., Maldonado, A., & Rodrı´ guez-Ferna´ndez, J. (2006a). Evolution of the Antarctic Peninsula Pacific margin offshore from Adelaide Island since the Late Miocene: An example of a glacial passive margin. Terra Antart. Rep., 12, 81–90. Herna´ndez-Molina, F. J., Larter, R. D., Rebesco, M., & Maldonado, A. (2006b). Miocene reversal of bottom water flow along the Pacific Margin of the Antarctic Peninsula: Stratigraphic evidence from a contourite sedimentary tail. Mar. Geol., 228, 93–116. Herron, M. J. (1988). Marine Geology and Geophysics of the Western South Orkney Plateau, Antarctica: Implications for Quaternary Glacial History Tectonics and Paleoceanography. M.A. Thesis, Rice University, Houston, TX, 256 pp.
202
A. K. Cooper et al.
Herron, M. J., & Anderson, J. B. (1990). Late Quaternary glacial history of the South Orkney Plateau, Antarctica. Quat. Res., 33, 265–275. Heroy, D. C., & Anderson, J. B. (2005). Ice-sheet extent of the Antarctic Peninsula region during the last glacial maximum (LGM) – Insights from glacial geomorphology. Geol. Soc. Am. Bull., 117, 1497–1512. Hilfinger, M., Franceschini, J., & Domack, E. W. (1995). Chronology of glacial marine lithofacies related to the recession of the West Antarctic Ice Sheet in the Ross Sea, Antarctica. Antarct. J. U.S., 30, 82–84. Hill, R. S. (1989). Fossil leaf. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. DSIR Bullettin, N.Z., pp. 143–144. Hinz, K., & Krause, W. (1982). The continental margin of Queen Maud Land/ Antarctica: Seismic sequences, structural elements and geological development. Geol. Jahrbuch, E23, 17–41. Hinz, K., & Block, M. (1984). Results of geophysical investigations in the Weddell Sea and in the Ross Sea, Antarctica. In: Proceedings of the 11th World Petroleum Congress, London 1983. Wiley, New York, Vol. 2, pp. 79–91. Hinz, K., & Kristoffersen, Y. (1987). Antarctica, Recent Advances in the Understanding of the Continental Shelf. Geol. Jahrbuch, E37, 3–54. Hollister, C. D., Craddock, C., et al. (1976). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 35, 930 pp. Hollister, C. D., & Elder, R. B. (1969). Contour Currents in the Weddell Sea. DeepSea Res., 16, 99–101. Houtz, R. E., & Davey, F. J. (1973). Seismic profiler and sonobuoy measurements in the Ross Sea, Antarctica. J. Geophys. Res., 78, 3448–3468. Houtz, R. E., & Meijer, R. (1970). Structure of the Ross Sea Shelf from profiler data. J. Geophys. Res., 75, 6592–6597. Igarashi, A., Harada, N., & Moriwaki, K. (1995). Marine fossils of 30–40 ka raised beach deposits and the late Pleistocene glacial history around Lu¨tzow-Holm Bay, East Antarctica. In: Proceedings of NIPR Symposium on Antarctic Geosciences. National Institute of Polar Research, pp. 219–229. Imbo, Y., De Batist, M., Canals, M., Prieto, M. J., & Baraza, J. (2003). The Gebra Slide: A submarine slide on the Trinity Peninsula Margin, Antarctica. Mar. Geol., 193, 235–252. Ishihara, T., Tanahashi, M., Sato, M., & Okuda, Y. (1996). Preliminary report of geophysical and geological surveys of the west Wilkes Land margin. Proc. NIPR Symp. Antarct. Geosci., 9, 91–108. Ishman, S. E., & Sperling, M. R. (2002). Benthic foraminiferal record of Holocene deep-water evolution in the Palmer Deep, western Antarctic Peninsula. Geology, 30, 435–438. Ishman, S. E., & Webb, P. N. (1988). Late Neogene foraminifera from the Victoria Land basin margin: Application to glacio-eustatic and tectonic events. Rev. Paleobiol., 2, 523–551.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
203
Ivanov, V. L. (1983). Sedimentary basins of Antarctica and their preliminary structural and morphological classification. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 539–541. Iwai, M., & Winter, D. (2002). Data report: Taxonomic notes of neogene diatoms from the western Antarctic Peninsula: Ocean Drilling Program Leg 178. In: P. F. Barker, A. Camerlenghi, G. D. Acton, A. T. S. Ramsay (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 178 [CD-ROM]. Available from Ocean Drilling Program, Texas A&M University, College Station, TX. Jabaloy, A., Balanya´, J. C., Barnolas, A., Galindo-Zaldiva´r, J., Herna´ndez-Molina, F. J., Maldonado, A., Martı´ nez-Martı´ nez, J. M., Rodrı´ guez-Ferna´ndez, J., Sanz de Gardeano, C., Somoza, L., Surin˜ach, E., & Va´squez, T. (2003). The transition from an active to a passive margin (southwestern end of the South Shetland Trench, Antarctic Peninsula). Tectonophysics, 366, 55–81. Jacobs, S. S., Amos, A. F., & Bruchhausen, P. M. (1970). Ross Sea oceanography and Antarctic bottom water formation. Deep Sea Res., 17, 935–962. Jahns, E. (1994). Evidence for a fluidized till deposit on the Ross Sea continental shelf. Antarct. J. U.S., 29, 139–141. Jeffers, J. D. (1988). Tectonics and Sedimentary Evolution of the Bransfield Basin, Antarctica. M.A. Thesis, Rice University, Houston, TX, 142 pp. Jeffers, J. D., & Anderson, J. B. (1990). Sequence stratigraphy of the Bransfield Basin, Antarctica: Implications for tectonic history and hydrocarbon potential. In: B. St. John (Ed.). Antarctica as an Exploration Frontier–Hydrocarbon Potential, Geology and Hazards. American Association of Petroleum Geologists, Tulsa, OK, Studies in Geology, Vol. 31, pp. 13–29. Jeffers, J. D., Anderson, J. B., & Lawver, L. A. (1991). Evolution of the Bransfield basin, Antarctic Peninsula. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 481–485. Jin, Y. K., & Kim, Y. (1998). A Crustal Model and its Tectonic Implication on the Evolution of the Pacific Margin of the Northern Antarctic Peninsula. Memoirs of NIPR Special Issue. National Institute of Polar Research, Tokyo, Vol. 53, pp. 203–213. Jin, Y. K., Kim, Y., Kim, H.-S., & Nam, S. H. (1996). Preliminary results of seismic survey in the Central Bransfield Strait, Antarctic Peninsula. In: Proceedings of NIPR Symposium on Antarctic Geosciences, Vol. 9. National Institute of Polar Research, Tokyo, pp. 141–149. Jin, Y. K., Larter, R. D., Kim, Y., Nam, S. H., & Kim, K. J. (2002). Postsubduction margin structures along Boyd Strait, Antarctic Peninsula. Tectonophysics, 346, 187–200. Jin, Y. K., Lee, M. W., Kim, Y., Nam, S. H., & Kim, K. J. (2003). Gas hydrate volume estimations on the South Shetland continental margin, Antarctic Peninsula. Antarct. Sci., 15, 271–282. Johnson, G. L., Vanney, J.-R., Elverhøi, A., & LaBrecque, J. L. (1981). Morphology of the Weddell Sea and southwest Indian Ocean. Dtsch. Hydrogr. Z., 34, 263–272.
204
A. K. Cooper et al.
Johnson, G. L., Vanney, J. R., & Hayes, D. (1982). The Antarctic continental shelf: review paper. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 995–1002. Kagami, H., Kuramochi, H., & Shima, Y. (1991). Submarine canyons in the Bellingshausen and Riiser-Larsen Seas around Antarctica. In: Proceedings of the NIPR Symposium on Antarctic Geosciences, No. 5, pp. 84–98. Kaharoeddin, F. A., Eggers, M. R., Graves, R. S., Goldstein, E. H., Hattner, J. G., Jones, S. C., & Ciesielski, P. F. (1979). ARA Islas Orcadas Cruise 1277 Sediment Descriptions. Sedimentology Research Laboratory Contribution, Vol. 47. Florida State University, Tallahassee, 108 pp. Kaharoeddin, F. A., Eggers, M. R., Goldstein, E. H., Graves, R. S., Watkins, D. K., Bergen, J. A., Jones, S. C., & Cassidy, D. S. (1980). ARA Islas Orcadas Cruise 1578; Sediment Descriptions. Sedimentology Research Laboratory Contribution, Vol. 48. Florida State University, Tallahassee, 162 pp. Kamenev, E. N., & Ivanov, V. L. (1983). Structure and outline of geologic history of the southern Weddell Sea Basin. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 194–196. Kaul, N. (1992). High resolution seismics and stratigraphy off Kapp Norvegia, Antarctica. Z. Geomorphol., 86, 105–112. Kavoun, M., & Vinnikovskaya, O. (1994). Seismic stratigraphy and tectonics of the northwestern Weddell Sea (Antarctica) inferred from marine geophysical surveys. Tectonophysics, 240, 299–323. Keany, J. (1978). Paleoclimatic trends in early and middle Pliocene deep-sea sediments of the Antarctic. Mar. Micropaleontol., 3, 35–49. Kellogg, D. E., & Kellogg, T. B. (1987). Microfossil Distributions in Modern Amundsen Sea Sediments. Mar. Micropaleontol., 12, 203–222. Kellogg, T. B. (1987). Glacial-interglacial changes in global deepwater circulation. Paleoceanography, 2, 259–271. Kellogg, T. B., & Kellogg, D. E. (1988). Antarctic cryogenic sediments: Biotic and inorganic facies of ice shelf and marine-based ice sheet environments. Palaeogeogr. Palaeoclimatol. Palaeoecol., 67, 51–74. Kellogg, T. B., Truesdale, R. R. S., & Osterman, L. E. (1979). Late Quaternary extent of the West Antarctic Ice Sheet: New evidence from Ross Sea cores. Geology, 7, 249–253. Kellogg, T. B., Kellogg, D. E., & Stuiver, M. (1990). Late Quaternary history of the southwestern Ross Sea: Evidence from debris bands on the McMurdo Ice Shelf, Antarctica. In: D. H. Elliot (Ed.). Contributions to Antarctic Research I. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 50, pp. 25–56. Kellogg, T. B., Hughes, T., & Kellogg, D. E. (1996). Late Pleistocene interactions of East and West Antarctic Ice Flow regimes: Evidence from the McMurdo Ice Shelf. J. Glaciol., 42(142), 486–500.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
205
Kemp, E. M., & Barrett, P. J. (1975). Antarctic glaciation and early Tertiary vegetation. Nature, 258, 507–508. Kemp, E. M., Frakes, L. A., & Hayes, D. E. (1975). Paleoclimatic significance of diachronous biogenic facies, Leg 28. In: D. E. Hayes, L. A. Frakes, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, pp. 909–917. Kennedy, D. S. (1988). Modern Sedimentary Dynamics and Quaternary Glacial History of Marguerite Bay, Antarctic Peninsula. M.A. Thesis, Rice University, Houston, TX, 203 pp. Kennedy, D. S., & Anderson, J. B. (1989). Glacial-Marine sedimentation and Quaternary glacial history of Marguerite Bay, Antarctic Peninsula. Quat. Res., 31, 255–276. Kennett, J., & Barker, P. (1990). Latest Cretaceous to Cenosoic climate and oceanographic developments in the Weddell Sea, Antarctica: An ocean-drilling perspective. In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 113, pp. 937–960. Kennett, J. P. (1966). Foraminiferal evidence of a shallow calcium carbonate solution boundary, Ross Sea, Antarctica. Science, 153, 191–193. Kennett, J. P. (1977). Cenozoic evolution of Antarctic Glaciation, the CircumAntarctic ocean, and their impact on Global paleoceanography. J. Geophys. Res., 82, 3843–3860. Kennett, J. P. (1978). Cenozoic microfossil datums in Antarctic–sub-Antarctic deepsea sedimentary sequences and the evolution of southern ocean planktonic biogeography. In: N. Ikebe (Ed.). Correlation of Tropical Through High Latitude Marine Neogene Deposits of the Pacific Basin. Stanford University Publications. Geol. Sci., 14, 30–31. Kennett, J. P., & Vella, P. (1975). Late Cenozoic planktonic foraminifera and paleoceanography at DSDP Site 284 in the cool subtropical South Pacific. In: J. P. Kennett, R. E. Houtz, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 29, pp. 769–799. Kennett, J. P., & Shackleton, N. J. (1976). Oxygen Isotopic evidence for the development of the Psychrosphere 38 Myr ago. Nature, 260, 513–515. Kennett, J. P., & Hodell, D. A. (1993). Evidence for relative climatic stability of Antarctica during the Early Pliocene: A marine perspective. Geogr. Ann., 75A, 205–220. Kim, Y., Kim, H.-S., Larter, R. D., Camerlenghi, A., Gamboˆa, L. A. P., & Rudowski, S. (1995). Tectonic deformation in the upper crust and sediments at the South Shetland Trench. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 157–166. Kimura, K. (1982). Geological and geophysical survey in the Bellingshausen Basin, off Antarctica. Antarct. Rec., 75, 12–24.
206
A. K. Cooper et al.
King, E. C., & Barker, P. F. (1988). The margins of the South Orkney microcontinent. J. Geol. Soc. London, 145, 317–331. King, E. C., & Bell, A. C. (1996). New seismic data from the Ronne Ice Shelf, Antarctica. In: B. C. Storey, E. C. King, & R. A. Livermore (Eds). Weddell Sea Tectonics and Gondwana Break-Up. Geological Society of London, London, UK, Geological Society Publication No. 108, pp. 213–226. King, E. C., Leitchenkov, G., Galindo-Zaldivar, J., Maldonado, A., & Lodolo, E. (1997). Crustal structure and sedimentation in Powell Basin. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, 2. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 71, pp. 75–93. Klepikov, V. V. (1963). Weddell sea hydrology; works of the Soviet Antarctic expedition. Hydrol. Antarct. Waters, 16, 45–93. Klepikov, V. V., & Grigor’yev, Y. A. (1966). Water circulation in the Ross Sea. Sov. Antarct. Expedition Rep., 6, 52–54. Kolobov, D. D., & Savatyugin, L. M. (1982). Bottom sediments under the Novolazarevskiy Ice Shelf. Polar Geog. Geol., 6, 267–271. Kristja´nsson, L., Gudmundsson, M. T., Smellie, J. L., McIntosh, W. C., & Esser, R. (2005). Palaeomagnetic, 40Ar/39Ar, and stratigraphical correlation of MiocenePliocene basalts in the Brandy Bay area, James Ross Island, Antarctica. Antarct. Sci., 17, 409–417. Kristoffersen, Y., Winterhalter, B., & Solheim, A. (2000). Shelf progradation on a glaciated continental margin, Queen Maud Land, Antarctica. Mar. Geol., 165, 109–122. Kuhn, G., Melles, M., Ehrmann, W. U., Hambrey, M. J., & Schmiedl, G. (1993). Character of clasts in glaciomarine sediments as an indicator of transport and depositional processes, Weddell and Lazarev seas, Antarctica. J. Sediment. Petrol., 63, 477–487. Kuhn, G., & Weber, M. E. (1993). Acoustical characterization of sediments by Parasound and 3.5 kHz Systems; related sedimentary processes on the southeastern Weddell Continental Slope, Antarctica. Mar. Geol., 113, 201–217. Kurtz, D. D., & Anderson, J. B. (1979). Recognition and sedimentologic description of recent debris flow deposits from the Ross and Weddell Seas, Antarctica. J. Sediment. Petrol., 49, 1159–1170. Kuvaas, B., & Kristoffersen, Y. (1991). The Crary Fan: A trough-mouth fan on the Weddell Sea continental margin, Antarctica. Mar. Geol., 97, 345–362. Kuvaas, B., & Leitchenkov, G. (1992). Glaciomarine turbidite and current controlled deposits in Prydz Bay, Antarctica. Mar. Geol., 108, 365–381. Kuvaas, B., Kristoffersen, Y. K., Guseva, J., Leitchenkov, G., Gandjukhin, V., Løva˚s, O., Sand, M., & Brekke, H. (2004a). Input of glaciomarine sediments along the East Antarctic continental margin: depositional processes on the Cosmonaut Sea continental slope and rise and a regional acoustic stratigraphic correlation from 401W to 801E. Mar. Geophys. Res., 25, 247–263.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
207
Kuvaas, B., Kristoffersen, Y. K., Leitchenkov, G., Guseva, J., & Gandjukhin, V. (2004b). Seismic expression of glaciomarine deposits in the eastern Riiser-Larsen Sea, Antarctica. Mar. Geol., 207, 1–15. Kuvaas, B., Kristoffersen, Y. K., Guseva, J., Leitchenkov, G., Løva˚s, O., Sand, M., & Brekke, H. (2005). Interplay of turbidite and contourite deposition along the Cosmonaut Sea/Enderby Land margin, East Antarctica. Mar. Geol., 217, 143–159. Larter, R. D. (2007). Margin architecture reveals the transition to the modern Antarctic Ice Sheet ca. 3 Ma: Comment. Geology Online Forum; doi:10.1130/ G23422C.1. Larter, R. D., & Barker, P. F. (1989). Seismic stratigraphy of the Antarctic Peninsula Pacific margin: A record of Pliocene–Pleistocene ice volume and paleoclimate. Geology, 17, 731–734. Larter, R. D., & Barker, P. F. (1991a). Effects of ridge crest–trench interaction on Antarctic–Phoenix spreading: Forces on a young subducting plate. J. Geophys. Res., 96, 19583–19607. Larter, R. D., & Barker, P. F. (1991b). Neogene interaction of tectonic and glacial processes at the Pacific margin of the Antarctic Peninsula. In: D. I. M. Macdonald (Ed.). Sedimentation, Tectonics and Eustasy. International Association of Sedimentologists, Oxford, Special Publication, Vol. 12, pp. 165–186. Larter, R. D., & Cunningham, A. P. (1993). The depositional pattern and distribution of glacial–interglacial sequences on the Antarctic Peninsula Pacific margin. Mar. Geol., 109, 203–219. Larter, R. D., & Vanneste, L. E. (1995). Relict subglacial deltas on the Antarctic Peninsula outer shelf. Geology, 23, 33–36. Larter, R. D., Rebesco, M., Vanneste, L. E., Gamboˆa, L. A. P., & Barker, P. F. (1997). Cenozoic tectonic, sedimentary and glacial history of the continental shelf west of Graham Land, Antarctic Peninsula. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, 2. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 71, pp. 1–27. Larter, R. D., Gohl, K., Hillenbrand, C.-D., Kuhn, K., Deen, T. J., Dietrich, R., Eagles, G., Johnson, J. S., Livermore, R. A., Nitsche, F. O., Pudsey, C. J., Schenke, H.-W., Smith, J. A., Udintsev, G., & Uenzelmann-Neben, G. (2007). West Antarctic Ice Sheet change since the last glacial period. Eos Trans. Am. Geophys. Union, 88, 189–190. Lawver, L., & Gahagan, L. M. (2003). Evolution of Cenozoic seaways in the circumAntarctic region. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 11–37. Lawver, L. A., Della Vedova, B., & Von Herzen, R. P. (1991a). Heat flow in Jane Basin, northwest Weddell Sea. J. Geophys. Res., 96, 2019–2038. Lawver, L. A., Royer, J.-Y., Sandwell, D. T., & Scotese, C. R. (1991b). Evolution of the Antarctic Continental Margins. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 533–539.
208
A. K. Cooper et al.
Lawver, L. A., Gahagan, L. M., & Coffin, M. F. (1992). The development of paleoseaways around Antarctica. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change 1. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 56, pp. 7–30. Lawver, L. A., Williams, T., & Sloan, B. (1994). Seismic stratigraphy and heat flow of Powell Basin. Terra Antart., 1, 309–310. Lawver, L. A., Sloan, B. J., Barker, D. H. N., Ghidella, M., Von Herzen, R. P., Keller, R. A., Kinkhammer, G. P., & Chin, C. S. (1996). Distributed, active extension in Bransfield Basin, Antarctic Peninsula: Evidence from multibeam bathymetry. GSA Today, 6(11), 1–6. Lazarus, D., Pallant, A., & Hays, J. D. (1987). Database of Antarctic Pre-Pleistocene Sediment Cores. Woods Hole Oceanographic Institution, Woods Hole, 42 pp. Lazarus, D., & Caulet, J. P. (1993). Cenozoic southern ocean reconstructions from sedimentologic, radiolarian, and other microfossil data. In: J. P. Kennettm, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change II. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 60, pp. 145–174. Leckie, R. M., & Webb, P. N. (1983). Late Oligocene–early Miocene glacial record of the Ross Sea, Antarctica: Evidence from DSDP Site 270. Geology, 11, 578–582. Leckie, R. M., & Webb, P. N. (1986). Late Paleogene and early Neogene foraminifers of deep sea drilling project site 270, Ross Sea, Antarctica. In: J. P. Kennett, C. C. von der Borch, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. Leg 90. US Government Printing Office, Washington, DC, pp. 1093–1118. Ledbetter, M. T. (1981). Palaeooceanographic significance of bottom-current fluctuations in the Southern Ocean. Nature, 294, 554–556. Ledbetter, M. T., & Ciesielski, P. F. (1986). Post-Miocene disconformities and paleoceanography in the Atlantic sector of the Southern Ocean. Palaeogeogr. Palaeoclimatol. Palaeoecol., 52, 185–214. Ledbetter, M. T., & Ellwood, B. B. (1982). Variations in particle alignment and size in sediments of the Vema channel record antarctic bottom-water velocity changes during the last 400,000 years. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 1033–1038. Ledbetter, M. T., Ciesielski, P. F., Osborn, N. I., & Allison, E. T. (1983). Bottomcurrent erosion in the Southeast Indian and Southwest Pacific Oceans during the last 5.4 million years. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 379–383. Ledford-Hoffman, P. A., DeMaster, D. J., & Nittrouer, C. A. (1986). Biogenic-Silica accumulation in the Ross Sea and the importance of Antarctic continental-shelf deposits in the Marine silica budget. Geochim. Cosmochim. Acta, 50, 2099–2110. Leitchenkov, G. L. (1991). Structure and evolution of the Prydz Bay. In: Abstracts of Sixth International Symposium on Antarctic Earth Sciences. National Institute of Polar Research, Japan, pp. 346–351.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
209
Leitchenkov, G. L., & Kudryavtzev, G. A. (2000). Structure and origin of the Earthus crust in the Weddel Sea embayment (beneath the front of the Filchner and Ronne ice shelves) from deep seismic sounding. Polarforschung, 67, 143–154. Leitchenkov, G. L., Guseva, Y. B., & Gandyukhin, V. V. (2007). Cenozoic environmental changes along the East Antarctic continental margin inferred from regional seismic stratigraphy. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th International Symposium on Antarctic Earth Sciences. USGS Open-File Report 2007-1047, Short Research Paper 055, 4, doi:10.3133/of2007-1047.srp005. LeMasurier, W. E., Melander, O., Grindley, G. W., & McIntosh, W. C. (1979). Tillite, Glacial Striae, and Hyaloclastite Associations on Hobbs Coast, Marie Byrd Land. Antarct. J. U.S., 14, 48–50. Lepley, L. K. (1966). Submarine Geomorphology of Eastern Ross Sea and Sulzberger Bay, Antarctica. Ocean Surveys Division Technical Report, No. 172. US Naval Oceanographic Office, Washington, DC, 34 pp. Leventer, A., Dunbar, R. B., & De Master, D. J. (1993). Diatom Evidence for Alte Holocene Climatic Events in Granite Harbor, Antarctica. Paleoceanography, 8, 373–386. Leventer, A., Domack, E. W., Ishman, S. E., Brachfeld, S., McClennen, C. E., & Manly, P. (1996). Productivity cycles of 200–300 years in the Antarctic Peninsula Region: Understanding linkages among the Sun, Atmosphere, Oceans, Sea Ice, and Biota. Geol. Soc. Am. Bull., 108, 1626–1644. Leventer, A., et al. (2001). Coring Holocene Antarctic Ocean Sediments. NBP1-1 Report, 190 pp. Leventer, A., Domack, E., Barkoukis, A., McAndrews, B., & Murray, J. (2002). Laminations from the Palmer Deep: A diatom-based interpretation. Paleoceanography, 17, 8002, doi:10.1029/2001PA000624. Levy, R. H., Stillwell, J. D., Harwood, D. M., Francis, J., & Pole, M. (1995). Eocene palaeobiology of East Antarctica – A report on evidence from fossiliferous glacial erratics in McMurdo Sound. In: Abstracts, Seventh International Symposium on Antarctic Earth Science, Siena, Italy, p. 243. Licht, K. J., Jennings, A. E., Andrews, J. T., & Williams, K. M. (1996). Chronology of late Wisconsin ice retreat from the western Ross Sea, Antarctica. Geology, 24, 223–226. Licht, K. J., Dunbar, N. W., Andrews, J. T., & Jennings, A. E. (1999). Distinguishing subglacial till and glacial marine diamictons in the western Ross Sea, Antarctica: Implications for last glacial maximum grounding line. Geol. Soc. Am. Bull., 111, 91–103. Lien, R. (1981). Sea bed features in the Blaaenga area, Weddell Sea, Antarctica. In: Proceedings of the Sixth International Conference on Port and Ocean Engineering under Arctic Conditions, No. 6. Universite´ Laval, Que´bec, pp. 706–716. Lisiecki, L. E., & Raymo, M. E. (2005). A Pliocene-Pleistocene stack of 57 globally distributed benthic d18O records. Palaeoceanography, 20, PA1003, doi:10.1029/ 2004PA001071.
210
A. K. Cooper et al.
Lisitsyn, A. P. (1962). Bottom sediments of the Antarctic. In: H. Wexler (Ed.). Antarctic Research. American Geophysical Union Monograph No. 7, pp. 81–88. Lisker, F., Belton, D. X., & Kroner, U. (2005). Thermochronological investigation around the Lambert Graben: Review of pre-existing data and field work during PCMEGA. Terra Antart., 12, 45–50. Livermore, R. A., Balanya´, J. C., Maldonado, A., Martı´ nez, J. M., Rodrı´ guezFerna´ndez, J., Sanz de Galdeano, C., Galindo Zaldı´ var, J., Jabaloy, A., Barnolas, A., Somoza, L., Herna´ndez-Molina, J., Surin˜ach, E., & Viseras, C. (2000). Autopsy on a dead spreading centre: The Phoenix Ridge, Drake Passage, Antarctica. Geology, 28, 607–610. Lodolo, E., Camerlenghi, A., & Brancolini, G. (1993). A bottom simulating reflector on the South Shetland margin, Antarctic Peninsula. Antarct. Sci., 5, 201–210. Lodolo, E., Camerlenghi, A., Madrussani, G., Tinivella, U., & Rossi, G. (2002). Assessment of gas hydrate and free gas distribution on the South Shetland margin (Antarctica) based on multichannel seismic reflection data. Geophys. J. Int., 148, 103–119. Lonsdale, M. J. (1990). The relationship between silica diagenesis, methane, and seismic reflections on the South Orkney Microcontinent. In: P. F. Barker, & J. P. Kennett (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 113, pp. 27–37. Lowe, A. L., & Anderson, J. B. (2002). Reconstruction of the West Antarctic Ice Sheet in Pine Island Bay during the last glacial maximum and its subsequent retreat history. Quat. Sci. Res., 21, 1879–1897. Lozano, J. A., & Hayes, J. D. (1976). Relationship of radiolarian assemblages to sediment types and physical oceanography in the Atlantic and western Indian Ocean sectors of the Antarctic Ocean. Geol. Soc. Am. Mem., 145, 303–336. Lyra, C. S. (1986). Tertiary sediment palynology at Fildes Peninsula, King George Island, South Shetland Islands, and some paleoenvironmental considerations. Anais Acad. Brasil. Ciencias, 58, 137–147. Macdonald, D. I. M., Barker, P. F., Garrett, S. W., Ineson, J. R., Pirrie, D., Storey, B. C., Whitham, A. G., Kinghorn, R. R. F., & Marshall, J. E. A. (1988). A preliminary assessment of the hydrocarbon potential of the Larsen Basin, Antarctica. Mar. Petrol. Geol., 5, 34–53. MacDonald, S. E., & Anderson, J. B. (1986). Paleoceanographic implications of terrigenous deposits on the Maurice Ewing Bank, southwest Atlantic Ocean. Mar. Geol., 71, 259–287. Macellari, C. E. (1988). Stratigraphy, sedimentology, and paleoecology of Upper Cretaceous/Paleocene shelf-deltaic sediments of Seymour Island. In: R. M. Feldman, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America, Boulder, CO, Memoir, Vol. 169, pp. 25–53. Macphail, M. K., & Truswell, E. M. (2004). Palynology of Site 1166. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean
Cenozoic Climate History from Seismic Reflection and Drilling Studies
211
Drilling Program, Scientific Results. Available from Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 188 [CD-ROM], pp. 1–43. Maldonado, A., Aldaya, F., Balanya, J. C., Galindo-Zaldivar, J., Jabaloy, A., Larter, R. D., Rodriguez-Fernandez, J., Roussanov, M., & Sanz de Galdeano, C. (1994a). Cenozoic continental margin growth patterns in the northern Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments, and Sea-Level Change. Terra Antart., 1 (2), 311–314. Maldonado, A., Larter, R. D., & Aldaya, F. (1994b). Forearc tectonic evolution of the South Shetland margin, Antarctic Peninsula. Tectonics, 13, 1345–1370. Maldonado, A., Zitellini, N., Leitchenkov, G., Balanya´, J. C., Coren, F., GalindoZaldı´ var, J., Lodolo, E., Jabaloy, A., Zanolla, C., Rodrı´ guez-Ferna´ndez, J., & Vinnikovskaya, O. (1998). Small ocean basin development along the Scotia–Antarctica plate boundary and in the northern Weddell Sea. Tectonophysics, 296, 371–402. Maldonado, A., Barnolas, A., Bohoyo, F., Escutia, C., Galindo-Zaldivar, J., Herna´ndez-Molina, J., Jabaloy, A., Lobo, F., Nelson, H., Rodrı´ guez-Ferna´ndez, J., Somoza, L., & Va´zquez, J.-T. (2005). Miocene to Recent contourite drifts development in the northern Weddell Sea, Antarctica. Global Planet. Change, 45, 99–129. Maldonado, A., Bohoyo, F., Galindo-Zaldı´ var, J., Herna´ndez-Molina, J., Jabaloy, A., Lobo, F. J., Rodrı´ guez-Ferna´ndez, J., Surin˜ach, E., & Va´squez, J. T. (2006). Ocean basins near the Scotia–Antarctic boundary: Influence of tectonics and paleoceanography on the Cenozoic deposits. Mar. Geophys. Res., 27, 83–107. Margolis, S. V., & Kennett, J. P. (1971). Cenozoic paleoglacial history of Antarctica recorded in sub-Antarctic deep-sea cores. Am. J. Sci., 271, 1–36. Mawson, D. (1940). Sedimentary rocks. Australasian Antarctic Expedition 1911-14. Scientific Rep. Ser. A, 4, 347–367. Mawson, D. (1942). Geographical Narrative and Cartography. Australian Antarctic Expedition 1911-14. Scientific Reports, Series A, Vol. 1, pp. 1–364. Mazzullo, J., & Anderson, J. B. (1987). Grain shape and surface texture analysis of till and glacial-marine sand grains from the Weddell and Ross seas, Antarctica. In: J. R. Marshall (Ed.). Clastic Particles: Scanning Electron Microscopy and Shape Analysis of Sedimentary and Volcanic Clasts. Van Nostrand Reinhold Co., New York, pp. 314–327. McCoy, F. W. (1991). Southern ocean sediments; Circum-Antarctic to 30 degrees S. In: D. E. Hayes (Ed.). Marine Geological and Geophysical Atlas of the CircumAntarctic to 30 degrees S. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 54, pp. 37–46. McGinnis, J. P., & Hayes, D. E. (1994). Sediment drift formation along the Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic
212
A. K. Cooper et al.
Records of Cenozoic Glaciation, Paleoenvironments, and Sea-Level Change. Terra Antart., 1 (2), 275–276. McGinnis, J. P., & Hayes, D. E. (1995). The roles of downslope and alongslope depositional processes: Southern Antarctic Peninsula continental rise. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 141–156. McGinnis, J. P., Hayes, D. E., & Driscoll, N. W. (1997). Sedimentary processes across the continental rise of the southern Antarctic Peninsula. Mar. Geol., 141, 91–109. McGinnis, L. D. (Ed.) (1981). Dry Valley Drilling Project. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 33. McKelvey, B. C. (1991). The Cenozoic glacial record in South Victoria Land: A geological evaluation of the McMurdo Sound drilling projects. In: R. J. Tingey (Ed.). The Geology of Antarctica. Oxford Monographs on Geology and Geophysics 17, pp. 434–454. McKelvey, B. C., & Stephenson, N. C. N. (1990). A geological reconnaissance of the Radok Lake area, Amery Oasis, Prince Charles Mountains. Antarct. Sci., 2, 53–66. McKelvey, B. C., Webb, P.-N., Harwood, D. M., & Mabin, M. C. G. (1991). The Dominion Range Sirius Group: A record of the late Pliocene-early Pleistocene Beardmore glacier. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 675–682. McLoughlin, S., & Drinnan, A. N. (1997a). Fluvial sedimentology and revised stratigraphy of the Triassic Flagstone Bench Formation, northern Prince Charles Mountains, East Antarctica. Geol. Mag., 134, 781–806. McLoughlin, S., & Drinnan, A. N. (1997b). Revised stratigraphy of the Permian Bainmedart Coal Measures, northern Prince Charles Mountains, East Antarctica. Geol. Mag., 134, 335–353. Melles, M., & Kuhn, G. (1993). Sub-bottom profiling and sedimentological studies in the southern Weddell sea, Antarctica; evidence for large-scale erosional/ depositional processes. Deep Sea Res. I, 40, 739–760. Mercer, J. H. (1981). Tertiary terrestrial deposits of the Ross Ice Shelf area, Antarctica. In: M. J. Hambrey, & W. B. Harland (Eds). Earth’s Pre-Pleistocene Glacial Records. Cambridge University Press, Cambridge, pp. 204–207. Michel, E., Crosta, X., & Shipboard Scientific Party, Les rapport de campagne a` la mer a` bord du Marion Dufresne-MD130/MD131. (2006) CADO/OISO/ AUSCAN/SOCOZA (IMAGES X). Institut Polaire Francais Paul-E´mile Victor (IPEV). Michels, K. H., Rogehnagen, J., & Kuhn, G. (2001). Recognition of contour current influence in mixed contourite-turbidite sequences of the Western Weddell Sea. Mar. Geophys. Res., 22, 465–485. Michels, K. H., Kuhn, G., Hillenbrand, C.-D., Diekmann, B., Fu¨tterer, D. K., Grobe, H., & Uenzelmann-Neben, G. (2002). The southern Weddell Sea:
Cenozoic Climate History from Seismic Reflection and Drilling Studies
213
Combined contourite-turbidite sedimentation at the southeastern margin of the Weddell Gyre. In: D. A. V. Stow, C. J. Pudsey, J. A. Howe, J. C. Fauge´res, & A. R. Viana (Eds). Deep-Water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics. Memoirs. Geological. Society, London, Vol. 22, pp. 305–323. Milam, R. W., & Anderson, J. B. (1981). Distribution and ecology of recent benthonic foraminifera of the Ade´lie-George V continental shelf and slope, Antarctica. Mar. Micropaleontol., 6, 297–325. Mildenhall, D. C. (1989). Terrestrial palinology. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole McMurdo Sound. DSIR Bulletin 245. Science Information Publishing Centre, Wellington. Miller, H., Henriet, J.-P., Kaul, N., & Moons, A. (1990). A fine-scale stratigraphy of the eastern margin of the Weddell Sea. In: U. Bleil, & J. Thiede (Eds). Geological History of the Polar Oceans: Arctic versus Antarctic. Kluwer Academic Publishers, Amsterdam, pp. 131–161. Miller, K. E., Fairbanks, G., & Mountain, G. S. (1987). Tertiary oxygen isotope synthesis, sea level history, and continental margin erosion. Paleoceanography, 2, 1–19. Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E., Sugarman, P. J., Cramer, B. S., Christie-Blick, N., & Pekar, S. F. (2005). The Phanerozoic record of sea-level change. Science, 310, 1293–1298. Miller, K. G., Wright, J. D., Katz, M. E., Browning, J. V., Cramer, B. S., Wade, B. S., & Mizintseva, S. F. (2008). A view of Antarctic Ice Sheet evolution from sea-level and deep-sea isotope changes during the Late Cretaceous-Cenozoic. In: A. K. Cooper, et al. (Eds). Antarctica: A Keystone in a Changing World. The National Academies Press, Washington, DC. Mizukoshi, T., Sunouchi, H., Saki, T., & Tanahshi, M. (1986). Preliminary report of the geological and geophysical surveys of Amery Ice Shelf, East Antarctica. Mem. Natl. Inst. Polar Res. Spec. Issue Jpn., 43, 48–62. Monechi, S., & Reale, V. (1997). Calcareous nannofossil assemblages of the CIROS1 Core (Ross Sea): Further assessments on age. Terra Antart., 4(2), Moons, A., De Batist, M., Henriet, J. P., & Miller, H. (1992). Sequence stratigraphy of the Crary Fan, southeastern Weddell Sea. In: Y. Yoshida, et al. (Eds). Recent Progress on Antarctic Earth Science. Terra Scientific Publishing Company, Toyko, pp. 613–618. Moriwaki, K. (1979). In: T. Nagata (Ed.). Submarine Topography of the Central Part of Lu¨tzow-Holm Bay and around Ongul Islands, Antarctica. Memoirs of National Institute of Polar Research, Special Issue, No. 14. National Institute of Polar Research, Tokyo, pp. 194–209. Mortlock, R. A., Charles, C. D., Froelich, P. N., Zibello, M. A., Saltzman, J., Hays, J. D., & Buckle, L. H. (1991). Evidence for lower productivity in the Antarctic Ocean during the last glaciation. Nature, 351, 220–223. Naish, T. R., Powell, R. D., & Levy, R. H. (2007). Initial science results from AND-1B, ANDRILL McMurdo Ice Shelf Project, Antarctica. Terra Antart., 14 (3), 111–328.
214
A. K. Cooper et al.
Naish, T. R., et al., (2001). Orbitally induced oscillations in the East Antarctic Ice Sheet at the Oligocene/Miocene boundary. Nature, 413, 719–723. Nayudu, Y. R. (1971). Lithology and chemistry of surface sediments in subAntarctic regions of the Pacific Ocean. In: J. L. Reid (Ed.). Antarctic Oceanology I. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 15, pp. 247–282. Nitsche, F. O., Gohl, K., Vanneste, K., & Miller, H. (1997). Seismic expression of glacially deposited sequences in the Bellingshausen and Amundsen Seas, West Antarctica. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, 2. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 71, pp. 95–108. Nitsche, F. O., Cunningham, A. P., Larter, R. D., & Gohl, K. (2000). Geometry and development of glacial continental margin depositional systems in the Bellingshausen Sea. Mar. Geol., 162, 277–302. O´ Cofaigh, C., Pudsey, C. J., Dowdeswell, J. A., & Morris, P. (2002). Evolution of subglacial bedforms along a paleo-ice stream, Antarctic Peninsula continental shelf. Geophys. Res. Lett., 29, 1199, doi:10.1029/2001GL014488. O´ Cofaigh, C., Larter, R. D., Dowdeswell, J. A., Hillenbrand, C.-D., Pudsey, C. J., Evans, J., & Morris, P. (2005). Flow of the West Antarctic Ice Sheet on the continental margin of the Bellingshausen Sea at the last glacial maximum. J. Geophys. Res., 110, B11103, doi:10.1029/2005JB003619. O’Brien, P. E., Truswell, E. M., & Burton, T. (1994). Morphology and late glacial history of Prydz Bay, Antarctica based on echo sounder data. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin: Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea Level Change. Terra Antart., 1 (2), 403–405. O’Brien, P. E., Harris, P. T., Quilty, P. G., Taylor, F., & Wells, P. (1995). Post-cruise report, Antarctic CRC marine geoscience, Prydz Bay, MacRobertson Shelf, and Kerguelen Plateau. Australian Geological Survey Record, 1995/29. O’Brien, P. E., & Harris, P. T. (1996). Patterns of glacial erosion and deposition in Prydz Bay and the past behaviour of the Lambert Glacier. Pap. Proc. R. Soc. Tas., 130, 79–86. O’Brien, P. E., & Leitchenkov, G. (1997). Deglaciation of Prydz Bay, East Antarctica, based on echo sounding and topographic features. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series Vol. 71, pp. 109–125. O’Brien, P. E., De Santis, L., Harris, P. T., Domack, E., & Quilty, P. G. (1999). Ice shelf grounding zone features of western Prydz Bay, Antarctica: Sedimentary processes from seismic and sidescan images. Antarct. Sci., 11, 78–91. O’Brien, P. E., Cooper, A. K., Richter, C., et al. (2001). Initial reports, Prydz BayCooperation Sea, Antarctica: Glacial history and paleoceanography. In: Proceedings of the Ocean Drilling Program, Scientific Results. Texas A & M University, College Station, TX, Vol. 188 [CD-ROM].
Cenozoic Climate History from Seismic Reflection and Drilling Studies
215
O’Brien, P. E., Cooper, A. K., Florindo, F., Handwerger, D. A., Lavelle, M., Passchier, S., Pospichal, J. J., Quilty, P. G., Richter, C., Theissen, K. M., & Whitehead, J. M. (2004). Prydz Channel Fan and the history of extreme ice advances in Prydz Bay. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. Available from Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 188 [CD-ROM], pp. 1–32. O’Brien, P. E., Stanley, S., & Parums, R. (2006). Post-rift continental slope and rise sediments from 381E to 1641E, East Antarctica. In: D. K. Fu¨tterer, D. Damaske, G. Kleinschmidt, H. C. H. Miller, & F. Tessensohn (Eds). Antarctica; Contributions to Global Earth Sciences, Proceedings of the IX International Symposium on Antarctic Earth Sciences, pp. 341–348. O’Brien, P. E., & Stagg, H. M. J. (2007). Tectonic elements of the continental margin of East Antarctica, 381–1641. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th International Symposium on Antarctic Earth Sciences. USGS Open-File Report 2007-1047, Short Research Paper 085, 4, doi:10.3133/of2007-1047.srp085. Orsi, A. H., Nowlin, W. D. Jr., & Whitworth, T., III. (1993). On the circulation and stratification of the Weddell Gyre. Deep-Sea Res. 1, 40, 169–203. Park, B.-K., Lee, M.-S., Yoon, H.-I., & Nam, S.-H. (1989). Marine geology and petrochemistry in the Maxwell Bay area, South Shetland Islands. In: H. T. Huh, B.-K. Park, & S.-H. Lee (Eds). Antarctic Science: Geology and Biology. Seo Deung Co. Ltd., Seoul, Korea, pp. 85–120. Passchier, S., O’Brien, P. E., Damuth, J. E., Janusczak, N., Handwerger, D. A., & Whitehead, J. M. (2003). Pliocene-Pleistocene glaciomarine sedimentation in eastern Prydz Bay and development of the Prydz trough-mouth fan, ODP Sites 1166 and 1167, East Antarctica. Mar. Geol., 199, 279–305. Payne, R. F., Conolly, J. R., & Abbott, W. H. (1972). Turbidite muds within diatom ooze off Antarctica: Pleistocene sediment variaton defined by closely spaced piston cores. Geol. Soc. Am. Bull., 83, 481–486. Payne, R. R., & Conolly, J. R. (1972). Turbidite sedimentation off the Antarctic continent. Antarct. Res. Ser., 19, 349–364. Pezzetti, T. F., & Krissek, L. A. (1986). Re-evaluation of the Eocene La Meseta Formation of Seymour Island, Antarctic Peninsula. Antarct. J. U.S., 21, 75. Pickard, J., Adamson, D. A., Harwood, D. M., Miller, G. H., Quilty, P. G., & Dell, R. K. (1988). Early Pliocene marine sediments, coastline, and climate of East Antarctica. Geology, 16, 158–161. Piper, D. J. W., & Brisco, C. D. (1975). Deep-water continental margin sedimentation, DSDP LEG 28, Antarctica. In: D. E. Hayes, L. A. Frakes, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, pp. 727–755. Piper, D. J. W., Swint, T. R., Sullivan, L. G., & McCoy, F. W. (1985). Manganese Nodules, Seafloor Sediment, and Sedimentation Rates of the Circum-Pacific
216
A. K. Cooper et al.
Region. Circum-Pacific Council for Energy and Mineral Resources. American Association of Petroleum Geologists, Tulsa, OK [Chart]. Pirrie, D., Whitham, A. G., & Ineson, J. R. (1991). The role of tectonics and eustasy in the evolution of a marginal basin: Cretaceous-Tertiary Larsen Basin, Antarctica. In: D. I. M. MacDonald (Ed.). Sedimentation, Tectonics, and Eustasy: Sea Level Changes at Active Margins. International Association of Sedimentologists, Special Publication, Oxford, Vol. 12, pp. 293–305. Pirrie, D., Crame, J. A., Riding, J. B., Butcher, A. R., & Taylor, P. D. (1997). Miocene glaciomarine sedimentation the northern Antarctic Peninsula region: The stratigraphy and sedimentology of the Hobbs Glacier Formation, James Ross Island. Geol. Mag., 134, 745–762. Pope, P. G., & Anderson, J. B. (1992). Late Quaternary glacial history of the northern Antarctic Peninsula’s western continental shelf: Evidence from the marine record. In: D. H. Elliot (Ed.). Contributions to Antarctic Research III. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 57, pp. 63–91. Porebski, S. J., Meischner, D., & Go¨rlich, K. (1991). Quaternary mud turbidites from the South Shetland trench (West Antarctica): Recognition and implications for turbidite facies models. Sedimentology, 38, 691–715. Pospichal, J. J. (2004). Calcareous nannofossils from continental rise Site 1165, ODP Leg 188, Prydz Bay, Antarctica. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. Available from Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 188 [CD-ROM], pp. 1–14. Powell, R. D. (1981). Sedimentation conditions in Taylor valley, Antarctica, inferred from textural analysis of DVDP cores. Antarct. Res. Ser., 33, 331–350. Prieto, M. J., Gra´cia, E., Canals, M., Ercilla, G., & de Batist, M. (1997). Sedimentary history of the central Bransfield Basin (NW Antarctic Peninsula). In: C. A. Ricci (Ed.). The Antarctic Region. Geological Evolution and Processes, Special Publication, Issue VII. ISAES. Prieto, M. J., Canals, M., Ercilla, G., & De Batist, M. (1998). Structure and geodynamic evolution of the Central Bransfield Basin (NW Antarctica) from seismic reflection data. Mar. Geol., 149, 17–38. Prieto, M. J., Ercilla, G., Canals, M., & De Batist, M. (1999). Seismic stratigraphy of the Central Bransfield Basin (NW Antarctic Peninsula): interpretation of deposits and sedimentary processes in a glacio-marine environment. Mar. Geol., 157, 47–68. Pudsey, C. J. (1992). Late Quaternary changes in Antarctic bottom water velocity inferred from sediment grain size in the northern Weddell Sea. Mar. Geol., 107, 9–33. Pudsey, C. J. (2002). The Weddell Sea: Contourites and hemipelagites at the northern margin of the Weddell Gyre. In: D. A. V. Stow, C. J. Pudsey, J. A. Howe, J.-C. Fauge`res, & A. R. Viana (Eds). Deep-Water Contourite
Cenozoic Climate History from Seismic Reflection and Drilling Studies
217
Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics. Memoirs. Geological Society, London, Vol. 22, pp. 289–303. Pudsey, C. J., & Evans, J. (2001). First survey of Antarctic sub-ice shelf sediments reveals mid-Holocene ice shelf retreat. Geology, 29, 787–790. Pudsey, C. J., Barker, P. F., & Hamilton, N. (1988). Weddell Sea abyssal sediments; a record of Antarctic bottom water flow. Mar. Geol., 81, 289–314. Pudsey, C. J., Barker, P. F., & Larter, R. D. (1994). Ice sheet retreat from the Antarctic Peninsula shelf. Cont. Shelf Res., 14, 1647–1675. Pudsey, C. J., Evans, J., Domack, E. W., Morris, P., & Del Valle, R. (2001). Bathymetry and acoustic facies beneath the former Larsen-A and Prince Gustav Ice Shelves, NW Weddell Sea. Antarct. Sci., 13, 312–322. Quilty, P. G. (1991). The geology of Marine Plain, Vestfold Hills, East Antarctica. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 683–686. Rathburn, A. E., Pichon, J. J., Ayress, M. A., & Deckker, P. (1997). Microfossil and stable–isotope evidence for changes in Late Holocene paleo productivity and paleoceanographic conditions the Prydz Bay region of Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 131, 485–510. Rebesco, M., Larter, R. D., Barker, P. F., Camerlenghi, A., & Vanneste, L. E. (1994). A history of sedimentation on the continental rise west of the Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, P. N. Webb, & G. Brancolini (Eds). The Antarctic Continental Margin. Geophysical and Geological Stratigraphic Records of Cenozoic Glaciation, Paleoenvironments and Sea Level Change. Terra Antart., 1 (2), 277–280. Rebesco, M., Larter, R. D., Camerlenghi, A., & Barker, P. F. (1996). Giant sediment drifts on the continental rise west of the Antarctic Peninsula. Geo Mar. Lett., 16, 65–75. Rebesco, M., Larter, R. D., Barker, P. F., Camerlenghi, A., & Vanneste, L. E. (1997). The history of sedimentation on the continental rise west of the Antarctic Peninsula. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, 2. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 71, pp. 29–49. Rebesco, M., Camerlenghi, A., & Zanolla, C. (1998). Bathymetry and morphogenesis of the continental margin west of the Antarctic Peninsula. Terra Antart., 5, 715–725. Rebesco, M., Pudsey, C., Canals, M., Camerlenghi, A., Barker, P., Estrada, F., & Giorgetti, A. (2002). Sediment drift and deep-sea channel systems, Antarctic Peninsula Pacific margin. In: D. A. V. Stow, C. J. Pudsey, J. A. Howe, J. C. Faugeres, & A. R. Viana (Eds). Deep-Water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics. Memoirs. Geological Society, London, Vol. 22, pp. 353–371. Rebesco, M., Camerlenghi, A., Geletti, R., & Canals, M. (2006). Margin architecture reveals the transition to the modern Antarctic Ice Sheet ca. 3 Ma. Geology, 34, 301–304.
218
A. K. Cooper et al.
Rex, R. W., Margolis, S. V., & Murray, B. (1970). Possible interglacial dune sands from 300 meters water depth in the Weddell Sea, Antarctica. Geol. Soc. Am. Bull., 81, 3465–3472. Robert, C., & Kennett, J. P. (1994). Antarctic subtropical humid episode at the Paleocene-Eocene boundary; clay-mineral evidence. Geology, 22, 211–214. Robert, C. M., & Maillot, H. (1990). Paleoenvironments in the Weddell Sea area and Antarctic climates, as deducted from clay mineral associations and geochemical data, ODP Leg 113. In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. College Station, TX, Vol. 113, pp. 51–70. Robin, G. D. Q. (1988). The Antarctic Ice Sheet, its history and response to sea level and climactic changes over the past 100 million years. Palaeogeogr. Palaeoclimatol. Palaeoecol., 67, 31–50. Robinson, P. H., Pyne, A. R., Hambrey, M. J., Hall, K. J., & Barrett, P. J. (1987). Core log, Photographs and Grain Size Analyses from the CIROS-1 Drillhole, Western McMurdo Sound, Antarctica. Antarctic Data Series, Vol. 14, Antarctic Research Centre Publication, Research School of Earth Sciences, Victoria University of Wellington, Wellington, New Zealand, 241 pp. Rogenhagen, J., & Jokat, W. (2000). The sedimentary structure in the western Weddell Sea. Mar. Geol., 168, 45–60. Rogenhagen, J., Jokat, W., Hinz, K., & Kristoffersen, Y. (2004). Improved seismic stratigraphy of the Mesozoic Weddell Sea. Mar. Geophys. Res., 25, 265–282. Rooney, S. T., Blankenship, D. D., Alley, R. B., & Bentley, C. R. (1987). Till beneath ice stream B, 2. Structure and continuity. J. Geophys. Res., 92, 8913–8920. Roqueplo-Brouillet, C. (1982). Seismic Stratigraphy of the Eastern Continental Shelf of the Weddell Sea, Antarctica. M.A. Thesis, Rice University, Houston, TX, 132 pp. Rutford, R. H., Craddock, C., & Bastien, T. W. (1968). Late Tertiary glaciation and sea-level changes in Antarctica. Palaeogeogr. Paleoclimatol. Paleoecol., 5, 15–39. Sadler, I. (1988). Geometry and stratification of uppermost Cretaceous and Paleogene units on Seymour Island, northern Antarctic Peninsula. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America, Boulder, CO, Memoir, Vol. 169, pp. 303–320. Sato, S., Asakura, N., Saki, T., Oikawa, N., & Kaneda, Y. (1984). Preliminary results of geological and geophysical survey in the Ross Sea and in Dumont D’urville Sea, off Antarctica. Mem. Natl. Inst. Polar Res. Spec. Issue Jpn., 33, 66–92. Savage, M. L., & Ciesielsky, P. F. (1983). Revised history of glacial sedimentation in the Ross Sea region. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 555–559. Sayers, J., Symonds, P. A., Direen, N. G., & Bernardel, G. (2001). Nature of the continent-ocean transition on the non-volcanic rifted margin of the central Great Australian Bight. In: R. C. L. Wilson, R. B. Whitmarsh, & N. Froitzheim (Eds). Non-volcanic rifting of continental margins: A comparison of evidence from land and sea. Geological Society of London, Special Publication 187, pp. 51–76.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
219
Schenke, H. W., Dijkstra, S., Niederjasper, F., Scho¨ne, T., Hinze, H., & Hoppmann, B. (1998). In: S. Jacobs, & R. Weiss (Eds). The New Bathymetric Charts of the Weddell Sea: AWI BCWS. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 75. Scheuer, C., et al., (2006a). Gridded isopach maps from the South Pacific and their use in interpreting the sedimentation history of the West Antarctic continental margin. Geochem. Geophys. Geosyst., 7, Q11015, doi:10.1029/2006GC001315. Scheuer, C., Gohl, K., Larter, R. D., Rebesco, M., & Udintsev, G. (2006b). Variability in Cenozoic sedimentation along the continental rise of the Bellingshausen Sea, West Antarctica. Mar. Geol., 227, 279–298. Scherer, R. (1991). Quaternary and Tertiary microfossils from beneath Ice Stream B: Evidence for a dynamic West Antarctic Ice Sheet. Palaeogeogr. Palaeoclimatol. Palaeoecol., 90, 395–412. Scherer, R. P., Aldahan, A., Tulaczyk, S., Possnert, G., Engelhardt, H., & Kamb, B. (1998). Pleistocene collapse of the West Antarctic Ice Sheet. Science, 281, 82–85. Scott, D. P. D. (1981). Cartographie du fond des oceans: La carte generale bathymetrique des oceans (GEBCO) (Cartography of ocean floors; the general bathymetric map of the oceans (GEBCO). Canadian Hydrographic Service. Nat. Res., 17, 6–10. Scott, K. M. (1965). Geology of the southern Gerlache Strait region, Antarctica. J. Geol., 73, 518–527. Seabrook, J. M., Hufford, G. L., & Elder, R. B. (1971). Formation of Antarctic bottom water in the Weddell Sea. J. Geophys. Res., 76, 2164–2178. Shabtaie, S., & Bentley, C. R. (1982). Tabular icebergs: Implications from geophysical studies of ice shelves. J. Glaciol., 28, 413–430. Shackleton, N. J., & Kennett, J. P. (1975). Paleotemperature History of the Cenozoic and the Initiation of Antarctic Glaciation: Oxygen and Carbon Isotope Analysis in DSDP Sites, 277, 279, and 281. Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 29, pp. 743–755. Shevenell, A. E., & Kennett, J. P. (2002). Antarctic Holocene climate change: A benthic foraminiferal stable isotope record from Palmer Deep. Paleoceanography, 17(2), 1019, doi:10.1029/2000PA000596. Shevenell, A. E., Domack, E. W., & Kerman, G. M. (1996a). Record of Holocene paleoclimate change along the Antarctic Peninsula: Evidence from glacial marine sediments, Lallemand Fjord. Pap. Proc. R. Soc. Tas., 130, 55–64. Shevenell, A. E., Lopiccolo, M. H., Straten, B. T., & Domack, E. W. (1996b). Holocene Paleoenvironmental Studies within Antarctic Fjords along the Western Side of the Antarctic Peninsula; Understanding Hypsithermal and Neoglacial Fluctuations, Abstracts of Programs. Geological Society of America, Vol. 28, p. 99. Shipboard Scientific Party (1988). Site 694. In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Results, Vol. 13, pp. 449–526.
220
A. K. Cooper et al.
Shipboard Scientific Party (1989). Site 740. In: J. Barron, B. Larson, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Results, Vol. 119, pp. 345–376. Shipboard Scientific Party (1999). Palmer Deep (Sites 1098 and 1099). In: P. F. Barker, A. Camerlenghi, G. D. Acton, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports. Available from Ocean Drilling Program, Texas A&M University, College Station, TX, Vol. 178 [CD-ROM]. Shipboard Scientific Party (2001a). Site 1166. In: P. E. O’Brien, A. K. Cooper, C. Richter, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports. Ocean Drilling Program, College Station, TX, Vol. 188 [CD-ROM], pp. 1–110. Shipboard Scientific Party (2001b). Site 1167. In: P. E. O’Brien, A. K. Cooper, C. Richter, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports. Ocean Drilling Program, College Station TX, Vol. 188 [CD-ROM], pp. 1–97. Shipboard Scientific Party (2001c). Site 1165. In: P. E. O’Brien, A. K. Cooper, C. Richter, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports. Ocean Drilling Program, College Station TX, Vol. 188, [CD-ROM], pp. 1–191. Shipboard Scientific Party (2005). SHALDRIL 2005 Cruise Report. http:// shaldril.rice.edu/PDFs/total.pdf, 189 pp. Shipboard Scientific Party (2006). SHALDRIL II 2006 NBP0602A Cruise Report. http://shaldril.rice.edu/PDFs/NBP0602A.pdf, 369 pp. Shipp, S., & Anderson, J. B. (1994). High-resolution seismic survey of the Ross Sea continental shelf: Implications for ice-sheet retreat behavior. Antarct. J. U.S., 29, 137–138. Shipp, S., & Anderson, J. B. (1997). Lineations of the Ross Sea continental shelf, Antarctica. In: T. A. Davies, H. Josenhans, L. Polyak, A. Solheim, M. S. Stoker, & J. A. Stravers (Eds). Glaciated Continental Margins, An Atlas of Acoustic Images. Chapman and Hall, London, pp. 54–55. Shipp, S., & Anderson, J. B. (1998a). Grounding zone wedges on the Antarctic continental shelf, Ross Sea. In: T. A. Davies, H. Josenhans, L. Polyak, A. Solheim, M. S. Stoker, & J. A. Stravers (Eds). Glaciated Continental Margins, An Atlas of Acoustic Images. Chapman and Hall, London, pp. 104–105. Shipp, S., & Anderson, J. B. (1998b). Drumlin field on the Ross Sea continental shelf, Antarctica. In: T. A. Davies, H. Josenhans, L. Polyak, A. Solheim, M. S. Stoker, & J. A. Stravers (Eds). Glaciated Continental Margins, An Atlas of Acoustic Images. Chapman and Hall, London, pp. 52–53. Shipp, S., & Anderson, J. B. (1998c). Paleo-ice streams and ice stream boundaries, Ross Sea, Antarctica. In: T. A. Davies, H. Josenhans, L. Polyak, A. Solheim, M. S. Stoker, & J. A. Stravers (Eds). Glaciated Continental Margins: An Atlas of Acoustic Images. Chapman and Hall, London, pp. 106–109. Shipp, S., & Anderson, J. B. (1998d). Till sheets on the Ross Sea continental shelf, Antarctica. In: T. A. Davies, H. Josenhans, L. Polyak, A. Solheim, M. S. Stoker, &
Cenozoic Climate History from Seismic Reflection and Drilling Studies
221
J. A. Stravers (Eds). Glaciated Continental Margins: An Atlas of Acoustic Images. Chapman and Hall, London, pp. 235–237. Shipp, S., Anderson, J., De Santis, L., Bartek, L. R., Alonso, B., & Zayatz, I. (1994). High to intermediate resolution seismic stratigraphic analysis of mid-lateMiocene to Pleistocene strata in eastern Ross Sea: Implications for changing glacial/climatic regime. Terra Antart., 1(2), 381–384. Shipp, S., Anderson, J. B., & Domack, E. (1999). Late Pleistocene–Holocene retreat of the West Antarctic Ice Sheet system in the Ross Sea: Part 1: Geophysical results. GSA Bull., 111(10), 1486–1516. Singer, J. K. (1987). Terrigenous, Biogenic, and Volcaniclastic Sedimentation Patterns of the Bransfield Strait and Bays of the Northern Antarctic Peninsula: Implications for the Quaternary Glacial History. Ph.D. Dissertation, Rice University, Houston, TX, 342 pp. Sloan, B. J., Lawver, L. A., & Anderson, J. B. (1995). Seismic stratigraphy of the Larsen Basin, eastern Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 59–74. Smellie, J. L. (1999). Lithostratigraphy of Miocene–Recent, alkaline volcanic fields in the Antarctic Peninsula and eastern Ellsworth Land. Antarct. Sci., 11, 362–378. Smith, M. J. (1985). Marine Geology of the Northwestern Weddell Sea and Adjacent Coastal Fjords and Bays: Implications for Glacial History. M.A. Thesis, Rice University, Houston, TX, 157 pp. Smith, W. H. F., & Sandwell, D. T. (1997). Global sea floor topography from satellite altimetry and ship depth soundings. Science, 277, 1956–1962. Solheim, A., Forsberg, C. F., Pittenger, A. (1991). Stepwise consolidation of glaciogenic sediments related to the glacial history of Prydz Bay. In: J. Barron, B. Larson, et al., Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 119, pp. 169–184. Solli, K., Kuvaas, B., Kristoffersen, Y. K., Leitchenkov, G., Guseva, J., & Gandjukhin, V. (2007a). Seismic morphology and distribution of inferred glaciomarine deposits along the East Antarctic continental margin, 201E–601E. Mar. Geol., 237, 207–223. Solli, K., Kuvaas, B., Kristoffersen, Y. K., Leitchenkov, G., Guseva, J., & Gandjukhin, V. (2007b). A seismo-stratigraphic analysis of glaciomarine deposits in the eastern Riiser-Larsen Sea (Antarctica). Mar. Geophys. Res., 28, 43–57. Solli, K., Kuvaas, B., Kristoffersen, Y. K., Leitchenkov, G., Guseva, J., & Gandjukhin, V. (2007c). The Cosmonaut Sea Wedge. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES. USGS Open-File Report 2007-1047, Short Research Paper 009, 4, doi:10.3133/of2007-1047.srp009. Stagg, H. M. J. (1985). Structure and origin of Prydz Bay and MacRobertson Shelf, East Antarctica. Tectonophysics, 114, 315–340.
222
A. K. Cooper et al.
Stagg, H. M. J., Ramsay, D. C., & Whitworth, R. (1983). Preliminary report of a marine geophysical survey between Davis and Mawson Stations, 1982. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 527–532. Stagg, H. M. J., Colwell, J. B., Direen, N. G., O’Brien, P. E., Bernardel, G., Borissova, I., Brown, B. J., Ishihara, T. (2004a). Geology of the continental margin of Enderby and Mac. Robertson Lands, East Antarctica: insights from a new dataset. In: W. Jokat (Ed.). East-west Antarctic tectonics and Gondwana break-up 60W to 60E. Mar. Geophys. Res., 25, 183–218. Stagg, H. M. J., Colwell, J. B., Direen, N. G., O’Brien, P. E., Brown, B. J., Bernardel, G., Borissova, I., Carson, L., & Close, D. B. (2004b). Geological Framework of the Continental Margin in the Region of the Australian Antarctic Territory. Geoscience Australia Record, 2004/25. Stetson, H. C., & Upson, J. E. (1937). Bottom deposits of the Ross Sea [Antarctic]. J. Sediment. Petrol., 7, 55–66. Stillwell, J. D., & Feldmann, R. M. (Eds). (2000). Paleobiology and Paleoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 76, pp. 161–181. Strand, K., Passchier, S., & Na¨si, J. (2003). Implications of quartz grain microtextures for onset of Eocene/Oligocene glaciation in Prydz Bay, ODP Site 1166, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 101–112. Stroeven, A. P., Prentice, M. L., & Kleman, J. (1996). On microfossil transport and pathways in Antarctica during the Neogene: Evidence from the Sirius Group at Mount Fleming. Geology, 24, 727–730. Sugden, D. E., & Clapperton, C. M. (1977). The maximum ice sheet extent on island groups in the Scotia Sea, Antarctica. Quat. Res., 7, 268–282. Sugden, D. E., & Clapperton, C. M. (1981). An ice-shelf moraine, George VI Sound, Antarctica. Ann. Glaciol., 2, 135–141. Sugden, D. E., Marchant, D. R., & Denton, G. H. (1993). The case for a stable East Antarctic Ice Sheet; the background. Geogr. Ann., 75A, 151–155. Tanahashi, M., Eittreim, S., & Wanneson, J. (1994). Seismic stratigraphic sequences of the Wilkes Land margin. Terra Antart., 1(2), 391–393. Tanahashi, M., Ishihara, T., Yuasa, M., Murakami, F., & Nishimura, A. (1997). Preliminary report of the TH95 geological and geophysical survey results in the Ross Sea and the Dumont d’Urville Sea. Proc. NIPR Symp. Antarct. Geosci., 10, 36–58. Tanahashi, M., Saki, T., Oikawa, N., & Sato, S. (1987). An interpretation of the multichannel seismic reflection profiles across the continental margin of the Dumont d’Urville Sea, off Wilkes Land, East Antarctica. In: S. L. Eittreim, & M. A. Hampton (Eds). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, CPCEMR Earth Science Series, Vol. 5A, pp. 1–14.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
223
Tanner, P. W. G., & Macdonald, D. I. M. (1982). Models for the deposition and simple shear deformation of a turbidite sequence in the South Georgia portion of the southern Andes back-arc basin. J. Geol. Soc. London, 139, 739–754. Taviani, M., Reid, D. E., & Anderson, J. B. (1993). Skeletal and isotopic composition and paleoclimatic significance of Late Pleistocene carbonates, Ross Sea, Antarctica. J. Sediment. Petrol., 63, 84–90. ten Brink, U. S., & Cooper, A. K. (1992). Modeling the bathymetry of the Antarctic continental shelf. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing Co, Tokyo, Japan, pp. 763–771. ten Brink, U. S., Schneider, C., & Johnson, A. H. (1995). Morphology and stratal geometry of the Antarctic Continental Shelf: Insights from models. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 1–24. Theissen, K. M., Dunbar, R. B., Cooper, A. K., Mucciarone, D. A., & Hoffman, D. (2003). The Pleistocene evolution of the East Antarctic Ice Sheet in the Prydz Bay area: stable isotope evidence from ODP Site 1167. Global Planet. Change, 39, 227–256. Thomson, M. R. A. (1981). Late Mesozoic stratigraphy and invertebrate palaeontology of the South Orkney Islands. Br. Antarct. Surv. Bull., 54, 65–83. Thomas, M. A., & Anderson, J. B. (1994). Sea-level controls on the facies architecture of the Trinity/Sabine incised-valley system, Texas continental shelf. In: R. W. Dalrymple, R. Boyd, & B. A. Zaitlin (Eds). Incised-Valley Systems: Origin and Sedimentary Sequences. Society of Economic Petrologists and Mineralogists, Special Publication No. 51, pp. 63–82. Thomas, R. H., & Bentley, C. R. (1978). A model for Holocene retreat of the West Antarctic Ice Sheet. Quat. Res., 2, 150–170. Tinivella, U., Lodolo, E., Camerlenghi, A., & Boehm, G. (1998). Seismic tomography study of a bottom simulating reflector off the South Shetland Islands (Antarctica). In: J.-P. Henriet, & J. Mienert (Eds). Gas Hydrates: Relevance to World Margin Stability and Climate Change. Geological Society, London, Special Publication, Vol. 137, pp. 141–151. Tinivella, U., Accaino, F., & Camerlenghi, A. (2002). Gas hydrate and free gas distribution from inversion of seismic data on the South Shetland margin (Antarctica). Mar. Geophys. Res., 23, 109–123. Tomlinson, J. S., Pudsey, C. J., Livermore, R. A., Larter, R. D., & Barker, P. F. (1992). Long-range sidescan sonar (GLORIA) survey of the Antarctic Peninsula Pacific margin. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing Co., Tokyo, Japan, pp. 423–429. Traube, V., & Rybnikov, A. (1990). Preliminary Results of the First Marine Geophysical Investigations in the North-Western Weddell Sea. International
224
A. K. Cooper et al.
Workshop on Antarctic Offshore Seismic Stratigraphy (ANTOSTRAT). Overview and Extended Abstracts, pp. 278–280. Troedson, A. L., & Smellie, J. L. (2002). The Polonez Cove formation of King George Island, Antarctica: Stratigraphy, facies and implications for midCenozoic cryosphere development. Sedimentology, 49, 277–301. Truswell, E. M. (1982). Palynology of seafloor samples collected by the 1911–14 Australiasian Antarctic Expedition: Implications for the geology of coastal East Antarctica. J. Geol. Soc. Aust., 29, 343–356. Truswell, E. M. (1983). Geological implications of recycled palynomorphs in continental shelf sediments around Antarctica. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Cambridge University Press, New York, pp. 394–399. Truswell, E. M. (1991). Data report: Palynology of sediments from Leg 119 drill sites in Prydz Bay, Antarctica. In: J. Barron, B. Larson, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 119, pp. 941–948. Truswell, E. M., & Anderson, J. B. (1984). Recycled palynomorphs and the age of sedimentary sequences in the eastern Weddell Sea. Antarct. J. U.S., 19, 90–92. Truswell, E. M., & Drewry, D. J. (1984). Distribution and provenance of recycled palynomorphs in surficial sediments of the Ross Sea, Antarctica. Mar. Geol., 59, 187–214. Tsumuraya, Y., Tanahashi, M., Saki, T., Machihara, T., & Asakura, N. (1985). Preliminary report of the marine geophysical and geological surveys off Wilkes Land, Antarctica, in 1983–1984. Mem. Natl. Inst. Polar Res. Spec. Issue, 37, 48–62. Tucholke, B. E. (1977). Sedimentation processes and acoustic startigraphy in the Bellingshausen basin. Mar. Geol., 25, 209–230. Tucholke, B. E., & Houtz, R. E. (1976). Sedimentary framework of the Bellingshausen Basin from seismic profile data. In: C. D. Hollister, C. Craddock, et al. (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 35, pp. 197–227. Tucholke, B. E., Hollister, C. D., Weaver, F. M., & Vennum, W. R. (1976). Continental rise and abyssal plain sedimentation in the southeast Pacific Basin – Leg 35 Deep Sea Drilling Project. In: C. D. Hollister, & C. Craddock (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 35, pp. 359–400. Turner, B. R. (1991). Depositional environment and petrography of preglacial continental sediments from Hole 740A, Prydz Bay, Antarctica. In: J. Barron, B. Larson, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 119, pp. 45–56. Uenzelmann-Neben, G. (2006). Depositional patterns at Drift 7, Antarctic Peninsula: Along-slope versus down-slope sediment transport as indicators for oceanic currents and climatic conditions. Mar. Geol., 233, 49–62. Uenzelmann-Neben, G., Gohl, K., Larter, R. D., & Schlu¨ter, P. (2007). Differences in Retreat Rates across Pine Island Bay, West Antarctica, since the Last Glacial
Cenozoic Climate History from Seismic Reflection and Drilling Studies
225
Maximum: Indications from Multichannel Seismic Data. US Geological Survey Open File Report, OF-2007-1047, doi:10.3133/of2007-1047.srp084. Vanneste, L. E., & Larter, R. D. (1995). Deep-tow boomer survey on the Antarctic Peninsula Pacific margin: An investigation of the morphology and acoustic characteristics of late Quaternary sedimentary deposits on the outer continental shelf and upper slope. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. American Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 68, pp. 97–121. Vanney, J. R., & Johnson, G. L. (1976). Geomorphology of the Pacific Continental margin of the Antarctic Peninsula. In: C. D. Hollister, & C. Craddock (Eds). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 35, pp. 279–289. Vanney, J. R., & Johnson, G. L. (1979). Wilkes Land continental margin physiography, East Antarctica. Polarforschung, 49, 20–29. Vanney, J. R., Falconer, R. K. H., & Johnson, G. L. (1981). Geomorphology of the Ross Sea and adjacent oceanic provinces. Mar. Geol., 41, 73–102. Veevers, J. J. (1987). The conjugate continental margin of Antarctica and Australia. In: S. L. Eittreim, & M. A. Hampton (Eds). The Antarctic Continental Margin: Geology and Geophysics of Offshore Wilkes Land. Circum-Pacific Council for Energy and Mineral Resources, Houston, TX, Earth Sciences Series 5A, pp. 45–74. Viseras, C., & Maldonado, A. (1999). Facies architecture, seismic stratigraphyand development of a high-latitude basin: The Powell Basin (Antarctica). Mar. Geol., 157, 69–87. Volpi, V., Camerlenghi, A., Hillenbrand, C.-D., Rebesco, M., & Ivaldi, R. (2003). The effects of biogenic silica on sediment compaction and slope stability on the Pacific margin of the Antarctic Peninsula. Basin Res., 15, 339–363. Wannesson, J. (1990). Geology and petroleum potential of the Adelie Coast margin, East Antarctica. In: B. St. John (Ed.). Antarctica as an Exploration Frontier. Am. Assoc. Petrol. Geol. Stud. Geol., 31, 77–88. Wannesson, J., Pelras, M., Petitperrin, B., Perret, M., & Segoufin, J. (1985). A geophysical transect of the Adelie Margin, East Antarctica. Mar. Petrol. Geol., 2, 192–201. Wardell, N., Childs, J. R., & Cooper, A. K. (2007). Advances through collaboration: Sharing seismic reflection data via the Antarctic Seismic Data Library System for Cooperative Research (SDLS). In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES. USGS Open-File Report 2007–1047, SRP 001, 4, doi:10.3133/of20071047.srp001. Warnke, D. A., Marzo, B., & Hodell, D. A. (1996). Major deglaciation of East Antarctica during the early late pliocene? Not likely from a marine perspective. In: R. Z. Poore, & L. C. Sloan (Eds). Climates and Climate Variability of the Pliocene. Mar. Micropaleontol., 27, 237–251.
226
A. K. Cooper et al.
Watkins, N. D., & Kennett, J. P. (1972). Regional sedimentary disconformities and Upper Cenozoic changes in bottom water velocities between Australasia and Antarctica. In: D. E. Hayes (Ed.). Antarctic Oceanology II: The Australian– New Zealand Sector. American Geophysical Union, Washington, DC, pp. 273–294. Watkins, N. D., Keany, J., Ledbetter, M. T., & Huang, T. (1974). Antarctic glacial history from analyses of ice-rafted deposits in marine sediments: New model and initial tests. Science, 186, 533–536. Watkins, N. D., & Kennett, J. P. (1977). Erosion of deep-sea sediments in the Southern Ocean between longitudes 70E and 190E and contrasts in manganese nodule development. Mar. Geol., 23, 103–111. Watkins, N. D., Ledbetter, M. T., & Huang, T. C. (1982). Antarctic glacial history using spatial and temporal variations of ice-rafted debris in abyssal sediments of the Southern Ocean. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 1013–1016. Webb, P. N. (1991). A review of the Cenozoic stratigraphy and paleontology of Antarctica. In: M. R. A. Thomson, J. A. Crane, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 599–607. Webb, P. N., & Harwood, D. M. (1991). Late Cenozoic glacial history of the Ross embayment, Antarctica. Quat. Sci. Rev., 10, 215–223. Webb, P. N., & Wrenn, J. H. (1982). Upper Cenozoic micropaleontology and biostratigraphy of eastern Taylor Valley, Antarctica. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, WI, pp. 1117–1122. Webb, P. N., Ronan, T. E. Jr., Lipps, J. H., & Delata, T. E. (1979). Miocene glaciomarine sediments from beneath the southern Ross Ice Shelf, Antarctica. Science, 203, 435–437. Webb, P. N., Leckie, R. M., & Ward, B. L. (1986). Foraminifera (late Oligocene): Antarctic Cenozoic history from the MSSTS-1 drillhole, McMurdo Sound. Science Information Publishing Centre, Wellington, pp. 115–125 Weber, M. E., Bonani, G., & Fu¨tterer, D. K. (1994). Sedimentation processes within channel-ridge systems, southeastern Weddell Sea, Antarctica. Paleoceanography, 9, 1027–1048. Wellner, J. S., Lowe, A. L., Shipp, S. S., & Anderson, J. B. (2001). Distribution of glacial geomorphic feature on the Antarctic continental shelf and correlation with substrate: Implications for ice behavior. J. Glaciol., 47, 397–411. Wellner, J. S., Anderson, J. B., & Wise, S. W. (2005). The inaugural SHALDRIL expedition to the Weddell Sea, Antarctica. Sci. Drill., 1, 40–43. Wellner, J. S., Heroy, D. C., & Anderson, J. B. (2006). The death mask of the Antarctic Ice Sheet: Comparison of glacial geomorphic features across the continental shelf. Geomorphology, 75, 157–171. Whitehead, J. M., Quilty, P. G., Harwood, D. M., & McMinn, A. (2001). Early Pliocene paleoenvironment of the Sørsdal Formation, Vestfold Hills, based on diatom data. Mar. Micropaleontol., 41, 125–152.
Cenozoic Climate History from Seismic Reflection and Drilling Studies
227
Whitehead, J. M., Harwood, D. M., & McMinn, A. (2003). Ice-distal Upper Miocene marine strat in inland Antarctica. Sedimentology, 50, 531–552. Whitehead, J. M., Harwood, D. M., McKelvey, B. C., Hambrey, M. J., & McMinn, A. (2004). Diatom biostratigraphy of the Cenozoic fjordal Pagodroma Group, northern Prince Charles Mountains, East Antarctica. Aust. J. Earth Sci., 51, 521–547. Williams, T., & Handwerger, D. (2005). A high-resolution record of early Miocene Antarctic glacial history from ODP Site 1165, Prydz Bay. Paleoceanography, 20, doi:10.1029/2004PA001067. Wilson, G. S., Roberts, A. P., Verosub, K. L., Florindo, F., & Sagnotti, L. (1998). Magnetobiostratigraphic chronology of the Eocene–Oligocene transition in the CIROS-1 core, Victoria Land margin, Antarctica: Implications for Antarctic glacial history. Geol. Soc. Am. Bull., 110, 35–47. Wise, S. W., Gombos, A. M., & Muza, J. P. (1985). Cenozoic evolution of polar water masses, southwest Atlantic Ocean. In: K. J. Hsue, & H. J. Weissert (Eds). South Atlantic Paleoceanography. Cambridge University Press, Cambridge, pp. 283–324. Wise, S. W., Jr., Breza, J. R., Harwood, D. M., & Wei, W. (1991). Paleogene glacial history of Antarctica. In: D. W. Muller, J. A. McKenzie, & H. Weissert (Eds). Controversies in Modern Geology; Evolution of Geological Theories in Sedimentology, Earth History and Tectonics. Academic Press, London, pp. 133–171. Wise, S. W. J., Breza, J. R., Harwood, D. M., Wei, W., & Zachos, J. C. (1992). Paleogene glacial history of Antarctica in light of Leg 120 drilling results. In: S. W. Wise. Jr., R. Schlich, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 120: College Station, TX, pp. 1001–1030. Wordie, J. M. (1921). Depths and deposits of the Weddell Sea, Shackleton Antarctic Expedition 1914–1917. Trans. R. Soc. Edinburgh, 52, 781–793. Wright, R. (1980). Sediment Gravity Transport on the Weddell Sea Continental Margin. M.A. Thesis, Rice University, Houston, TX, 96 pp. Wright, R., & Anderson, J. B. (1982). The importance of sediment gravity flow to sediment transport and sorting in a glacial marine environment: Eastern Weddell Sea, Antarctica. Geol. Soc. Am. Bull., 93, 951–963. Wright, R., Anderson, J. B., & Fisco, P. P. (1983). Distribution and association of sediment gravity flow deposits and glacial/glacial-marine sediments around the continental margin of Antarctica. In: B. F. Molnia (Ed.). Glacial-Marine Sedimentation. Plenum Press, New York, pp. 265–300. Wright, J. D., & Miller, K. G. (1992). Miocene stable isotope stratigraphy, site 747, Kerguelen Plateau. In: S. W. Wise Jr., & R. Schlich, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results. College Station, TX, Vol. 120, pp. 855–866. Wright, J. D., & Miller, K. G. (1993). Southern Ocean influences on late Eocene to Miocene deepwater circulation. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change II. American
228
A. K. Cooper et al.
Geophysical Union, Washington, DC, Antarctic Research Series, Vol. 60, pp. 1–25. Yamaguchi, K., Tamura, Y., Mizukoshi, I., & Tsuru, T. (1988). Preliminary report of geophysical and geological surveys in the Amundsen Sea, West Antarctica. In: Proceedings of NIPR Symposium on Antarctic Geosciences. National Institute of Polar Research, Tokyo, Vol. 2, pp. 55–67. Yoon, H. I., Han, M. W., Park, B. K., Oh, J. K., & Chang, S. K. (1994). Depositional environment of near-surface sediments, King George Basin, Bransfield Strait, Antarctica. Geo Mar. Lett., 14, 1–9. Zachos, J. C., Breza, J. R., & Wise, S. W. (1992). Early Oligocene ice-sheet expansion on Antarctica: Stable isotope and sedimentological evidence from Kerguelen Plateau, southern Indian Ocean. Geology, 20, 569–573. Zachos, J. C., Pagani, M., Sloan, L., Thomas, E., & Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693. Zayatz, I., Kavun, M., & Traube, V. (1990). The Soviet geophysical research in the Ross Sea. In: A. K. Cooper, & P. N. Webb (Conveners). International Workshop on Antarctic Offshore Seismic Stratigraphy (ANTOSTRAT), Overview and Extended Abstracts. US Geological Survey Open-File Report 90-309, pp. 283–290. Zhivago, A. V., & Lisitsin, A. P. (1957). New data on the bottom and submarine deposits in the eastern Antarctic. Izvestiya Akad. Nauk SSSR, 1, 19–35. Zinsmeister, W. J. (1982). Review of the Upper Cretaceous-Lower Tertiary sequence on Seymour Island, Antarctica. J. Geol. Soc. London, 139, 779–785.
Foldout RS-1: ‘Type section’ from the Eastern basin (modified from Cooper et al., 1995). The section is a compilation of seismic lines BGR-7 and IFP-208. RSS-1 are inferred pre-ice deposits, RSS-2 and -3 are early-glacial glaciomarine deposits, RSS-4, -5 and -6 are glaciomarine deposits of the ice-sheet growth phase, and RSS-7 and -8 are glaciomarine deposits of the polar ice-sheet phase. See Fig. RS-1 for location.
Foldout RS-2: ‘Type section’ from the Northern basin (modified from Cooper et al., 1995). The stratigraphy and origin of sediments are similar to those for the ‘type section’ in the Eastern basin (Foldout RS-1). RSS-1 is a basin-fill sequence, RSS-2 and -3 mostly aggrade, RSS-4, -5 and -6 mostly prograde and have with eroded topset beds, RSS-7 and -8 have aggrading geometries. See Fig. RS-1 for location.
50 km
B
Line TH99-35
Line TH99-35-1
Line GA 229/29
B’
ODP 742
N
S
Basement
Seconds
Two-way time
K
Mixed turbidite-contourite sequence
Fault?
Turbidites?
Basement
Foldout PB-1: Seismic sections from eastern Prydz Bay (Section B-Bu). In the bay, Cenozoic sediments overlie Cretaceous (Surface K) and older sequences resting on basement. The shelf edge and upper slope prograded seaward through time becoming steeper. The slope and rise overlie a thick post-rift section, the latest parts of which are turbidite (inferred) and contourite (drilled) sequences deposited since the onset of glaciation. Location shown in Fig. PB-1.
50 km
A
Line AGSO 149/0901
A’
Line GA 229/31 ODP Site 1167
S
N
Two-way time seconds
Prydz Bay
Prydz Channel Fan
Pp-12 Mud waves
Mounded contourite-turbidite sequences
Channel-levee sequences (Turbidites)
Basement
Foldout PB-2: Seismic sections from western Prydz Bay (Section A-Au). The very thick post-rift section includes thick contourite (drilled) and turbidite (inferred) mounds with mudwaves in places. The upper slope comprises the Prydz Channel Fan and the shelf shows progradation but only limited topset thickness. The base of Prydz Channel Fan is Reflector A of Mizukoshi et al. (1986) and PP-12 of O’Brien et al. (2004). Location shown in Fig. PB-19.
Foldout WS-1: Seismic line NARE-8508 along the front of the Crary Trough Mouth Fan, showing fan architecture (modified from Kuvaas and Kristoffersen, 1991). Profile location is in Fig. WS-1.
Foldout AP-1: Seismic line BAS878-19 across the Antarctic Peninsula Pacific margin, showing the shelf basin, prograding wedge and continental rise sediments between the continental slope and DSDP Site 325. A buried, early Miocene sediment drift is observed beneath the central rise (Herna´ndez-Molina et al., 2004, 2006b), but this line lies between two of the large drifts that have developed since the start of the middle Miocene (Rebesco et al., 1996, 1997, 2002). Oceanic basement ages are from Larter et al. (1997). Vertical exaggeration at the seafloor is 23:1. Inset in upper right shows an interpretation of major features. Inset in lower left shows line location.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00006-2
Chapter 6
Numerical Modelling of the Antarctic Ice Sheet Martin J. Siegert School of GeoSciences, University of Edinburgh, Grant Institute, West Mains Road, Edinburgh EH9 3JW, UK
ABSTRACT Studies on Antarctic climate evolution have benefited increasingly over the last 20 years from numerical ice-sheet modelling. Such activities have led to the testing of geological hypotheses concerning past ice-sheet changes and a better understanding of macro-scale glaciological processes through space and time. Here, numerical ice-sheet models are reviewed, and their use in understanding former changes in Antarctica is outlined.
6.1. Introduction Analyses of glacial geological data allow hypotheses concerning Antarctic Ice Sheet history to be formed. Testing these hypotheses can be undertaken in two ways. First, additional geological data can be acquired and, second, numerical ice-sheet modelling experiments can be performed. There are several advantages in taking the latter option over the former. Ice-sheet modelling allows time-dependent assessments of ice-sheet form and flow across the entire continent. Furthermore, model results can be analysed to determine where the ice sheet is most sensitive to changes and, in doing so, where the best sedimentary records of such changes may occur. Finally, the cost of undertaking numerical ice-sheet modelling is cheap relative to Antarctic data gathering. Hence, the use of ice-sheet modelling to test geologically based hypotheses has increased in recent years, and indeed is the Corresponding author. Tel.: +44(0)131 650 7543; Fax: +44(0)131 668 3184;
E-mail:
[email protected] (M.J. Siegert).
236
M. J. Siegert
focus of the Antarctic Climate Evolution programme. Ice-sheet models have limitations, however, concerning the necessary simplification of ice flow processes, and incomplete model input data (such as bed topography and climate forcing). Researchers need to comprehend these limitations when interpreting results. In this chapter, generic numerical ice-sheet modelling is reviewed, and ways in which Antarctic models may be validated are assessed. The chapter provides necessary background information to later chapters that discuss results from ice-sheet modelling.
6.2. Ice-Sheet Processes 6.2.1. Flow of Ice Direct observations of glacier motion and laboratory studies of ice rheology (the flow and deformation of ice) have identified three main mechanisms by which ice masses move. These are by internal deformation of the ice itself, and through two processes which take place at the base of the glacier; basal sliding and the deformation of water-saturated weak sediments (Fig. 6.1). Flow by internal ice deformation takes place in all ice masses, and generally accounts for motion of a few metres per year. However, basal motion occurs only where the bed is at the pressure melting point such that water is present. Where basal motion takes place, glaciers may move at tens to hundreds, and sometimes a few thousand, metres per year. The flow of ice sheets can be organized in general terms as follows. At the centre of an ice sheet, the flow speed is very low (of the order of metres per year), and controlled by internal deformation. A particle of ice on the ice-sheet surface will be buried by subsequent snowfall and so will possess a relatively significant vertical velocity component downwards into the ice. The flow of ice radiates from the ice divide, where there is no lateral flow. Ice-sheet interiors are characterized by a series of divides that define the margins of ice drainage basins. Ice sheets are effectively ‘‘drained’’ by fast-flowing rivers of ice, known as ice streams, transporting ice from the interior to the ice margin (Bennett, 2003). The velocity of ice streams is often several hundred metres per year. Ice streams flow quickly because water at their bases causes a reduction in subglacial friction allowing them to effectively slide across the subglacial topography, with ice deformation contributing only a small amount to the total velocity. The transition between the slow-moving ice sheet and the fastflowing ice streams has been shown recently to occur in ‘‘tributaries’’ several hundred kilometres inland from the margin (Bamber et al., 2000).
Numerical Modelling of the Antarctic Ice Sheet
237
rface
ier su
Glac
(a) Warm-based glacier
Internal flow
Basal sliding
resting on bedrock
Glacier bed
a
c
Bedrock
ce
r surfa
Glacie
Ice flow
(b) Cold--based glacier resting on bedrock
Glacier bed Bedrock
rface Glacier su Ice flow Sliding
on deformable sediment
Glacier bed Deforming horizon
Subglacial deformation
(c) Warm-based glacier resting
Stable bed
Figure 6.1: Processes controlling the flow of large ice sheets. Ice will flow through internal deformation in all situations. If the ice rests on bedrock and is warm based (a) then basal sliding can occur. If the base is frozen (b) then only ice deformation will take place. If the base is warm and loose unconsolidated sediments are present (c) then their own deformation can add to ice deformation and basal sliding as a contribution to ice flow. Each of these processes can be accounted for in ice-sheet models. Adapted from Siegert (2001). 6.2.2. Mass Balance Antarctica gains mass through accumulation of snow in two ways. Direct precipitation of snow occurs close to the ice-sheet margin, where accumulation rates can reach B1 m per year. This is in contrast to accumulation rates in the interior, which are far lower (centimetres per year) and are associated
238
M. J. Siegert
with solid precipitation direct from water vapour. Broadly speaking, ice loss in Antarctica is through two main processes: iceberg calving and sub-iceshelf melting (Jacobs et al., 1992). Consequently, ice-shelf processes are critical to the overall mass balance of the Antarctic Ice Sheet. On grounded ice, surface sublimation, which forms blue ice zones, and basal melting, responsible for at least 145 subglacial lakes (Siegert et al., 2005a), are not thought to contribute significantly to the overall ice-sheet mass balance.
6.2.3. Isostasy When a glacial load is placed on the Earth’s crust, the weight of ice will act to displace the crust into the viscous asthenosphere (beneath the lithosphere), which adjusts towards isostatic equilibrium according to Archimedes Principle, on time scales of a few thousand years. The process by which this is done involves flow in the asthenosphere and elastic deflection of the lithosphere. When the ice load is removed (deglaciation), the asthenosphere and lithosphere will relax back to their original state (i.e. isostatic uplift or recovery). There are several ways in which glacial isostasy can be accounted for in ice-sheet models, and a thorough description of the problem is provided in Le Meur and Huybrechts (1996).
6.3. Ice-Sheet Models Numerical ice-sheet models have been used extensively since the mid-1980s to aid the reconstruction of former ice sheets, to comprehend existing ice sheets and to forecast future behaviour. The principle behind numerical icesheet modelling is that an ice sheet can be divided into a number of ‘‘ice columns’’. Each of these columns represents a ‘‘cell’’ in the model’s 2D horizontal ‘‘grid’’. Ice-sheet models are usually arranged in a ‘‘loop’’ that begins by applying a series of algorithms, determining the flow of ice, mass balance and interaction with the Earth, in each cell. The loop is completed by a final equation (widely known as a ‘‘continuity’’ equation), used on the full grid, to calculate the interaction and flow of ice between cells. Each iteration of the loop advances the model through one time step. The accuracy of the model depends in part on the width of the grid cells, and the time step length. For continental-scale ice-sheet models, where the time-dependent change in ice-sheet behaviour needs to be calculated over several thousand years, the grid cell width is usually between five and twenty kilometres, and the loop
Numerical Modelling of the Antarctic Ice Sheet
239
time step is between 1 and 10 years. Because ice-sheet modelling is a valuable tool for palaeo-glaciologists, it is appropriate here to outline briefly how icesheet models are organized. Most ice-sheet models are centred on the continuity equation for ice (Mahaffy, 1976), where the time-dependent change in ice thickness is associated with the specific net mass budget as follows: dH ¼ bs ðx; tÞ r F ðu; H Þ dt where F(u, H) is the net flux of ice (m2 per year) (the flux of ice being the product of ice velocity, u, and ice thickness, H), r the divergence operator and bs the specific mass budget term (involving surface mass balance, iceberg calving and ice-shelf basal melting where necessary). In the simplest models, the depth-averaged ice velocity, u (m s1), is calculated by the sum of depthaveraged internal ice deformation and basal motion (Fig. 6.1). Such models make the so-called ‘‘shallow-ice’’ approximation and assume that ice flow can be described by a single horizontal vector. More complex models involving the full three-dimensional flow of ice require a sophisticated solution to the stresses and strains which govern the flow of ice (named ‘‘high-order stress models’’). Numerical ice-sheet models calculate glacial processes in a discrete manner (regardless of their sophistication). These processes are linked, however, such that feedback mechanisms can exist between the calculations. The relationship between distinct ice-sheet processes can be summarized by a flow diagram (Fig. 6.2). A good example of a feedback mechanism involves glacial isostasy, which acts to adjust bedrock elevation due to ice loading and, hence, the ice-sheet surface elevation. Surface mass balance and ice velocity will be modified by surface elevation changes, which in turn feed back on ice thickness and, hence, ice loading. Another feedback loop acts to change thermal conditions at the base of ice sheets. This is important because the ice velocity due to rapid basal motion (e.g. sliding) is influenced by the thermal regime. In a cold-based ice-sheet, ice velocities are generally low and so, theoretically, the ice sheet is allowed to build up with a minimum of basal sliding. However, as the ice sheet thickens, the basal temperature is likely to increase (because both surface temperatures and vertical temperature gradients remain relatively constant). Once the base of the ice sheet starts to melt, rapid basal motion becomes possible. Fast ice flow leads to more ice being advected to the ice-sheet margins and, so to ice-sheet thinning. Thinner ice then causes a reduction in the temperature of the icesheet base and possibly to the curtailment of rapid basal motion. This feedback process is complicated because the thermal regime is influenced by
240
M. J. Siegert
Basal marine melt
Initial precipitation
Precipitation
Isostasy Initial bedrock
Net mass balance
Ice thickness
Calving and maritime melt
Bedrock elevation Eustasy
Ablation and effectiveness
Latitude
Ice velocity
Surface slope
Ice surface elevation
Imposed temperature change
Figure 6.2: Flow diagram of the operation of an ice-sheet model. Adapted from Siegert (2001). the basal heat gradient, which is in part related to the heat due to basal sliding. Thus, rapid basal motion may maintain warm-based conditions even if the ice sheet thins. One might consider when it is appropriate to employ simple flow models versus complex models. Both have a role to play in understanding the behaviour of large ice sheets. The benefit of simple models, such as those using the ‘‘shallow-ice approximation’’, is twofold. First, in central regions of the ice sheet, the approximation works well at a horizontal scale greater than five ice thicknesses. In other words, there is little change in output at this scale between a simple and a complex model. Second, they can be run quickly, which means multiple runs and sensitivity experiments can be undertaken with minimum computing costs. More complex models are required to solve particular ice flow issues, such as the dynamic change between floating and grounded ice that occurs at the ice-sheet/shelf transition (Schoof, 2007) and over subglacial lakes (e.g. Pattyn, 2003). Given that recent changes to the Antarctic Ice Sheet have been observed at the marine ice-sheet margin (for example, in the recent IPCC report in 2007), understanding ice-sheet response to future warming scenarios is likely to require complex models. High-order models are also able to run at
Numerical Modelling of the Antarctic Ice Sheet
241
greater spatial accuracy, meaning that important features such as ice streams are resolved. However, given the uncertainty in model inputs regarding palaeo-ice sheets (see below), it is arguable whether the complexity accounted for in high-order stress models adds any value to the output of simpler models.
6.4. Model Inputs In numerical modelling (ice sheet, ocean or atmospheric), the quality of model output is dependent on the quality of model input. The most important data used as boundary and forcing conditions in an ice-sheet model are subglacial topography and surface mass balance. For present-day ice sheets, the quality of these model inputs is dependent on field data which, for some areas, are noticeably absent. Airborne radar surveying of large ice sheets provides the only viable method of acquiring information on the icesheet base at a continental scale. Such surveys are organized as a series of gridded flightlines. Although the technique is sound, there are currently two limitations to datasets collected by radar surveying. The first is that information between flightlines is absent. Interpolation software is used to infer the topography in such regions. The second limitation, which is acute in Antarctica, is that large portions of some ice sheets remain to be surveyed. Thus, subglacial topography in these regions has to be estimated, often from just a few seismic measurements, and is subject to a high level of potential error. To date, Antarctic Ice Sheet models have been run over a topographic DEM based on an incomplete coverage of bed elevation data. There have been several recent advances to alleviate this problem. Since the 1970s, when radar data from over 400,000 km worth of flightlines were collected over a wide area of Antarctica (albeit at a coarse line spacing), there have been several surveys over much smaller regions with close flightline spacing. The most recent depiction of Antarctic topography was assembled by the BEDMAP consortium (Lythe et al., 2000). While this database quantifies topography well in many places, it remains restricted in two ways. First, there are several large ‘‘data gaps’’ where very little is known about sub-ice topography. Second, even in regions where coverage is good, radar transects are often separated by several kilometres (often tens of kilometres) across which interpolation of data remains necessary. Comparison of the topography measured by recently acquired radar data with the interpolated information from BEDMAP reveals that, in some places, the errors in BEDMAP are up to the order of hundreds of metres (Welch and Jacobel, 2003).
242
M. J. Siegert
Mass balance of large ice masses is also very difficult to establish accurately over a wide area. Although it is possible to measure annual mass balance at specific sites, there is no quick and easy method by which large-scale data can be obtained. Satellite observations, used in conjunction with direct measurements, provide a means of estimating mass balance across a wide area, and have been used to assess Antarctic mass balance (Vaughan et al., 1999; Arthern et al., 2006). Such analysis only provides an estimate of the current situation, however, which is of less use when trying to model past or future scenarios. As this is a real problem for Antarctic Ice Sheet models, many are coupled with models of surface mass balance. The accuracy of ice-sheet model results will then be dependent on the quality of the surface balance results. Similar problems exist when accounting for the processes of iceberg calving and ice-shelf basal melting in ice-sheet models.
6.5. EISMINT The majority of ice-sheet models have been developed by individual researchers or within small institution-based groups. This ‘‘do-it-yourself’’ history is in contrast to that of large community-based models, which dominate research in the fields of meteorology and oceanography. The resulting diversity of ice-sheet models is both a strength, in that specific models were honed to specific applications, and a weakness, in that no consensus on the basic behaviour of the models existed. The European Ice Sheet Modelling Initiative (EISMINT) was developed to address this diversity. Siegert and Payne (2001) provide details of how the EISMINT programme has assisted the development of ice sheet models. The following section is summarised from Siegert and Payne (2001) to illustrate the purpose of EISMINT and its achievements. EISMINT had three main aims: first, to define and perform intercomparison exercises to help establish consensus on basic model prediction; second, to identify good practice in ice-sheet modelling; third, to develop the next generation of ice-sheet models. The EISMINT project was funded by the European Science Foundation in two phases, which ran from January 1993 to December 1995 and from January 1996 to December 1997. The funding allowed four major international meetings to concentrate on model intercomparison. It also allowed a series of smaller meetings on basal processes, ice rheology, ice–climate interactions, ice–oceanic interactions, ice–lithosphere interactions and former ice sheets.
Numerical Modelling of the Antarctic Ice Sheet
243
The aims of model intercomparisons were threefold. The first was to test the effects of numerical implementation on model prediction. This required the physics incorporated into models to be tightly constrained, along with the values of model parameters and boundary conditions. A wide variety of numerical techniques has been employed by ice-sheet modellers. For example, several finite-element models existed, although the majority of ice-sheet models employed finite difference schemes. Within the latter class, a variety of methods were available to deal with the basic nonlinearity of ice flow. The EISMINT intercomparisons explained, for the first time, the effect of the type of numerical solution on model prediction. The second was to determine the effect which the many poorly constrained physical parameters had on the overall prediction of the models. The third was to model ice masses through a given time-dependent climate change scenario. Examples included the response of ice sheets to stepped changes in forcing, glacial– interglacial growth and decay, and the response to future, anthropogenic change. Particular attention was paid to how models replicated and forecasted Antarctic Ice Sheet dynamics. 6.5.1. EISMINT Level One and Level Two Ice-sheet model intercomparison was designed to be as inclusive as possible, in order that results from models with one and two horizontal dimensions, which may or may not have included internal-temperature evolution, could be evaluated (Huybrechts et al., 1996, and papers within Annals of Glaciology, Vol. 23). Model boundary conditions and parameter values were prescribed as much as possible so that observed differences in output could be interpreted purely in terms of the numerics of the model. The intercomparison exercise involved several iterations in which model results were submitted for analysis. A consensus set of results gradually emerged from this process. Many of the differences identified initially between models arose for non-scientific reasons (such as ambiguity in the experiment descriptions). One problem encountered was that results tended to converge towards median values as outlying models were modified. The modification process may have been genuine (for instance, in finding coding errors) but may also have been driven by a perceived need to conform with the bulk of models regardless of whether they were correct. In this way, genuine differences could have become obscured. This problem was offset somewhat by recognizing analytical solutions to ice flow equations, taken as ‘‘truth’’. While the number of analytical solutions available is quite limited for ice-sheet models, their use was an important feature of the intercomparison process.
244
M. J. Siegert
In general, most models performed similarly in terms of their predictions of ice-thickness evolution, internal-temperature evolution and ice velocity patterns, but some differences were observed. Interpretation of these differences showed the effect of numerical solution on model predictions. Two groups of models were identified: those which employed a more stable scheme to solve the continuity equation were found to be less accurate in comparison to a more precise formulation which was, however, less stable numerically. In this way, the EISMINT programme was able to establish firm benchmarks for model types and their performance. 6.5.2. EISMINT Level Three Simulations of the Antarctic Ice Sheet took place in the third level of the EISMINT exercise during 1997. The aim of this level was to compare model predictions of the past and future evolution of these ice masses with the minimum amount of constraint on the details of the individual models. Three experiments were undertaken. The first compared results from simulations of the present-day ice sheet. The second compared results from glacial–interglacial experiments. The third analysed differences between forecasts of future icesheet behaviour (over the next 500 years) (see http://homepages.vub.ac.be/ Bphuybrec/eismint.html for details). Results for the first experiment were encouraging, as all four models revealed similar output. However, major differences were observed in the second and third experiments. Specifically, ice-sheet volumes predicted for both the Eemian interglacial (started at ca. 125 ka and ended at ca. 110 ka) and Last Glacial Maximum were noticeably different. Although only two of the four models yielded similar results, they were recommended as a benchmark for subsequent studies, presumably because other models had fatal flaws that became apparent in the experiment, allowing a level of ‘‘verification’’ for Antarctic Ice Sheet models. 6.5.3. Outcomes from EISMINT There is no doubt that the EISMINT programme has led to agreement and commonality regarding ice-sheet modelling. One important outcome is the establishment of a community within the field of glaciology, which can develop ice-sheet modelling in future. Importantly, the EISMINT programme has led to the availability of free software, in the form of the GLIMMER model (http://forge.nesc.ac.uk/projects/glimmer/), based on the
Numerical Modelling of the Antarctic Ice Sheet
245
‘‘best practice’’ elements established through model intercomparison exercises. The full development, documentation and release of GLIMMER was a consequence of the GENIE Earth System Model project (www. genie.ac.uk), which used an ice-sheet model built on the validated procedures established by EISMINT. It is very likely that this model will lead to an increase in ice-sheet modelling activity and, thus, its use in understanding past changes in Antarctica. The next phase of ice-sheet model intercomparisons has recently been established in the ice sheet Model Intercomparison Programme (ISMIP). In ISMIP, the performance of next generation of icesheet models (accounting for the full three-dimensional flow of ice) will be analysed. It will also help to establish community agreement on how icesheet models predict the cryosphere to respond to future climate change scenarios (details of ISMIP can be found at the following website: homepages.vub.ac.be/Bphuybrec/ismip.html).
6.6. Comparing Ice-Sheet Models with Antarctic Glaciological Data Despite most ice-sheet models being based on well-established physical assumptions about the flow of ice, these assumptions, and the data used as input, may oversimplify the actual glaciological situation. The result is mismatch between model results and real glaciological measurements. When comparing model output with glaciological data, the scale of the datasets needs to be similar to the scale of the ice-sheet model output (e.g. averaged over 5–20 km) for the exercise to be meaningful. Real data at a finer resolution than this show how ice-sheet models simplify the actual flow of ice sheets. Determining the difference between model output and these real data represents an important way in which to assess the validity of ice-sheet models. As ice-sheet models become more sophisticated, it will be necessary to test their output against a variety of ice-sheet measurements for validation purposes. This section outlines datasets currently available in Antarctica that have yet to be compared fully with ice-sheet model output, but should be used for the future validation of ice-sheet models (Siegert and Payne, 2001). 6.6.1. Surface and Sub-Ice Morphology, and the Flow of Ice in Central Regions The advent of satellite altimetry has resulted in the determination of the surface morphology of ice sheets to a high degree of accuracy. At present,
246
M. J. Siegert
most ice-sheet models of Antarctica work at a 5–20 km scale. These models replicate the broad shape of the ice sheet well. However, real morphology at this or a finer scale cannot be modelled at present. ERS-1 altimetry of the Antarctic Ice Sheets reveals that there is a complex morphology on the ice surface related to the flow of ice. A good example of where ice-sheet model results (Huybrechts, 1990) match well with large-scale ice-sheet features, yet poorly with sub-grid cell morphology, is at Dome C, East Antarctica (Fig. 6.3). The ice surface around Dome C has a generalized surface slope of about 0.081. However, there are a number of regions where the slope reduces to less than 0.011. These ‘‘flat’’ surfaces are caused by ice flow over subglacial trenches and/or subglacial lakes (e.g. Lake Vostok). In either case, the flow of ice is altered from the base-parallel shearing that most ice-sheet models account for, by longitudinal extension. In order to solve this problem, an accurate representation of subglacial topography is required. Although there is a
Figure 6.3: The location of Antarctic subglacial lakes. Adapted from Siegert et al. (2005a).
Numerical Modelling of the Antarctic Ice Sheet
247
major programme aimed at establishing an up to date bedrock elevation for Antarctica (e.g. BEDMAP), the resolution of these new data is not fine enough to solve the problem of complex flow of ice over subglacial topography highlighted here. 6.6.2. Radar Layering and the Internal Flow of Ice Some ice-sheet models are capable of calculating the flow of ice in three dimensions. Ice penetrating radar information can be used to verify calculated flow paths because it reveals internal layering which is assumed to be isochronous. Internal layering is detected by changes in the electrical properties of ice. However, there are three different ways by which these changes can occur, yielding three main types of internal layering. The first type of internal layer is when there are changes in the density of ice. This is a dominant process at ice depths less than 700 m. However, below this level, the density of ice does not change very much and so internal reflections must be caused by other processes. The second form of layering is caused by the acidity of ice. Layers with an acidity in excess of the normal background level of glacier ice are formed when the aerosol products of ancient volcanic activity are incorporated in the snow chemistry on the former ice surface. This acid snow is subsequently buried by later snow fall to its present-day position, several tens of thousand years later. These acid layers are therefore isochronous surfaces. The third type of layering is where there are changes in the crystallography of ice. Such layering is thought to develop from acidic layers in the presence of enhanced stress across the stoss face of subglacial hills (Fujita et al., 1999). Internal layers are often continuous across large sections of the Antarctic Ice Sheet (Siegert, 1999). Because they are isochronous, their patterns can be used to match the 3D flow of ice calculated in numerical models. However, as yet, very few ice-sheet modelling investigations have utilized this potentially powerful measurement. Leysinger Vieli et al. (2007) have recently identified a means by which internal layering can assist large-scale ice-sheet modelling, and this marks an important new area of model validation. 6.6.3. Location of Subglacial Lakes and Ice-Sheet Thermal Conditions Subglacial lakes are evidence for melting at the ice-sheet base. They can, therefore, be used to validate the thermal conditions calculated by ice-sheet models. Over 145 subglacial lakes have been identified at the base of the
248
M. J. Siegert
13 million km2 Antarctic Ice Sheet (Siegert et al., 2005a). The bulk of the lakes are located in the interior of the Antarctic Ice Sheet, at or close to ice divides in the form of domes and ridges, where ice velocity is low (Dowdeswell and Siegert, 1999). Almost 60 per cent of lakes are found within 200 km of an ice crest, remembering that ice flowlines from divide to coastal margin are often over 1,000 km long in Antarctica (Fig. 6.3). Only about 15 per cent of subglacial lakes are positioned more than 500 km from an ice divide. At least 16 subglacial lakes occur at locations which are near the onset of enhanced ice flow, some hundreds of kilometres from the ice-sheet crest. An example is provided by three subglacial lakes near the onset of fast flow into Byrd Glacier. Byrd Glacier is fast flowing and drains a very large interior ice-sheet drainage basin into the Ross Ice Shelf. These subglacial lakes are similar in size and depth to the small and probably shallow lakes found in major subglacial basins in the ice-sheet interior. Siegert and Bamber (2000) recognized subglacial lakes at the heads of ice stream tributaries (some of which are several hundred kilometres in the ice-sheet interior) as being evidence of warm-based conditions along the entire length of the enhanced flow unit. Huybrechts (1990) used the location of subglacial lakes, as evidence of warm-based conditions, to verify the broad-scale thermal character of the icesheet base across East Antarctica. Basal melting conditions were predicted across both the centre of East Antarctica and at the base of ice streams. However, because of the smoothed bedrock topography used as model input, individual lakes, or the outline of specific lake regions, were not able to be matched well. There is, therefore, plenty of scope to better use subglacial lakes in the future verification of ice-sheet model thermal conditions. Further validation of the ice-sheet thermal regime can be made by comparing results with temperature profiles from ice cores. However, this comparison is limited by the small number of deep ice cores and the lack of penetration to the lower layers of the ice sheet in most cases. 6.6.4. Satellite Interferometry One aspect of ice-sheet modelling that is in need of independent datasets for comparison and validation is the calculated velocity of ice. Until recently, such data have been scarce over large ice sheets. However, recent interferometric Synthetic Aperture Radar (SAR) InSAR techniques have been applied to glaciers and ice sheets to reveal the surface velocity continuously across wide areas. Work to date has shown that development of ice streams in West Antarctica seen through InSAR data (Joughin et al., 1999) compares well with
Numerical Modelling of the Antarctic Ice Sheet
249
calculations of Antarctic ice flux (Bamber et al., 2000). Maps of surface ice velocities are currently being assembled for the whole of Antarctica, which may be a vital dataset for model verification in future. InSAR data revealing the surface velocity of ice sheets also have an application in modelling subglacial sliding. In areas where the internal icesheet deformation is known to occur in a simple manner, the difference between the InSAR velocity and velocity due to internal deformation of ice will be equivalent to the component due to basal processes. Currently, models are incapable of accurately predicting subglacial sliding and, so, InSAR data may be of future use in establishing reliable algorithms for this process.
6.7. Ice-Sheet Reconstructions Once ice-sheet models have been verified and validated against the modern glaciological regime, they can be used to test geological hypotheses regarding past changes in Antarctica. Whereas the boundary conditions of the present ice sheet can be established reasonably well, past conditions are difficult to quantify. This difficulty obviously increases with past time. Researchers should be aware, therefore, of the limitations of ice-sheet models, and the reliability of information they are able to provide, given necessary assumptions about the past environment. Despite this problem, ice-sheet modelling has been used in innovative ways to help understand past ice-sheet changes in Antarctica. Two examples of such activities are provided below.
6.7.1. Pliocene Ice Sheets Huybrechts (1993) provided the first detailed modelling investigation into Pliocene Antarctic Ice Sheets. He tackled the problem of whether the East Antarctic Ice Sheet was stable at this time, or susceptible to significant changes (both hypotheses have been developed from the geological record; see Miller and Mabin, 1998). Running a standard ice-sheet model, using the shallow-ice approximation under a selection of climate forcing parameters, Huybrechts showed that the East Antarctic Ice Sheet could withstand a mean annual temperature rise of around 101C and that several degrees in excess of this amount of warming was required to significantly alter the ice-sheet configuration (Fig. 6.4). When the ice sheet was forced to change, it first did so predominantly across the Dome C region, where topography is lowest in East Antarctica. While the model was not used to understand the form and
250
M. J. Siegert
Figure 6.4: Numerical modelling results of the Antarctic Ice Sheet under a variety of climate warming scenarios. Taken from Huybrechts (1993) with permission from The University of Edinburgh. flow of the Pliocene ice sheet, it did demonstrate that even with substantial climate warming, a large semi-continental-scale ice mass was predicted. Air temperatures needed to rise by around 151C of modern values in order to reduce the ice sheet to a series of small isolated ice caps. As such warming is unlikely for the Pliocene, Huybrechts (1993) concluded that the idea of widespread decay of ice at this time is not supported by ice-sheet modelling. Huybrechts’ results are relatively easy to replicate within a simple ice-sheet model. Siegert et al. (2005b) ran a basic shallow-ice approximation EISMINT model with a simplistic climate control. By raising the equilibrium
Numerical Modelling of the Antarctic Ice Sheet
251
line altitude (ELA) to 400 m above sea level (i.e. a climate warming scenario of over 151C), the East Antarctic Ice Sheet reduced in size across the Dome C region. Further increases in the ELA resulted in the complete decay of ice at Dome C, but retention of a large ice sheet on the higher ground focused at Dome A and Ridge B.
6.7.2. Pleistocene Ice Sheets There have been several models of recent (last glacial cycle) ice-sheet behaviour (e.g. Ritz et al., 2001; Huybrechts, 2002). Huybrechts’ model demonstrated the differences in behaviour between East and West Antarctica over the last 200,000 years, and showed these differences to be controlled to a large degree by topography (Fig. 6.5). The West Antarctic Ice Sheet is known as a marine-based ice sheet, with a bed that is well below sea level in most places (mean bed elevation B400 m). This is in contrast to the terrestrial ice sheet in East Antarctica, where the bed elevation is generally above sea level (mean bed elevation Bþ50 m). Marine ice sheets are thought to be susceptible to greater change than terrestrial ice sheets as a consequence of (1) grounding line retreats caused by the disintegration of ice shelves, (2) the deepening of the bed inland of ice-sheet margins, which can lead to a positive feedback of ice decay, (3) the level to which the ice sheet is held buoyant by the surrounding sea, which may be enhanced if sea level rises, and (4) the nature and stability of ice streams, which may be underlain by weak marine sediments leading to enhanced flow speeds and complex dynamics. In the Pleistocene, the lower topography in West Antarctica was associated with significant fluctuations in ice-sheet size in contrast with a relatively stable East Antarctic Ice Sheet. While the ice stream configurations in East Antarctica remain largely similar (although they may reduce in flux), significant adjustments to the ice flow regime are predicted in West Antarctica. Consequently, ice flow in East Antarctica is associated strongly with topography, in which ice streams are most often constrained in topographic channels. In West Antarctica, the association between flow and topography is less clear. Here, ice streams flow over relatively flat terrain where weak water-saturated sediments dominate. Lateral migration of ice stream margins, without topographic controls, is entirely feasible in such places. The outcome is that the spatial pattern of glacial erosion and deposition in East Antarctica appears to have been relatively consistent over the past few glacial cycles, and indeed back through several million years, whereas substantial changes are likely to have occurred in West Antarctica.
Figure 6.5: Numerical modelling of Antarctic ice surface elevation during the last glacial–interglacial, including the ice-sheet configuration at the last glacial maximum. Taken from Huybrechts (2002) with permission.
252 M. J. Siegert
Numerical Modelling of the Antarctic Ice Sheet
253
Although Huybrechts’ (2002) investigation represents the most advanced numerical analysis of Pleistocene ice sheets in Antarctica, the model used has several limitations that prevent it from resolving ice streams well, or ice shelves at all. Increasing the complexity of the model to solve this issue might not lead to improved results, however, due to greater uncertainties that remain in terms of bed configuration and accumulation rates. The lesson to be learned is that glaciologists have an important choice concerning model complexity and the robustness of the model’s outputs, which must be made through consideration of the aims of the investigation. A broad-scale reconstruction at a continental level may still be best suited to a simple model, whereas information concerning time-dependent changes in discrete ice-sheet processes might be better served through a high-order model.
6.8. Summary Ice-sheet modelling allows quantitative predictions of how large ice masses behave and respond to environmental change. As many different ice-sheet models were developed in the early 1990s, the EISMINT programme was set up to establish best practice and commonality within the field. Through intercomparison exercises, the EISMINT programme established appropriate controls on ice-sheet modelling. One of the results is the development of freely available ice-sheet modelling software, which can be used in future to help test geological hypotheses concerning past changes in Antarctica. Although the large-scale behaviour of ice sheets can be reproduced well by numerical models, there are limitations to the accuracy of numerical ice-sheet modelling at a finer scale. This is because (1) ice-sheet models are derived from assumptions about the flow of ice which may not be accurate at all places in the ice sheet and (2) subglacial topography and other model inputs may be insufficiently well known to allow the model to work at a fine scale. The latter problem can be solved by obtaining high-resolution geophysical datasets of existing ice sheets to establish the subglacial topography, internal structure and surface elevation of the ice sheet. Such data acquisition has been ongoing for several decades. Ice-sheet models have been used to comprehend past changes, and test hypotheses based on geological data. For example, Huybrechts (1993) showed that climate warming of over 151C was required to decay the East Antarctic Ice Sheet into small discrete ice masses. As a result, the idea of restricted ice cover during the Pliocene was difficult to support by ice-sheet modelling. In additional exercises, Huybrechts (2002) revealed the sensitivity
254
M. J. Siegert
of the East and West Antarctic Ice Sheets to climate changes was noticeably different. In East Antarctica, the ice sheet appears resistant to even quite large changes, whereas modification to the size and shape of the West Antarctic Ice Sheet can be predicted with only small adjustments to boundary conditions. This sensitivity may characterize Antarctic glacial history throughout the Pleistocene and possibly much earlier. Ice-sheet models are used to predict how the cryosphere responds to and affects climate change. From climate information gathered from ice cores, there is reason to believe that ice sheets, oceans and the atmosphere interact with each other. It is therefore important that ice-sheet models be coupled with models of atmospheric and ocean circulation in order to predict how the ice–ocean–atmosphere system behaves in the future and has behaved in the past (e.g. DeConto and Pollard, 2003). Examples of such coupling are provided in later chapters, and these activities mark a clear way for future reconstructions of past ice sheets in Antarctica.
REFERENCES Arthern, R. J., Winebrenner, D. P., & Vaughan, D. G. (2006). Antarctic snow accumulation mapped using polarization of 4.3-cm wavelength microwave emission. J. Geophys. Res., 111, D06107. Bamber, J. L., Vaughan, D. G., & Joughin, I. (2000). Widespread complex flow in the interior of the Antarctic Ice Sheet. Science, 287, 1248–1250. Bennett, M. R. (2003). Ice streams as the arteries of an ice sheet: Their mechanics, stability and significance. Earth Sci. Rev., 61, 309–339. DeConto, R. M., & Pollard, D. (2003). Rapid Cenozoic glaciation of Antarctica induced by declining atmospheric CO2. Nature, 421, 245–249. Dowdeswell, J. A., & Siegert, M. J. (1999). The dimensions and topographic setting of Antarctic subglacial lakes and implications for large-scale water storage beneath continental ice sheets. Geol. Soc. Am. Bull., 111, 254–263. Fujita, S., Maeno, H., Uratsuka, S., Furukawa, T., Mae, S., Fujii, Y., & Watanabe, O. (1999). Nature of radio-echo layering in the Antarctic Ice Sheet detected by a two-frequency experiment. J. Geophys. Res., 104(B6), 13013–13024. Huybrechts, P. (1990). A 3-D model for the Antarctic Ice Sheet: A sensitivity study on the glacial interglacial contrast. Climate Dynamics, 5, 79–92. Huybrechts, P. (1993). Glaciological modelling of the Late Cenozoic East Antarctic Ice Sheet: Stability or dynamism? Geografiska Annaler, 75, 221–238. Huybrechts, P. (2002). Sea-level changes at the LGM from ice – Dynamic reconstructions of the Greenland and Antarctic Ice Sheets during the glacial cycles. Quaternary Sci. Rev., 21, 203–231. Huybrechts, P., Payne, A., & The EISMINT Intercomparison Group (1996). The EISMINT benchmarks for testing ice-sheet models. Ann. Glaciol., 23, 1–12.
Numerical Modelling of the Antarctic Ice Sheet
255
IPCC (2007). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Inter governmental Panel on Climate Change. In: S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, & H. L. Miller (Eds). Cambridge University Press, Cambridge, UK, 996 pp. Jacobs, S. S., Hellmer, H. H., Doake, C. S. M., Jenkins, A., & Frolich, R. M. (1992). Melting of ice shelves and the mass balance of Antarctica. J. Glaciol., 38, 375–387. Joughin, I., Gray, L., Bindschadler, R., Price, S., Morse, D., Hulbe, C., Mattar, K., & Werner, C. (1999). Tributaries of West Antarctic ice streams revealed by RADARSAT interferometry. Science, 286, 283–286. Le Meur, E., & Huybrechts, P. (1996). A comparison of different ways of dealing with isostasy: Examples from modelling the Antarctic Ice Sheet during the last glacial cycle. Ann. Glaciol., 23, 309–317. Leysinger Vieli, G. J. M. C., Hindmarsh, R. C. A., & Siegert, M. J. (2007). Three dimensional flow influences on radar layer stratigraphy. Ann. Glaciol., 46, 22–28. Lythe, M. B., Vaughan, D. G., & The BEDMAP Consortium (2000). BEDMAP – Bed Topography of the Antarctic. 1:10,000,000 Map. BAS (Misc.) 9, Cambridge. Mahaffy, M. W. (1976). A three dimensional numerical model of ice sheets: Tests on the Barnes Ice Cap, northwest territories. J. Geophys. Res., 81, 1059–1066. Miller, M. F., & Mabin, M. C. G. (1998). Antarctic Neogene landscapes – In the refrigerator or in the deep freeze? GSA Today, 8(4), 1–2. Pattyn, F. (2003). A new three-dimensional higher-order thermomechanical ice-sheet model: Basic sensitivity, ice-stream development and ice flow across subglacial lakes. J. Geophys. Res., 108(B8), 2382, doi:10.1029/2002JB002329. Ritz, C., Rommelaere, V., & Dumas, C. (2001). Modeling the evolution of the Antarctic Ice Sheet over the last 420,000 years: Implications for altitude changes in the Vostok region. J. Geophys. Res., 106, 31943–31964. Schoof, C. (2007). Ice sheet grounding line dynamics: Steady states, stability and hysteresis. J. Geophys. Res., 112, F03S28, doi:10.1029/2006F000664. Siegert, M. J. (1999). On the origin, nature and uses of Antarctic Ice-Sheet radioecho layering. Prog. Phys. Geogr., 23, 159–179. Siegert, M. J. (2001). Ice Sheets and Late Quaternary Environmental Change. John Wiley and Sons Ltd., Chichester, UK, 231 pp. Siegert, M. J., & Bamber, J. L. (2000). Subglacial water at the heads of Antarctic icestream tributaries. J. Glaciol., 46, 702–703. Siegert, M. J. & Payne, T. (2001). Validation of ice sheet models. In: Anderson, M. G. & Bater, P. D. (Eds). Model validation in hydrological science. John Wiley and Sons Ltd. Chichester, England, pp. 439–460. Siegert, M. J., Carter, S., Tabacco, I., Popov, S., & Blankenship, D. (2005a). A revised inventory of Antarctic subglacial lakes. Antarct. Sci., 17, 453–460.
256
M. J. Siegert
Siegert, M. J., Taylor, J., & Payne, A. J. (2005b). Spectral roughness of subglacial topography and implications for former ice-sheet dynamics in East Antarctica. Global Planet. Change, 45, 249–263. Vaughan, D. G., Bamber, J. L., Giovinetto, M., Russell, J., & Cooper, AP. R. (1999). Reassessment of net surface mass balance in Antarctica. J. Climate, 12, 933–946. Welch, B. C., & Jacobel, R. W. (2003). Analysis of deep-penetrating radar surveys of West Antarctica, US-ITASE 2001. Geophys. Res. Lett., 30(8), 1444, doi:10.1029/ 2003/GL017210.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00007-4
Chapter 7
The Antarctic Continent in Gondwanaland: A Tectonic Review and Potential Research Targets for Future Investigations F. M. Talarico1, and G. Kleinschmidt2 1
Dip. Scienze della Terra, Universita` di Siena, Via del Laterino 8, I-53100 Siena, Italy Institut f. Geowissenschaften d. Univesita¨t Frankfurt, Altenho¨ferallee 1, D-60438 Frankfurt/Main, Germany
2
ABSTRACT The geological record of the Antarctic continent has, for a long time, had a key role in supercontinent reconstructions. More recently, because of the welldocumented relevance of the polar regions’ processes in influencing the global changes of both ocean circulation and climate patterns, Antarctica has increasingly been of similar importance in the context of palaeoenvironmental and palaeoclimatic investigations, particularly those focused on the Cenozoic glacial evolution. After a description of the present-day geotectonic setting and of the main geological units before Gondwana amalgamation, the chapter focuses on the tectonic evolution of the Antarctic continent from its inclusion as part of the Gondwana supercontinent to the break-up of this landmass and the repositioning of Antarctica at southern polar latitudes since the Early Cretaceous. The geological evolution of the Antarctic continent is reviewed considering two main time periods: (i) c. 600–450 Ma, covering the processes which were active immediately before and during the amalgamation of Gondwana; and (ii) c. 450– 180 Ma, including all the major events that occurred after the final stage of Gondwana amalgamation to the time immediately before the break-up phase. A subsequent section addresses the last 180 Ma during which present-day Antarctica and the other southern continents and surrounding oceanic basins formed as a consequence of the fragmentation of Gondwana. After a general overview of the most significant plate tectonic stages and coeval magmatic products, the attention Corresponding author. Tel.: þ39-0577-233812; Fax: þ39-0577-233938;
E-mail:
[email protected] (F.M. Talarico).
258
F. M. Talarico and G. Kleinschmidt
is then devoted to one of the most investigated regions in Antarctica; the Transantarctic Mountains and the Ross Sea sector of the Western Antarctic Rift System. Some of the main persisting problems and potential research themes for on- and off-shore activities are discussed in the conclusions, which also include some of the most interesting palaeoclimatic issues that are essential to improving our understanding of the polar climate, ice ages and their influences on Earth’s climate system in the Cenozoic to present time.
7.1. Introduction The geological literature has for a long time demonstrated the fundamental role of Antarctica as a key continent for providing data and constraints to supercontinent reconstructions and the plate tectonic evolution of the Southern Hemisphere. Du Toit’s (1937) Gondwana reconstruction could be hypothesised on the basis of the results of the heroic period first expeditions (i.e. the Glossopteris-bearing Beacon sandstones collected during the last Scott’s expedition). The Mesoproterozoic Rodinia supercontinent (Dalziel, 1991; Moores, 1991) provides another example with fundamental pieces of evidence linked to the central position envisaged for Antarctica with respect to the other continental blocks, albeit the configurations differ for the proposed supercontinent. Progressively revealed by increased research activities of national and multinational expeditions following the 1957–1958 International Geophysical Year to those planned for the International Polar Year 2007/2008, the geological record of Antarctica has now a similarly valuable interest and importance in the context of palaeoenvironmental and palaeoclimatic investigations, particularly those focused on the initiation of Cenozoic glaciation, the stability of the polar ice sheets and the complex interaction among tectonic, sedimentary and climatic processes. In the following sections, the Antarctic geological record is reviewed in the context of the major phases of the late Neoproterozoic–Cenozoic plate tectonic evolution of the Southern Hemisphere. The chapter describes the evolution of the Antarctic continent from its inclusion as part of the Gondwana supercontinent to the break-up of this landmass and the repositioning of Antarctica at southern polar latitudes since the Early Cretaceous (c. 120 Ma). The chapter also highlights some of the most interesting palaeoclimatic issues, which are considered essential to understanding the polar climate, ice ages and their influences on Earth’s climate system throughout the Cenozoic.
The Antarctic Continent in Gondwanaland
259
The aim is to give a general overview which should be complementary to the more detailed information included in the accompanying chapters devoted to specific aspects of the geological record. Moreover, the review will highlight some of the main persisting open problems and those aspects of Antarctic tectonic evolution that are still not completely investigated or are under debate and may potentially be understood as major research themes for future on- and off-shore research.
7.2. The Present-Day Geotectonic Setting of Antarctica The Antarctic continent comprises three primary tectonic regions: (i) East Antarctica; (ii) West Antarctica and the associated West Antarctic Rift System (WARS); and (iii) the Transantarctic Mountains (Fig. 7.1). East Antarctica is thought to feature Precambrian continental lithosphere c. 35–45 km thick (Bentley, 1991), stable, coherent and topographically high (Cogley, 1984), that held a central position in the Palaeozoic supercontinent of Gondwana (Tingley, 1991) as it did in the Mesoproterozoic supercontinent Rodinia (Dalziel, 1991; Moores, 1991). In contrast, West Antarctica is an amalgamation of low-lying, 20–35 km thick, younger crustal blocks (Janowski and Drewry, 1981; Dalziel and Elliot, 1982). The West Antarctic Rift lithosphere compares to major Cenozoic continental rift systems (Behrendt, 1999). In the Ross Sea, the WARS borders the Transantarctic Mountains on their eastern side as a broad region of thinned continental crust associated with Cretaceous and episodic Cenozoic extension (Behrendt et al., 1991a,b; Behrendt, 1999). The WARS has high heat flow (83–126 m W m2) (Blackman et al., 1987; Berg et al., 1989) and thick (5–14 km) sedimentary basins with recent faulting (Cande and Leslie, 1986; Cooper et al., 1987; Hamilton et al., 2001). The Ross Archipelago (LeMasurier and Thomson, 1990) is currently active with fumarolic activity associated with alkaline volcanism at Mt. Erebus and Mt. Melbourne. The crust in the WARS is currently B1972 km thick (Trehu, 1989; Cooper et al., 1997). The Transantarctic Mountains are approximately 2,500 km long and 200 km wide, dividing East Antarctica from West Antarctica with peaks that rise over 4 km above sea level. Crustal thickness estimates under the Transantarctic Mountains vary between 20 and 45 km (ten Brink et al., 1993, 1997; Cooper et al., 1997; Busetti et al., 1999; Kanao et al., 2002; Bannister et al., 2003). These mountains sharply differ from most mountain ranges of similar size and lateral extent because their formation did not reflect any compressional orogenic phases, but rather a process thought to have been
260
F. M. Talarico and G. Kleinschmidt
Figure 7.1: Map showing the main geographic subdivision of Antarctica in three domains which also correspond to distinct tectonic regions: West Antarctica, Transantarctic Mountains and East Antarctica. The location of the two active volcanoes in the Ross Sea area is also shown (ME: Mt. Erebus; MM: Mt. Melbourne). Antarctic bed topography is based on the BEDMAP dataset (Lythe et al., 2000).
directly linked and causally related to the development of the WARS (e.g. ten Brink et al., 1997; Studinger et al., 2004, and references therein). The Transantarctic Mountains uplifted B6–10 km in an asymmetric tilt block formation and underwent denudation from the Cenozoic to the Cretaceous (Fitzgerald, 1992, 1995; Studinger et al., 2004). The later part of the uplift and denudation phases occurred under persisting spreading/extension within the western Ross Sea (Cande et al., 2000) and has been concomitant with voluminous sediment infilling from the Late Eocene/Oligocene to present (Barrett et al., 2000, 2001; Hamilton et al., 2001).
The Antarctic Continent in Gondwanaland
261
7.3. The Main Geological Units of Antarctica Before Gondwana Amalgamation As with all other continents, Antarctica consists of a number of Archean/ Early Proterozoic cratons (older than 1.5 Ga) surrounded by successively younger belts that formed and/or accreted to the continental margins, as products of convergent plate tectonic events such as subduction of oceanic crust underneath continental crust and/or collision of two former separated continents (Fig. 7.2). Cratons are confined to East Antarctica and, at least two among several, are generally accepted: the small Grunehogna Craton (ca. 3.0 Ga basement covered by flat-lying and undeformed sediments older than 1 Ga) interpreted as a fragment of the African Kalahari Craton, and the East Antarctic Craton s.s. The East Antarctic Craton s.s. is exposed only in limited areas including the southern Shackleton Range, the eastern Thiel Mountains (Ford, 1963), the Miller and Geologists Ranges of the central Transantarctic Mountains, near the continental margin in Enderby Land, in the southern Prince Charles Mountains, in the Vestfold Hills, in Wilkes Land, in Terre Ade´lie and in George V Land. The exposed basement, only occasionally covered by almost flat-lying platform sediments (e.g. Thiel Mountains), consists of high-grade gneisses and granulite facies metamorphic complexes showing radiometric ages of older than 1,500 Ma up to just over 3,500 Ma in Enderby Land. Recently, parts of Enderby, Kemp and Mac. Robertson Lands have been separated from the East Antarctic Craton and attributed to younger orogenic belts, i.e. the Grenvillian and Pan-African belts s.l. (Fitzsimons, 2000a; Boger et al., 2001, 2002; Boger and Miller, 2004), so that northern Enderby Land is now considered a separate craton (Napier Complex). Moreover, the results of more recent investigations suggest that, in the next few years, our appreciation of the present configuration of the East Antarctic Craton could change drastically and be subdivided in much smaller cratons including: (1) the Napier Craton (considered a fragment of Dwarhai Craton in India); (2) a craton comprising the southern Prince Charles Mountains, the southern Shackleton Range and perhaps the eastern Thiel Mountains; (3) the Vestfold Hills Craton; (4) the Mawson Craton (Fitzsimons, 2000a, 2003), probably consisting of the Miller and Geology ranges of the central Transantarctic Mountains, the shield areas of eastern Wilkes Land, Terre Ade´lie, George V Land, and their Australian counterpart, the Gawler Craton. Since the relationships between the Mawson Craton, the Vestfold Hills Craton, the southern Shackleton Range and the southern Prince Charles
Figure 7.2: Schematic geological map of Antarctica (modified from Kleinschmidt, 2007). The map shows the possible subice extent of the cratonic areas, the three ‘‘Grenvillian’’ belts, the Pan-African orogenic belts and the Ross Orogen. The map also shows the distribution of Beacon Supergroup and Ferrar Supergroup outcrops and of the two main Mesozoic orogenesis (Ellsworth or Weddell Orogeny and Antarctic Andean Orogen) which formed along the palaeo-pacific margin of Gondwana in West Antarctica, and the three intra-plate fracture zones in East Antarctica (Lambert Graben or Lambert Rift, West Antarctic Rift System, Jutul Penck Graben and the Rennick Graben). Abbreviations – BLM: Bertrab, Littlewood, Moltke nunataks, Bu: Bunger Hills, Ele: Elephant Island, Ge: Geologists Range, JPG: Jutul Penck Graben, KS: Kuunga Suture, LHB: Lu¨tzow-Holmbukta, MF: Matusevich strike–slip fault, Mi: Miller Range, PCM: Prince Charles Mts., Pe: Pensacola Mts., RG: Rennick Graben, Sh: Shackleton Range, Sn: Snow Island, Sø: Sør Rondane, Th: Thiel Mts., Wi: Windmill Islands; Lands – DML: Dronning Maud Land, EL: Enderby Land, GVL: George V Land, KL: Kemp Land, KWL: Kaiser-Wilhelm-II-Land, MBL: Marie Byrd Land, MRL: Mac.Robertson Land, OL: Oates Land, PEL: Princess Elizabeth Land, TA: Terre Ade´lie, VL: Victoria Land, WL: Wilkes Land.
The Antarctic Continent in Gondwanaland
263
Mountains are still totally unknown, a tectonic reconstruction is largely speculative and the cratonic areas could be even more reduced and further increased in number if, for example, 1.6–1.7 Ga-old orogens are separated. The orogenic belts separating the cratons as internal belts in East Antarctica or as elongated peripheral belts at the palaeo-Pacific margin of Gondwana developed during five major orogenic cycles spanning in time from c. 1.3–0.9 Ga (Grenvillian-aged orogens), through 500–600 Ma (Ross and the Pan-African orogens), 250–200 Ma (Ellsworth or Weddell Orogen) and 150–90 Ma (Antarctic Andean Orogen) to the period between 50 Ma and recent, documented only by the active plate-margin processes in the area north-west of the Antarctic Peninsula. As in eastern North America and other continents, the Grenvillian-aged orogens in Antarctica are considered to record the fundamental orogenic event marking the amalgamation of the Rodinia supercontinent in Meso-/ Neoproterozoic times. Since polyphase deformation, high-grade metamorphism and magmatism resembles somewhat those in the cratons, the areas formed during the Grenvillian orogeny have been often undifferentiated from the cratons. Moreover, until the early 1990s, only one very long Antarctic Grenvillian orogen was assumed, following the Antarctic coast as a 250 km wide strip from Coats Land in the west up to George V Land in the east, occasionally specially named the ‘‘Circum East Antarctic Mobile Belt’’ (Yoshida, 1992). Such an extension has been demonstrated to be incorrect for Terre Ade´lie and George V Land and unproven for Coats Land, where the tiny outcrops represented by three little groups of nunataks (Littlewood, Bertrab and Moltke) consist of 1,100 Ma rhyolites and granophyres (Storey et al., 1994), but absolutely undeformed, thus not consistent with the existence of an orogenic belt. The existence of a Circum East Antarctic Mobile Belt is no more supported by new geological evidence which, instead, indicates the presence of three distinct Grenvillian-aged orogens (Fitzsimons, 2000a) (Fig. 7.2): (1) the Maud Belt, extended from western to central Dronning Maud Land to possibly parts of Sør Rondane, interpreted as the trace of welding the Grundhogna Craton and the Shackleton Range/southern Prince Charles Mountains craton; (2) the Rayner Belt (Yoshida and Kizaki, 1983) that in Enderby, Kemp and Mac. Robertson Lands separates the Napier craton from the southern Prince Charles Mountains craton, and was verified in the northern Prince Charles Mountains and the Rayner Complex (e.g. Boger et al., 2002); (3) the Wilkes province belt (Fitzsimons, 2000a) of Wilkes Land that is exposed in the Bunger Hills and Windmill Islands and merges the Vestfold Hills and the Mawson cratons.
264
F. M. Talarico and G. Kleinschmidt
The Antarctic Continent in Gondwanaland
265
7.4. Antarctica in the Gondwana Supercontinent The fit of the large continental blocks within Gondwana is well established (e.g. Lawver et al., 1998) as three continents, Australia, India and Africa, fit well against the present-day rifted margins of Antarctica (Fig. 7.3). In contrast, due to considerable uncertainty over the number and position of some of the microplates within the West Antarctic region (e.g. Dalziel and Elliot, 1982; Storey, 1996; Storey et al., 1998), the Gondwana reconstruction is less well constrained along the Transantarctic margin of East Antarctica. In this region, the Ellsworth–Whitmore Mountains crustal block, Antarctic Peninsula, Thurston Island and Marie Byrd Land (subdivided into west and east portions, Divenere et al., 1996) are generally accepted as the main West Antarctic microplates, but other minor microplates (i.e. The Haag Nunatak, Berkner and Filchner microplates) are also used in reconstructions (e.g. De Wit et al., 1988). On the basis of palaeomagnetic data, geometry and geological constraints (particularly the presence of a large proportion of subduction-related rocks), the main West Antarctic microplates are generally retained in approximately their present location along the palaeo-Pacific margin of Gondwana in most models (e.g. Lawver et al., 1992). To obtain their Gondwana pre-break-up positions, both the Antarctic Peninsula and the Thurston Island block (Grunow et al., 1991) are rotated anticlockwise from their present-day position with respect to East Antarctica, the two microplates of Marie Byrd Land are restored along a strike–slip fault (Divenere et al., 1995), and Campbell Plateau, Chatham Rise, and North and South Islands of New Zealand can be reconstructed to Marie Byrd Land.
Figure 7.3: (A) Gondwana tight fit reconstruction (Early Jurassic). (B and C) Major initial stage in the break-up of Gondwana: (A) initial rifting stage (Late Jurassic); (B) change in the Gondwana break-up stress regime from dominantly north–south between east and west Gondwana to dominantly east–west with the two-plate system being replaced by a multiple-plate system (Early Cretaceous) (after Fitzgerald, 2002, with permission from the Royal Society of New Zealand). AP, Antarctic Peninsula; TI, Thurston Island; MBL, Marie Byrd Land; CR, Chatham Rise; CP, Campbell Plateau; SNZ, Southern New Zealand; NNZ, northern New Zealand; LHR, Lord Howe Rise; WS, Weddell Sea. Continent and microplate positions are from Lawver et al. (1992, 1998) with other information from Storey (1996).
266
F. M. Talarico and G. Kleinschmidt
Detailed information about the different reconstructions and tectonic scenarios for Antarctica in Gondwana can be found in Lawver et al. (1998) and Larter et al. (2002). We will here focus on the major Antarctic geological constraints, key regions and datasets which represent a well-established basis for reconstructing the southern continents in Gondwana. In this context, our review will also face some controversial aspects, which are presently under debate, concerning the reconstruction of the main phases during the amalgamation of the supercontinent. The geological evolution of the Antarctic continent is reviewed considering two main time periods: (i) c. 600–450 Ma, covering the processes which were active immediately before and during the amalgamation of Gondwana; and (ii) c. 450–180 Ma, including all the major events that occurred after the final stage of Gondwana amalgamation to the time immediately before the break-up phase. 7.4.1. Antarctica in the Late Precambrian–Early Palaeozoic (c. 600–450 Ma) Evolution of Gondwana The connections between Antarctica and the other southern hemisphere continents at the time of Gondwana are clearly documented by the so-called ‘‘Gondwanian’’ sequences which correlate with analogous stratigraphic successions in South America, Africa, India, Sri Lanka, Australia and New Zealand and by palaeontological evidence, such as the significant first appearance in Triassic time of the herbivorous reptile Lystrosaurus found in all Gondwanian continents, Antarctica included (Collinson, 1991; Retallack et al., 2005). Additional geological evidence for Gondwana’s reconstruction is provided by a number of older Antarctic geological provinces (such as the Archean Cratons and the Palaeozoic orogenic belts, e.g. the Ross– Delamerian Orogen) which fit tightly across a closed Southern Ocean. In comparison to this well-established evidence, the reconstruction of tectonic models for the initial stages of the formation of Gondwana is more problematic. The amalgamation phase necessarily involved the aggregation of various continental fragments which derived from the fragmentation and dispersal of a former supercontinent – variously named Ur-Gondwana (Hartnady, 1991), Katania (Young, 1995), Palaeopangea (Piper, 2000) or Rodinia (McMenamin and McMenamin, 1990). However, the precise configuration and modality of break-up of this supercontinent are yet not completely known and these uncertainties obviously propagate to the formulation of tectonic models for the constructive phase of Gondwana.
The Antarctic Continent in Gondwanaland
267
Since the focus of this paper is on the Cretaceous–Cenozoic record, we will here avoid reviewing Rodinia models (e.g. SWEAT: Dalziel, 1991; Hoffman, 1991; Moores, 1991; AUSWUS: Karlstrom et al., 1999; see also reviews by Dalziel, 1997 and Meert and Torsvik, 2003) or alternative models (e.g. Palaeopangea, Piper, 1982, 2000) and we will briefly summarize the main recent results in the reconstruction of the Late Precambrian–Early Palaeozoic tectonic evolution of Antarctica in Gondwana.
7.4.1.1. From Rodinia to Gondwana In spite of significant uncertainties about the precise reconstruction of the global palaeogeography in Neoproterozoic time and the still incomplete geochronological framework, most authors agree that the break-up of Rodinia led to the development of extensive passive continental margins which are documented in the late Neoproterozic (c. 750–600 Ma) record of most present-day continents (Dalziel, 1991, 1992, 1997; Powell et al., 1993; Meert and Van der Voo, 1997; Cawood, 2005). In Antarctica, following Dalziel’s hypothesis (1991), the first stage of Rodinia break-up involved a rifting phase which started at c. 750–725 Ma leading to the separation of Laurentian from the East AntarcticaþAustralia block and the formation of the intervening proto-Pacific ocean. This process was accompanied by the drift of the cratonic blocks, presently exposed in South America and Africa (Amazonian, Rio de la Plata and Western African) (i.e. West Gondwana) which eventually collided with India, Sri Lanka and East Antarctica (East Gondwana). Most workers agree that key evidence of this evolution is stored in the Mozambique Belt (Holmes, 1951) or East African Orogen (Stern, 1994) and that this extensive orogenic belt formed as a result of the closure of a ‘‘Mozambique Ocean’’ and subsequent collision and amalgamation of East and West Gondwana during the Pan-African event (Dalziel, 1992; Stern, 1994; Shackleton, 1996). A review of structural and geochronological data from East Antarctica (Dronning Maud Land and Lu¨tzow Holm Bay) and comparison with the adjacent (in Gondwana) Falkland Microplate and south-eastern Africa led Jacobs and Thomas (2002) to corroborate the proposal by Jacobs et al. (1998) of a southward continuation of the Mozambique belt into Dronning Maud Land in Antarctica. A Mozambique suture zone in the Shackleton Range was suggested by Grunow et al. (1996) and conclusive evidence was found by Talarico et al. (1999) who described relics of ophiolites consisting of serpentinites and amphibolites with N-type MORB to OIB geochemistry and a maximum
268
F. M. Talarico and G. Kleinschmidt
Sm–Nd age of c. 900 Ma. Metamorphic reworking of these rocks occurred under variable high P to medium P conditions in the eastern Shackleton Range; a stage of eclogite-facies metamorphism in the area has been recently reported by Schma¨dicke and Will (2006) in the central Shackleton Range. On the basis of these discoveries and thrust patterns in both the Shackleton Range and Western Dronning Maud Land, Kleinschmidt et al. (2002) proposed that the ophiolites may have formed part of a connection between the Palaeo-Pacific and the Mozambique oceans, in the way the Drake Passage links the present Pacific and Atlantic Oceans. The Mozambique Belt and its continuation in Antarctica shows a general N–S trending (Jacobs and Thomas, 2002). East of Dronning Maud Land, structural and petrological data indicate that the Lu¨tzow Holm Complex (Kriegsman, 1995) can be interpreted as representing the western end of a c. E–W, subperpendicular branch of the Mozambique Belt that might have extended to Prydz Bay, where Pan-African high-grade rocks have also been reported (Dirks and Wilson, 1995; Fitzsimons, 1997). The Lu¨tzow Holm–Prydz Bay Pan-African orogenic belt has provided sound evidence in contrast to the classical assumption that Eastern Gondwana formed during the consolidation of Rodinia in the Mesoproterozoic time and it remained tectonically stable until the modern continents rifted from Gondwana in the Mesozoic. The Lu¨tzow Holm–Prydz Bay–PanAfrican structures were initially considered part of a wider belt termed the Kuunga Orogen by Meert et al. (1995) or Kuunga Suture by Boger and Miller (2004), and interpreted as the result of the collision between East Antarctica (þAustralia) and India (þMadagascarþSri Lanka) at c. 535–520 Ma, after the amalgamation of India with the rest of Gondwana along the Mozambique suture (Fig. 7.4). More recent data actually suggest that there are three orogenic belts formed at about the same Pan-African period of 500–600 Ma in East Antarctica (Fig. 7.2): (1) the belt in the Shackleton Range–Dronning Maud Land–southern Sør Rondane–Lu¨tzow-Holmbukta region (‘‘East Antarctic Orogen’’ or ‘‘East Antarctic Belt’’ of Jacobs et al., 1998, renamed ‘‘Lu¨tzow Holm Belt’’ by Fitzsimons, 2000b); (2) the belt in the southern Prince Charles Mountains–Grove Mountains (‘‘Kuunga Suture’’ according to Boger et al. 2002); and (3) the belt in the Denman Glacier region, interpreted as prolongation of the Leeuwin Complex of Australia’s Pinjarra Orogen (e.g. Fitzsimons, 2000b). However, the exact extent of these orogens is doubtful because of the extensive ice cover, and this applies especially to the Denman Glacier belt. The Lu¨tzow Holm Belt is characterized by extensive thrust and nappe tectonics (Shackleton Range), by widespread and distinct late-orogenic collapse
The Antarctic Continent in Gondwanaland
269
Figure 7.4: Main tectonic stages of the amalgamation of Gondwana (modified after Boger and Millar, 2004, with permission from Elsevier). Abbreviated cratons – D: Dwarhai craton; G: Gawler craton; K: Kalahary craton; P/Y: Pilbara/Yilgarn craton; sPCM: southern Prince Charles Mountains.
270
F. M. Talarico and G. Kleinschmidt
structures (Shackleton Range), by thick molasse formations (Blaiklock Glacier Group: Shackleton Range; Urfjell Group: Dronning Maud Land), and by syn- and post-orogenic magmatism (Dronning Maud Land; Paech, 2004). Therefore, these Antarctic portions, formed during late Neoproterozoic to Cambrian, show typical characteristics of a collisional orogen. The three belts indicate the amalgamation of West Gondwana (South America, Africa and Grunehogna Craton) and East Gondwana (India, Australia and main East Antarctica). As proposed by Boger et al. (2001, 2002), Gondwana’s amalgamation may have taken place in two steps: the first before 550 Ma and the second after 550 Ma. The first step involved the amalgamation of West Gondwana with ‘‘Indo-Antarctica’’ (i.e. India and the northern Prince Charles MountainsþNapier Complex) documented by the Mozambique Belt and as its Antarctic prolongation, the Lu¨tzow Holm Belt. The second step led to the aggregation of these terranes to the rest of East Gondwana, i.e. the rest of East Antarctica and Australia and thus producing the Kuunga Suture. This model could explain the diacronous development of the Lu¨tzow Holm Belt, mainly older than 550 Ma in Dronning Maud Land and o550 Ma in the Shackleton Range, as well as the existence of an interleaved old alien element – the Grenvillian granophyres of Coats Land – maybe just an exotic terrane, or a mini-craton (Kleinschmidt, 2007).
7.4.1.2. The Antarctic record of the late precambrian–early palaeozoic evolution of the palaeo-pacific margin of Gondwana The Transantarctic Mountains, at the margin of the East Antarctic shield, represent a key element in providing an important but still cryptic record of Proterozoic and Early Palaeozoic supercontinent history in the period from c. 800 to 500 Ma. The existing dataset and data that can be provided by future research are from sedimentary, plutonic and volcanic assemblages that potentially reflect different tectonic events integral to the Rodinian– Gondwanan transformation: from the break-up of Rodinia to the development of a transitional margin and plate-margin activity during Gondwanan assembly. In the Transantarctic Mountains, constraints on the timing of Rodinia break-up are provided by mafic magmatism and siliciclastic sedimentation of Neoproterozoic age from two areas. In the Skelton Glacier area, basalts interlayered with Skelton Group sediments yielded a Sm–Nd model age of 700–800 Ma (Rowell et al., 1993). In the Nimrod Glacier area (Cotton Plateau), gabbros and basalts interlayered with sediments of the Beardmore
The Antarctic Continent in Gondwanaland
271
Group previously dated at 762 Ma (Sm–Nd isochron age, Borg et al., 1990) are now considered to have been emplaced at 668 Ma (U–Pb zircon age, Goodge et al., 2002). Thick siliciclastic successions were long interpreted as deep-water turbidities deposited in Proterozoic time along a rifted margin. Although the older parts of some successions may relate depositionally to the rifting process, recent investigations have demonstrated that some units are nearshore deposits and the depositional age of several major successions must be revised upward. For example, in the central Transantarctic Mountains, sandstones formerly included in the Beardmore Group and considered Neoproterozoic are now assigned to the Middle Cambrian or younger (Goodge et al., 2002). A transformation from drifting to active subducting mode is inferred for the late Neoproterozoic, starting at c. 700 Ma and continuing, most probably as a protracted contractional tectonic cycle including several discrete deformation events, with a tectonic climax during the Ross Orogeny, until the Ordovician (Goodge, 2001). The first stage of the conversion from a passive margin to an active convergent margin is signalled by the structural inversion of cratonmargin sedimentary successions in central Transantarctic Mountains. Younger tectonic events between 560 and 480 Ma are generally referred to a broadly defined Ross Orogeny, a long-lived tectonic process which developed along the active Gondwanan margin involving episodic deformation, calc-alkaline magmatism and syn-orogenic deposition of arcderived detritus in a sinistral-transpressive, continental-margin arc setting (Fig. 7.2). The orogenic belt is exposed from northern Victoria Land at the Pacific end up to the Pensacola Mountains at the Atlantic end. Westernmost Marie Byrd Land (Edward VII Peninsula) has to be considered as part of the same orogenic belt, from which it became isolated only much later, during a major phase of the evolution of the WARS around the end of the Cretaceous. The Ross Orogen is characterized by folds, thrusts, very low- to high-grade metamorphism, granitoids, terranes and flysch- and molasse-type sediments. Remarkable thrusts have been reported from Oates Land, which could be traced into the Australian continuation of the Ross Orogen, the Delamerian Orogen (Flo¨ttmann et al., 1993). The systematic distribution of high- and low-pressure types of metamorphism (e.g. Talarico et al., 2004) and of S- and I-type granitoids (e.g. Vetter and Tessensohn, 1987) led to the model of subduction of the palaeo-Pacific beneath East Antarctica. The igneous activity occurred as early as 560–550 Ma ago and is considered to be related to initial subduction of palaeo-Pacific lithosphere beneath the East Antarctic cratonal margin. In southern Victoria Land, the older plutons also include a peculiar suite of highly alkaline rocks (nepheline syenites and carbonatites)
272
F. M. Talarico and G. Kleinschmidt
(Koettlitz Glacier Alkaline Province, Cooper et al., 1997) and a stage of Precambrian rift-related magmatism and sedimentation has been documented by Cook (2007) in the Skelton Glacier area. In northern Victoria Land, the Ross Orogen is made up of three so-called terranes: the high- to medium-grade and granite-dominated Wilson Terrane to the west, the low-grade turbiditic Robertson Bay Terrane to the east, and the low-grade and volcanic-rich Bowers Terrane in between. These terranes are considered to be allochthonous or just adjacent palaeogeographic domains including an intra-oceanic island arc (Bowers Terrane) and an accretionary wedge (Robertson Bay Terrane) (Tessensohn and Henjes-Kunst, 2005). In northern Victoria Land, a unique occurrence of ultra-high-pressure rocks, including well-preserved mafic eclogites as lenses and pods within metasedimentary gneisses and quartzites, decorates the tectonic boundary between the inboard Wilson Terrane and the Bower Terrane in the Lanterman Range. Geological, petrological and geochronological studies indicate that mafic, ultra-mafic and felsic host rocks in this region underwent a common metamorphic evolution with an eclogite facies stage about 500 Ma ago at temperatures of up to about 8501C and pressures greater than 2.6 GPa (Di Vincenzo et al., 1997; Palmeri et al., 2003, 2007). 7.4.2. Antarctica in the Upper Palaeozoic–Mesozoic (c. 450–180 Ma) Evolution of Gondwana In this time window, the main geological events so far known in the Antarctic geological record include the formation of a continental basin where a dominantly fluvial succession, the Beacon Supergroup, was deposited in Devonian to Triassic times, and the development of two main deformational events generally known as the middle Palaeozoic Borchgrevink Orogeny and the Permo-Triassic Gondwanian Orogeny.
7.4.2.1. Borchgrevink Orogeny The occurrence in northern Victoria Land of a calc-alkaline magmatic suite of middle Palaeozoic (c. 350 Ma) age is considered as evidence of an orogenic event which is temporally well separated from the waning phase of the Ross Orogeny (Craddock, 1970). This suite includes dominant granitoids (Admiralty Intrusives) and minor felsic volcanic (Gallipoli Volcanics). The granitoids form several plutons which are mainly concentrated in the
The Antarctic Continent in Gondwanaland
273
Robertson Bay Terrane, although some cut across the tectonic boundary between the Bowers and Wilson Terrane, where the coeval volcanics are also concentrated. All the intrusions are characterized by isotropic fabrics and discordant contacts with respect to the surrounding metasediments suggesting a post-tectonic emplacement. These features and meagre evidence of radiometric data for concomitant metamorphism and deformation have so far limited a comprehensive reconstruction of the tectonic setting and development of the Borchgrevink Orogeny in northern Victoria Land. Nevertheless, the continental arc geochemical signature of the Devonian– Carboniferous magmatic suite indicates that the plutonism most likely occurred as the result of a renewed period of subduction activity along the palaeo-Pacific margin (Kleinschmidt and Tessensohn, 1987; Borg and DePaolo, 1991). Similar suites are known in Tasmania, New Zealand and the Campbell Plateau (Gibson and Ireland, 1996; Bradshaw et al., 1997) and in the Ford Range in Marie Byrd Land. A correlation between the Borchgrevink Orogeny of northern Victoria Land with the Tasmanian Orogeny was put forward by Findlay (1987). Elsewhere in Antarctica, marine sediments from the Ellsworth Mountains and Pensacola Mountains are considered to have been deposited in the same time window, within a basin that according to Elliot (1975) could have extended from the Weddell Sea to the Ross Sea.
7.4.2.2. Beacon Supergroup The Beacon Supergroup consists of dominantly continental sedimentary deposits which constitute a generally flat-lying, 2.5–3.5 km thick cover developed over a marked unconformity (Kukri Peneplane) above the Ross orogenic belt throughout most part of the Transantarctic Mountains (Fig. 7.2). Similar sequences are known in limited outcrops in East Antarctica (Prince Charles Mountains, Dronning Maud Land and Ellsworth Mountains) (Barrett, 1991) and as bedrock of the Cenozoic glaciomarine sediments in the Victoria Land Basin (VLB) in the Ross Sea as documented by the CRP-3 drill-hole (Cape Roberts Science Team, 2000). Outside Antarctica, similar sedimentary rocks, collectively called Gondwanian sequences, occur in South Africa, Australia and South America. The deposition of these sediments in Antarctica started in Devonian time with the Taylor Group, consisting dominantly of quartz-arenites and conglomerates. Deposited as the result of erosion and fluvial processes under arid and semiarid conditions (Campbell and Claridge, 1987), the Taylor Group accumulated in a series of basins along the palaeo-Pacific margin of
274
F. M. Talarico and G. Kleinschmidt
Gondwana. After the deposition of fossiliferous siltites, the subsequent stage corresponds to an erosional phase, probably related to a glacial event which led to the deposition of glaciogenic sediments in late Carboniferous–Early Permian time. In the Transantarctic Mountains, all the Carboniferous to Triassic sediments are formally known as the Victoria Group, which includes carbonaceous layers and feldspathic sandstones. The low sedimentation rate (c. 12.5 m/My) and absence of concomitant compressional deformation suggest that deposition occurred over a thick continental crust (Barrett, 1991) but different tectonic models have been proposed, including a passive margin (Isbell, 1999), intra-cratonic (Barrett, 1991; Woolfe and Barrett, 1995) or marginal/back-arc (Bradshaw and Webers, 1988).
7.4.2.3. Ellsworth or Weddell Orogeny The Ellsworth or Weddell Orogeny (or – in a larger context – also known as the Gondwanide Orogeny including components in southern Africa – Cape Fold Belt – and South America – Sierra de la Ventana Fold Belt) occurred in Permo-Triassic time ca. 250–200 Ma. As noted by Cawood (2005), this orogeny overlaps with the end Palaeozoic assembly of Pangea (Li and Powell, 2001), through ocean closure and accretion of Gondwana, Laurussia (LaurentiaþBaltica) and Siberia, as well as completion of terrane accretion in the Altaids. Stratigraphic and geochronological data (Dalziel, 1982; Dalziel and Elliot, 1982; Storey et al., 1987; Trouw and De Wit, 1999; Johnston, 2000) indicate that Permo-Triassic deformation of variable intensity and distribution occurs throughout West Antarctica and the adjoining Cape Fold Belt of southern Africa. In Antarctica, this orogenic event is well documented in the Ellsworth–Whitmore Mountains and in the Pensacola Mountains, where upright to inclined folds with axial planar cleavage are inferred to have formed in a dextral transpressive environment (Curtis, 1998) (Fig. 7.2). The Ellsworth–Pensacola Mountains chain represents the fold belt, while the Whitmore Mountains represent the magmatic arc of an Andean-type orogen. It merges with the Ross Orogen in the Pensacola Mountains, where it partly overprints the older Ross-aged structures. Elsewhere, deformation is heterogeneously distributed, with Storey et al. (1987) noting that in the Antarctic Peninsula, unconformities previously ascribed to the Gondwanide Orogeny are younger and that the only event related to Gondwanide deformation s.s. is regional metamorphism at 245 Ma of parts of the Trinity Peninsula Group.
The Antarctic Continent in Gondwanaland
275
The Ellsworth Orogen trends noticeably crossways to the palaeo-Pacific margin of Antarctica as indicated by the Ross Orogen. This obliqueness is due to secondary rotation, as proven by palaeomagnetic investigations by Funaki et al. (1991) and confirmed by Randall and MacNiocaill (2004).
7.4.2.4. The Antarctic Andean Orogen The orogen of the Antarctic Andes occupies the entire Antarctic Peninsula down to the Walgreen Coast (Fig. 7.2). It formed mainly in three episodes: (i) Late Jurassic through Early Cretaceous (150–140 Ma), (ii) mid-Cretaceous (B105 Ma) and (iii) BTertiary (B50 Ma to recent), and is partly still active (e.g. Birkenmajer, 1994; Vaughan and Storey, 1997). Thus, it represents the youngest growth zone of the continent. The Antarctic Andes are a typical subduction orogen accompanied by orogenic magmatism in the form of granitic plutons and volcanic rocks. In detail, the deformation and metamorphism are very complicated, because they are polyphase. Folding and thrust faulting is reported mainly from the southern portion of the Antarctic Peninsula (Palmer Land and Alexander Island) and from the extreme north (Trinity Peninsula and eastern South Shetland Islands). The distribution of related metamorphism is also heterogeneous, including high-pressure metamorphism with blueschists characteristic of subduction complexes, e.g. on Elephant Island (Trouw et al., 1991).
7.4.2.5. The plate-tectonically active parts of Antarctica The only plate-tectonically active part of Antarctica is situated north-west of the Antarctic Peninsula, in the South Shetland Islands (from Snow Island in the south-west up to Elephant Island in the north-east) and the Bransfield Strait (Fig. 7.2). North-west of the South Shetland Islands, a small section of the ocean floor, called the Drake Plate (i.e. the remnant of the older, but largely subducted Phoenix Plate), is being subducted at the South Shetland Trench beneath the Antarctic Plate. Related, mainly andesitic, volcanism forms the island arc of the South Shetland Islands. Parts of the South Shetland Islands (part of Livingston Island and Elephant Island) belong – as does the Peninsula itself – to earlier stages of the Antarctic Andean orogeny and consist of strongly deformed Jurassic trench sediments.
276
F. M. Talarico and G. Kleinschmidt
The Bransfield Strait is located south of the subduction-related volcanic island arc and forms an active extensional basin accompanied by tholeiitic volcanism, partly submarine, partly as active island volcanoes (e.g. Penguin Island and Deception Island; Smellie and Lo´pez-Martı´ nez, 2002). The Bransfield Strait often is regarded as a classic example of a back-arc basin, but recently this has been disputed (Gonza´les-Casado et al., 2000).
7.5. Antarctic Record of Gondwana Break-Up and Dispersal of the Southern Hemisphere Continents This section addresses the 180 Ma to recent time window during which present-day Antarctica and the other southern continents and surrounding oceanic basins formed as a consequence of the fragmentation of Gondwana, and the tectonic processes involved in the drift and dispersion of the various continental fragments. After a general overview of the most significant plate tectonic stages and coeval magmatic products, the attention is devoted to one of the most investigated regions in Antarctica, the Transantarctic Mountains and the Ross Sea sector of the Western Antarctic Rift System.
7.5.1. Summary of Main Tectonic Stages of Gondwana Break-up The first major tectonic stage in the break-up of Gondwana corresponds to an initial rifting phase that started in the Weddell Sea, initially as a back-arc basin, in the Late Jurassic (Fig. 7.4) (Lawver et al., 1991). This stage involved right-lateral transtension as East Gondwana (Antarctica, Australia, India and New Zealand) and West Gondwana (South America and Africa) moved apart with stretching beginning in the north and propagating southward (Lawver et al., 1992). Initial break-up involved complex geodynamic evolution characterized by the rotation and translation of several microplates, such as the Ellsworth Mountains block, a displaced part of the Gondwanide fold belt. The original position of the various microplates is still controversial, but the Ellsworth–Whitmore Mountains crustal block most likely originated from near south-eastern Africa (e.g. Curtis and Storey, 1996). Rotation of West Antarctic microplates must have been accomplished by c. 165 Ma (the time of opening of the Weddell Sea), and rotation of the Ellsworth–Whitmore Mountains crustal block was finished before translation into its present position by 175 Ma (Grunow et al., 1987). Rotation of microplates did not apparently involve the production of oceanic crust
The Antarctic Continent in Gondwanaland
277
(Marshall, 1994) but may have occurred as block rotation with controlling faults concealed beneath Mesozoic sedimentary basins in the Weddell Sea (Storey, 1996). The break-up has been explained by several authors as the result of a hot mega-plume, which, according to Storey and Kyle (1999), could have promoted domal uplift and formation of a triple junction. According to Dalziel et al. (1999), the Gondwana plume may have also caused or expedited formation of the early Mesozoic Gondwanide fold belt, due to the buoyancy of a hot plume acting on the downgoing slab and causing it to flatten. In both cases, plume activity led to the production of magma batches reflecting different degrees of plume–lithosphere interaction, which migrated along crustal shear zones to ultimately form the various large igneous provinces. The plume-related magmatic products are represented by huge withinplate mafic and felsic magmatic provinces in many Gondwana continents as well as Antarctica (Cox, 1988; White and MacKenzie, 1989; Storey, 1996). In the Transantarctic Mountains (LeMasurier and Thomson, 1990), Jurassic mafic rocks are generally known as the Ferrar Supergroup or the Jurassic Ferrar Large Igneous Province (FLIP). They are divided into a volcanic component, the Kirkpatrick Basalt – preceded by extensive phreatomagmatic volcanoclastic rocks (Elliot et al., 2006; Viereck-Go¨tte et al., 2007) – and the intrusive dolerite (diabase) sills and dikes of the Ferrar Dolerite. Geochemically, the FLIP is unusual with upper crustal-like characteristics such as high large ion lithophile element concentrations and enriched isotopic signatures (87Sr/86SrW0.709). The crust-like signature suggests derivation from an enriched lithosphere, possibly connected to subduction along the Pacific margin of Gondwana during the Palaeozoic and Mesozoic. The mafic rocks are concentrated within a long linear belt exposed in Tasmania, Antarctica and South Africa. Cox (1988) considered that the linear pattern of the Ferrar and Tasman provinces could not be compatible with classic circular plumes and proposed a hot line rather than a hot spot. A number of rifts that intersected at the Dufek Massif (Elliot, 1992) could have favoured the development of zones of weakened lithosphere, which acted as pathways for lateral migration of magmas derived from lithospheric sources. In spite of the still not completely understood tectonic setting, it is important to note that, similarly to other continental flood basalt provinces, all the mafic products formed during a short period of eruption (Tasman province: 175718 Ma; Ferrar Supergroup: 180–183 Ma; Dronning Maud Land province: 17772 Ma; Karoo province: 18272 Ma) (Hergt et al., 1989; Hooper et al., 1993; Heimann et al., 1994). Coeval felsic intrusions considered to be rift-related are known in West Antarctica (Storey et al., 1988) and southern South America (Chon-Aike or Tobı´ fera province) (Gust et al., 1985).
278
F. M. Talarico and G. Kleinschmidt
The second major geodynamic stage occurred in the Early Cretaceous (Fig. 7.4), as a consequence of the change of the Gondwana break-up stress regime from dominantly north–south between East and West Gondwana to dominantly east–west with the two-plate system being replaced by a multiple-plate system (e.g. Lawver et al., 1992). This modification in stress regime is thought to have induced large-scale ductile deformation concentrated along shear zones in the Antarctic Peninsula (Storey et al., 1996) and thin-skinned deformation and inversion of existing sedimentary basins such as the Latady Basin (Kellogg and Rowley, 1989). By c. 110 Ma, the microplates of West Antarctica had nearly reached their present location with respect to East Antarctica. In the same time period, separation also began between India and Antarctica (Lawver et al., 1991). Initial stretching between Australia and Antarctica began as early as 125 Ma (Stagg and Willcox, 1992), but sea-floor spreading was delayed to c. 95 Ma (Cande and Mutter, 1982; Veevers et al., 1990; Royer and Rollet, 1997), in the Ross embayment (Elliot, 1992), as well as extension between the Lord Howe Rise and northern New Zealand (Lawver et al., 1992). By the Late Cretaceous, Antarctica had reached its final polar location and configuration, and the final stage of break-up was completed when New Zealand (Campbell Plateau) rifted from Marie Byrd Land at 84 Ma (e.g. Stock and Molnar, 1987; Lawver et al., 1991). In this geodynamic context, four particularly large and conspicuous intraplate fracture zones have been investigated and all related to extensive and prolonged extensional regimes spanning in time from Mesozoic to present (Fig. 7.2). These major extensional zones include: the Lambert Graben or Lambert Rift (East Antarctica); the WARS and its main part, the Ross Sea Rift (Pacific sector); the graben of Jutulstraumen and Penckmulde (occasionally called Jutul Penck Graben; Atlantic sector); the Rennick Graben as the main element of a strike–slip fault system in Victoria and Oates Lands. The Lambert Graben developed in the East Antarctic Craton and is filled by sediments of the Permo-Triassic Beacon Supergroup. Faulting started already during the early Palaeozoic, reached its peak in the Permian and continued to the Early Cretaceous (Hofmann, 1996). Possibly, subglacial Lake Vostok belongs to the same rift system, but somewhat offset. The continuation of the Lambert Graben is the Indian Mahanadi Rift south-west of Calcutta in the state of Orissa (Hofmann, 1996), filled with sediments of the same type and age as the Lambert Graben. The reconstruction of the Gondwanan India–Antarctica fit by these graben systems coincides with
The Antarctic Continent in Gondwanaland
279
reconstructions by Archean to Early Proterozoic elements for Rodinia. That means, interestingly enough, that the relationships of Antarctica and India during Rodinian and Gondwanan times do not differ substantially. The Ross Sea Rift is extremely wide (about 1,000 km). Its subsidence started during the late Mesozoic (about 140 Ma ago), reached its main activity in the Early Tertiary (about 40 Ma ago) and produced an enormous relief at its western shoulder. The difference in altitude between the tops of the Transantarctic Mountains and the floor of the adjacent Ross Sea exceeds 14 km. The crustal extension is combined with alkaline intra-continental volcanism, which is still active at Mt. Erebus (3,794 m) and at Mt. Melbourne (2,732 m), both located in Victoria Land (Kyle and Cole, 1974; Wo¨rner et al., 1989; Kyle, 1990a,b; Tessensohn and Wo¨rner, 1991). The Jutul Penck Graben of western Dronning Maud Land originated probably around 140 Ma ago or a little bit later (Jacobs and Lisker, 1999). The graben marks the boundary of the Grunehogna Craton towards the south-east and thus it follows a much older geological structure. A possible continuation of the Jutul Penck Graben into the still active East African rift system is under discussion (Grantham and Hunter, 1991). By no means is this out of the question, because parts of the East African rift system were already active in the Jurassic (Ring and Betzler, 1993). The strike–slip fault system of Victoria and Oates Lands runs obliquely to the Ross Sea Rift and is cut by it. The principle element of the system is the Rennick Graben, which is presently active, as demonstrated by earthquakes in 1952, 1974 and 1998. The graben contains downfaulted Ferrar volcanics and sediments of the Beacon Supergroup, which have been spectacularly folded and squeezed onto the graben shoulders (Rossetti et al., 2003). This demonstrates alternating dextral transpression and transtension (with formation of pull-apart basins), being parts of a complicated strike–slip system in detail (Rossetti et al., 2003). The 1974 quake happened some 120 km to the west at the parallel structure of the Matusevich Glacier. There is an ongoing discussion, as to whether this strike–slip fault system constitutes the continuation of oceanic fracture zones between Australia and Antarctica (e.g. the Tasman Fracture Zone) into the continental crust of Antarctica (Salvini et al., 1997; Salvini and Storti, 2003; Kleinschmidt and La¨ufer, 2006). 7.5.2. Tectonic Evolution and Magmatism in the Ross Sea Sector during the Cenozoic With a similar form to the Weddell Sea, the Ross embayment is one of the most striking morphological expressions of the WARS, a region of thin, and,
280
F. M. Talarico and G. Kleinschmidt
by inference, extended continental crust whose regional boundaries are difficult to define precisely and may have been different for different episodes of extension (Fig. 7.5). Three major episodes of extension have been proposed, with rifting starting in the Middle Jurassic, in coincidence with the onset of Gondwana break-up and the associated Middle Jurassic magmatism (Ferrar Group), and with subsequent episodes in the Cretaceous and late Cenozoic. The active rifting may continue to the present as suggested by the active volcanoes along the western margin of the Ross Sea. However, most extension and thinning of the Ross Sea crust is considered to have taken place during the Mesozoic rift period and the tectonic connection between the Ross and Weddell embayments is largely unclear (Behrendt et al., 1992; LeMasurier, 2007, and references therein). In particular, our knowledge of the Jurassic episode is problematic since, although the location of rifting is well constrained in the South Africa – Queen Maud Land region, neither the location nor the amount of extension is well known in the Transantarctic Mountains and Ross Sea. By comparison, the Cretaceous episode is better documented (LeMasurier, 2007, and references therein) and is considered to be closely related to the initial stage of stretching and rifting in the Weddell Sea region, where disruption of the plate margin involved the rotation and translation of the several crustal blocks forming the present West Antarctica, including the Ellsworth/Whitmore Block, which was translated to the boundary between the Ross and Weddell embayments. Since movement of the West Antarctic crustal blocks was largely completed by 110 Ma, later tectonic activity in the WARS has been restricted to the Ross sector of the TAM and the margin of the Ellsworth–Whitmore Mountains, with the continuation of the TAM into the Weddell Sea region considered a remnant Jurassic rift (Schmidt and Rowley, 1986; LeMasurier, 2007). The first direct evidence of rifting in the Ross embayment is documented in Marie Byrd Land where a voluminous suite of mafic dikes and anorogenic A-type granites (the so-called ‘‘Byrd Coast Granites’’) have been dated as early as 10775 to 102–95 Ma (Storey et al., 1999). Considering Bradshaw’s (1989) suggestion that a sea-floor spreading centre was subducted beneath Mesozoic New Zealand around 105 Ma, the production of the mafic dykes could have been linked to the subduction of such a spreading centre at 10775 Ma, followed by the emplacement of the anorogenic silicic rocks at 102–95 Ma. On the basis of the age of Anomaly 33 (83 Ma) identified off Campbell Plateau, the earliest sea-floor spreading between Campbell Plateau and Marie Byrd Land must be B85 Ma. The unequivocal reconstruction of
Figure 7.5: (A) Map of the Transantarctic Mountains and West Antarctic Rift System (after Fitzgerald, 2002, with permission from the Royal Society of New Zealand). Structural features along the TAM and in the West Antarctic Rift System are from Tessensohn and Wo¨rner (1991), Fitzgerald (1992), Fitzgerald and Baldwin (1997) and Salvini et al. (1997). DAZ: Discovery accommodation zone from Wilson (1999). Basin positions in the Ross Sea (light grey) are from Davey and Brancolini (1995) and include North Basin (NB), Eastern Basin (EB), Central Trough (CT) and the Terror Rift (TR) lying within the Victoria Land Basin (VLB). The approximate outline of the Wilkes subglacial basin is shown inboard of the TAM in a dot pattern. West Antarctic microplates are after Dalziel and Elliot (1982); see Fig. 7.3 for abbreviations. (B) Generalized crustal profile A–A (in A) across the Ross Sea is after Cooper et al. (1991). Early rift grabens lie beneath regional unconformity U6 (heavy line). Late rift-faults and intrusive rocks deform the VLB and some small basement grabens. Note intra-basement reflector (dashed line) on eastern edge of VLB, possibly a detachment fault extending under the entire basin. DSDP site 270 is projected onto this crosssection.
282
F. M. Talarico and G. Kleinschmidt
the Campbell Plateau against Marie Byrd Land and the remarkable match of the deduced ocean–continent boundary as well as the alignment of the eastern edge of the Eastern Ross Basin with the northern edge of the Campbell Basin on the Campbell Plateau are evidence that the entire Eastern Ross Basin should have experienced extension prior to initiation of the Pacific–Antarctic Ridge at B85 Ma. The Pacific–Antarctic spreading centre played a major role in the region in late Cenozoic time. With its initiation, a triple junction developed with three extensional arms (Australian–Antarctic spreading centre, Australian–Pacific arm and the Pacific–Antarctic spreading). Geochronological constraints are provided by early sea-floor spreading anomalies identified between Tasmania and the South Tasman Rise. Key tectonic events include (i) the jumping of the Australia–Antarctica spreading centre about Anomaly 30 (B65 Ma) to between the South Tasman Rise and northern Victoria Land, (ii) the completion of sea-floor spreading in the Tasman Sea at Anomaly 24 time (B54 Ma) as a consequence of the reformation of the ridge–ridge–ridge triple junction as a new spreading centre that went between Campbell Plateau and the Tasman Sea anomalies and (iii) the opening of the B40 km wide Adare Rift within ‘‘older’’ sea-floor off the western edge of the Ross Sea sometime after Anomaly 13 (B33 Ma) (Cande et al., 2000; Cande and Stock, 2006). Satellite gravity data suggest that the Adare Rift steps east, and then is transformed westward along the southern margin of the Coulman High to the region of the Terror Rift, a comparable structure in width, located along the western edge of the VLB in the western Ross Sea. Striking products of the large-scale Cenozoic geodynamic evolution in the Ross Sea area are four major sedimentary basins which developed by rifting of previously thinned continental crust, possibly along pre-existing crustal faults, and coincide with major localized crustal thinning (to less than 10 km), and, in the case of the VLB (the westernmost one), with high heat flow and alkaline volcanism (Blackman et al., 1987; White, 1989; Cooper et al., 1991; Behrendt et al., 1993; LeMasurier, 2007). Extensive late Cenozoic volcanism is also inferred from aeromagnetic data. These basins include from E to W: the Eastern Basin – underlying most of eastern Ross Sea; the Central Trough – running north–south discontinuously through central Ross Sea; the Northern Basin – underlying the north-eastern Ross Sea margin; and the VLB underlying south-western Ross Sea, adjacent to the Transantarctic Mountains. The region has been extensively investigated through a number of geophysical surveys. But, due to the lack of appropriate age data for the main seismostratigraphical units, inferences on the timing of Cenozoic rifting in the entire Ross Sea remain largely speculative. Based on the results of the
The Antarctic Continent in Gondwanaland
283
Cape Roberts Drilling Project (Barrett et al., 2001), the VLB is now considered to be mostly late Eocene or younger in age. Recent drilling on the western margin of this basin has indeed indicated an onset of subsidence at about 34 Ma, significantly younger than the onset of uplift of the adjacent Transantarctic Mountains (55 Ma) (Fitzgerald, 1992) and thus identifying an issue in the relationship of the uplift of rift margin mountains to the subsidence of the adjacent rift basin. However, the effect of uplift episodes at this margin on the sedimentary sequences recorded may need to be considered. The fact that the basin extends only from the Ross Island region to Terra Nova Bay, whereas the Transantarctic Mountains continue further north and south, would indicate that the basin may have originated from other processes in addition to simple extension. A transtension or ‘‘pull-apart’’ process has been suggested, limited by transfer mechanisms associated with the igneous activity recorded in the north by the ‘‘Polar Three’’ anomaly off Coulman Island and in the south by the magnetic anomalies south of the Ross Sea Fault (Bosum et al., 1989). Another significant geological component in the Ross Sea area tectonic evolution is represented by an extensive alkalic magmatic province, one of the largest in the world and including two active volcanoes (Mt. Erebus at Ross Island and Mt. Melbourne on the Ross Sea coast in northern Victoria Land; LeMasurier and Thomson, 1990). These rocks, either exposed or suggested by aeromagnetic studies in the area under the Ross Sea and West Antarctic Ice Sheet (Behrendt et al., 2002; Bell et al., 2006), occur on either side of the WARS in Marie Byrd Land (West Antarctica) and in the western Ross Sea. In the western Ross Sea, the oldest rocks are restricted to coastal areas in northern Victoria Land where they consist of Eocene to Oligocene alkali intrusive rocks (Meander Intrusive), interpreted as the eroded remnants of subvolcanic magmatic complexes (Mu¨ller et al., 1991; Rocchi et al., 1999, 2002). Elsewhere, the alkaline magmatism is predominantly volcanic and has been subdivided into the informal, but geographic and petrologic distinct, Hallett, Melbourne and Erebus volcanic provinces (Kyle, 1990b). Two small isolated occurrences of basalts dated at 16–20 Ma occur around the head of the Scott Glacier (Stump et al., 1980). In the Erebus volcanic province (Kyle, 1990a), there is a continuous eruptive sequence from 19 Ma to the present day, with the Mt. Erebus crater showing a persistent anorthoclase phonolite lava lake. The oldest exposed rocks of this province occur on the northern slopes of Mt. Morning where 19 Ma trachyandesites lavas are intruded by 16–18 Ma trachyte dikes. Volcanic ash (tephra) layers in CRP drill cores demonstrate evidence for older Miocene trachytic eruptions. The tectonic setting suggests that the southern extent of the volcanism may be controlled along a major transfer
284
F. M. Talarico and G. Kleinschmidt
fault, which coincides with the southern boundary of the Terror Rift (Wilson, 1999). Many of the larger volcanic centres have a radial distribution around Mt. Discovery or Mt. Erebus. The radial distribution is interpreted as a result of upwelling of a mantle plume (the Erebus plume), which was centred under Mt. Discovery prior to 4 Ma and then migrated to its present position under Mt. Erebus (Kyle, 1990a). 7.5.3. Tectonic Evolution of the Transantarctic Mountains The Transantarctic Mountains, one of the major young mountain chains on Earth, separate East Antarctica from the West Antarctica Rift System and the grounded East Antarctic Ice Sheet from the marine-based West Antarctic Ice Sheet over a substantial portion of the Antarctic interior along the margin of the Ross embayment (Fig. 7.5). The regional structural architecture of the Transantarctic Mountains remains poorly known in most regions because of the extensive ice cover. The East Antarctic Ice Sheet hides the structure of the mountains along the Polar Plateau, preventing the identification of the extent of mountain structures into East Antarctica. In the McMurdo Sound coastal area of the Ross embayment, thin but extensive piedmont glaciers obscure the structural boundary (‘‘Transantarctic Mountains Front’’) with the off-shore Victoria Land rift basin. The main drainage for the East Antarctic Ice Sheet into the Ross embayment of West Antarctica is through outlet glaciers carved through the mountains. It has long been inferred that these outlet glaciers developed where faults cut transverse to the mountain trend (e.g. Gould, 1935; Gunn and Warren, 1962; Grindley and Laird, 1969; Davey, 1981; Wrenn and Webb, 1982; Cooper et al., 1991; Tessensohn and Wo¨rner, 1991; Fitzgerald, 1992), and the occurrence of pseudotachylites has been proved a valuable tool to constrain the age of some faults (Di Vincenzo et al., 2004). In most cases, however, there is little direct evidence either for the existence of a fault or of its age. This is evidently a key issue to be addressed in order to determine if appropriate structures are present to allow differential uplift of discrete mountain blocks (e.g. van der Wateren et al., 1999), to understand the localization of valley incision and how erosion has contributed to mountain uplift, and to provide constraints on the development of the pathways for drainage of the East Antarctic Ice Sheet. In contrast to other young mountain belts, such as the Alpine–Himalayan system and the North American and Andean Cordilleras, which formed at convergent plate boundaries, there is a tight genetic link between the uplift of the Transantarctic Mountains and intra-plate processes associated with the
The Antarctic Continent in Gondwanaland
285
rifting within the Antarctic plate. These mountain ranges are generally interpreted as a high-relief rift flank uplift (van der Wateren et al., 1999), which occurred during the Mesozoic–Cenozoic break-up of the Gondwana supercontinent (Cooper et al., 1987, 1991; Tessensohn and Wo¨rner, 1991; Davey and Brancolini, 1995; Fitzgerald and Stump, 1997). The development of the Transantarctic Mountains in an extensional, rather than a contractional, tectonic regime was already recognized by pioneering Antarctic geologists, who interpreted the mountain chain as a fault-bounded horst block (David and Priestley, 1914; Gould, 1935). More recent structural investigations indicate that the Transantarctic Mountains consist of a linear to curvilinear chain of asymmetric tilt blocks bounded on the West Antarctic edge by a major normal fault zone and subdivided by transverse faults (Fitzgerald et al., 1986; Tessensohn and Wo¨rner, 1991; Fitzgerald, 1992; Tessensohn, 1994a,b; Fitzgerald and Baldwin, 1997). Active rift tectonism and mountain uplift have been inferred from the presence of active volcanism and Neogene–Quaternary age faulting in the western portion of the rift and the Transantarctic Mountains (Behrendt and Cooper, 1991; Davey and Brancolini, 1995; Jones, 1997). The Cenozoic–Cretaceous asymmetric uplift and subsequent erosion exposed basement rock and older sediments along the coastward side of the Transantarctic Mountains, leaving younger sediments only on the inland side. Apatite fission track analysis in the McMurdo Sound area indicate an uplift of B6 km since c. 55 Ma, though other sectors of the Transantarctic Mountains record denudation events in the Late Cretaceous also (Fitzgerald, 1992, 1995; Studinger et al., 2004) (Fig. 7.6) and in the Early Cretaceous (e.g. Scott Glacier area: Stump and Fitzgerald, 1992; Fitzgerald and Stump, 1997). The cause of this uplift and denudation is the subject of continuing debate, the reconstruction of Cenozoic tectonic processes being complicated by the complex interplay between a number of factors including the regional plate geodynamics, rifting style, erosion rates, subsidence and formation of thick sedimentary layers, the volcanic activity and the glacial processes. The possible mechanisms for the uplift include thermal buoyancy due to conductive or advective heating from the extended upper mantle of the hotter West Antarctic lithosphere (Stern and ten Brink, 1989; ten Brink and Stern, 1992), simple shear extension (Fitzgerald et al., 1986), isostatic rebound due to stretching of the lithosphere through normal faulting (Bott and Stern, 1992), plastic necking (Chery et al., 1992), elastic necking (van der Beek et al., 1994) and rebound response to erosion (Stern and ten Brink, 1989). In the McMurdo Sound area, some of these mechanisms are based on specific assumptions about the crustal and upper mantle structure beneath the ‘‘Transantarctic Mountains Front’’, as well as about the timing of the
286
F. M. Talarico and G. Kleinschmidt
Figure 7.6: Schematic diagram showing the variation of exhumation events along the TAM at different localities (after Fitzgerald, 2002, with permission from the Royal Society of New Zealand). SCG, Scott Glacier area; BDM, Beardmore Glacier Area; SHG, Shackleton Glacier area; SVL, Southern Victoria Land; TNB, Terra Nova Bay; NVL, Northern Victoria Land. A relative scale only is shown for exhumation as the amount of exhumation at any one locality will vary across the range. Early or late Cretaceous exhumation is not always present throughout an area (e.g. Scott Glacier region; see Fitzgerald and Stump, 1997).
rift-related processes in the nearby VLB. Some constraints have already been provided by gravity studies (e.g. Davey and Cooper, 1987; Reitmayr, 1997), and by seismic reflection data (O’Connell and Stepp, 1993; Della Vedova et al., 1997). Data from the ACRUP seismic experiment indicate thickening crust beneath the Transantarctic Mountains to a depth of 38 km, and quite low P-wave velocities (7.6–7.7 km/s) in the mantle beneath the VLB (Della Vedova et al., 1997), while slow S-wave velocities are inferred at 60–160 km depth from surface wave analysis (Bannister et al., 1999), suggesting that the upper mantle is anomalously warm at that depth. More recently, the findings of drilling projects in the McMurdo Sound area (CIROS, Cape Roberts Project) have resolutely constrained the age of
The Antarctic Continent in Gondwanaland
287
the onset of subsidence at the westernmost margin of the VLB. The first direct geological evidence of a major pre-Oligocene uplift phase of the Transantarctic Mountains comes from the oldest strata cored in the CIROS-1 and CRP-3 drill-holes (Barrett et al., 1989, 2001). These include granitic clasts eroded from exposed basement to the west, implying that the Transantarctic Mountains were at least half of their present height by then, for erosion had cut through the more than 2,000 m of Devonian–Jurassic Gondwana cover beds to basement (Barrett et al., 1989, 2001). In the Cape Roberts drill core, the presence of the Devonian Arena Formation (Beacon Supergroup) as bedrock beneath the Cenozoic sediments indicates that significant uplift and unroofing of the Transantarctic Mountains must have occurred prior to the Oligocene (Barrett et al., 2001).
7.6. Open Problems and Potential Research Themes for Future Geoscience Investigations in Antarctica In contrast to all other continents, 99% of Antarctica (including its rocks and geological structures) is covered by a major ice sheet. Less than 1% of the continent provides our geological knowledge and even this has not been investigated thoroughly in all places. On the other hand, about half of the covered area has been surveyed geophysically, through mainly aeromagnetic and gravimetric techniques. Therefore, further investigations are needed and the image of the Antarctic geological structure and its history will likely change, improve and be completed in the future. Several mountain ranges in the Antarctic are difficult to access and therefore have been seldom visited in the past. However, their study would decidedly improve the understanding of geologic and geotectonic connections. It is evident that there are still conspicuous gaps in our knowledge of Marie Byrd Land, the Pensacola Mountains, eastern Dronning Maud Land and East Antarctica between 601 and 1201E. Moreover, the unknown ice-covered interior needs targeted studies starting from well-known areas, for example, tracing known geological rock complexes from their exposed areas under the ice with the help of suitable geophysical methods. The confidence we can have in geophysical interpretations declines with increasing distance from directly accessible rock complexes. Even a few spot checks of rock samples from isolated deep drill-holes would provide a better reliability of interpretations, i.e. calibration of airborne geophysical data. The concept of the formation of supercontinents represents an important step in higher level global cycles in order to unravel the history of
288
F. M. Talarico and G. Kleinschmidt
development of continents and oceans, and to improve our knowledge on global plate tectonic and geodynamic processes. Since Antarctica, and particularly East Antarctica, had a central position in both Rodinia, between 1,300 and 700 million years ago, and Gondwana, between 550 and 200 million years ago (today’s southern continents), geological information archived in this ‘‘keystone’’ region is therefore important not only for a wellfounded analysis of local conditions, but also in relation to our understanding of Earth system processes in general. Abundant evidence exists for the former connection of both Antarctic cratonic areas (the East Antarctic and Grunehogna cratons) to geologically similar provinces on neighbouring continents. The study of East Antarctica is a prerequisite for the reconstruction of the assembly and break-up processes of both supercontinents, Rodinia and Gondwana. This pertains to each of the plate configurations, including the orogens as well as the processes in general, which have led to the assembly of the continental lithosphere. Geological and palaeomagnetic data show that East Antarctica includes at least two older cratonic fragments, and Grenvillian and Pan-African structures in coastal exposures support this notion. The larger part (East Antarctic Craton s.s.) continues into India and Australia, while a small fragment (Grunehogna Craton) connects to the Kalahari Craton of Africa. Newest results show that even the East Antarctic Craton s.s. consists of a number of cratonic nuclei (Fitzsimons, 2000a; Boger and Miller, 2004). These results confirm the importance of Antarctica for the reconstruction of continental distribution in early Earth history. Significant contributions to the understanding of global plate tectonic and geodynamic processes are also stored in the orogenic belts of Pan-African age (600–500 million years ago) which contain important information about the juxtaposition of West and East Gondwana. In Antarctica, the Shackleton Range and parts of East Antarctica between Dronning Maud Land, Lu¨tzow Holm Bukta and Prydz Bay belong to these belts (Buggisch et al., 1990; Jacobs et al., 1998; Tessensohn et al., 1999; Fitzsimons, 2000b; Boger et al., 2002; Paech, 2005; Buggisch and Kleinschmidt, 2007). Equivalent rocks have been found in the African Mozambique Belt (Paech, 1985; Jacobs et al., 1998; Paech et al., 2005). So far, nothing is known about how these fragments continue under the ice and how they can be connected with the better researched mountains in Dronning Maud Land and the Transantarctic Mountains. Equally unknown is the formation, age and relation of subglacial mountains in the East Antarctic interior, e.g. the Gamburtsev Subglacial Mountains. At approximately the same time as the development of internal Pan-African orogenic belts, the geographic domain corresponding to the present Transantarctic Mountains experienced the Ross orogeny, the
The Antarctic Continent in Gondwanaland
289
result of dominant accretionary tectonic processes produced by the subduction of the palaeo-Pacific plate under the palaeo-Pacific Gondwana margin. There is widespread consensus in considering the end of the Ross orogeny as a nearly synchronous event along the different sectors of the belt (e.g. Stump, 1995; Encarnacion and Grunow, 1996). Voluminous granitoids (Granite Harbour Intrusive Complex) intruded as batholiths at c. 530– 480 Ma (Encarnacion and Grunow, 1996) represent a unifying feature throughout the length of the Transantarctic Mountains (Borg et al., 1990; Stump, 1995). But although the general tectonic history of the Ross Orogen is fairly well known within each of the major segments of the Transantarctic Mountains, significant variations in lithostratigraphic, structural and metamorphic patterns, as well as in granitoid geochemical affinity, are evident between the different segments. Considerable uncertainty still remains about the onset of the subduction, the tectonic setting of the early granitoids with variable chemical affinities (from calc-alkaline to alkaline and carbonatite) of southern Victoria Land, the nature of the contact between the orogenic belt and the East Antarctic Craton and the relations with the Pan-African structures of the Shackleton Range. Until our knowledge of the relationships of the tectono-metamorphic histories and of the detailed chronology of the magmatic episodes between the segments is improved a comprehensive tectonic model of the development of the Ross improved remains to be formulated. The importance of Antarctica is also well recognized with regard to the destructive processes of plate tectonics, including the fragmentation of supercontinents. The young continental margins and rift structures of Antarctica, as well as their development, document the break-up of Gondwana and its meaning for the present-day habitats and environmental conditions. Aside from a small section along the Antarctic Peninsula, the continental margins of the Antarctic depict the fault structures of the breakup of Gondwana leading to the formation of the present southern continents and oceans. Antarctica has lain in its south polar position since at least the Late Cretaceous (approximately 130 Ma). Its isolation from the neighbouring continents through the opening of the Southern Ocean and the formation of the Antarctic Circumpolar Current began at this time. The further break-up process led to the formation of the Southern Ocean. The exact opening processes and their effects on palaeoceanography have not yet been satisfactorily reconstructed because high-quality magnetic data are lacking in many key areas of the oceans. This is especially the case for the South Pacific and areas between Antarctica and Africa and Australia, respectively.
290
F. M. Talarico and G. Kleinschmidt
For an understanding of the plate tectonic development of the Antarctic continent, as well as climatic and biological developments, several events must be considered: 130–100 million years ago: the opening of the Weddell, Lazarev and Riiser-Larsen Seas and thus the opening of the southern Atlantic and Indian Oceans; 110–80 million years ago: the genesis of the southern Kerguelen Plateau; 80–40 million years ago: the separation of Tasmania/Australia from Antarctica; approximately 40 million years ago: the main phase of development of the Ross Sea Rift; 30–20 million years ago: the separation of South America from Antarctica, and thus opening of the Drake Passage and Scotia Sea with the formation of the Circumpolar Current. Aside from the basic questions about the mechanisms and the consequences of these global processes, e.g. of climatic nature, the breakup of Gondwana also led to the current mosaic of continental plates. It was the starting point for many current processes and patterns, for example, the glaciation of the polar regions, the present-day oceanic and atmospheric circulation, the distribution of climatic zones and biota, and finally the global conditions for our existence. The present state of knowledge indicates that the distinct steps of the break-up of Gondwana must be determined more precisely. In particular, the phase since the total isolation of Antarctica is globally relevant for the development of ocean circulation and climate patterns and it must therefore be further deciphered. More specifically, an increased research effort would be essential to provide new and more complete on- and off-shore (mainly from drill-holes) data to fill the numerous knowledge gaps on the Cenozoic Antarctic glacial history and to allow a better timing of the Antarctic Circumpolar Current onset (Barker and Thomas, 2004; Barker et al., 2007). Indeed, the important role of the polar regions is emphasized by general circulation models (Sloan et al., 1996; Oglesby, 1999). The presence of thick ice sheets in the continents influences latitudinal gradients and sea-ice formation, and it drives the formation of cold Antarctic bottom water which through deep ocean currents reaches the lower latitudes. The oxygen isotope record from deep-sea cores, and eustatic changes inferred from sequence stratigraphic records on passive continental margins, have led our knowledge of the long-term and broad history of the Antarctic Ice Sheet. However, interpretations based on these proxy records of glacio-eustasy have little
The Antarctic Continent in Gondwanaland
291
direct confirmation from geologic records in Antarctica, and in numerous cases have led to conflicting interpretations (Harwood et al., 1991, 1993; Moriwaki et al., 1992; Wilson, 1995; Miller and Mabin, 1998). In this context, since direct data from the Antarctic region are essential to validate these models, a series of drilling projects have targeted the Antarctic continental margin to retrieve high-resolution stratigraphic records of Antarctica’s glacial and climatic history (Cooper and Webb, 1992; Barker et al., 1998; Hambrey et al., 1998; Barrett et al., 2000, 2001; Hambrey, 2002) and more work in this area is planned for the next decade e.g. (ANDRILL, Harwood et al., 2002). Palaeoclimatic reconstructions based on the results of the Cape Roberts drilling project in the Ross Sea region for the early Oligocene to early Miocene time show two major climatic phases both documenting a warmer climate in the Antarctic than that which prevails today (Hambrey et al., 1998; Barrett et al., 2000, 2001; Naish et al., 2001). Also in the early Quaternary, which was characterized especially in the northern hemisphere by cool climatic conditions, the Cape Roberts core proved that there were times when the Antarctic experienced higher temperatures than today (Hambrey et al., 1998). However, despite the successes of drilling to date, there remain major unresolved questions concerning Cenozoic tectonic and palaeoenvironmental evolution of the Antarctic region. Chief among these is the still controversial problem of how stable the Antarctic Ice Sheets were during the last 20 My and the timing of the onset of glaciation. The former has been one of the main scientific targets of the first ANDRILL drilling seasons in 2006 and 2007 austral summers (Harwood et al., 2002) and the ongoing investigations are attempting to determine the responses of past ice sheets and shelfs to climate forcing, including variability at a range of time-scales over the last 20 My. The latter has remained an elusive target for both CIROS and Cape Roberts drilling. Indeed, in CRP-3, Oligocene strata passed via an unconformity into the pre-glacial Devonian–Triassic Beacon Supergroup rocks. The search goes on for sites which will recover an Eocene and earlier Cenozoic, or even a Cretaceous record. This issue, along with others, will likely be addressed by future drilling programs that will also complement data from the Ross Sea with data from other coastal regions of Antarctica, and will provide a deeper insight into the climatic relevance of this region.
ACKNOWLEDGEMENTS This chapter is a contribution to the SCAR programme Antarctic Climate Evolution (ACE). Acknowledgements for funding and logistical and
292
F. M. Talarico and G. Kleinschmidt
technical support with the field campaigns and drilling projects in the McMurdo Sound area are made to the US Office of Polar Programs (NSFOPP), Antarctica New Zealand, the Italian National Antarctic Program (PNRA), the Alfred Wegener Institute for Polar and Marine Research (AWI) and the Deutsche Forschungsgemeinschaft (DFG). Valuable comments by F. Florindo and M. Siegert, and careful corrections and reviews by C.A. Ricci and E. Stump are also much appreciated.
REFERENCES Bannister, S., Snieder, R. K., & Passier, M. L. (1999). Shear-wave velocities under the Transantarctic Mountains and terror rift from surface wave inversion. Geophys. Res. Lett., 27, 281–284. Bannister, S., Yu, J., Leitner, B., & Kennett, B. L. N. (2003). Variations in crustal structure across the transition from west to east Antarctica, southern Victoria Land. Geophys. J. Int., 155, 870–884. Barker, P. F., Barrett, P. J., Camerlenghi, A., Cooper, A. K., Davey, F. J., Domack, E. W., Escutia, C., Kristoffersen, Y., & O’Brien, P. E. (1998). Ice sheet history from Antarctic continental margin sediments: The ANTOSTRAT approach. Terra Antartica, 5(4), 737–760. Barker, P. F., Filippelli, G. M., Florindo, F., Martin, E. E., & Scher, H. D. (2007). Onset and role of the Antarctic circumpolar current. Deep-Sea Res. II, 54, 2388–2398. Barker, P. F., & Thomas, E. (2004). Origin, signature and palaeoclimatic influence of the Antarctic circumpolar current. Earth Sci. Rev., 66, 143–162. Barrett, P. J. (1991). The Devonian to Triassic Beacon Supergroup of the Transantarctic Mountains and correlatives in other parts of Antarctica. In: R. J. Tingey (Ed.). The Geology of Antarctica. Oxford Monographs on Geology and Geophysics. Clarendon Press, Oxford, pp. 120–152. Barrett, P. J., Davey, F. J., Ehrmann, W. U., Hambrey, M. J., Jarrard, R., van der Meer, J. J. M., Raine, J., Roberts, A. P., Talarico, F., & Watkins, D. K. (Eds). (2000). Studies from the Cape Roberts Project, Ross Sea, Antarctica, Scientific Results of CRP-2/2A, Parts I and II. Terra Antartica, 7 (4/5), 665 pp. Barrett, P. J., Hambrey, M. J., Harwood, D. M., Pyne, A. R., & Webb, P. N. (1989). Synthesis. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. DSIR Bull., 245, pp. 241–252. Barrett, P. J., Ricci, C. A., Bucker, C. J., Davey, F. J., Ehrmann, W. U., Laird, M. G., van der Meer, J. J. M., Raine, J. I., Smellie, J. L., Talarico, F., Thomson, M. R. A., Verosub, K. L., & Villa, G. (Eds). (2001). Studies from the Cape Roberts Project, Ross Sea, Antarctica, Scientific Report of CRP-3, Parts I and II. Terra Antartica, 8 (3/4), 620 pp.
The Antarctic Continent in Gondwanaland
293
Behrendt, J. C. (1999). Crustal and lithospheric structure of the west Antarctic rift system from geophysical investigations – A review. Global Planet. Change, 23, 25–44. Behrendt, J. C., Blankenship, D. D., Morse, D. L., Finn, C. A., & Bell, R. E. (2002). Removal of subglacially erupted volcanic edifices beneath the divide of the west Antarctic Ice Sheet interpreted from aeromagnetic and radar ice-sounding surveys. In: J. A. Gamble, D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium. R. Soc. N.Z. Bull., 35, 579–587. Behrendt, J. C., & Cooper, A. K. (1991). Evidence of rapid Cenozoic uplift of the shoulder escarpment of the Cenozoic west Antarctic rift system and a speculation on possible climate forcing. Geology, 19, 315–319. Behrendt, J. C., Damaske, D., & Fritsch, J. (1993). Geophysical characteristics of the west Antarctic rift system. Geologisches Jahrbuch, E47, 49–101. Behrendt, J. C., Duerbaum, H. J., Damaske, D., Saltus, R., Bosum, W., & Cooper, A. (1991a). Extensive volcanism and related tectonism beneath the Ross Sea continental shelf, Antarctica: Interpretation of an aeromagnetic survey. In: M. R. A. Thompson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, UK, pp. 299–304. Behrendt, J. C., LeMasurier, W., & Cooper, A. K. (1992). The west Antarctic rift system – A propagating rift ‘‘captured’’ by a mantle plume? In: Y. Yoshida (Ed.). Recent Progress in Antarctic Earth Science. Terrapub, Tokyo, pp. 315–322. Behrendt, J. C., LeMasurier, W. E., Cooper, A. K., Tessensohn, F., Trehu, A., & Damaske, D. (1991b). Geophysical studies of the west Antarctic rift system. Tectonics, 10, 1257–1273. Bell, R. E., Studinger, M., Karner, G., Finn, C. A., & Blankenship, D. D. (2006). Identifying major sedimentary basins beneath the west Antarctic Ice Sheet from aeromagnetic data analysis. In: D. K. Fu¨tterer, D. Damaske, G. Kleinschmidt, H. Miller, & F. Tessensohn (Eds). Antarctica: Contributions to Global Earth Sciences. Springer, Berlin, pp. 117–122. Bentley, C. R. (1991). Configuration and structure of the subglacial crust. In: R. J. Tingey (Ed.). The Geology of Antarctica. Clarendon, Oxford, pp. 335–364. Berg, J. H., Moscati, R. J., & Herz, D. L. (1989). A petrologic geotherm from a continental rift in Antartica. Earth Planet. Sci. Lett., 93, 98–108. Birkenmajer, K. (1994). Evolution of the Pacific margin of the northern Antarctic peninsula: An overview. Geol. Rundsch., 83, 309–321. Blackman, D. K., Von Herzen, R. P., & Lawver, L. A. (1987). Heat flow and tectonics in the western Ross Sea, Antarctica. In: A. K. Cooper, & F. J. Davey (Eds). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea. Earth Science Series. Circum-Pacific Council for Energy and Natural Resources, Houston, TX, Vol. 5B, pp. 179–189. Boger, S. D., Carson, C. J., Fanning, C. M., Hergt, J. M., Wilson, C. J. L., & Woodhead, J. D. (2002). Pan-African intraplate deformation in the northern Prince Charles Mountains, east Antarctica. Earth Planet. Sci. Lett., 195, 195–210.
294
F. M. Talarico and G. Kleinschmidt
Boger, S. D., & Miller, J. M. (2004). Terminal suturing of Gondwana and the onset of the Ross–Delamerian orogeny: The cause and effect of an early Cambrian reconstruction of plate motions. Earth Planet. Sci. Lett., 219, 35–48. Boger, S. D., Wilson, C. J. L., & Fanning, C. M. (2001). Early Paleozoic tectonism within the east Antarctic craton: The final suture between east and west Gondwana? Geology, 29, 463–466. Borg, S. G., & DePaolo, D. J. (1991). A tectonic model of the Antarctic Gondwana margin with implications for southeastern Australia: Isotopic and geochemical evidence. Tectonophysics, 196, 339–358. Borg, S. G., DePaolo, D. J., & Smith, B. M. (1990). Isotopic structure and tectonics of the central Transantarctic Mountains. J. Geophys. Res., 95, 6647–6669. Bosum, W., Damaske, D., Roland, N. W., Behrendt, J., & Saltus, R. (1989). The GANOVEX IV Victoria Land/Ross Sea aeromagnetic survey: Interpretation of anomalies. Geol. Jb., E38, 153–230. Bott, M. H. P., & Stern, T. A. (1992). Finite element analysis of Transantarctic Mountain uplift and coeval subsidence in the Ross embayment. Tectonophysics, 201, 341–356. Bradshaw, J. D. (1989). Cretaceous geotectonic patterns in the New Zealand region. Tectonics, 8, 803–820. Bradshaw, J. D., Weaver, S. D., Pankurst, R. J., Storey, B. C., & Muir, R. J. (1997). New Zealand superterranes recognized in Marie Byrd Land and Thurston Island. Terra Antartica, 3, 429–436. Bradshaw, M. A., & Webers, G. F. (1988). The Devonian rocks of Antarctica. In: N. J. McMillan, A. F. Embry, & D. J. Glass (Eds). Devonian of the World. Canadian Society of Petroleum Geologists, Calgary, pp. 783–795. Buggisch, W., & Kleinschmidt, G. (2007). The Pan-African nappe tectonics in the Shackleton Range. In: A. K. Cooper, C. R. Raymond, & the 10th ISAES Editorial Team Antarctica (Eds). A Keystone in a Changing World. Online Proceedings of the 10th ISAES, USGS Open-File Report 2007-1047, Short Research Paper, 058, p. 4. Buggisch, W., Kleinschmidt, G., Kreuzer, H., & Krumm, S. (1990). Stratigraphy, metamorphism and nappe-tectonics in the Shackleton Range. Geoda¨tische Geophysikalische Vero¨ff. Reihe I, 15, 64–86. Busetti, M., Spadini, G., Van der Wateren, F. M., Cloetingh, S., & Zanolla, C. (1999). Kinematic modelling of the west Antarctic rift system, Ross Sea, Antarctica. Global Planet. Change, 23, 79–103. Campbell, I. B., & Claridge, G. G. C. (1987). Antarctica: Soils, Weathering Processes and Environment. Elsevier, Amsterdam, 368 pp. Cande, S. C., & Leslie, R. B. (1986). Late Cenozoic tectonics of the southern Chile trench. J. Geophys. Res., 91, 451–496. Cande, S. C., & Mutter, J. C. (1982). A revised identification of the oldest sea-floor spreading anomalies between Australia and Antarctica. Earth Planet. Sci. Lett., 58, 151–160.
The Antarctic Continent in Gondwanaland
295
Cande, S. C., & Stock, J. M. (2006). Constraints on the timing of extension in the northern basin, Ross Sea. In: D. K. Fu¨tterer, D. Damaske, G. Kleinschmidt, H. Miller, & F. Tessensohn (Eds). Antarctica: Contributions to Global Earth Sciences. Springer, Berlin, pp. 319–326. Cande, S. C., Stock, J. M., Mu¨ller, R. D., & Ishihara, T. (2000). Cenozoic motion between east and west Antarctica. Nature, 404, 145–150. Cape Roberts Science Team (2000). Studies from the Cape Roberts Project, Ross Sea, Antarctica. Initial Report on CRP-3. Terra Antartica, 7, 1–209. Cawood, P. A. (2005). Terra Australis orogen: Rodinia breakup and development of the Pacific and Iapetus margins of Gondwana during the Neoproterozoic and Paleozoic. Earth Sci. Rev., 69, 249–279. Chery, J., Lucazeau, F., Daignie`res, M., & Vilotte, J. P. (1992). Large uplift of rift flanks: A genetic link with lithospheric rigidity? Earth Planet. Sci. Lett., 112, 195–211. Cogley, J. G. (1984). Deglacial hypsometry of Antarctica. Earth Planet. Sci. Lett., 67, 151–177. Collinson, J. W. (1991). The palaeo-Pacific margin as seen from East Antarctica. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 199–204. Cook, Y. A. (2007). Precambrian rift-related magmatism and sedimentation, south Victoria Land, Antarctica. Antarctic Sci., 8, 1–14. Cooper, A. K., Davey, F. J., & Behrendt, J. C. (1987). Seismic stratigraphy and structure of the Victoria Land Basin, Western Ross Sea, Antarctica. In: A. K. Cooper, & F. J. Davey (Eds). The Antarctic Continental Margin. Geology and Geophysics of the Western Ross Sea, Earth Science Series. CircumPacific Council for Energy and Mineral Resources, Houston, TX, Vol. 5B, pp. 27–76. Cooper, A. K., Davey, F. J., & Hinz, K. (1991). Crustal extension and origin of sedimentary basins beneath the Ross Sea and Ross ice shelf, Antarctica. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 285–292. Cooper, A. K., Trey, H., Pellis, G., Cochrane, G., Egloff, F., Busetti, M., & ACRUP Working Group (1997). Crustal structure of the southern central trough, western Ross Sea. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 637–642. Cox, K. G. (1988). The Karoo Province. In: J. D. MacDougall (Ed.). Continental Flood Basalts. Kluwer Academic Publishers, Dordrecht, pp. 239–271. Craddock, C. (1970). Tectonic map of Gondwana. In: V. C. Bushnell, & C. Craddock (Eds). Geologic Maps of Antarctica, Antarctic Map Folio Series, Folio 12, Plate XXIII. American Geographic Society, New York. Curtis, M. L. (1998). Development of kinematic partitioning within a pure-shear dominated dextral transpression zone: The southern Ellsworth Mountains, Antarctica. In: R. A. Strachan, & J. F. Dewey (Eds). Continental Transpressional
296
F. M. Talarico and G. Kleinschmidt
and Transtensional Tectonics. Geological Society, London, pp. 289–306. Special Publication. Curtis, M. L., & Storey, B. C. (1996). A review of geological constraints on the prebreakup position of the Ellsworth Mountains within Gondwana: Implications for Weddell Sea evolution. In: B. C. Storey, E. C. King, & R. A. Livermore (Eds). Weddell Sea Tectonics and Gondwana Breakup. Geological Society, London, pp. 11–30. Special Publication. Dalziel, I. W. D. (1982). The early (pre-middle Jurassic) history of the Scotia Arc region: A review and progress report. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, pp. 111–126. Dalziel, I. W. D. (1991). Pacific margins of Laurentia and east Antarctica–Australia as a conjugate rift pair: Evidence and implications for an Eocambrian supercontinent. Geology, 19, 598–601. Dalziel, I. W. D. (1992). Antarctica: A tale of two supercontinents. Annu. Rev. Earth Planet. Sci., 20, 501–526. Dalziel, I. W. D. (1997). Neoproterozoic–Paleozoic geography and tectonics: Review, hypothesis, environmental speculation. Geol. Soc. Am. Bull., 109, 16–42. Dalziel, I. W. D., & Elliot, D. H. (1982). West Antarctica: Problem child of Gondwanaland. Tectonics, 1, 3–19. Dalziel, I. W. D., Lawver, L. A., & Murphy, J. B. (1999). Plume, orogenesis, supercontinental fragmentation and ice-sheets. In: D. N. B. Skinner (Ed.). 8th Symposium on Antarctic Earth Sciences, 5–9 July, Programme and Abstracts. Victoria University of Wellington, Wellington, New Zealand, p. 78. Davey, F. J. (1981). Geophysical studies in the Ross Sea region. J. R. Soc. N. Z., 11, 465–479. Davey, F. J., & Cooper, A. K. (1987). Gravity studies of the Victoria Land Basin and Iselin Bank. In: A. K. Cooper, & F. J. Davey, (Eds). The Antarctic Continental Margin: Geology and Geophysics of the Western Ross Sea, Earth Science Series. Circum-Pacific Council for Energy and Natural Resources, Houston, TX, Vol. 5B, pp. 119–137. Davey, F. J., & Brancolini, G. (1995). The Late Mesozoic and Cenozoic structural setting of the Ross Sea region. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. AGU Ant. Res. Ser., 68, 167–182. David, T. E., & Priestley, R. E. (1914). British Antarctic Expedition 1907–09. Rep. Sci. Invest. Geol., v. 1, Heinemman, p. 312. De Wit, M., Jeffery, M., Bergh, H., & Nicolaysen, L. (1988). Geological map of sectors of Gondwana reconstructed to their dispositionB150 Ma. American Association of Petroleum Geologists, Tulsa, OK 74101, scale 1:10,000,000. Della Vedova, B., Pellis, G., Trey, H., Zhang, J., Cooper, A. K., Makris, J., & the ACRUP working group (1997). Crustal structure of the Transantarctic Mountains, western Ross Sea. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 609–618.
The Antarctic Continent in Gondwanaland
297
Di Vincenzo, G., Palmeri, R., Talarico, F., Andriessen, P. A. M., & Ricci, C. A. (1997). Petrology and geochronology of eclogites from the Lanterman Range, Antarctica. J. Petrol., 38, 1391–1417. Di Vincenzo, G., Rocchi, S., Rossetti, F., & Storti, F. (2004). 40Ar–39Ar dating of pseudotachylytes: The effect of clast-hosted extraneous argon in Cenozoic faultgenerated friction melts from the west Antarctic rift system. Earth Planet. Sci. Lett., 223, 349–364. Dirks, P. H. G. M., & Wilson, C. J. L. (1995). Crustal evolution of the east Antarctic mobile belt in Prydz Bay: Continental collision at 500 Ma? Precambrian Res., 75(3), 189–207. Divenere, V., Kent, D. V., & Dalziel, I. W. D. (1995). Summary of paleomagnetic results from west Antarctica: Implications for the tectonic evolution of the Pacific margin of Gondwana during the Mesozoic. In: B. C. Storey, E. C. King, & R. A. Livermore (Eds). Weddell Sea Tectonics and Gondwana Breakup. Geological Society, London, pp. 31–43. Special Publication. Divenere, V., Kent, D. V., & Dalziel, I. W. D. (1996). Early Cretaceous paleomagnetic results from Marie Byrd Land, west Antarctica: Implications for the Weddellia collage of crustal blocks. J. Geophys. Res., 100, 8133–8152. Du Toit, A. L. (1937). Our Wandering Continents. Oliver and Boyd, Edinburgh, 366 pp. Elliot, D. H. (1975). Tectonics of Antarctica: A review. Am. J. Sci., 275, 45–106. Elliot, D. H. (1992). Jurassic magmatism and tectonism associated with Gondwanaland breakup: An Antarctic perspective. In: B. C. Storey, T. Alabaster, & R. J. Pankhurst (Eds). Magmatism and the Causes of Continental Breakup. Geological Society, London, pp. 165–184. Special publication Elliot, D. H., Fortner, E. H., & Grimes, C. B. (2006). Mawson Breccias intrude Beacon strata at Allan Hills, South Victoria Land: Regional implications. In: D. K. Fu¨tterer, D. Damaske, G. Kleinschmidt, H. Miller, & F. Tessensohn (Eds). Antarctica – Contributions to Global Earth Sciences. Springer, Berlin/Heidelberg, pp. 291–298. Encarnacion, J., & Grunow, A. (1996). Changing magmatic and tectonic styles along the paleo-Pacific margin of Gondwana and the onset of early Paleozoic magmatism in Antarctica. Tectonics, 15, 1325–1341. Findlay, R. H. (1987). A review of the problems important for the interpretation of the Cambro-Ordovician paleogeography of northern Victoria Land (Antarctica), Tasmania, and New Zealand. In: G. D. McKenzie (Ed.). Gondwana Six: Structure, Tectonics, and Geophysics. American Geophysical Union Monographs. Vol. 40, pp. 49–66. Fitzgerald, P. G. (1992). The Transantarctic Mountains of southern Victoria Land: The application of apatite fission track analysis to a rift shoulder uplift. Tectonics, 11, 634–662. Fitzgerald, P. G. (1995). Cretaceous and Cenozoic exhumation of the Transantarctic Mountains: Evidence from the Kukri Hills of southern Victoria Land compared to fission track data from Gneiss at DSDP site 270. In: C. A. Ricci (Ed.). 7th International Symposium on Antarctic Earth Sciences, Siena, Italy, p. 133.
298
F. M. Talarico and G. Kleinschmidt
Fitzgerald, P. G. (2002). Tectonics and landscape evolution of the Antarctic plate since the breakup of Gondwana, with an emphasis on the west Antarctic rift system and the Transantarctic Mountains. In: J. Gamble, & D. A. Skinner (Eds). Proceedings of the 8th International Symposium on Antarctic Earth Science. R. Soc. N. Z. Bull., 35, 453–469. Fitzgerald, P. G., & Baldwin, S. L. (1997). Detachment fault model for the evolution of the Ross embayment. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 555–564. Fitzgerald, P. G., Sandiford, M., Barrett, P. J., & Gleadow, A. J. W. (1986). Asymmetric extension associated with uplift and subsidence of the Transantarctic Mountains and Ross embayment. Earth Planet. Sci. Lett., 81, 67–78. Fitzgerald, P. G., & Stump, E. (1997). Cretaceous and Cenozoic episodic denudation of the Transantarctic Mountains, Antarctica: New constraints from apatite fission track thermochronology in the Scott Glacier region. J. Geophys. Res., 102, 7747–7765. Fitzsimons, I. C. W. (1997). The Brattstrand paragneiss and the Sostrene orthogneiss: A review of Pan-African metamorphism and Grenvillian relics in southern Prydz Bay. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica, Siena, pp. 121–130. Fitzsimons, I. C. W. (2000a). Grenville-age basement provinces in East Antarctica: Evidence for three separate collisional orogens. Geology, 28, 879–882. Fitzsimons, I. C. W. (2000b). A review of tectonic events in the east Antarctic shield and their implications for Gondwana and earlier supercontinents. J. Afr. Earth Sci., 31, 3–23. Fitzsimons, I. C. W. (2003). Proterozoic basement provinces of southern and south-western Australia, and their correlation with Antarctica. In: M. Yoshida, B. F. Windley, & S. Dasgupta (Eds). Proterozoic East Gondwana: Supercontinent Assembly and Breakup. Geological Society, London, pp. 93–129. Flo¨ttmann, T., Gibson, G. M., & Kleinschmidt, G. (1993). Structural continuity of the Ross and Delamerian orogens of Antarctica and Australia along the margin of the paleo-Pacific. Geology, 21, 319–322. Ford, A. B. (1963). Cordierite-bearing hypersthene–quartz–monzonite–porphyry in the Thiel Mountains and its regional importance. In: R. J. Adie (Ed.). Antarctic Geology. North-Holland Publishing Company, Amsterdam, pp. 429–441. Funaki, M., Yoshida, M., & Matsueda, H. (1991). Palaeomagnetic studies of palaeozoic rocks from the Ellsworth Mountains. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 257–260. Gibson, G. M., & Ireland, T. R. (1996). Extension of Delamerian (Ross) orogen into western New Zealand: Evidence from Zircon ages and implications for crustal growth along the Pacific margin of Gondwana. Geology, 24, 1087–1090. Gonza´les-Casado, J. M., Giner-Robles, J. L., & Lo´pez-Martinez, J. (2000). Bransfield Basin, Antarctic Peninsula: Not a normal Backarc Basin. Geology, 28, 1043–1046.
The Antarctic Continent in Gondwanaland
299
Goodge, J. W. (2001). From Rodinia to Gondwana: Supercontinent evolution in the Transantarctic Mountains. In: J. Gamble, & D. Skinner (Eds). Proceedings of the 8th International Symposium on Antarctic Earth Science. R. Soc. N. Z. Bull., 35, 61–74. Goodge, J. W., Myrow, P., Williams, I. S., & Bowring, S. A. (2002). Age and provenance of the Beardmore Group, Antarctica: Constraints on Rodinia supercontinent breakup. J. Geol., 110, 393–406. Gould, L. M. (1935). Structure of the Queen Maud Mountains, Antarctica. Geol. Soc. Am. Bull., 46, 973–984. Grantham, G. H., & Hunter, D. R. (1991). The timing and nature of faulting and jointing adjacent to the Penckso¨kket, Western Dronning Maud Land, Antarctica. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 47–51. Grindley, G. W., & Laird, M. G. (1969). Geology of the Shackleton Coast. Antarctic Map Folio Series, v. Folio 12, XIV. Grunow, A., Hanson, R., & Wilson, T. (1996). Were aspects of Pan-African deformation linked to Iapetus opening? Geology, 24, 1063–1066. Grunow, A. M., Kent, D. V., & Dalziel, I. W. D. (1987). Evolution of the Weddell Sea Basin, new paleomagnetic constraints. Earth Planet. Sci. Lett., 86, 16–26. Grunow, A. M., Kent, D. V., & Dalziel, I. W. D. (1991). New paleomagnetic data from Thurston Island: Implications for the tectonics of West Antarctica and Weddell Sea opening. J. Geophys. Res., 96 (B11), 17935–17954 Gunn, B. M., & Warren, G. (1962). Geology of Victoria Land between the Mawson and Mulock Glaciers, Antarctica. N. Z. Geol. Surv. Bull., 71, 157. Gust, D. A., Biddle, K. T., Phelps, D. W., & Uliana, M. A. (1985). Associated middle to late Jurassic volcanism and extension in southern South America. Tectonophysics, 116, 223–253. Hambrey, M.J. (2002). Historical review of drilling in the western Ross Sea. In: D. M. Harwood, L. Lacy, & R. H. Levy (Eds). Future Antarctic Margin Drilling: Developing a Science Plan for McMurdo Sound. ANDRILL Science Management Office (SMO) Contribution 1, University of Nebraska-Lincoln, Lincoln, NE. Hambrey, M. J., Wise, S. W., Jr., Barrett, P. J., Davey, F. J., Ehrmann, W. U., Smellie, J. L., Villa, G., & Woolfe, K. J. (Eds). (1998). Scientific Report on CRP-1, Cape Roberts Project, Antarctica. Terra Antartica, 5, 713 pp. Hamilton, R. J., Luyendyk, B. P., Sorlien, C. C., & Bartek, L. R. (2001). Cenozoic tectonics of the Cape Roberts Rift Basin and Transantarctic Mountains front, southwestern Ross Sea, Antarctica. Tectonics, 20, 325–342. Hartnady, C. J. H. (1991). About turn for supercontinents. Nature, 352, 476–478. Harwood, D. M., Lacy, L., & Levy, R. H. (Eds). (2002). Future Antarctic Margin Drilling: Developing a Science Plan for McMurdo Sound. ANDRILL Science Management Office (SMO) Contribution 1, University of Nebraska-Lincoln, Lincoln, NE, 301 pp. Harwood, D.M., Webb, P.-N., & Barrett, P.J. (1993). The search for consistency between indices of Antarctic Cenozoic glaciation: Cenozoic glaciations and
300
F. M. Talarico and G. Kleinschmidt
deglaciations. Geological Society of London, Stratigraphic Committee and the Quaternary Research Association. Quaternary Newslett., 69, 61–63. Harwood, D. M., Webb, P.-N., Barrett, P. J., & McKelvey, B. C. (1991). The changing style of Antarctic glaciations. In: 6th International Antarctic Earth Science Symposium, Tokyo, August, p. 20. Heimann, A., Fleming, T. H., Elliot, D. H., & Foland, K. A. (1994). A short interval of Jurassic continental flood basalt volcanism in Antarctica as demonstrated by 40Ar/39Ar geochronology. Earth Planet. Sci. Lett., 121, 19–41. Hergt, J. M., Chappell, M. T., McCullock, M., MacDougall, I., & Chivas, A. R. (1989). Geochemical and isotopic constraints on the origin of the Jurassic dolerites of Tasmania. J. Petrol., 30, 841–883. Hoffman, P. F. (1991). Did the breakout of Laurentia turn Gondwanaland insideout? Science, 252, 1409–1412. Hofmann, J. (1996). Fragmente Intragondwanischer Rifte als Werkzeug der GondwanaRekonstruktion-das Beispiel des Lambert-Mahanadi-Riftes (OstantarktikaPeninsular Indien). Neues Jb. Geol. Pala¨ontol. Abhandlungen, 199, 33–48. Holmes, A. (1951). The sequence of pre-Cambrian orogenic belts in south and central Africa. In: Proceedings of the 18th International Geological Congress, London, Vol. 24, pp. 254–269. Hooper, P. R., Rehacek, J., & Duncam, R. A. (1993). The basalts of Lesotho, Karoo Province, southern Africa. EOS Trans. Am. Geophys. Union, 74, 553. Abstract Isbell, J. L. (1999). The Kukri erosion surface: A reassessment of its relationship to rocks of the Beacon Supergroup in the central Transantarctic Mountains, Antarctica. Antarctic Sci., 11, 228–238. Jacobs, J., Fanning, C. M., Henjes-Kunst, F., Olesch, M., & Paech, H. J. (1998). Continuation of the Mozambique belt into east Antarctica: Grenville-age metamorphism and polyphase Pan-African high-grade events in central Dronning Maud Land. J. Geol., 106, 385–406. Jacobs, J., & Lisker, F. (1999). Post Permian tectono-thermal evolution of western Dronning Maud Land, east Antarctica: An apatite fission-track approach. Antarctic Sci., 11, 451–460. Jacobs, J., & Thomas, R. J. (2002). The Mozambique belt from an east Antarctic perspective. In: J. A. Gamble, D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium. Proceedings of the 8th International Symposium on Antarctic Earth Sciences. R. Soc. N. Z. Bull., 35, 3–18. Janowski, E. J., & Drewry, D. J. (1981). The structure of west Antarctic from geophysical studies. Nature, 291, 17–21. Johnston, S. T. (2000). The Cape Fold Belt and syntaxis and the rotated Falklands Islands: Dextral transpressional tectonics along the southwest margin of Gondwana. J. Afr. Earth Sci., 31, 51–63. Jones, S. (1997). Late quaternary faulting and neotectonics, south Victoria Land, Antarctica. J. Geol. Soc. Lond., 154(4), 645–652. Kanao, M., Shibutani, T., Negishi, H., & Tono, H. (2002). Crustal structure around the Antarctic margin by teleseismic receiver function analyses. In: J. A. Gamble,
The Antarctic Continent in Gondwanaland
301
D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium, Wellington. R. Soc. N. Z., 35, 485–491. Karlstrom, K. E., Williams, M. L., McLelland, J., Geissman, J. W., & Ahall, K.-I. (1999). Refining Rodinia: Geologic evidence for the Australia–western US connection in the Proterozoic. GSA Today, 9(10), 1–7. Kellogg, K. S., & Rowley, P. D. (1989). Structural Geology and Tectonics of the Orville Coast Region, Southern Antarctic Peninsula, Antarctica. US Geological Survey, 25 pp. Kleinschmidt, G. (2007). Geological evolution and structure of Antarctica. In: B. Riffenburgh (Ed.). Encyclopedia of the Antarctic. New York/London (Routledge/Taylor & Francis), Vol. 1, pp. 430–437. Kleinschmidt, G., Buggisch, W., La¨ufer, A. L., Helferich, S., & Tessensohn, F. (2002). The ‘‘Ross orogenic’’ structures in the Shackleton Range and their meaning for Antarctica. In: J. A. Gamble, D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium, Wellington. R. Soc. N. Z., 35, 75–83. Kleinschmidt, G., & La¨ufer, A. L. (2006). The Matusevich Fracture Zone, in Oates Land, east Antarctica. In: D. K. Fu¨tterer, D. Damaske, G. Kleinschmidt, H. Miller, & F. Tessensohn (Eds). Antarctica: Contributions to Global Earth Sciences. Berlin/Heidelberg, pp. 175–180. Kleinschmidt, G. and Tessensohn, F. (1987). Early Paleozoic westward directed subduction at the Pacific margin of Antarctica. In: G. D. McKenzie (Ed.). Gondwana Six: Structure, Tectonics, and Geophysics. American Geophysical Union Monographs, Vol. 40, pp. 89–105. Kriegsman, L. M. (1995). The Pan-African event in east Antarctica: A view from Sri Lanka and the Mozambique Belt. Precambrian Res., 75, 263–277. Kyle, P. R. (1990a). Erebus Volcanic Province. In: W. E. LeMasurier, & J. W. Thompson (Eds). Volcanoes of the Antarctic Plate and Southern Oceans, Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 48, pp. 81–135. Kyle, P. R. (1990b). McMurdo Volcanic Group, western Ross embayment. Introduction. In: W. E. LeMasurier, & J. W. Thompson (Eds). Volcanoes of the Antarctic Plate and Southern Oceans: Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 48, pp. 19–25. Kyle, P. R., & Cole, J. W. (1974). Structural control of volcanism in the McMurdo Volcanic Group, Antarctica. Bull. Volcanologique, 38, 16–25. Larter, R. D., Cunningham, A. P., Barker, P. F., Gohl, K., & Nitsche, F. O. (2002). Tectonic evolution of the Pacific margin of Antarctica 1. Late Cretaceous tectonic reconstructions. J. Geophys. Res., 107(B12), 2345. doi: 10.1029/2000JB000052. Lawver, L. A., Gahagan, L. M., & Coffin, M. F. (1992). The development of Paleoseaways around Antarctica. In: The Antarctic Paleoenvironment: A Perspective on Global Change. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 56, pp. 7–30. Lawver, L. A., Gahagan, L. M., & Dalziel, I. W. D. (1998). A tight fit – Early Mesozoic Gondwana: A plate tectonic perspective. In: Origin and Evolution of
302
F. M. Talarico and G. Kleinschmidt
Continents. Memoirs of the National Institute of Polar Research. National Institute of Polar Research, Tokyo, pp. 214–229. Special issue. Lawver, L. A., Royer, J.-Y., Sandwell, D. A., & Scotese, C. T. (1991). Evolution of the Antarctic continental margins. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 533–539. LeMasurier, W. E. (2007). West Antarctic rift system. In: B. Riffenburgh (Ed.). Encyclopedia of the Antarctic. Routledge, New York, Vol. 2, pp. 1060–1066. LeMasurier, W. E., & Thomson, J. W. (1990). Volcanoes of the Antarctic Plate and Southern Oceans. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 48, 487 pp. Li, Z. X., & Powell, C. M. (2001). An outline of the palaeogeographic evolution of the Australasian region since the beginning of the Neoproterozoic. Earth Sci. Rev., 53, 237–277. Lythe, M. B., Vaughan, D. G., & the BEDMAP Consortium (2000). BEDMAP – Bed topography of the Antarctic, 1:10 000000 scale map. BAS (MISC) 9, British Antarctic Survey, Cambridge. Marshall, J. E. A. (1994). The Falkland Islands: A key element in Gondwana paleogeography. Tectonics, 13, 449–514. McMenamin, M. A. S., & McMenamin, D. L. S. (1990). The Emergence of Animals: The Cambrian Break Through. Columbia University Press, New York, 217 pp. Meert, J. G., & Torsvik, T. H. (2003). The making and unmaking of a supercontinent: Rodinia revisited. Tectonophysics, 375, 261–288. Meert, J. G., & Van der Voo, R. (1997). The assembly of Gondwana 800–550 Ma. J. Geodyn., 23, 223–235. Meert, J. G., Van der Voo, R., & Ayub, S. (1995). Paleomagnetic investigation of the Neoproterozoic Gagwe Lavas and Mbozi complex, Tanzania and the assembly of Gondwana. Precambrian Res., 74, 225–244. Miller, M. F., & Mabin, M. C. G. (1998). Antarctic Neogene landscapes – In the refrigerator or in the deep freeze? GSA Today, 8, 1–3. Moores, E. M. (1991). Southwest US–east Antarctica (SWEAT) connection: A hypothesis. Geology, 19, 425–428. Moriwaki, K., Yoshida, Y., & Harwood, D. M. (1992). Cenozoic glacial history of Antarctica – A correlative synthesis. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing Company, Tokyo, pp. 773–780. Mu¨ller, P., Schmidt-Thome´, M., Kreuzer, H., Tessensohn, F., & Vetter, U. (1991). Cenozoic peralkaline magmatism at the western margin of the Ross Sea, Antarctica. Mem. Della Soc. Geol. Italiana, 46, 315–336. Naish, T. R., Woolfe, K. J., Barrett, P. J., Wilson, G. S., Atkins, C., Bohaty, S. M., Bucker, C., Claps, M., Davey, F. J., Dunbar, G., Dunn, A., Fielding, C. R., Florindo, F., Hannah, M. J., Harwood, D. M., Watkins, D. K., Henrys, S., Krissek, L. A., Lavelle, M., van der Meer, J., McIntosh, W. C., Niessen, F., Passchier, S., Powell, R. D., Roberts, A. P., Sagnotti, L., Scherer, R. P., Strong,
The Antarctic Continent in Gondwanaland
303
C. P., Talarico, F., Verosub, K. L., Villa, G., Webb, P.-N., & Wonik, T. (2001). Orbitally induced oscillations in the east Antarctic Ice Sheet at the Oligocene/Miocene boundary: Direct evidence from Antarctic margin drilling. Nature, 413, 719–723. O’Connell, D. R. H., & Stepp, T. M. (1993). Structure and evolution of the crust at the Transantarctic Mountains–Ross Sea crustal transition: Results from the Tourmaline Plateau seismic array of the GANOVEX V ship-to-shore seismic refraction experiment. Geol. Jb., 47, 229–276. Oglesby, R. (1999). Use of climate models to extend paleoclimatic data. In: P. J. Barrett, & G. Orombelli (Eds). Geological Records of Global and Planetary Change. Terra Antarctica Rep., 3, 131–149. Paech, H.-J. (1985). Comparison of the geologic development of southern Africa and Antarctica. Z. Geol. Wiss., 13, 399–415. Paech, H.-J. (2004). GEOMAUD, Part 1, Geologisches Jahrbuch, B96. Paech, H.-J. (2005). Present knowledge of the geology of central Dronning Maud Land, east Antarctica: Main results of the 1995/96 GeoMaud expedition. Geol. Jb., B97, 341–408. Paech, H.-J., Bauer, W., Piazolo, S., Jacobs, J., & Markl, G. (2005). Comparison of the geology of central Droning Maud Land, east Antarctica, with other areas of the Gondwana supercontinent. Geol. Jb., B97, 309–340. Palmeri, R., Ghiribelli, B., Ranalli, G., Talarico, F., & Ricci, C. A. (2007). Ultrahigh-pressure metamorphism and exhumation of garnet-bearing ultramafic rocks from the Lanterman Range (northern Victoria Land, Antarctica). J. Metamorphic Geol., 25(2), 225–243. Palmeri, R., Ghiribelli, B., Talarico, F., & Ricci, C. A. (2003). Ultra-high-pressure metamorphism in felsic rocks: the garnet-phengite gneisses and quartzites from the Lanterman Range, Antarctica. Eur. J. Mineral., 15, 513–525. Piper, J. D. A. (1982). The Precambrian palaeomagnetic record: The case for the Proterozoic supercontinent. Earth Planet. Sci. Lett., 59, 61–89. Piper, J. D. A. (2000). The Neoproterozoic supercontinent: Rodinia or palaeopangea? Earth Planet. Sci. Lett., 176, 131–146. Powell, C. M., Li, Z. X., McElhinny, M. W., Meert, J. G., & Park, J. K. (1993). Paleomagnetic constraints on timing of the Neoproterozoic breakup of Rodinia and the Cambrian formation of Gondwana. Geology, 21, 889–892. Randall, D. E., & Mac Niocaill, C. (2004). Cambrian palaeomagnetic data confirm a natal embayment location for the Ellsworth–Withmore Mountains, Antarctica, in Gondwana reconstructions. Geophys. J. Int., 157, 105–116. Reitmayr, G. (1997). Gravity studies of Victoria Land and adjacent oceans, Antarctica. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antarctica, Siena, pp. 597–602. Retallack, G. J., Jahen, A. H., Sheldon, N. D., Chakrabarti, R., Metzger, C. A., & Smith, R. M. H. (2005). The Permian–Triassic boundary in Antarctica. Antarctic Sci., 17, 241–258.
304
F. M. Talarico and G. Kleinschmidt
Ring, U., & Betzler, C. (1993). Architecture, tectonics and sedimentology of the Malawi Rift (east Africa). Z. Dtsch. Ges. Geowiss., 144, 30–44. Rocchi, S., Armienti, P., D’Orazio, M., Tonarini, S., Wijbrans, J., & Di Vincenzo, G. (1999). Eocene–Oligocene rift related plutonic to subvolcanic magmatism in the western Ross embayment, Antarctica. In: D. N. B. Skinner (Ed.). 8th International Symposium on Antarctic Earth Sciences, 5–9 July, Wellington, Programme and Abstracts. The Royal Society of New Zealand, Wellington, p. 269. Rocchi, S., Fioretti, A. M., & Cavazzini, G. (2002). Petrography, geochemistry, and geochronology of the Cenozoic Cape crossfire, Cape King, and no ridge igneous complexes (northern Victoria Land, Antarctica). In: J. A. Gamble, D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium. R. Soc. N. Z. Bull., 35, 215–225. Rossetti, F., Lisker, F., Storti, F., & La¨ufer, A. L. (2003). Tectonic and denudation history of the Rennick Graben (north Victoria Land): Implications for the evolution of rifting between east and west Antarctica. Tectonics, 22, 11–18. Rowell, A. J., Rees, M. N., Duebendorfer, E. M., Wallin, E. T., Van Schmus, W. R., & Smith, E. I. (1993). An active Neoproterozoic margin: Evidence from the Skelton Glacier area, Transantarctic Mountains. J. Geol. Soc. Lond., 150, 677–682. Royer, J.-Y., & Rollet, N. (1997). Plate-tectonic setting of the Tasmanian region. In: N. F. Exon, & A. J. Crawford (Eds). West Tasmanian Margin and Offshore Plateaus: Geology, Tectonic and Climatic History, and Resource Potential. Aust. J. Earth Sci., 44, 543–560. Salvini, F., Brancolini, G., Busetti, M., Storti, F., Mazzarini, F., & Coren, F. (1997). Cenozoic geodynamics of the Ross Sea region, Antarctica: crustal extension, intraplate strike–slip faulting, and tectonic inheritance. J. Geophys. Res., 102, 24669–24696. Salvini, S., & Storti, F. (2003). Do transform faults propagate and terminate in east Antarctica continental lithosphere? Terra Nostra, 4, 283–284. Schma¨dicke, E., & Will, T. M. (2006). First evidence of eclogite facies metamorphism in the Shackleton Range, Antarctica: Trace of a suture between east and west Gondwana? Geology, 34(3), 133–136. doi: 10.1130/G22170.1. Schmidt, D. L., & Rowley, P. D. (1986). Continental rifting and transform faulting along the Jurassic Transantarctic Rift, Antarctica. Tectonics, 5, 279–291. Shackleton, R. M. (1996). The final collision zone between east and west Gondwana: Where is it? J. Afr. Earth Sci., 23(3), 271–287. Sloan, L. C., Crowley, T. J., & Pollard, D. (1996). Modeling of middle Pliocene climate with the NCAR GENESIS general circulation model. Mar. Micropaleontol., 27(1–4), 51–61. Smellie, J. L., & Lo´pez-Martı´ nez, J. (2002). Geology and geomorphology of Deception Island. In: BAS GEOMAP Series Sheets 6-A and 6-B, 1:25.000. British Antarctic Survey, Cambridge.
The Antarctic Continent in Gondwanaland
305
Stagg, H. M. J., & Willcox, J. B. (1992). A case for Australia–Antarctica separation in the Neocomian (ca. 125 Ma). Tectonophysics, 210, 21–32. Stern, R. J. (1994). Arc assembly and continental collision in the Neoproterozoic east African Orogeny – Implications for the consolidation of Gondwana. Annu. Rev. Earth Planet. Sci., 22, 319–351. Stern, T., & ten Brink, U. S. (1989). Flexural uplift of the Transantarctic Mountains. J. Geophys. Res., 94, 10315–10330. Stock, J., & Molnar, P. (1987). Revised history of early tertiary plate motion in the southwest Pacific. Nature, 325, 495–499. Storey, B. C. (1996). Microplates and mantle plumes in Antarctica. Terra Antartica, 3, 91–102. Storey, B. C., Dalziel, I. W. D., Garrett, S. W., Grunow, A. M., Pankhurst, R. J., & Vennum, W. R. (1998). West Antarctica in Gondwanaland: Crustal blocks, reconstruction and breakup processes. Tectonophysics, 155, 381–390. Storey, B. C., Hole, M. J., Pankhurst, R. J., Millar, I. L., & Vennum, W. R. (1988). Middle Jurassic within-plate granites in west Antarctica and their bearing on the breakup of Gondwanaland. J. Geol. Soc. Lond., 145, 999–1007. Storey, B. C., & Kyle, P. R. (1999). The Weddell Sea region: A plume impact site and source for Gondwana magmas during the initial stages of Gondwana breakup. In: D. N. B. Skinner (Ed.). 8th International Symposium on Antarctic Earth Sciences, 5–9 July, Programme and Abstracts. Victoria University of Wellington, Wellington, New Zealand, p. 292. Storey, B. C., Leat, P. T., Weaver, S. D., & Pankhurst, R. J. (1999). Mantle plumes and Antarctica–New Zealand rifting: Evidence from mid-Cretaceous mafic dykes. J. Geol. Soc. Lond., 156, 659–671. Storey, B. C., Pankhurst, R. J., & Johnson, A. C. (1994). The Grenville Province within Antarctica: A test of the SWEAT hypothesis. J. Geol. Soc., 151, 1–4. Storey, B. C., Thompson, M. R. A., & Meneilly, A. W. (1987). The Gondwanian orogeny within the Antarctic Peninsula: A discussion. In: G. D. McKenzie (Ed.). Gondwana Six: Structure, Tectonics, and Geophysics. Geophysical Monograph. American Geophysical Union, Washington, DC, Vol. 40, pp. 191–198. Storey, B. C., Vaughan, A. P. M., & Millar, I. L. (1996). Geodynamic evolution of the Antarctic Peninsula during Mesozoic times and its bearing on Weddell Sea history. In: B. C. Storey, E. C. King, & R. A. Livermore (Eds). Weddell Sea Tectonics and Gondwana Breakup. Geological Society, London, pp. 87–103. Special publication. Studinger, M., Bell, R. E., Buck, W. R., Karner, G. D., & Blankenship, D. D. (2004). Subglacial geology inland of the Transantarctic Mountains in light of new aerogeophysical data. Earth Planet. Sci. Lett., 220, 391–408. Stump, E. (1995). The Ross Orogen of the Transantarctic Mountains. Cambridge University Press, Cambridge, 284 pp. Stump, E., & Fitzgerald, P. G. (1992). Episodic uplift of the Transantarctic Mountains. Geology, 20, 161–164.
306
F. M. Talarico and G. Kleinschmidt
Stump, E., Sheridan, M. F., Borg, S. G., & Sutter, J. E. (1980). Early Miocene subglacial basalts, the east Antarctic Ice Sheet and uplift of the Transantarctic Mountains. Science, 207, 757–759. Talarico, F., Kleinschmidt, G., & Henjes-Kunst, F. (1999). An ophiolitic complex in the northern Shackleton Range, Antarctica. Terra Antartica, 6, 293–315. Talarico, F., Palmeri, R., & Ricci, C. A. (2004). Regional metamorphism and P-T evolution of the Ross orogen in northern Victoria Land (Antarctica): A review. Periodico Mineralogia, 73, 185–196. ten Brink, U. S., Bannister, S., Beaudoin, B. C., & Stern, T. A. (1993). Geophysical investigations of the tectonic boundary between east and west Antarctica. Science, 261, 45–50. ten Brink, U. S., Hackney, R. I., Bannister, S., Stern, T. A., & Makovsky, Y. (1997). Uplift of the Transantarctic Mountains and the bedrock beneath the east Antarctic Ice Sheet. J. Geophys. Res., 102, 27603–27621. ten Brink, U. T., & Stern, T. (1992). Rift flank uplifts and hinterland basins: Comparison of the Transantarctic Mountains with the Great Escarpment of southern Africa. J. Geophys. Res., 97, 569–585. Tessensohn, F. (1994a). Structural evolution of the northern end of the Transantarctic Mountains. In: F. M. van der Wateren, A. L. L. M. Verbers, & F. Tessensohn (Eds). Landscape Evolution in the Ross Sea Area, Antarctica. Rijks Geologische Dienst, Haarlem, pp. 57–61. Tessensohn, F. (1994b). The Ross Sea region, Antarctica: Structural interpretation in relation to the evolution of the Southern Ocean. Terra Antartica, 1, 553–558. Tessensohn, F., & Henjes-Kunst, F. (2005). Northern Victoria Land terranes, Antarctica: Far-travelled or local product? In: A. P. M. Vaughan, P. T. Leat, & R. J. Pankhurst (Eds). Terrane Processes at the Margins of Gondwana. Geol. Soc. Lond., 246, 275–291. Special publication. Tessensohn, F., Kleinschmidt, G., Talarico, F., Buggisch, W., Brommer, A., HenjesKunst, H., Kroner, U., Millar, I. L., & Zeh, A. (1999). Ross-age amalgamation of east and west Gondwana: Evidence from the Shackleton Range, east Antarctica. Terra Antartica, 6, 317–325. Tessensohn, F., & Wo¨rner, G. (1991). The Ross Sea rift system, Antarctica: Structure, evolution, and analogues. In: M. R. A. Thompson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 273–277. Tingley, R. J. (1991). The Geology of Antarctica. Oxford University Press, Oxford. Trehu, A. M. (1989). Crustal structure in the Ross Sea, Antarctica: Preliminary results from GANOVEX V. EOS Trans. Am. Geophys., 70, 1344. Trouw, R. A. J., & De Wit, M. J. (1999). Relation between Gondwanide orogen and contemporaneous intracratonic deformation. J. Afr. Earth Sci., 28, 203–213. Trouw, R. A. J., Ribeiro, A., & Paciullo, F. V. P. (1991). Structural and metamorphic evolution of the Elephant Island Group and Smith Island, south
The Antarctic Continent in Gondwanaland
307
Shetland Islands. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge. van der Beek, P., Cloetingh, S., & Andriessen, P. (1994). Mechanisms of extensional basin formation and vertical motions at rift flanks: Constraints from tectonic modeling and fission track thermochronology. Earth Planet. Sci. Lett., 121, 417–433. Van der Wateren, F. M., Dunaia, T. J., Van Balena, R. T., Klasc, W., Verbersd, A. L. L. M., Passchiere, S., & Herpersc, U. (1999). Contrasting neogene denudation histories of different structural regions in the Transantarctic Mountains rift flank constrained by cosmogenic isotope measurements. Global Planet. Change, 23(1–4), 145–172. Vaughan, A. P. M., & Storey, B. C. (1997). Mesozoic geodynamic evolution of the Antarctic Peninsula. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 373–382. Veevers, J. J., Powell, C. M., & Rotts, S. R. (1990). Review of seafloor spreading around Australia. I. Synthesis of the patterns of spreading. Aust. J. Earth Sci., 38, 373–389. Vetter, U., & Tessensohn, F. (1987). S- and I-type granitoids of north Victoria Land, Antarctica, and their inferred geotectonic setting. Geol. Rundsch., 76, 233–243. Viereck-Go¨tte, L., Scho¨ner, R., Bomfleur, B., & Schneider, J. (2007). Multiple shallow level sill intrusions coupled with hydromagmatic explosive eruptions mark the initial phase of ferrar magmatism in north Victoria Land, Antarctica. In: A. K. Cooper, A. C. Raymond, & ISAES Editorial Team (Eds). 10th International Symposium on Antarctic Earth Sciences Program Book, Santa Barbara, p. 60. White, P. (1989). Downhole logging. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. DSIR Bull., 245, 7–14. White, R. S., & MacKenzie, D. (1989). Magmatism at rift zones: The generation of volcanic continental margins and flood basalts. J. Geophys. Res., 94, 7685–7729. Wilson, G. S. (1995). The Neogene Antarctic Ice Sheet. A dynamic or stable feature? Quaternary Sci. Rev., 14, 101–123. Wilson, T. J. (1999). Cenozoic structural segmentation of the Transantarctic Mountains rift flank in southern Victoria Land. Global Planet. Change, 23, 105–127. Woolfe, K. J., & Barrett, P. J. (1995). Constraining the Devonian to Triassic tectonic evolution of the Ross Sea sector. Terra Antartica, 2, 7–21. Wo¨rner, G., Viereck, L., Hertogen, J., & Niephaus, H. (1989). The Mt. Melbourne volcanic field (Victoria Land, Antarctica) II. Geochemistry and magma genesis. Geol. Jb., E38, 395–433. Wrenn, J. H., & Webb, P.-N. (1982). Physiographic analysis and interpretation of the Ferrar Glacier–Victoria Valley area, Antarctica. In: C. Craddock
308
F. M. Talarico and G. Kleinschmidt
(Ed.). Antarctic Geosci. University of Wisconsin Press, Madison, WI, pp. 1091–1099. Yoshida, M. (1992). Precambrian tectonothermal events in east Gondwanian crustal fragments and their correlation (IGCP-288). In: Japan Contribution to the IGCP. IGCP National Committee of Japan, Tokyo, pp. 51–62. Yoshida, M., & Kizaki, K. (1983). Tectonic situation of Lu¨tzow-Holm Bay in east Antarctica and its significance in Gondwanaland. In: R. L. Oliver, P. R. James, & L. B. Jago (Eds). Antarctic Earth Science. Australian Academy of Science, Canberra, pp. 36–39. Young, G. M. (1995). Are Neoproterozoic glacial deposits pre erved on the margins of Laurentia related to the fragmentation of two supercontinents? Geology, 23, 153–156.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00008-6
Chapter 8
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica J. E. Francis1, S. Marenssi2, R. Levy3, M. Hambrey4,, V. C. Thorn1, B. Mohr5, H. Brinkhuis6, J. Warnaar6, J. Zachos7, S. Bohaty7 and R. DeConto8 1
School of Earth & Environment, University of Leeds, Leeds, LS2 9JT, UK Instituto Anta´rtico Argentino, Universidad Buenos Aires and CONICET, Buenos Aires 1010, Argentina 3 Geosciences, University of Nebraska Lincoln, Lincoln, NE 68588, USA 4 Institute of Geography and Earth Sciences, Aberystwyth University, Ceredigion SY23 3DB, Wales, UK 5 Humboldt-Universita¨t zu Berlin, Museum fu¨r Naturkunde, D-10099 Berlin, Germany 6 Palaeoecology, Department of Biology, Faculty of Sciences, Laboratory of Palaeobotany and Palynology, Utrecht University, Budapestlaan 4, 3584CD Utrecht, The Netherlands 7 Earth and Planetary Sciences Department, University of California, Santa Cruz, CA 95064, USA 8 Department of Geosciences, University of Massachusetts, Amherst, MA 01003, USA 2
ABSTRACT The change from a warm, ice-free greenhouse world to the glacial Antarctic icehouse occurred during the latest Eocene–earliest Oligocene. Prior to this, during the Early–Middle Eocene, Antarctica experienced warm climates, at least on the margins of the continent where geological evidence is present. Climates appear to have been warm and wet, the seas were warm and plants flourished in a frost-free environment, although there is some suggestion of valley glaciers on King George Island. Climate signals in the geological record show that the climate then cooled but not enough to allow the existence of significant ice until the latest Eocene. Glacial deposits on Seymour Island indicate that ice was present there at Eocene/Oligocene boundary times. Further south, ice-rafted clasts in drill cores Corresponding author. Tel.: +44(0)1970 621860;
E-mail:
[email protected] (M. Hambrey).
310
J. E. Francis et al.
from the Ross Sea region and deposits of tidewater glacier origin in Prydz Bay confirm the presence of ice at the continental shelf by the earliest Oligocene. This matches the major Oi-1 oxygen isotope event in the marine record. On land, vegetation was able to persist but the warmth-loving plants of the Eocene were replaced by shrubby vegetation with the southern beech Nothofagus, mosses and ferns, which survived in tundra-like conditions. Throughout the Oligocene, glaciation waxed and waned until a major glacial phase in the Miocene. Coupled climate–ice sheet modelling indicates that changing levels of atmospheric CO2 controlled Antarctica’s climate. Factors such as mountain uplift, vegetation changes and orbital forcing all played a part in cooling the polar climate, but only when CO2 levels reached critical thresholds was Antarctica tipped into its icy glacial world.
8.1. Introduction One of the most intriguing challenges in Antarctic Earth history is to understand the fundamental climate change from the past greenhouse world with no major polar ice caps to our present icehouse that is dominated by the vast ice sheets on the Antarctic continent. This change across a major climate threshold holds many clues that will help us understand the potential changes our world may undergo in future. Geological evidence from rocks and fossils from the Antarctic continent and from marine oxygen isotopes that record changes in temperature, ice volume and water masses indicate that ice sheets built up on Antarctica from about Eocene/Oligocene (E/O) boundary times, approximately 34 million years ago. This is named the Oi-1 event in the marine oxygen isotope record and is represented by the appearance of several glacial deposits in the rock record. However, the actual pattern of climate cooling and the causes of glaciation are far from understood, and there are hints of the presence of ice in the Late Eocene. Indeed, it is possible that ice existed on Antarctica even during the Cretaceous (Miller et al., 2005; Tripati et al., 2005). This chapter reviews our current understanding of the greenhouse– icehouse transition in Antarctica. It covers the interval of climate change from the warm greenhouse climates of the Early Eocene through to the first appearance of ice and the establishment of glacial conditions during the Oligocene. Several lines of evidence are presented for climate and environmental change: the sedimentary rock record on the continent provides clues to the nature of the cooling climate during the Eocene; latest
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
311
Eocene and Oligocene sediments recovered from marginal basins in drill cores contain the earliest undisputed glacial deposits; fossil plants and palynomorphs from both the continent and marginal basins have yielded information about cooling climates in the terrestrial realm; marine microfossils hold clues to ocean circulation and the significance of ocean gateways; the marine isotope record of the open oceans tells us about changes in deep ocean temperatures as the result of climate cooling and ice growth; and finally, the cause of cooling and the birth of the icehouse world is explored through computer modelling. The final section summarizes our current understanding of the greenhouse–icehouse transition in Antarctica.
8.2. Climate Signals from the Sedimentary Record Sedimentary strata of Eocene and Oligocene age are exposed in the Antarctic Peninsula region and provide an important record of environmental conditions on land and in shallow marine settings during this period. In addition, there are intriguing hints about environments at higher latitudes, extracted from erratic boulders composed of Eocene and Oligocene fossiliferous sediments in the Ross Sea region, derived from sub-glacial outcrops. A summary of the important outcrops and their environmental signal is given below. Sedimentary rocks of Palaeogene age are exposed around the northern part of the Antarctic Peninsula, on the South Shetland Islands and on Seymour Island (Fig. 8.1). The sediments were deposited in very different tectonic settings and environments – the South Shetland sequence is of terrestrial volcanic and sedimentary deposits that represent an outer-arc (Birkenmajer, 1995) or fore-arc (Elliot, 1988) succession; the sequence on Seymour Island consists of marine clastics deposited in a back-arc basin, the uppermost beds of a regressive megasequence (Hathway, 2000). Both contain evidence of Palaeogene cooling and the first appearance of ice but in different settings.
8.2.1. Antarctic Peninsula Region The Palaeogene back-arc deposits exposed on Seymour Island and Cockburn Island comprise more than 1,000 m of shallow marine to coastal fossiliferous clastic sedimentary rocks mainly of Palaeocene and Eocene age (Elliot, 1988; Sadler, 1988; Marenssi et al., 1998). The Maastrichtian–Danian sequence forms a simple BN–S homocline dipping gently to the east while the Late Palaeocene and the Eocene sediments fill incised valleys trending NW–SE.
312
J. E. Francis et al.
58°
South Shetland Islands
55°W
Fildes South Shetland Islands Peninsula
South Orkney Islands (Bruce Bank)
King George Island Admiralty Bay Nelson Island
Antarctic Peninsula
Prince Charles Mtns
N
it
ra
ld
St
fie
63°
s
Prydz Bay
an
Br
Joinville Island
A
ity
Pe
ni
ns
a ul
64°S
°S
Cockburn Island
60
c Mtns
Ross Sea
°
rcti
McMurdo erratics
in Tr
70
ta an
80
°
Tra ns
Seymour Island James Ross Island
B
Figure 8.1: Map of the Antarctic continent (A) and the Antarctic Peninsula (B) showing some of the locations mentioned in the text. The Early Eocene to latest Late Eocene La Meseta Formation (Elliot and Trautman, 1982) is an unconformity-bounded unit (La Meseta Alloformation of Marenssi et al., 1996, 1998) of maximum composite thickness of 720 m, which fills a 7 km wide valley cut down into older strata after the regional uplift and tilting of the Palaeocene and Cretaceous older beds. The La Meseta Formation comprises mostly poorly consolidated siliciclastic fine-grained sediments deposited in deltaic, estuarine and shallow marine environments as part of a tectonically controlled incised valley system (Marenssi, 1995; Marenssi et al., 2002). The richly fossiliferous Eocene sediments have yielded the only fossils of land mammals in the whole Antarctic continent (Reguero et al., 2002) along with fossil wood, fossil leaves, a rare flower, plus marine vertebrates (including giant penguins) and invertebrates (Stilwell and Zinsmeister, 1992; Gandolfo et al., 1998a,b,c; Francis et al., 2004; Tambussi et al., 2006). Overall, evidence from fossils, sediments and geochemistry from Seymour Island indicates generally warm and ice-free conditions during the earliest part of the Eocene but followed by gradual cooling. Dingle et al. (1998), based on chemical analysis (chemical index of alteration, CIA) and clay mineralogy, recorded a climatic deterioration from warm, non-seasonally wet conditions during the early Middle Eocene (smectite-dominated clay assemblage and CIA values o0.7 to W0.6) to a latest Eocene cold, frostprone and relatively dry regime (illite-dominated clay association and CIA
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
313
values o0.6). Gazdzicki et al. (1992) showed a 6 m decrease in d13C values in biogenic carbonates and Tatur et al. (2006) recorded an increase in the Cd/Ca ratio in bivalve shells from the upper part of the La Meseta Formation. They interpreted these results as a change from stratified to vigorously mixed oceanic conditions related to the cryosphere development in the Southern Ocean by the end of the Eocene. Unpublished stable isotope (d18O and d13C) measurements (Feldmann and Marenssi) obtained from molluscan shells of the La Meseta Formation show a period of warmer seawater temperatures during 51–47 Ma and a drop of 1.51C during 35–34 Ma. Further, Ivany et al. (2008), also undertaking isotopic analysis of bivalves from the La Meseta Formation, propose the possibility of winter sea-ice formation in this region during the late Middle Eocene (c. 37 Ma). They suggest mean ocean temperatures have approached freezing at this time by assuming a sea water d18O value of 3m as predicted for the east Antarctic Peninsula by Huber et al. (2003) in their early Eocene climate model. This information is consistent with ice-rafted debris reported for the upper part of the La Meseta Formation (Doktor et al., 1988) and palaeoclimatic evidence of a severe climatic deterioration towards the end of the Eocene. It is possible that by the end of the Eocene, limited ice, perhaps as valley glaciers, was already present in the area. By the end of the Eocene, it is possible that an ice sheet extended over much of the peninsula, although average Seymour Island Shelf temperatures did not reach below zero (Ivany et al., 2008). Ivany et al. (2006) reported 5–6 m thick glacial deposits that conformably overlie the typical marine sandstones of the La Meseta Formation but are beneath the glacial diamictites of the younger Weddell Sea Formation. Based on dinocyst stratigraphy and strontium isotopes, these authors suggested an age of 33.57–34.78 Ma for these glaciomarine deposits (supported by marine Sr dates from Dingle and Lavelle, 1998; Dingle et al., 1998; Dutton et al., 2002), at or very close to the E/O boundary. There is therefore a small but intriguing window into the early stages of the icehouse world in the James Ross Basin. 8.2.2. King George (25 de Mayo) Island, South Shetland Islands The South Shetland Islands contain a different story of early glacial events within a fore-arc/outer-arc terrestrial setting. King George Island and neighbouring Nelson Island consist of several tectonic blocks bounded by two systems of strike–slip faults of Tertiary (54–21 Ma) age (Birkenmajer, 1989). Thus, considerable differences in
314
J. E. Francis et al.
stratigraphic succession, age and character of the rocks occur between particular blocks. The stratigraphic sequence includes mainly Late Cretaceous to Early Miocene island-arc extrusive and intrusive rocks comprising mainly terrestrial lavas, pyroclastic and volcaniclastic sediments often with terrestrial plant fossils. Hypabyssal dykes and plutons intrude the latter. Fossiliferous marine and glaciomarine sediments are also represented that provide clues to palaeoclimates. Several sequences of tillites crop out within these complicated sequences, representing glacial and interglacial events. Reports of supposed Eocene-age tillites at Magda Nunatak (Birkenmajer, 1980a,b), named the Krakow Glaciation and dated at 49 Ma, have been disproved by Sr dating (Dingle and Lavelle, 1998). However, Birkenmajer et al. (2005) describe valley-type tillites between lava sequences in the E/O Point Thomas Formation in Admiralty Bay. K–Ar dating gives ages of 41–45 Ma for the lavas below the tillites and 45–28 Ma in the lavas above. Birkenmajer et al. thus propose that a glacial period occurred at 45–41 Ma during the Middle Eocene, being the oldest record of alpine glaciers in West Antarctica. A clear record of glacial activity is present as diamictites and ice-rafted deposits within the Polonez Cove Formation, of mid-Oligocene age (26–30 Ma) (Troedson and Smellie, 2002). This is called the Krakowiak Glacial Member. At its maximum extent, ice was grounded on a shallow marine shelf. Interestingly, exotic clasts within this sequence may represent ice-rafted debris that was derived from as far away as the Transantarctic Mountains, suggesting marine-based glaciation in the Weddell Sea region. Non-glacial sediments that overlie the Polonez Cove Formation signal an interglacial period in the Late Oligocene (26–24.5 Ma), before another glacial phase in the Miocene. The only other terrestrial evidence for Oligocene ice, possibly representing local alpine glaciation, is from Mount Petras in Marie Byrd Land, West Antarctica, where deposits indicate volcanic eruptions beneath ice (Wilch and McIntosh, 2000).
8.2.3. Eocene Environments in the Ross Sea Region There is an intriguing record of Eocene pre-glacial environments in the Ross Sea region, around McMurdo Sound. Although the location of the Eocene outcrop is presumably under the ice sheet, there are no major exposures. A glimpse of Eocene environments is, however, provided by several hundred erratic boulders and cobbles recovered from coastal moraines around the shores of Mount Discovery, Brown Peninsula and Minna Bluff (Fig. 8.1).
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
315
These fossiliferous glacial erratics (called the McMurdo Erratics), recovered from moraine around the northwest coast of Mount Discovery and Minna Bluff in southern McMurdo Sound, provide a window into the environment that may have existed along the coast of the gradually rising Transantarctic Mountains during the Eocene. The erratics are most likely derived from sub-glacial basins, such as Discovery Deep, that lie along the coast of the Transantarctic Mountains or basement highs situated to the east of the discovery accommodation zone (Wilson, 1999; Wilson et al., 2006). The erratics were distributed into their distinctive pattern of terminal and lateral retreat moraines during relatively recent advance and retreat of grounded ice into southern McMurdo Sound (Wilson, 2000). Subsequent basal adfreezing and surface ablation has transported the erratics to the surface of the McMurdo Ice Shelf. Although currently out of their original stratigraphic position, this suite of erratics provides us with a mechanism to obtain geologic data that are otherwise buried beneath the Antarctic Ice Sheets and fringing ice shelves. The McMurdo Erratics comprise a range of lithotypes and ages. Eocene rocks contain a rich suite of fossil floras and faunas including marine and terrestrial palynomorphs, diatoms, ebridians, marine vertebrates and invertebrates, terrestrial plant remains and a bird humerus. Biostratigraphic data from dinoflagellate cyst (dinocyst), ebridian and mollusc assemblages recovered from many of the erratics indicate that the majority of fossiliferous rocks range from Middle to Late Eocene, B43–34 Ma. Erratics collected between 1993 and 1996 (Stilwell and Feldmann, 2000) include several hundred samples of Oligocene, Miocene and Pliocene sediment. Although relatively rich dinocyst assemblages have been described from Oligocene Miocene Sequences from the Cape Roberts Cores, assemblages in post Eocene erratics comprise few taxa (typically o5 species). This general paucity of dinocyst species in late Palaeogene and Neogene sequences is observed in several other sites from the southern high latitudes (Wilson, 1989; Mao and Mohr, 1995) and may be a reflection of geographical and thermal isolation of Antarctica (McMinn, 1995; Williams et al., 2004) or preservation. The majority of the Eocene erratics record a suite of lithofacies that were deposited in coastal–terrestrial to inner shelf marine environments (Levy and Harwood, 2000a,b). These sediments were probably deposited within fan deltas that formed along the rugged coastline of the rapidly rising Transantarctic Mountains. Abundant macroinvertebrate faunas, including bivalves, gastropods, scaphopods, cirripeds, bryzoans, decapods and brachiopods, indicate that many of these sediments were deposited in a spectrum of predominantly shallow marine environments. The presence of terrestrial plant material and palynomorphs also suggests that the
316
J. E. Francis et al.
majority of the rocks were formed in nearshore environments. However, the occurrence of open marine dinocyst species and the absence of benthic diatom taxa in many of the fine-grained lithofacies indicate that outer shelf open marine environments were also present in the source region. The Eocene erratics contain no direct or unequivocal sedimentological evidence for the presence of ice close to the basins in which the sediments were originally deposited. It is notable that erratics composed of diamictites recovered from the coastal moraines are all Oligocene and younger in age. Although rare, fossil leaves, wood and pollen recovered from several erratics provide a glimpse of the Eocene climate for the region. One erratic contains wood and leaves from Araucaria and Nothofagus trees, which suggests that the climate was not extreme. Cool temperate conditions with some winter snow may have been possible but temperatures were probably not cold enough to allow extensive ice at sea level (Francis, 2000; Pole et al., 2000). Spore and pollen assemblages recovered from the erratics reflect Nothofagus-podocarpaceous conifer-Proteaceae vegetation with other angiosperms growing in temperate climate conditions (Askin, 2000). Oligocene and younger erratics show a major drop in species richness, which is also noted in sequences recovered in CIROS-1 and the Cape Roberts Project (CRP) cores (Mildenhall, 1989; Raine and Askin, 2001). Fossil invertebrate remains recovered from the erratics include a humerus shaft from a pseudodontorn (giant bony-toothed sea bird) (Jones, 2000), a probable crocodile tooth (Willis and Stilwell, 2000) and teeth from two species of shark (Long and Stilwell, 2000). The small but significant record of East Antarctic invertebrate fauna indicates a temperate to cool temperate marine environment. 8.2.4. Climate Evidence from Drilling on the Antarctic Margin The onset of glaciation in Antarctica is not yet well constrained, largely because no cores have yet been obtained that unequivocally provide a continuous transition from no-ice to ice-sheet scale glaciation. This is because (i) most cores terminated before the base of the glacigenic sediments was reached, (ii) a hiatus exists at the base of the glacigenic strata, as in CRP-3 in the Ross Sea (Barrett and Ricci, 2001a,b) and Ocean Drilling Program (ODP) Site 1166 in Prydz Bay (Shipboard Scientific Party, 2001a,b,c,d), or (iii) the age models based on multiple criteria have been revised several times (e.g. CIROS-1; Wilson et al., 1998a). On land, there is evidence for ice proximal-fjordal sedimentation possibly dating back to Oligocene time in the Prince Charles Mountains (Hambrey and McKelvey,
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
317
2000a; McKelvey et al., 2001) and some exposures of the Sirius Group in the Transantarctic Mountains could also be this old (Sugden and Denton, 2004).
8.2.4.1. Drill cores in the western Ross Sea – CIROS-1 and Cape Roberts Two drill cores have been recovered from the western Ross Sea that approach or even cross the E/O boundary (Fig. 8.2). Drilling was undertaken from a sea-ice platform in spring-time, and was characterized
Figure 8.2: Location of the drill sites in the western Ross Sea. CIROS-1 and CRP-3 are those that bear on the E/O question (from Hambrey et al., 2002). Reproduced with permission of The Geological Society Publishing House, Bath, UK.
318
J. E. Francis et al.
by exceptionally high recovery (up to 98%). In the 702 m deep CIROS-1 hole (Barrett, 1989; Barrett et al., 1991), the lower part of the core was originally regarded as Late Eocene, with a breccia passing up into mudstone and sandstone. The boundary with the Oligocene was originally placed at about 570 m (Barrett et al., 1989), but magnetobiostratigraphic data (Wilson et al., 1998a) suggest that the E/O boundary is much higher at about 410–420 m. In either case, there is no obvious lithological transition, these finer grained facies including alternations of weakly stratified sand and mud, with intraformational conglomerate and occasional diamictite. These facies are strongly bioturbated and contain exotic clasts, as well as intraclasts, while some beds are graded. Moving up core, a major hiatus exists at 366 m, which coincides with the Early/Late Oligocene boundary. Above is a suite of fining-upwards sandstone beds, followed by alternating massive and stratified diamictites and thin interbeds of sand and mud (Hambrey et al., 1989). The CIROS-1 core was originally interpreted in terms of depositional setting, ice proximity and water depth (Hambrey et al., 1989; Hambrey & Barrett, 1993; Barrett, 1996). The breccia at the base of the hole is interpreted as a fault-brecciated conglomerate. The overlying sandstone/mudstone/ diamictite succession is marine, influenced to varying degrees by resedimentation and iceberg-rafting. Above the Early/Late Oligocene hiatus, the sandstones were regarded as fluvial, and the diamictites as basal glacial deposits, indicating ice overriding the site. However, Fielding et al. (1997) argued, based on a sequence stratigraphic analysis, that the Late Oligocene diamictite was also glaciomarine. In contrast, Hiemstra (1999) reverted in part to the original view of grounded ice on the basis of microstructural studies. Whichever solution is the correct one, there is no clear evidence for a major environmental shift at the E/O boundary, but there is one towards lower sea level at the Early/Late Oligocene transition. A record of climate change through the E/O transition has also been determined from the environmental magnetic record in the CIROS-1 core (Sagnotti et al., 1998). Variations in magnetite were related to the concentration of detrital material transported into the Victoria Land Basin, influenced by climate and weathering rates on the Antarctic continent (especially of the Ferrar Group). Sagnotti et al. (1998) determined, from changes in the abundance of magnetite, that although there were some cold dry intervals (35–36 and W36.5 Ma) alternating with warm humid climates during the Late Eocene, a stable cold dry climate was not established in Antarctica until the E/O boundary, with major ice-sheet growth occurring at the Early/Late Oligocene boundary. This pattern matches clay mineral history, which shows a shift from smectite-rich to smectite-poor assemblages
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
319
in Antarctica at the E/O boundary (Ehrmann and Mackensen, 1992; Ehrmann, 1997). The second core that contains the E/O transition is the Cape Roberts Project Core CRP-3; Barrett (2008; Chapter 3) has synthesized the wealth of data and numerous papers from this and other Cape Roberts cores. The strata were deposited in the same rift basin, the Victoria Land Basin, as CIROS-1. The basin floor comprises Early Devonian sandstone, above which is about 1,500 m of Cenozoic sediment. CRP-3 records a dolerite conglomerate and a basal sandstone breccia at the base of the Cenozic succession, believed to be of latest Eocene age (34 Ma). Above lies nearly 800 m of sandstone with thin beds of conglomerate, all of Early Oligocene age. Diamictite and sandstone occur towards the top of the hole, while outsized clasts are scattered through much of the core. Core CRP-2A/2A almost follows on directly above CRP-3 and spans the Late Oligocene/Early Miocene interval. The CRP-1 core overlaps with the top of CRP-2/2A core and extends further up into the Miocene. Alternating sequences of mudstone, sandstone, diamictite and conglomerate occur within this part of the Cenozoic record. The facies in CRP-3 and CRP-2A/2A represent a marine sedimentary record (Barrett, 2007). Conglomerate, sandstone and mudstone are typical of the coastal margin of a subsiding sedimentary basin, influenced by iceberg rafting. The diamictite beds additionally record tidewater glaciers that extended periodically beyond the coast. In these respects, the combined Cape Roberts cores represent an expanded record of that from CIROS-1, although there is little evidence of ice-grounding. Nevertheless, the facies fluctuations are thought to reflect glacioeustatic changes in sea level on a wave-dominated coast, in parallel with tidewater glacier fluctuations. In this context, diamicton and sand represent nearshore sedimentation, and this association grades upwards into shelf mud and then to inner shelf/shoreface sand. A conceptual depositional model for the Late Oligocene/Miocene interval was developed by Powell et al. (2000), based on facies associations and comparison with modern glaciomarine environments, such as those in Alaska and Greenland. This shows that during an advance and still-stand, a grounding-line fan develops, and this is followed by rapid recession until another fan develops. The sequence becomes even more complex when the glacier overrides previously formed fans. Figure 8.3 is a simplified version of this model. The huge volume of palaeoecological, mineralogical and geochemical data generated by these drilling projects enables us to gain insight into the broader environmental and climatic evolution through the early Cenozoic (Hambrey et al., 2002; Barrett, 2007). A temperate glacier regime is suggested for the Early Oligocene, followed by cooling into the Miocene, typified by
320
J. E. Francis et al.
Figure 8.3: Grounding-line fan model of glaciomarine sedimentation for Late Oligocene/Early Miocene time (from Hambrey et al., 2002; Simplified from Powell et al., 2000). Reproduced with permission of The Geological Society Publishing House, Bath, UK.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
321
Figure 8.4: Cartoon depicting the Victoria Land coast in Early/Late Oligocene time, with glacier- and river-influenced coast, and vegetated mountainsides and lowlands. This scenario is based on a combination of sedimentological evidence and floral data from CIROS-1 and Cape Roberts cores (from Hambrey et al., 2002). Reproduced with permission of The Geological Society Publishing House, Bath, UK. polythermal glaciers. The cold frigid regime of today only began at the end of Pliocene time. The Early Oligocene landscape was characterized by temperate glaciers flowing from the early East Antarctic Ice Sheet, some terminating in the sea, and others on braided outwash plains (Fig. 8.4).
8.2.5. The Prydz Bay Region Drilling in Prydz Bay was undertaken by two ship-borne legs of the ODP (Fig. 8.5). In contrast to the western Ross Sea cores, core-recovery rates here were much less satisfactory; hence interpreting depositional processes is more questionable. Nevertheless, plausible scenarios have been derived, albeit lacking precise constraints owing to core loss. Prydz Bay represents the continuation of the Lambert Graben, which contains the Lambert Glacier–Amery Ice Shelf System, an ice drainage basin covering approximately 1 M km2, draining 13% of the East Antarctic Ice Sheet by surface
322
J. E. Francis et al.
Figure 8.5: Location of drill sites in Prydz Bay, East Antarctica, from ODP legs 119 and 188. Continental shelf sites 742 and 1166 include strata the cross the E/O transition. From shipboard Scientific Party (2001a). Reproduced with permission of the Ocean Drilling Program, College Station, Texas. area. Thus, the record in Prydz Bay provides a signal of the ice sheet as a whole since its inception, and complements the Oligocene (?) to Pliocene uplifted glaciomarine record in the Prince Charles Mountains (see Haywood et al., 2008; Chapter 10). Prydz Bay itself is dominated by a trough-mouth fan that prograded during phases of glacier advance to the shelf break. Like the western Ross Sea, large data-sets are available covering all aspects of core analysis from ODP Legs 119 and 188 (Barron, Larsen, & Shipboard Scientific Party, 1991; Cooper and O’Brien, 2004; Cooper et al., 2004) and a convenient summary has been provided by Whitehead et al. (2006). ODP Leg 119 obtained two cores, 739 (480 m) and 742 (316 m), the lower parts of which were loosely dated as Middle Eocene to Early Oligocene. The dominant facies recovered was massive diamictite, with minor stratified diamictite and sand (Hambrey et al., 1991, 1992). Poorly consolidated finegrained facies may well have been washed away during the drilling, however, since core-recovery rates were less than 50%. A few broken shell fragments are present, but there is a dearth of material suitable for precise dating. The base of
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
323
Site 742 is represented by a zone of soft-sediment deformation. The Oligocene succession forms part of a prograding unit as defined in seismic profiles, but is truncated by a regional unconformity. Above lies a flat-lying sequence of more diamictite, some with preferred clast orientation and overcompaction, of late Miocene to Pliocene age (Cooper et al., 1991). Leg 188 drilled Site 1166 on the continental shelf near Sites 739 and 742 in order to obtain a more complete record of the E/O transition, but again core recovery was poor (19%). From the base upwards, Late Eocene matrixsupported sand was followed by a transgressive surface and the Late Eocene to Early Oligocene graded sand and diatom-bearing claystone with dispersed granules. These facies were capped, above an unconformity, by ‘‘clast-rich clayey sandy silt’’ (diamicton/ite) of Neogene age (Shipboard Scientific Party, 2001a,b,c,d). The interpretation of the Leg 119 facies is as follows: the deformed bed at the base of Site 742 may represent the first stages of glaciation, with the ancestral Lambert Glacier extending across the continental shelf for the first time. Then the bulk of the recovered facies in Sites 739 and 742 (diamictite) records deposition from the grounding-line of a tidewater glacier margin, by debris rain-out and submarine sediment gravity flow beyond the shelf break, conditions which characterize much of Early Oligocene time. The overlying Miocene succession is quite different, and represents successive advances across the already prograded shelf to produce sub-glacial, ice-proximal and ice-distal facies in alternation, forming the flat-lying sequence (Hambrey et al., 1991). Leg 188’s Site 1166 begins at the base with Late Eocene sand of fluvio-deltaic character, and is inferred to be pre-glacial. The overlying Late Eocene to Early Oligocene sand and claystone represent shallow marine and open marine conditions, respectively, but in a proglacial setting as indicated by ice-rafted granule-sized material. The Neogene strata that lie unconformably above represent full glacial conditions. Combining the sedimentary and seismic records, along with bathymetric data from the over-deepened Lambert Graben, a conceptual model of erosion and deposition has been developed (Fig. 8.6). Considering the whole Prydz Bay region in 3D, it is apparent that Early Oligocene time was characterized by sedimentation at the shelf break. A major change then took place in the Late Miocene to mid-Pliocene following the development of a fast-flowing ice stream that now occupied a more constrained channel, leading to the growth of the trough-mouth fan (Shipboard Scientific Party, 2001a,b,c,d; Taylor et al., 2004). Comparing data-sets from the western Ross Sea and Prydz Bay, it is apparent that ice first reached the continental shelf edge in earliest Oligocene time in Prydz Bay, while the Victoria Land coast was influenced by iceberg
324
J. E. Francis et al.
Figure 8.6: Conceptual model of styles of glacial sedimentation during phases of advance and recession of the ancestral Lambert Glacier (from Hambrey and McKelvey, 2000b). (A) ‘‘Cold polar’’ glaciation as occurs today. (B) Expanded ice in the Neogene Period. (C) Recessed ice in the Neogene period. Both B and C are characterized by polythermal glaciation (comparable to the High Arctic today), and may also have been a feature of the Palaeogene period. Data are too limited, however, to define the palaeoclimate for this earlier period. Reproduced with permission of The Geological Society of America, Boulder, Colorado.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
325
rafting. Late Oligocene time saw repeated ice expansions recorded in the CIROS-1 core by a succession of basal tills, and probably by the major unconformity in Prydz Bay. Subsequently, frequent expansions took place to the shelf edges in both areas, resulting in a fragmented stratigraphic record.
8.3. Climate Signals from the Terrestrial Realm – Fossil Plants and Palynomorphs 8.3.1. Plant Macrofossils An important source of palaeoclimate information on land comes from fossil plants (both macro- and microfossils) (Fig. 8.7). The fossil plant record suggests that during the Late Palaeocene to Early Eocene moist, cool temperate rainforests were present, similar to modern low to mid-altitude Valdivian rainforests in southern Chile. These forests were dominated by Nothofagus and conifer trees, with ferns, horsetails and some less-prominent angiosperm groups. Assemblages of Eocene fossil plants and palynomorphs signal a generally warm climate during the Early Eocene but conditions deteriorated throughout the Antarctic Peninsula through the latter part of the Eocene when cold, seasonal climates developed (Francis, 1991; Francis and Poole, 2002; Francis et al., 2004, 2008; Poole et al., 2005). Fossil plant assemblages of younger age have been found in glacial sediments in the Transantarctic Mountains, representing vegetation that grew on Antarctica under icehouse conditions. These include the Sirius Group flora from the Transantarctic Mountains (the age of which is problematic, Francis and Hill, 1996; Wilson et al., 1998a,b; Askin et al., 1999), and a new flora of Miocene age discovered in the Dry Valleys (Ashworth et al., 2008). Surprisingly, leaf fossils, although single leaves, have also been discovered in drill cores within Oligocene and Miocene glacial sequences. The early glacial world of Antarctica was clearly not totally barren of vegetation. Palaeogene fossil plants have been discovered from around the Antarctic margin in outcrop, in sea-floor cores and in glacial erratic boulders from the Antarctic Peninsula (King George Island and Seymour Island) and Ross Sea (McMurdo Sound and Minna Bluff) regions. The collections consist of compressions, impressions and petrifactions of leaves, seeds, flowers and wood indicating a high southern latitude flora of variable diversity but dominated by fossils comparable to modern Nothofagus (the southern beech) and conifer trees. Ferns, horsetails and additional significant Southern
PatagonianMagellanic rainforests, Chile
Valdivian rainforests, Chile
NothofagusPodocarpus dominated communities
Temperate rainforest
the small leaves suggest this may be too warm (Francis, 1999).
Warm (frost-free, MAT 11.7-15°C)
Recent fern-bush communities of southern oceanic islands, e.g. Gough and Auckland Islands
Nothofagus forest with well-developed undergrowth, where ferns (including tree ferns) are important
MAT c.10°C MAP c.1000mm
Moist (600-4300mm)
Cool temperate (MAT 5-8°C)
Moist (MAP 1220-3225mm)1
Inferred palaeoclimate
Modern analogue
Vegetation model
Figure 8.7: Antarctic Peninsula Late Palaeocene–Late Eocene fossil floras: composition, modern analogue vegetation and palaeoclimate interpretations. Refer to text for relevant references. MAT: mean annual temperature; MAP: mean annual precipitation.
1 Although
Stratigraphic Flora Composition age King George Island, South Shetland Islands: Cytadela/Platt Ferns Late Eocene Cliffs Nothofagus-type (mostly small) and other dicotyledon types ?Podocarpaceae Petrified Forest Fagus-Nothofagus Creek Araucaria Mount Wawel Late Equisetum Ferns PalaeoceneSeveral Nothofagus spp. Middle (microphyllous leaves) and a Eocene few other angiosperms Podocarpaceae Dragon Glacier Equisetum Moraine Ferns Angiosperms (dominated by Nothofagus spp.) Conifers (Araucariaceae, Cupressaceae, Podocarpaceae)
326 J. E. Francis et al.
Composition
40 taxa: Ferns Mixed broadleaf angiosperms (dominated by Nothofagus spp.) Conifers (podocarp, araucarian, cupressacean)
Ferns Angiosperms (including Nothofagoxylon spp., Weinmannioxylon eucryphioides, Myceugenelloxylon antarcticus Conifers (Araucaria, Cupressinoxylon, Podocarpoxylon fildesense)
Flora
Fossil Hill
Collins Glacier
3
Figure 8.7: (Continued ).
Temperate – warm temperate3 Slightly above 10 C with a low annual temperature range (MAT c.9 C), Abundant precipitation (c.1000mm) Cool temperate
Inferred palaeoclimate
Valdivian and Cool to cold temperate Magellanic rainforests, (MAT 11-13°C or Chile 10.8°C) with abundant rainfall (MAP 10003000mm) Seasonal.
Low-mid altitude Valdivian rainforests, Chile
Cool temperate rainforest
Cool – cold temperate rainforest
Tropical Latin America and southern South American rainforests
Modern analogue
Rainforest – mixed neotropical and subantarctic elements2
Vegetation model
Perhaps altitude related? The Nothofagus leaves in the Fossil Hill flora are much bigger than their modern relatives suggesting a warmer and more humid climate during the Middle Eocene.
2
Seymour Island, James Ross Basin: La Meseta Fm. Ferns, angiosperms and conifers Early-latest including: Nothofagus spp. Late Eocene (notophyllous at older Middle Eocene locality), Dilleniaceae, Myricaceae, Myrtaceae, Lauraceae, Proteaceae, Podocarpaceae, Araucariaceae
Stratigraphic age Late PalaeoceneMiddle Eocene (continued)
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica 327
328
J. E. Francis et al.
Hemisphere angiosperm families, including the Proteaceae, Myrtaceae and Lauraceae, are also represented. The macrofossil record has largely been described from isolated collections, even of single leaves in cores, and although reasonably dated as a whole, a comprehensive understanding of stratigraphic relationships between the floras, particularly in the Antarctic Peninsula region, is at present hampered by differing stratigraphic interpretations. 8.3.2. Antarctic Peninsula: King George Island Many macrofloras have been discovered on King George Island in the South Shetland Island group, north of the Bransfield Strait in the Antarctic Peninsula region, currently dated between Late Palaeocene and Late Eocene in age. The floras may have lived at a palaeolatitude of B621S, similar to its present-day location (Lawver et al., 1992). The stratigraphy is complex because Birkenmajer (1981, 1989, 1990) and Birkenmajer et al. (1986) erected many local formations in comparison to a simpler scheme created by Smellie et al. (1984). No single stratigraphic scheme exists and so the relationship between the floras is confusing. The stratigraphic framework used here includes both schemes (also reviewed by Hunt, 2001). Leaf macrofloras, currently understood to be of Late Palaeocene to Middle Eocene in age, have been described in varying completeness from the Admiralty Bay and Fildes Peninsula areas of the island. In the Admiralty Bay area, the Middle Eocene Mt. Wawel Formation (Point Hennequin Group) contains the macroflora deposits collectively known as the Point Hennequin Flora with individual localities named Mount Wawel and Dragon Glacier Moraine floras (Zastawniak et al., 1985; Birkenmajer and Zastawniak, 1989a; Askin, 1992; Hunt, 2001; Hunt and Poole, 2003). The Mount Wawel flora comprises macrofossils of Equisetum (horsetail), ferns and several Nothofagus species as microphyllous leaves (a leaf-size category of 2.5–7.5 cm long), in addition to a few other angiosperms and Podocarpaceae. The Dragon Glacier Moraine flora is similar, the angiosperm leaves being dominated by Nothofagus and the conifers including Araucariaceae and Cupressaceae, in addition to Podocarpaceae. The Middle Eocene Petrified Forest Creek flora from the Arctowski Cove Formation and the Late Eocene Cytadela flora from the Point Thomas Formation are both within the Ezcurra Inlet Group. The former is a wood flora requiring revision, but intermediate Fagus– Nothofagus-type species are recorded. The Cytadela leaf flora includes ferns (including a Blechnum-affinity species), mostly small Nothofagus-type
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
329
leaves with pinnately veined leaves of other dicotyledonous types and possible Podocarpaceae (Birkenmajer and Zastawniak, 1989a; Askin, 1992; Birkenmajer, 1997). Birkenmajer and Zastawniak (1989a) considered this flora to be E/O boundary in age. In this region, therefore, Nothfagus-dominated forests were the norm in the Middle to Late Eocene with ferns and tree ferns becoming increasingly important. Estimated mean annual temperatures of 5–81C are slightly cooler than those on Seymour Island to the east during the Middle Eocene, and the vegetation was similar to the southernmost Patagonian–Magellanic forests of southern Chile (Zastawniak et al., 1985; Birkenmajer and Zastawniak, 1989a; Askin, 1992; Hunt, 2001). By the Late Eocene, vegetation was more comparable to the recent fern-bush communities of southern oceanic islands (e.g. the Auckland Islands), interpreted from the Cytadela and Petrified Forest Creek floras (Birkenmajer and Zastawniak, 1989a; Askin, 1992; Birkenmajer, 1997). However, mean annual temperature estimates of 11.7–151C appear too high, especially considering the small-sized leaves (Francis, 1999). In the Fildes Peninsula area in the southwest of King George Island, the Fildes Peninsula Group contains the contemporary Middle Eocene Collins Glacier and Rocky Cove floras within the Fildes Formation, and the diverse Late Palaeocene–Middle Eocene Fossil Hill flora (Fossil Hill Formation). The latter is a leaf flora containing 40 recognized taxa, including mixed broadleaf angiosperms (with large-leaved Nothofagus species), conifers (podocarp, araucarian and cupressacean) and ferns (Birkenmajer and Zastawniak, 1989a,b; Li, 1992; Haomin, 1994; Francis, 1999; Reguero et al., 2002). Neotropical and sub-Antarctic elements appear to be mixed perhaps indicating a collection derived from communities at different altitudes (Li, 1992), although this mixed signature may be a feature of early Tertiary polar vegetation (Francis et al., 2004). The Nothofagus leaves are much larger than their modern relatives, suggesting a warmer and more humid climate during the early part of the Eocene. Estimates of mean annual temperature suggest W10 1C (from 40% entire-margined leaves) and a small annual temperature range (Li, 1992). Fossil leaves remain undescribed from the Rocky Cove flora; however, wood from this locality has been identified as Nothofagoxylon antarcticus (Shen, 1994; Hunt, 2001). The Collins Glacier deposit is primarily a wood flora that includes wood of both coniferous (Cupressinoxylon sp. and Podocarpoxylon fildesense) and angiospermous (Nothofagoxylon spp., Weinmannioxylon eucryphioides (Cunoniaceae) and Myceugenelloxylon antarcticus (Myrtaceae)) affinity (Hunt, 2001; Poole et al., 2001, 2005). The mean annual temperature had dropped to 91C by the Middle Eocene and
330
J. E. Francis et al.
precipitation appears to have increased; however, the latter is probably due to changes in the environmental setting rather than climate change because of no immediate change in angiosperm wood from a semi-ring to ring porous condition (Poole et al., 2005). 8.3.3. Antarctic Peninsula: Seymour Island The Late Eocene La Meseta Formation of Seymour Island contains floras that document climate change. Leaves and wood of both angiosperm and conifer affinity occur with fern fossils and a flower (Case, 1988; Reguero et al., 2002; Francis et al., 2004). Middle Eocene Nothofagus leaves were found to be notophyllous (a leaf-size category 7.5–12.5 cm long). Angiosperm fossils affiliated to families including Nothofagaceae, Dilleniaceae, Myricaceae, Myrtaceae, Elaeocarpaceae, Moraceae, Cunoniaceae, Winteraceae and Lauraceae have been described (Gandolfo et al., 1998a,b; Reguero et al., 2002; Francis et al., 2004). Doktor et al. (1996) also described leaves affiliated with Podocarpaceae, Araucariaceae, Nothofagaceae and Proteaceae. Gothan (1908) was the first to describe fossil wood from Seymour Island, which has subsequently been re-examined by several authors and identified as having both angiosperm and coniferous affinities (Francis, 1991; Torres et al., 1994; Brea, 1996, 1998; Francis and Poole, 2002; Reguero et al., 2002). Decrease in leaf sizes during the Late Eocene suggests that climate deteriorated towards the end of the Eocene, as observed in studies of the La Meseta Formation by Case (1988) and Reguero et al. (2002). Gandolfo et al. (1998a,b) suggested a MAT of 11–131C for the Cucullaea I Allomember during the early Late Eocene. Further climate data were provided by leaf margin analyses of a Late Palaeocene flora (Cross Valley Formation) and of the early Late Eocene Cucullaea 1 flora, which indicate a decrease in floral diversity and a change from mean annual temperatures of 141C in the Late Palaeocene to 111C in the early Late Eocene, with signs of freezing winters in the Late Eocene (Francis et al., 2004). 8.3.4. Ross Sea Region: McMurdo Sound Only two E/O floras have been described from East Antarctica – a significant, but ex situ, Middle–Late Eocene flora within glacial erratic boulders found at Minna Bluff (mentioned above), and a single Nothofagus leaf within the Early Oligocene strata of the CRP-3 core in McMurdo Sound
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
331
(Cantrill, 2001; Florindo et al., 2005). An additional single Nothofagus leaf was also found in the CIROS-1 core, originally thought to be Oligocene in age (Hill, 1989), but after recent refinement of the age model, it is now considered Early Miocene (Roberts et al., 2003). The McMurdo erratic flora comprises leaves, wood, seeds and Araucaria-type scale-leaves (Pole et al., 2000). Two species of Nothofagus leaves were found, some of which are interpreted as deciduous based on the interpretation of plicate vernation (venation with distinct folds). Fossil wood was identified as Araucarioxylon, Phyllocladoxylon and Nothofagoxylon (Francis, 1999, 2000). The CRP-3 Nothofagus leaf (compared to Nothofagus beardmorensis from the Sirius Group; Francis and Hill, 1996; Cantrill, 2001) was small and also had plicate vernation, but differs to the leaf in the CIROS-1 core thought to be of N. gunnii affinity, an alpine species from Tasmania (Hill, 1989; Francis, 1999). The floras in the erratics indicate the presence of large forest trees comparable to those in South American araucarian (emergent) – Nothofagus forests in the Valdivian Andes of Chile or the Phyllocladus- and Nothofagusdominated cool, sclerophyll forests of temperate New Zealand and Tasmania. Francis (1999, 2000) suggested a cool temperate climate, with a mean annual temperature of o131C from the wood flora, considering some winter snow likely, but temperatures not cold enough to allow extensive ice to form at sea level. Single Nothofagus leaves found in the CRP-3 and CIROS-1 cores, from Early Oligocene and Early Miocene intervals, respectively, indicate, in conjunction with the palynomorphs, a cold temperate and periglacial climate at those times. The CRP-3 Early Oligocene vegetation is compared to low Nothofagus woodland in the Magellanic region of southern Chile (Francis and Hill, 1996; Cantrill, 2001). Cantrill (2001) compared it to N. beardmorensis, known from the Sirius Group and considered to have a minimum requirement of 221C (which is probably conservative) with several weeks at least 51C during the growing season. The Early Miocene Nothofagus leaf is different and may have come from a smallto medium-sized tree in a sub-alpine rainforest or shrub community probably o1 m tall and living in exposed conditions (Hill, 1989; Francis, 1999). 8.3.5. Environmental Signals from Palynomorphs While macrofossils of plants, found usually in distinct levels in geological sections, allow a fair understanding of the biodiversity and ecology of fossil vegetation at specific times, palynological assemblages can deliver a higher resolution picture of vegetation and climate change through time, especially
332
J. E. Francis et al.
due to their presence in drill cores. In addition, pollen and spores can be recovered from areas in which macrofossils are unknown (e.g. Prydz Bay). Palynological studies have been undertaken on various Palaeogene sections in West and East Antarctica, but complete recovery is rare, especially over the crucial time interval spanning the E/O boundary when climate dramatically deteriorated. Cranwell (1959) carried out the earliest palynological studies in the Antarctic realm on a single sample of probable Palaeogene age from Seymour Island in the Antarctic Peninsula area. Subsequently, several early Tertiary stratigraphic sections on Seymour, Cockburn and King George Islands, and cores from the South Orkney Islands (ODP Leg 113, Site 696) and South Scotia Ridge (Bruce Bank, Eltanin Core IO 1578-59) have been subjected to detailed palynological analyses (Mohr, 1990; Askin et al., 1991, 1997; Grube, 2004; Grube and Mohr, 2004). Data used to reconstruct Tertiary vegetation and climate in East Antarctica have been derived from drilling campaigns in southern McMurdo Sound in the Ross Sea (CIROS-1, CRP-2/2A and CRP 3: Mildenhall, 1989; Askin and Raine, 2000; Raine and Askin, 2001; Prebble et al., 2006) and in Prydz Bay (MacPhail and Truswell, 2004). ODP Leg 189, in the Tasman Sea, cored the E/O boundary (Grube and Mohr, 2008). The glacial erratics of Eocene-aged sediments in southern McMurdo Sound have also yielded terrestrial palynomorphs. Pollen assemblages in Antarctic Palaeogene strata contain many taxa comparable to those still found today in southern high latitudes, including areas of southern South America, Tasmania, Australia, New Zealand and New Caledonia. They vary quite substantially in the abundance of their major components, and consist of moss and fern spores, gymnosperm and angiosperm pollen. Ferns were species-rich until the Middle to early Late Eocene (Mohr, 2001) and include genera that live today under humid subtropical conditions, such as Cnemidaria (Mohr and Lazarus, 1994). At Bruce Bank (c. 46–44 Ma), fern spores dominated the assemblage, at intervals comprising more than 50% of the sporomorphs. During the Late Eocene, and even more so during the Oligocene, fern diversity and abundance dropped dramatically. During the Early Miocene, some of the taxa seem to return, but recycling of older Eocene spores cannot be completely ruled out, a problem prevalent in glaciogene sediments all around the Antarctic. Gymnosperms were important components of the southern high latitude Palaeogene forests. Cycads seem to have been present until the Middle Eocene (Cycadopites), while Araucariaceae (Araucaria), Cupressaceae and especially Podocarpaceae pollen (Podocarpus, Phyllocladus, Lagarostrobus,
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
333
Dacrydium and Microcachrys) play a major role in the palyno-associations well into the Oligocene. Angiosperms were relatively species-rich in the Palaeogene Antarctic pollen spectra with a dominance of various types of Nothofagidites (pollen comparable to that of extant Nothofagus, the southern beech), but diversity declined during the Late Eocene. Middle to Late Eocene assemblages from the Antarctic Peninsula area (ODP Leg 113, Site 696) show, in earliest sections, relatively large amounts of angiosperm pollen with a clear dominance of Nothofagidites, while in assemblages of latest Eocene and Early Oligocene age, moss spores become more common, and, except for Nothofagidites, almost no angiosperms are registered. The McMurdo Sound cores (CIROS-1, CRP-2/2A and CRP-3, Late Eocene to Oligocene) and glacial erratics (Middle to Late Eocene) are characterized by prominent Nothofagidites, particularly Nothofagidites lachlaniae and the N. fusca group. Various Podocarpus taxa are also abundant (Askin, 2000; Raine and Askin, 2001). In a relatively short sequence of CRP-3 (glaciomarine cycle 26) of Early Oligocene age, N. fusca-type, N. flemingii and N. lachlaniae contribute each about 23% of the total count (Prebble et al., 2006). In glaciomarine cycle 11 of Late Oligocene age, N. fusca-type pollen dominates with about 50%, followed by N. flemingii and N. lachlaniae. In Prydz Bay sections dated Late Eocene, Nothofagidites is clearly dominant at 41–57% of sporomorphs; the second largest group are conifer pollen that reach in a few samples up to 50% and more. Fern spores comprise 6–20% and cryptogams 3–6% (Macphail and Truswell, 2004) (Figs. 8.8 and 8.9). During the warmer periods of the Palaeogene, the following families have been identified from pollen: Aquifoliaceae (includes holly), Casuarinaceae (she-oak), Cunoniaceae/Elaeocarpaceae, Epacridaceae (southern heath, now included within Ericaceae), Euphorbiaceae (spurge), Gunneraceae, Liliaceae, Myrtaceae, Nothofagaceae (including all four morphotype groups brassiitype, fusca-type, menziesii-type and ancestral-type; Dettmann et al., 1990) Olacaceae, Proteaceae (Gevuina/Hicksbeachia, Adenanthos, Carnarvonia, Telopea and Beauprea; Dettmann and Jarzen, 1991), Restionaceae (rush), Sapindaceae (soapberry) and Trimeniaceae (Prebble et al., 2006). In Prydz Bay sections (Late Eocene) and within the La Meseta Formation on Seymour Island (Eocene), Fischeripollis and Droseridites, that belong to Droseraceae (sundew), which are today restricted to moors or damp sites, are excellent ecological markers. During the Late Eocene to Oligocene, Apiaceae, Asteraceae (daisy), possibly Campanulaceae (bellflower), Caryophyllaceae (carnation), Chenopodiaceae (now in the Amaranthaceae), Onagraceae (willowherb; Corsinipollenites) and Gramineae (grasses) seem to play a role as members of
334
J. E. Francis et al.
Figure 8.8: Relative abundance (%) of major plant groups in ODP Leg 189, Site 1168. Asterisk marks the E/O boundary. the local vegetation (Mildenhall, 1989; Mohr, 1990; Askin, 1997; Askin and Raine, 2000; Raine and Askin, 2001; Prebble et al., 2006). During the Early Oligocene, a low shrub or closed Nothofagus-podocarp forest of small stature may have developed, occupying warmer sites on the Antarctic continent (Prebble et al., 2006). In colder phases, a tundra-like vegetation, evidenced by moss spores, few but relatively diverse herb pollen and a few Nothofagidites pollen, derived possibly from dwarfed southern beech, may have grown near the coast. Palynological studies by Grube and Mohr (2008) of cores from ODP Leg 189, Site 1168 in the Tasman Sea show a clear response to E/O climate change. The abundance-time-chart (Figs. 8.8 and 8.9) for the Tasman Sea samples shows that during the latest Eocene, the pollen flora was dominated by the Nothofagaceae (especially the evergreen type Brassospora), with araucarian and podocarp conifers (gymnosperms) and typical fern families (cryptogams). Near the E/O boundary itself, there is a short peak in the occurrence of araucarian and some other gymnosperm pollen, as well as an increase in ferns, in response to a decline in Nothofagaceae. Surprisingly, however, there was no sustained change in terrestrial pollen after this that might reflect a major change in climatic regime. Vegetation typical of latest Eocene composition seems to have been restored during the earliest Oligocene, significant changes
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
335
Figure 8.9: Relative abundance (%) of sporomorphs in the Tasman Sea samples from ODP Leg 189, Site 1168. Asterisk marks the E/O boundary; Lines represent palynological ‘‘events’’. being the decrease in Casuarinaceae angiosperms, and the gradual replacement of Osmundaceae ferns by Schizaeaceae and Gleicheniaceae. Nothofagus fusca- and menziesii-type pollen increase from about 33.8 Ma on, while other angiosperm pollen show a slight decline, which Grube and Mohr (2008) interpret as a gradual response to long-term cooling. The pollen record in Figs. 8.8 and 8.9 also highlights a further short-lived episode of vegetational change at about 32.9 Ma. The pattern is similar to that at 33.7 Ma, with an increase in araucarian conifer pollen and in fern spores at the expense of the angiosperms, especially the Nothofagaceae – does this represent a later episode of cold climate? The pollen diagram in Fig. 8.9 also hints at cyclical changes, possibly at intervals of 0.8 m.y or even 0.4 m.y.
8.4. Environmental Changes Documented by Marine Microfossils Early Palaeogene marine microfossil associations from the circum-Antarctic realm are typically characterized by the dominance of largely endemic
336
J. E. Francis et al.
Antarctic organic walled dinocysts and siliceous groups like diatoms and radiolarians over calcareous microfossils (e.g. Brinkhuis et al., 2003; Stickley et al., 2004; Warnaar, 2006). The characteristic endemic dinocyst assemblages reported from Middle and Late Eocene deposits around Antarctica are often referred to as the ‘‘Transantarctic Flora’’ (cf. Wrenn and Beckmann, 1982). In fact, Late Cretaceous to (early) Palaeogene organic walled dinocysts from the circum-Antarctic realm are comparatively well known, notably from southern South America, from James Ross and Seymour islands, but also from southeastern Australia and New Zealand, from erratics along the Antarctic margin, from the Ross Sea continental shelf (CIROS; CRP) and from several ocean drill sites (see, e.g. Haskell and Wilson, 1975; Wilson, 1985, 1988; Askin, 1988a,b; Wrenn and Hart, 1988; papers in Duane et al., 1992; Pirrie et al., 1992; Mao and Mohr, 1995; Hannah, 1997; Truswell, 1997; Hannah et al., 2000; Levy and Harwood, 2000a,b; Guerstein et al., 2002). Meaningful chronostratigraphic calibration of (sub-)Antarctic dinocyst events was a classic problem due to the general absence of other ageindicative biotas and/or magnetostratigraphy or other means of dating in sections in which dinocysts are encountered. The first integrated Oligocene to earliest Miocene biomagnetostratigraphy, including dinocysts, was achieved only relatively recently on the basis of successions drilled during the CRP (e.g. Hannah et al., 1998, 2000). Even more recently, the first magnetostratigraphically calibrated Late Maastrichtian to earliest Oligocene dinocyst succession was established on the basis of records drilled during ODP Leg 189, offshore Tasmania (e.g. Brinkhuis et al., 2003; Sluijs et al., 2003; Huber et al., 2004; Stickley et al., 2004). Building on these studies, Warnaar (2006) produced higher resolution records for the ODP 189 holes, and (re)analysed critical intervals from other circum-Antarctic sites like 696, 739, 1090 and 1166. In the following section, the dinocyst record for the Eocene and Oligocene is documented and the implications for our understanding of palaeoceanography at this time are discussed. 8.4.1. Palaeocene–Middle Eocene Dinocysts Circum-Antarctic early Palaeogene dinocyst associations were recovered from the Tasmanian Gateway (between Australia and East Antarctica, e.g. Brinkhuis et al., 2003), New Zealand (e.g. Wilson, 1978, 1984, 1988; Willumsen, 2000; Crouch, 2001) and Seymour Island (e.g. Askin, 1988a,b; Elliot et al., 1994). The associations are characterized by nearly identical
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
337
composition and stratigraphic succession. While circum-Antarctic endemic taxa were present at least since the Maastrichtian (e.g. Riding and Crame, 2002; Brinkhuis et al., 2003), the taxa often referred as the ‘‘Transantarctic Flora’’ established itself in the early Palaeogene in an otherwise largely cosmopolitan assemblage. Since the late Early Eocene, the influence of the ‘‘Transantarctic Flora’’ (constituted by species such as Deflandrea antarctica, Octodinium askiniae, Enneadocysta partridgei, Vozzhennikovia spp., Spinidinium macmurdoense and Arachnodinium antarcticum) increases until the middle Late Eocene. While the younger part of the circum-Antarctic Middle Eocene is comparatively less well studied, with records only from New Zealand (e.g. Wilson, 1988; Strong et al., 1995), DSDP Sites 280 and 281 (Crouch and Hollis, 1996), Seymour Island (Wrenn and Hart, 1988), the Scotia Sea (Mao and Mohr, 1995), ODP Site 1172 (Brinkhuis et al., 2003) and southern Argentinean successions (e.g. Guerstein et al., 2002), a broad similarity is apparent. The younger part of the Southern Ocean Middle Eocene appears to be characterized by several important last occurrences (LOs), including those of Membranophoridium perforatum, Hystrichosphaeridium truswelliae, Hystrichokolpoma spinosum and Hystrichokolpoma truncatum (cf. Wilson, 1988; see Brinkhuis et al., 2003). 8.4.2. Late Eocene–Early Oligocene Dinocysts At the Tasmanian Gateway, the early Late Eocene dinocyst distribution forms a continuation of the Middle Eocene pattern. ‘‘Transantarctic Flora’’ species predominate, and final acmes of Enneadocysta partridgei, the Deflandrea antarctica group and Spinidinium macmurdoense are recorded (Fig. 8.10). Important first occurrences (FOs) in this phase include those of Schematophora speciosa, Aireiana verrucosa, Hemiplacophora semilunifera and Stoveracysta ornata. Towards the middle Late Eocene, FOs of Achomosphaera alcicornu, Reticulatosphaera actinocoronata and Alterbidinium distinctum and the LO of S. speciosa appear important for interregional correlation, as is the FO of Stoveracysta kakanuiensis. Vozzhennikovia spp. continues to be a common constituent of the associations (Sluijs et al., 2003). Typically, sediments representing the E/O transition are barren of organic microfossils in all ODP Leg 189 records; dinocysts briefly reappear in the Early Oligocene (assigned to Chron C11-1r; Stickley et al., 2004). In this single productive sample thus far from the Early Oligocene, virtually all Transantarctic Palaeogene dinocysts have disappeared (only a single, poorly preserved, probably reworked specimen of E. partridgei was recovered;
Figure 8.10: Circum-Antarctic geographical distribution maps (Late Paleocene–Miocene) showing dinocyst endemism. Maps derived from the Ocean Drilling Stratigraphic Network (ODSN). Black areas indicate (continental) blocks that are mostly sub-aerial. Note that several blocks shown in black were partly submerged (e.g., the Ross Sea, the southern Australian margin and parts of Argentina). Shaded areas indicate mostly submerged (continental) blocks (e.g., Brown et al., 2006).
338 J. E. Francis et al.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
339
Brinkhuis et al., 2003). The association in this sample is characterized by the abundance of taxa more typical for Tethyan waters, including an occurrence of Hystrichokolpoma sp. cf. Homotryblium oceanicum (e.g. Brinkhuis and Biffi, 1993; Wilpshaar et al., 1996; Brinkhuis et al., 2003). Some of the Late Eocene dinocyst events have previously been reported from the South Australian margin, e.g. from the Browns Creek section (Cookson and Eisenack, 1965; Stover, 1975). For example, the ranges of S. speciosa, A. verrucosa, H. semilunifera and S. ornata appear useful for regional and even global correlation. Many of the ‘‘Browns Creek’’ Late Eocene dinocysts have been recorded from locations around the world, also in otherwise well-calibrated sections in central and northern Italy, including the Priabonian Type Section (Brinkhuis and Biffi, 1993; Brinkhuis, 1994). It appears that these index species have slightly earlier LOs in this region than they have in Italy (Tethyan Ocean), if the records of Cookson and Eisenack (1965) and Stover (1975) are combined with more recent nannoplankton and magnetostratigraphic studies from the same section (Shafik and Idnurm, 1997). This aspect may be related to the progressive global cooling during the latest Eocene (Fig. 8.11). In the rare sediments covering the E/O transition, more specifically the Oi-1 event, from the Weddell Sea and near the Drake Passage (between South America and the Antarctic Peninsula), dinocysts are typically not preserved (Gradstein et al., 2004). However, in the oldest Oligocene sediments bearing dinocysts, the Transantarctic Palaeogene dinocysts (dominant in the latest recovered Eocene sediments) are replaced by cosmopolitan taxa (Gradstein et al., 2004). This suggests that the changes in dinocyst associations in this area were at least broadly similar compared to those in the Tasman Sector. In contrast, it seems that sediments covering the E/O transition are preserved at Prydz Bay (ODP Site 739). Here a gradual change is observed from the typical Transantarctic Palaeogene dinocysts to the taxa typically found in (post-) Oligocene near-Antarctic records (Warnaar, 2006). However, the (cosmopolitan) taxa that are useful for correlation, as mentioned above, are apparently not present in the Prydz Bay records. 8.4.3. Palaeoceanography During the early Palaeogene, Antarctica was less glacierized than it is now, or not glacierized at all (e.g. Zachos et al., 2001; Pagani et al., 2005; see above). South America and Australia were still not fully separated from the Antarctic continent (e.g. Livermore et al., 2005; Brown et al., 2006), which
340
J. E. Francis et al.
Figure 8.11: Generalized (Middle Eocene) dinocyst geographical distribution map overlain with the ocean circulation pattern inferred from GCM results. Maps derived from the Ocean Drilling Stratigraphic Network (ODSN). Shaded areas indicate mostly submerged (continental) blocks (e.g., Brown et al., 2006). Abbreviations: TA-SW, Trans-Antarctic Seaway (hypothetical; see Wrenn and Beckmann, 1982); TSA-SW, Trans-South American Seaway (hypothetical; see Kohn et al., 2004); EAC, East Australian Current; p-LC, proto-Leeuwin Current; p-RG, proto-Ross Gyre; TC, Tasman Current.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
341
prevented the development of a (proto-) Antarctic Circumpolar Current (ACC). It was hypothesized that during the early Palaeogene, warm ocean currents from lower latitudes could reach and warm Antarctica. The opening and subsequent deepening of critical conduits (i.e. Drake Passage and the Tasmanian Gateway) towards the end of the Eocene have long been thought to have played a central role in ACC establishment and Antarctic cooling (e.g. Kennett et al., 1975; Kennett, 1977, 1978; Murphy and Kennett, 1975). However, recent advances through ocean drilling (e.g. ODP Leg 189) and coupled Global Circulation Model (GCM) experiments suggest that Eocene Southern Ocean surface circulation patterns were fundamentally different than previously thought, and that the opening and deepening of oceanic gateways were of little climatic consequence (Sloan and Huber, 2001; Huber et al., 2004; Warnaar, 2006). Instead, it is nowadays argued that the ‘‘greenhouse–icehouse’’ transition was caused by changes in greenhouse gas concentrations, rather than oceanographic changes (e.g. DeConto and Pollard, 2003a,b; Huber et al., 2004; Pagani et al., 2005). The high degree of endemism in the circum-Antarctic marine microfossil associations denies the existence of a southward-bound, warm, proto-East Australian Current as proposed by Kennett et al. (1975) and Exon et al. (2004), according to Huber et al. (2004) (Fig. 8.11). For example, a colder northward flowing western boundary current, designated the ‘‘Tasman Current’’ (see Huber et al., 2004) existed off southeast Australia. GCM experiments indicate that the Eocene Southern Ocean, including the southern Pacific, was dominated by clockwise gyres (Sloan and Huber, 2001; Huber et al., 2004). Moreover, several studies show that the Tasmanian Gateway had already been open to neritic water depths (i.e. o200 m) since at least the Middle Eocene (Stickley et al., 2004). Deepening to bathyal water depths (i.e. 200–4,000 m) occurred during the early Late Eocene (B35.5 Ma). The Drake Passage had possibly been open to (upper) bathyal water depths by the Middle Eocene (Eagles et al., 2006; Scher and Martin, 2006). Both tectonic events thus seem to have occurred too early to be related to the Antarctic glaciation in the Early Oligocene (i.e. the Oi-1 stable-isotope event (33.3 Ma), e.g. Miller et al., 1998; Zachos et al., 2001). Given the similar continent–ocean configuration, a corollary of the GCM experiments is that the Palaeogene circum-Antarctic surface circulation should not have been fundamentally different from the Cretaceous situation. If this were the case, and as long as substantial equator–pole temperature gradients existed, then it may be expected that, throughout the Late Cretaceous to early Palaeogene time interval, circum-Antarctic waters were consistently dominated by endemic biota, particularly in environments influenced by the proposed western boundary currents. Both hypotheses
342
J. E. Francis et al.
were recently tested by Warnaar (2006) by mapping distribution patterns of circum-Antarctic dinocysts through Palaeogene times and comparing them with coupled GCM results. Warnaar (2006) conceived a model termed the ‘‘refrigerator trap’’ wherein it is hypothesized that cosmopolitan and endemic dinoflagellates were taken through the cold and darkness along the Antarctic continent, transported by the proto-Ross Gyre. Conceivably, taxa normally living in warmer waters (e.g. the East Australian Current) that were trapped in the gyre were unable to survive such conditions. It is conceivable that the endemic taxa (notably taxa of the ‘‘Transantarctic Flora’’ and bi-polar Phthanoperidinium echinatum group) were specifically adapted to tolerate cold conditions (41C to the freezing point), prolonged darkness and possibly seasonal sea ice.
8.5. Evolution of Ocean Temperatures and Global Ice Volume During the Eocene to Oligocene from the Ocean Isotope Record The evolution of climate during the Eocene and Oligocene can be determined from the deep-sea isotope and trace element records of ocean temperatures and ice volume. Earlier isotope work suggests that the primary transition from greenhouse to icehouse world took place during the Late Eocene and Early Oligocene, with large, permanent ice sheets appearing on Antarctica at 34 Ma (Zachos et al., 1992, 1996, 2001; Miller et al., 1998; Coxall et al., 2005). This transition was preceded by a period of long-term cooling which initiated near the Early–Middle Eocene boundary, roughly 50 Ma, following a sustained period of Early Eocene warmth. The Eocene cooling trend was not monotonic, but followed a somewhat step-like pattern with several reversals, the most substantial of which was the Middle Eocene climatic optimum (MECO) (Bohaty and Zachos, 2003; Jovane et al., 2007). By the Late Eocene, the climate on Antarctica appears to have cooled sufficiently to allow for the formation of small, ephemeral ice sheets, a state that persisted until B34 Ma, when most of East Antarctica became glaciated by a large ice sheet (Fig. 8.12). From that time forward, the ice sheet was a permanent feature of Antarctica. For the remainder of the Oligocene, this ice sheet waxed and waned, most likely in response to orbital forcing (Naish et al., 2001). The long-term cooling trend that facilitated the formation of continental ice sheets has been attributed to either changes in palaeogeography or the concentration of greenhouse gases. Geographical isolation of
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
δ18O 5
4
3
2
Plt.
Plio. Miocene
10
20 Oligocene Eocene
40
0
-1
Oi-1 Glaciation
Antarctic Ice-Sheets
Mi-1Glaciation
30
1 N. Hemisphere Ice-sheets
Age (Ma) 0
343
Mid-Eocene Climatic Optimum
50
60
Paleocene
Paleocene-Eocene Thermal Maximum
Partial or Ephemeral Full Scale and Permanent
70 0° 4° 8° 12° Ice-free Temperature (°C)
Figure 8.12: Global compilation of oxygen isotope records for the Cenozoic (after Zachos et al., 2001). Solid bars span intervals of ice-sheet activity in the Antarctic and Northern Hemisphere. Reproduced with permission of the Geological Society Publishing House, Bath, UK (Zachos et al., 2001; Billups and Schrag, 2003; Bohaty and Zachos, 2003; Lear et al., 2004). the Antarctic continent with the tectonic widening of ocean gateways is often cited as one means of driving long-term cooling (see above). A second possible mechanism is a decline in atmospheric CO2 . Recent modelling studies (see below) suggest that the former would have had little impact on heat fluxes and mean annual temperatures on Antarctica, while the latter would be a more effective way of cooling the continent (Huber and Sloan, 2001). However, the record of Eocene CO2 has until
344
J. E. Francis et al.
recently (Pagani et al., 2005) lacked sufficient resolution to fully test this possibility. Recent investigations of marine cores have largely focused on improving two aspects of the Eocene and Oligocene climate reconstructions: (1) the resolution of proxy records of ocean temperature and ice volume, which has promoted the development of high-resolution, orbitally tuned records, and (2) quantifying changes in ice volume, which has spurred the development and application of palaeotemperature proxies. These studies have benefited in part from efforts of the ODP to recover highly expanded, stratigraphically intact sediment sequences spanning the E/O. The latest high-resolution records show that ice volume and ocean temperatures varied in a periodic fashion with power concentrated in the long eccentricity and obliquity bands (Fig. 8.13; Coxall et al., 2005). The latter is consistent with the presence of a large polar ice sheet. Prior to the Late Eocene, however, power in the obliquity band is relatively weak (Palike et al., 2001), suggesting little to no ice volume on Antarctica. These records have also revealed distinct climatic variability coherent with the lower frequency components of obliquity with periods of 1.25 m.y. The recent development of seawater temperature proxies, Mg/Ca and TEX86 (Schouten et al., 2003), has improved estimates of ice volume from oxygen isotope records. In particular, the first low-resolution benthic Mg/Ca records suggest that much of the d18O increase just after the E–O boundary (33.4 Ma) was the result of a substantial increase in ice volume as deep-sea temperature was fairly constant (Lear et al., 2000; Billups and Schrag, 2002). Most recently, Lear et al. (2008) have interpreted a marine cooling of B2.51C associated with this ice growth from exceptionally well preserved foraminifera well above the calcite compensation depth. Though still controversial, the magnitude of ice-volume increase would have exceeded that of the present-day Antarctic Ice Sheet.
8.6. Connection of CO2 and Ice-Sheet Inception at the E/O Boundary – Computer Modelling While the onset of major, continental-scale glaciation in the earliest Oligocene has long been attributed to the opening of Southern Ocean gateways (Kennett and Shackleton, 1976; Kennett, 1977; Robert et al., 2001), recent modelling studies suggest declining atmospheric CO2 was the most important factor in Antarctic cooling and glaciation.
C12r
C13n E. formosa
C13rC13n
C13n
C13n.1
?
35
34
33
32
31
0.5
1
5
4
3
MAR CaCO3 Bulk sediment
Oi-1
10 20 30 40 50 Age (Ma)
δ18O δ13C
0
Inset Mio. Oli. Eoc.
0
2
1
0 0.5
1.5
δ13 C‰ VPDB 1 1.5
2
MAR g/cm2/kyr
245
240
230
220
210
22 23 24 1 3 5
Climatic Eccentricity precession Obliquity (%) (deg)
% CaCO3 0 20 40 60 80 by mass-spectrometry by physical property measurements Bulk sediment % CaCO 3
1.5
0.5 0.04 0 -0.04
Figure 8.13: High-resolution isotope and % CaCO3 records of the E/O boundary (Coxall et al., 2005). Age model is derived from orbital tuning.
C13r
EARLY OLIGOCENE
Mag. chrons Epoch Age (Ma)
C12n Paleodepth (km)
LATE EOCENE
δ 18 O‰ VPDB
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica 345
346
J. E. Francis et al.
As the Drake Passage and Tasmanian Gateway widened and deepened during the late Palaeogene and early Neogene (Lawver and Gahagan, 1998), the ACC and Polar Frontal Zone (APFZ) presumably cooled the Southern Ocean by limiting the advection of warm subtropical surface waters into high latitudes (Kennett, 1977). While the opening of the Tasmanian Gateway does broadly coincide with the earliest Oligocene glaciation event (Oi-1) (Stickley et al., 2004), the tectonic history of the Scotia Sea remains equivocal. Estimates for the timing of the opening of Drake Passage range between 40 and 20 Ma (Barker and Burrell, 1977; Livermore et al., 2004; Scher and Martin, 2006), clouding the direct ‘‘cause and effect’’ relationship between the gateways and glaciation. A number of ocean modelling studies have shown that the opening of both the Drake Passage and Tasmanian gateways reduces poleward heat convergence in the Southern Ocean and cools sea surface temperatures by up to several degrees (Mikolajewicz et al., 1993; Nong et al., 2000; Toggweiler and Bjornsson, 2000). More recent, coupled atmosphere–ocean GCM simulations suggest a more modest effect, however. Huber et al. (2004) showed that the Tasmanian Gateway likely had a minimal effect on oceanic heat convergence and sea surface temperatures around the continent, because the warm East Australia Current does not travel any further south if the gateway is open or closed. The gateway’s effect on East Antarctic climate and snowfall was also shown to be minimal, pointing to some other forcing (perhaps decreasing atmospheric CO2 concentrations) as the primary cause of Antarctic cooling and glaciation. The recent development of coupled climate–ice sheet models capable of running long (W106 years), time-continuous simulations of specific climate events and transitions (DeConto and Pollard, 2003a) has allowed simulations of the Oi-1 event that account for decreasing CO2 concentrations, orbital variability and prescribed changes in ocean transport (DeConto and Pollard, 2003b; Pollard and DeConto, 2005). These simulations support the conclusions of Huber et al. (2004) as to the likely importance of CO2 by showing that, even if significant, tectonically forced changes in ocean circulation and heat transport had occurred around the E/O boundary, they would have had only a small effect on temperature and glacial mass balance in the Antarctic interior. Therefore, Southern Ocean gateways could only have triggered glaciation if the climate system was already close to a glaciation threshold. Considering the sensitivity of polar climate to the range of CO2 concentrations likely to have existed over the Palaeogene–Neogene (Pagani et al., 2005), CO2 likely played a fundamental role in controlling Antarctica’s climatic and glacial sensitivity to a wide range of forcing mechanisms. This conclusion is supported by a number of additional
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
347
modelling studies exploring the role of orbital variability (DeConto and Pollard, 2003b), mountain uplift in the continental interior (DeConto and Pollard, 2003a), geothermal heat flux (Pollard et al., 2005), Antarctic vegetation dynamics (Thorn and DeConto, 2006) and Southern Ocean sea ice (DeConto et al., in press) in the E/O climatic transition. The results of these studies can be summarized as follows. The timing of glaciation on East Antarctica was shown to be sensitive to orbital forcing, mountain uplift and continental vegetation, but only within a very narrow range of atmospheric CO2 concentrations around 2.8 times modern levels, close to the model’s glaciation threshold. Once the glaciation threshold is approached, astronomical forcing can trigger sudden glaciation through non-linear height/mass balance and albedo feedbacks that result in the growth of a continental-scale ice sheet within 100 kyrs (Fig. 8.14). The timing of glaciation appears to be insensitive to both expanding concentrations of seasonal sea ice and changes in geothermal heat flux under the continent; however, a doubling of the background geothermal heat flux (from 40 to 80 mW m 2) does have a significant effect on the area under the ice sheet at the pressure-melt point (where liquid water is present), which may have had some influence on the distribution and development of sub-glacial lakes and subsequent ice-sheet behaviour. While these modelling studies have certainly improved our understanding of the importance of atmospheric CO2 concentrations relative to other Cenozoic forcing factors, several important model-data inconsistencies remain unresolved. For example, long, time-continuous GCM-ice-sheet simulations of an increasing CO2 (warming) scenario show strong hysteresis once a continental ice sheet has formed (Pollard and DeConto, 2005). In these simulations, orbital forcing alone is not sufficient to produce the range of Palaeogene–Neogene ice-sheet variability (B50–120% modern Antarctic ice volumes) inferred by marine oxygen isotope records and sequence stratigraphic reconstructions of eustasy (Zachos et al., 2001; Pekar and DeConto, 2006; Pekar et al., 2006), pointing to the importance of additional feedbacks (possibly related to the marine carbon cycle and atmospheric CO2) in controlling Cenozoic ice-sheet variability. Furthermore, several recent isotopic analyses of deep-sea cores imply ice volumes during the peak Oligocene and Miocene glacial intervals that are too big to be accommodated by East Antarctica alone (Lear et al., 2004; Coxall et al., 2005; Holbourn et al., 2005). This suggests that either our interpretations of the proxy data are faulty, or episodic, bipolar glaciation occurred much earlier than currently accepted. These, among other unresolved controversies related to the climatic and glacial evolution of the high southern latitudes, will be the focus of future ACE modelling exercises and model-data comparisons.
Time (10 model years)
2
4
6
8
10
0
0
0.0
10
20
30
15 6
40
Δ sea level
20
50
3
25
60
0.6
Ice volume (10 km )
5
10
0.4
18 Δδ Ο
0.2
3.6
3.4
3.2
3.0
2.8
2.6
2.4
2.2
2.0
CO2 (PAL)
0
500
1000
1500
2000
2500
3000
3500
4000
meters
0
500
1000
1500
2000
2500
3000
3500
4000
meters
Figure 8.14: Ice volume (left) and corresponding ice-sheet geometries (right) simulated by a coupled GCM-ice sheet model in response to a slow decline in atmospheric CO2 and idealized orbital cyclicity across the E/O boundary. The sudden, two-step jump in ice volume (left panel) corresponds to the Oi-1 event. The left panel shows simulated ice volume (red line), extrapolated to an equivalent change in sea level and the mean isotopic composition of the ocean (top). Arbitrary model years (left axis) and corresponding, prescribed atmospheric CO2 (right axis) are also labelled. CO2 is shown as the multiplicative of pre-industrial (280 ppmv) levels. Ice-sheet geometries (right panels) show ice-sheet thickness in metres. Black arrows correlate the simulated geometric evolution of the ice sheet through the Oi-1 event (modified from DeConto and Pollard, 2003b). Reproduced with permission of The Geological Society Publishing House, Bath, UK.
6
348 J. E. Francis et al.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
349
8.7. Summary Although no one single source of evidence or locality yields the complete story of climate change and the onset of glaciation in Antarctica, piecing together information from a range of sources, as presented above, does provide a picture of how climate cooled and glaciation became established on the continent. A summary is presented below (for references, see text above). Paradoxically, the present ice sheet hides its own history, certainly at the highest latitudes in the middle of the continent, so most of what is known about past environments is from the lower latitude marginal sites.
8.7.1. Early–Middle Eocene Polar Warmth Evidence from fossil plants, sediments and isotopes indicates that the Late Palaeocene and Early Eocene experienced warm climates at high latitudes, at least on the margins of Antarctica where strata of this age crop out. Climates appear to have been warm and wet, seas were warm and plants flourished in a frost-free environment. The oldest record of glacial activity (if the dating is correct in this problematic region) is of valley-type tillites of Middle Eocene age on King George Island, indicating the presence of alpine glaciers. However, floras of Middle Eocene age from King George and Seymour islands suggest warm to cool temperate climates, generally moist and probably frost-free. The ocean isotope record also suggests that climates were generally warm until the Middle Eocene, although the climate trend was towards cooling.
8.7.2. Late Eocene Cooling A variety of sources, particularly fossil plants, suggest that during the early Late Eocene, climates cooled but perhaps not to the extent of significant ice build-up. The Late Eocene sediment record in the Ross Sea region (McMurdo Erratics, magnetic and clay mineral record) and in the Prydz Bay area could be indicative of cold climates but the coastal/open marine shelf and fluvial-deltaic environments in these two areas, respectively, do not show signs of the presence of significant ice. By the latest Eocene, however, glacial deposits are apparent. Glacial deposits on Seymour Island, close to the E/O boundary, may indicate the presence of valley glaciers in that region, situated at about 651 South Palaeolatitude.
350
J. E. Francis et al.
The ocean record, especially marine microfossils, provides information about the climate and currents in the oceans at critical times during this interval. South America and Australia were still not separated from Antarctica during the Early Eocene, so the Antarctic Circumpolar Current was unable to develop. Instead, warm equatorial currents may have fed warmth to the continent. Palaeoceanographic changes related to the deepening of the Tasmanian Gateway and the opening of the Drake Passage are still debated but do not seem to have been strictly related to Early Oligocene climate cooling and intensification of glaciation, so these oceanographic changes are not considered to be major drivers of polar cooling. Recent atmosphere–ocean modelling has also shown that changes in oceanography related to tectonic events were not likely to have driven the climate cooling that led to glaciation. 8.7.3. Latest Eocene/Earliest Oligocene Glaciation By the E/O boundary times, there is no doubt that ice was present on Antarctica. In the Ross Sea region, drill cores show evidence of relatively uniform marine sedimentation through the latest part of the Eocene and into the Oligocene but sediments include exotic clasts indicative of iceberg rafting. There does not appear to be a major environmental shift at this time but more of an intensification of cooling. In the Prydz Bay region, tidewater glaciers were present in the Early Oligocene, with ice reaching the continental shelf edge. In the oceans, the oxygen isotope record and other geochemical indicators signal a strong cooling at the boundary, the Oi-1 event, which has been interpreted as a time of major build-up of ice. Even though climates were cold, vegetation was able to persist but by this time the higher diversity and warmth-loving plants of the Early and Middle Eocene forests had disappeared, to be replaced by vegetation that was dominated by several species of the southern beech, Nothofagus. Along with mosses, a few ferns and some podocarp conifers, southern beech trees probably grew as shrubby vegetation in the most hospitable areas. 8.7.4. Oligocene Ice Sheets A hiatus at the Early/Late Oligocene boundary marks a change to fluvial conditions, with grounded ice or possibly glaciomarine conditions. A distinct drop in sea level is noted at this time. Facies changes and diamictite beds in the Cape Roberts core are indicative of the periodic expansion of tidewater
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
351
glaciers, typical of a temperate glacier regime with glaciers flowing from the young East Antarctic Ice Sheet across the continental margin. Throughout the Oligocene, glaciation seems to have waxed and waned. A distinct record of glacial activity is recorded by glacial sediments in the Polonez Cove Formation on King George Island of mid-Oligocene age (Krakowiak Glacial Member). At that time, ice was grounded on a shallow marine shelf. More extensive ice sheets may have been present further south in the Weddell Sea region, from which clasts of rock from the Transantarctic Mountains may have been derived to be incorporated as exotic clasts in the Polonez Cove Formation. Sediments without a glacial signature that overlie these glacial deposits suggest a phase of climate warming and glacial retreat until the next glacial pulse in the Miocene. Why did the climate cool during the Eocene and Oligocene, causing such a major change in Antarctic environments? The influence of palaeoceanographic changes is now considered less critical; instead, coupled climate–ice sheet modelling indicates that it was changing levels of atmospheric CO2 that controlled Antarctica’s climate. Factors such as mountain uplift, vegetation changes and orbital forcing all played a part in cooling the polar climate, but only when CO2 levels fell to critical threshold levels (2.8 times present-day levels) did orbital forcing tip Antarctica into its icy glacial world.
ACKNOWLEDGEMENTS Brinkhuis and Warnaar would like to thank Catherine Stickley, Matt Huber and Appy Sluijs. Francis would like to thank ACE and SCAR for funds to attend ACE meetings. B. Mohr wishes to thank for funds provided by the German Funding agency (DFG), which enabled R. Grube to study Antarctic material from various sources, mainly from Ocean Drilling Program (ODP) cores.
REFERENCES Ashworth, A. C., Lewis, A. R., Marchant, D. R., Cantrill, D. J., Francis, J. E., Leng, M. J., Raine, J. I., Newton, A. E., Williams, M., & Wolfe. A. P. (2008). The Neogene Biota of the Transantarctic Mountains, US Geological Survey and The National Academies, USGS OFR-2007. Askin, R. A. (1988a). Campanian to Paleocene palynological succession of Seymour and adjacent islands, northeastern Antarctic Peninsula. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir, Vol. 169, pp. 131–153.
352
J. E. Francis et al.
Askin, R. A. (1988b). The palynological record across the Cretaceous/Tertiary transition on Seymour Island, Antarctica. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir, Vol. 169, pp. 155–162. Askin, R. A. (1992). Late Cretaceous–Early Tertiary Antarctic outcrop evidence for past vegetation and climate. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 56, pp. 61–75. Askin, R. A. (1997). Eocene to ?Earliest Oligocene terrestrial palynology of Seymour Island. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antartica Publication, Siena, pp. 993–996. Askin, R. A. (2000). Spores and pollen from the McMurdo Sound erratics, Antarctica. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Paleoenvironments of Eocene rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 161–181. Askin, R. A., Ashworth, A. C., Harwood, D. H., Webb, P. N., & Wilson, G. S. (1999). Palynological survey of Sirius Group and related Neogene sediments in the transantarctic mountains. In: D. N. B. Skinner (Ed.). 8th International Symposium on Antarctic Earth Sciences, Victoria University of Wellington, New Zealand, p. 30, Programme and Abstracts. Askin, R. A., & Raine, J. I. (2000). Oligocene and Early Miocene Terrestrial palynology of the Cape Roberts Drill Hole CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7(4), 493–501. Barker, P. F., & Burrell, J. (1977). The opening of Drake Passage. Mar. Geol., 25, 15–34. Barrett, P. (2008). A history of Antarctic Cenozoic glaciation – View from the continent. In: F. Florindo, & M. Siegert (Eds). Antarctic Climate Evolution. Elsevier, Amsterdam, Vol. 8, pp. 33–83. Barrett, P. J. (Ed.) (1989). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. New Zealand DSIR Bulletin. Wellington, Vol. 245, 251 pp. Barrett, P. J. (1996). Antarctic paleoenvironments through Cenozoic time – A review. Terra Antartica, 3, 103–119. Barrett, P. J. (2007). Cenozoic climate and sea level history from glaciomarine strata off the Victoria Land coast, Cape Roberts Project, Antarctica. In: M. J. Hambrey, P. Christoffersen, N. F. Glasser, & B. Hubbard (Eds). Glacial Sedimentary Processes and Products. International Association of Sedimentologists, Special Publication, Vol. 39, pp. 259–287. Barrett, P. J., Hambrey, M. J., Harwood, D. M., Pyne, A. R., & Webb, P.-N. (1989). Synthesis. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. New Zealand DSIR Bulletin. Wellington, Vol. 245, 241–251. Barrett, P. J., Hambrey, M. J., & Robinson, P. H. (1991). Cenozoic glacial and tectonic history from CIROS-1, McMurdo Sound. In: M. R. A. Thomson,
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
353
A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 651–656. Barron, J., Larsen, B., & Shipboard Scientific Party (1991). In: Proceedings of the Ocean Drilling Program, Scientific Results, Leg 119, available from http://wwwodp.tamu.edu/publications/119_SR/119TOC.HTM Barrett, P. J., & Ricci, C. A. (Eds) (2001a). Studies from the Cape Roberts Project Ross Sea, Antarctica. Scientific Results of CRP-3, Part I, Geophysics and Tectonic Studies, Sedimentary Environments. Terra Antartica, 8(3), 120–308. Barrett, P. J., & Ricci, C. A. (Eds) (2001b). Studies from the Cape Roberts Project Ross Sea, Antarctica. Scientific results of CRP-3, Part II, palaeontological studies, provenance and climate from petrology, chronology and chronostratigraphy. Terra Antartica, 8(4), 309–620. Billups, K., & Schrag, D. P. (2002). Paleotemperatures and ice volume of the past 27 Myr revisited with paired Mg/Ca and O-18/O-16 measurements on benthic foraminifera. Paleoceanography, 17(1), 1003, doi:10.1029/2000PA000567. Billups, K., & Schrag, D. P. (2003). Application of benthic foraminiferal Mg/Ca ratios to questions of Cenozoic climate change. Earth Planet. Sci. Lett., 209, 181–195. Birkenmajer, K. (1980a). A revised lithostratigraphic standard for the tertiary of King George Island, South Shetland Islands (West Antarctica). Bull. Pol. Acad. Sci. Earth Sci., 27(1–2), 49–57. Birkenmajer, K. (1980b). Tertiary volcanic–sedimentary succession at Admiralty Bay, King George Island (South Shetland Islands, Antarctica). Studia Geologica Pol., 64, 7–65. Birkenmajer, K. (1981). Lithostratigraphy of the Point Hennequin Group (Miocene volcanics and sediments) at King George Island (South Shetland Islands, Antarctica). Studia Geologica Pol., 74, 175–197. Birkenmajer, K. (1989). A guide to tertiary geochronology of King George Island, West Antarctica. Pol. Polar Res., 10, 555–579. Birkenmajer, K. (1990). Geochronology and climatostratigraphy of tertiary glacial and interglacial successions on King George Island, South Shetland Islands (West Antarctica). Zbl. Geol. Pala¨ont., 1, 141–151. Birkenmajer, K. (1995). Mesozoic–Cenozoic magmatic arcs of Northern Antarctic Peninsula: Subduction, rifting and structural evolution. In: R. K. Srivastava, & R. Chandra (Eds). Magmatism in Relation to Diverse Tectonic Settings. Oxford/ IBH Publishing, New Delhi, pp. 329–344. Birkenmajer, K. (1997). Tertiary glacial/interglacial palaeoenvironments and sea-level changes, King George Island, West Antarctica. An overview. Bull. Pol. Acad. Sci. Earth Sci., 44, 157–181. Birkenmajer, K., Delitala, M. C., Narebski, W., Nicoletti, M., & Petrucciani, C. (1986). Geochronology of tertiary island-arc volcanics and glacigenic deposits, King George Island, South Shetland Islands (West Antarctica). Bull. Pol. Acad. Sci. Earth Sci., 34(3), 257–273.
354
J. E. Francis et al.
Birkenmajer, K., Gazdzicki, A., Krajewski, K. P., Przybycin, A., Solecki, A., Tatur, A., & Yoon, H. I. (2005). First Cenozoic glaciers in West Antarctica. Pol. Polar Res., 26, 3–12. Birkenmajer, K., & Zastawniak, E. (1989a). Late Cretaceous–Early Tertiary floras of King George Island, West Antarctica: Their stratigraphic distribution and palaeoclimatic significance, origins and evolution of the Antarctic biota. Geological Society of London Special Publication, 147, 227–240. Birkenmajer, K., & Zastawniak, E. (1989b). Late Cretaceous–Early Neogene vegetation history of the Antarctic Peninsula sector, Gondwana breakup and tertiary glaciations. Bull. Pol. Acad. Sci. Earth Sci., 37, 63–88. Bohaty, S. M., & Zachos, J. C. (2003). Significant Southern Ocean warming event in the Late Middle Eocene. Geology, 31, 1017–1020. Brea, M. (1996). Ana´lisis de los anillos de crecimiento de len˜os fo´siles de conı´ feras de la Formacio´n La Meseta, Isla Seymour, Anta´rtida. Congreso Paleo´geno de Ame´rica del Sur. Resu´menes, Santa Rosa, p. 28. Brea, M. (1998). Ana´lisis de los Anillos de Crecimiento en Len˜os Fo´siles de Conı´ feras de la Formacio´n La Meseta, Isla Seymour (Marambio), Anta´rtida. In: S. Casadı´ o (Ed.). Paleo´geno de Ame´rica del Sur y de la Penı´nsula Anta´rtica, Asociacio´n Paleontolo´gica Argentina, pp. 163–175, Publicacio´n Especial, Buenos Aires. Brinkhuis, H. (1994). Late Eocene to Early Oligocene dinoflagellate cysts from the priabonian type-area (Northeast Italy); biostratigraphy and palaeoenvironmental interpretation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 107, 121–163. Brinkhuis, H., & Biffi, U. (1993). Dinoflagellate cyst stratigraphy of the Eocene/ Oligocene transition in Central Italy. Mar. Micropaleontol., 22, 131–183. Brinkhuis, H., Munsterman, D. K., Sengers, S., Sluijs, A., Warnaar, J., & Williams, G. L. (2003). Late Eocene–quaternary dinoflagellate cysts from ODP site 1168, off Western Tasmania. In: N. F. Exon, J. P. Kennett, & M. J. Malone (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, 189, available from http://www-odp.tamu.edu/publications/189_SR/105/105.htm Brown, B., Gaina, C., & Dietmar Mu¨ller, R. (2006). Circum-Antarctic palaeobathymetry: Illustrated examples from Cenozoic to recent times. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 158–168. Cantrill, D. (2001). Early Oligocene Nothofagus from CRP-3, Antarctica: Implications for the vegetation history. Terra Antartica, 8(4), 401–406. Case, J. A. (1988). Paleogene floras from Seymour Island, Antarctic Peninsula, geology and paleontology of Seymour Island, Antarctic Peninsula. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir, Vol. 169, pp. 523–530. Cookson, I. C., & Eisenack, A. (1965). Microplankton from the Browns Creek Clays, SW Victoria. Proc. R. Soc. Victoria, 79(1), 119–137. Cooper, A. K., & O’Brien, P. E. (2004). Leg 188 synthesis: Transitions in the glacial history of the Prydz Bay region, East Antarctica, from ODP drilling. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
355
Drilling Program, Scientific Results, 188, available from http://www-odp.tamu. edu.publications/188_SR/synth/synth.htm Cooper, A. K., O’Brien, P. E., & Shipboard Scientific Party (2004). Prydz Bay – Co-operation Sea, Antarctica: Glacial history and paleoceanography sites 1165–1167. Proceedings of the Ocean Drilling Program, Scientific Results, 188, available from http://www-odp.tamu.edu.publications/188_SR/188TOC.HTM Cooper, A. K., Stagg, H., & Geist, E. L. (1991). Seismic stratigraphy and structure of Prydz Bay, Antarctica: Implications for Leg 119 drilling. In: Proceedings of the Ocean Drilling Program, scientific results, 119, available from http://wwwodp.tamu.edu/publications/119_SR/119TOC.HTM Coxall, H. K., Wilson, P. A., Palike, H., Lear, C. H., & Backman, J. (2005). Rapid stepwise onset of Antarctic glaciation and deeper calcite compensation in the Pacific Ocean. Nature, 433, 53–57. Cranwell, L. M. (1959). Fossil pollen from Seymour Island, Antarctica. Nature, 186, 1782–1785. Crouch, E. M. (2001). Environmental change at the Paleocene–Eocene biotic turnover, Ph.D. Thesis, Utrecht University, The Netherlands, LPP Foundation Series 14, 216 pp. Crouch, E. M., & Hollis, C. J. (1996). Paleocene palynomorph and radiolarian biostratigraphy of DSDP Leg 29, sites 280 and 281, South Tasman Rise, Institute of Geological and Nuclear Sciences, Scientific Report 96/19, 46 pp. DeConto, R. M., & Pollard, D. (2003a). A coupled climate–Ice sheet modelling approach to the Early Cenozoic history of the Antarctic Ice Sheet. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 39–53. DeConto, R. M., & Pollard, D. (2003b). Rapid Cenozoic glaciation of Antarctica induced by declining atmospheric CO2. Nature, 421, 245–249. DeConto, R. M., Pollard, D., & Harwood, D. (2007). Sea ice feedback and Cenozoic evolution of Antarctic climate and ice sheets. Paleoceanography, 22, PA3214, doi:10.1029/2000PA000567. Dettmann, M. E., & Jarzen, D. M. (1991). Pollen evidence for Late Cretaceous differentiation of proteaceae in Southern Polar forests. Can. J. Bot., 69, 901–906. Dettmann, M. E., Pocknall, D. T., Romero, E. J., & Zamaloa, M. del C. (1990). Nothofagidites Erdtman ex Potonie´, 1960; a catalogue of species with notes on the phytogeographic distribution of Nothofagus Bl. (Southern Beech). N. Z. Geol. Surv. Palaeontol. Bull., 60, 79 pp. Dingle, R., & Lavelle, M. (1998). Late Cretaceous–Cenozoic climatic variations of the Northern Antarctic Peninsula: New geochemical evidence and review. Palaeogeogr. Palaeoclimatol. Palaeoecol., 141(3), 215–232. Dingle, R., Marenssi, S. A., & Lavelle, M. (1998). High latitude Eocene climatic deterioration: Evidence from the Northern Antarctic Peninsula. J. S. Am. Earth Sci., 11(6), 571–579. Doktor, M., Gazdzicki, A., Marenssi, S., Porebski, S., Santillana, S., & Vrba, A. (1988). Argentine–Polish geological investigations on Seymour (Marambio) Island, Antarctica, 1988. Pol. Polar Res., 9, 521–541.
356
J. E. Francis et al.
Doktor, M., Gazdzicki, A. J., Jermanska, A., Porebski, S., & Zastawaniak, E. (1996). A plant–fish assemblage from the Eocene La Meseta Formation of Seymour Island (Antarctic Peninsula) and its environmental implications. Acta Palaeontol. Pol., 55, 127–146. Duane, A. M., Pirrie, D., & Riding, J. B. (Eds) (1992). Palynology of the James Ross Island Area, Antarctic Peninsula. Antarctic Sci., 4(3), 257–362, Special Issue. Dutton, A., Lohmann, K., & Zinsmeister, W. J. (2002). Stable isotope and minor element proxies for Eocene climate of Seymour Island, Antarctica. Paleoceanography, 17(2), 1016, doi:10.1029/2000PA000593. Eagles, G., Livermore, R., & Morris, P. (2006). Small basins in the Scotia Sea; the Eocene Drake Passage gateway. Earth Planet. Sci. Lett., 242(3–4), 343–353. Ehrmann, W. (1997). Smectite concentrations and crystallinities: Indications for Eocene Age of glaciomarine sediments in the CIROS-1 Drill Hole, McMurdo Sound, Antarctic. In: E. A. Ricci (Ed.). The Antarctic Region, Geological Evolution and Processes. Museo Nazionale dell’Antartide, Siena, Italy, pp. 771–780. Ehrmann, W. U., & Mackensen, A. (1992). Sedimentological evidence for the formation of an East Antarctic Ice Sheet in Eocene/Oligocene time. Palaeogeogr. Palaeoclimatol. Palaeoecol., 93, 85–112. Elliot, D. H. (1988). Tectonic setting and evolution of the James Ross Basin, Northern Antarctic Peninsula. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir, Vol. 169, pp. 541–555. Elliot, D. H., Askin, R. A., Kyte, F., & Zinsmeister, W. J. (1994). Iridium and dinocysts at the Cretaceous/Tertiary boundary on Seymour Island, Antarctica. Geology, 22, 675–678. Elliot, D. H., & Trautman, T. A. (1982). Lower tertiary strata on Seymour Island, Antarctic Peninsula. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, pp. 287–298. Exon, N. F., Kennett, J. P., & Malone, M. J. (Eds) (2004). Scientific results of the Ocean Drilling Program, 189, College Station, Texas, available from http://wwwodp.tamu.edu/publications/189_SR/106/106.htm Fielding, C. R., Woolfe, K. J., Purdon, R. G., Lavelle, M. A., & Howe, J. A. (1997). Sedimentological and stratigraphical re-evaluation of the CIROS-1 Core, McMurdo Sound, Antarctica. Terra Antartica, 4, 149–160. Florindo, F., Wilson, G. S., Roberts, A. P., Sagnotti, L., & Verosub, K. L. (2005). Magnetostratigraphic chronology of a Late Eocene to Early Miocene glacimarine succession from the Victoria Land Basin, Ross Sea, Antarctica. Global Planet. Change, 45, 207–236. Francis, J. E. (1991). Palaeoclimatic significance of Cretaceous–Early Tertiary fossil forests of the Antarctic Peninsula. In: M. R. A. Thomson, A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, New York, pp. 623–627.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
357
Francis, J. E. (1999). Evidence from fossil plants for Antarctic palaeoclimates over the past 100 million years. In: P. J. Barrett, & G. Orombelli (Eds). Geological Records of Global and Planetary Changes. Terra Antartica Report 3, Siena, Italy, pp. 43–52. Francis, J. E. (2000). Fossil wood from Eocene high latitude forests, McMurdo Sound, Antarctica. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Palaeoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 253–260. Francis, J. E., Ashworth, A., Cantrill, D. J., Crame, J. A., Howe, J., Stephens, R., Tosolini, A.-M., & Thorn, V. (2008). 100 Million Years of Antarctic Climate Evolution: Evidence from Fossil Plants, US Geological Survey and The National Academies; USGS OFR-2007. Francis, J. E., & Hill, R. S. (1996). Fossil plants from the Pliocene Sirius Group, Transantarctic Mountains: Evidence for climate from growth rings and fossil leaves. Palaios, 11(4), 389–396. Francis, J. E., & Poole, I. (2002). Cretaceous and Early Tertiary climates of Antarctica: Evidence from fossil wood. Palaeogeogr. Palaeoclimatol. Palaeoecol., 182(1–2), 47–64. Francis, J. E., Tosolini, A.-M., & Cantrill, D. (2004). Biodiversity and climate change in Antarctic Palaeogene floras. In: VII International Organisation of Palaeobotany Conference, Bariloche, Argentina, pp. 33–34, Abstract Volume. Gandolfo, M. A., Hoc, P., Santillana, S., & Marenssi, S. A. (1998a). Una Flor Fo´sil Morfolo´gicamente Afı´ n a las Grossulariaceae (Orden Rosales) de la Formacio´n La Meseata (Eoceno medio), Isla Marambio, Anta´rtida. In: S. Casadı´ o (Ed.). Paleo´geno de Ame´rica del Sur y de la Penı´nsula Anta´rtica, Asociacio´n Paleontolo´gica Argentina, pp. 147–153, Publicacio´n Especial, Buenos Aires. Gandolfo, M. A., Marenssi, S. A., & Santillana, S. N. (1998b). Flora y Paleoclima de la Formacio´n La Meseta (Eoceno medio), Isla Marambio (Seymour), Anta´rtida. In: S. Casadı´ o (Ed.). Paleo´geno de Ame´rica del Sur y de la Penı´nsula Anta´rtica, Asociacio´n Paleontolo´gica Argentina, pp. 155–162, Publicacio´n Especial, Buenos Aires. Gandolfo, M. A., Marenssi, S. A., & Santillana, S. N. (1998c). Flora y Paleoclima de la Formacio´n La Meseta (Eoceno-Oligoceno Inferior?) Isla Marambio (Seymour), Anta´rtida. In: I Congreso del Paleo´geno de Ame´rica del Sur, La Pampa, Argentina, Actas, pp. 31–32. Gazdzicki, A., Gruszczynski, M., Hoffman, A., Malkowski, K., Marenssi, S., Halas, S., & Tatur, A. (1992). Stable carbon and oxygen isotope record in the Paleogene La Meseta Formation, Seymour Island, Antarctica. Antarctic Sci., 4(4), 461–468. Gothan, W. (1908). Die Fossilen Holzer von der Seymour und Snow Hill Insel. In: O. Nordenskjold (Ed.). Wissenschaftliche Erbegnesse Schwedischen Sudpolar Expedition 1901–1903, Stockholm, pp. 1–33. Gradstein, F. M., Ogg, J. G., Smith, A. G., Agterberg, F. P., Bleeker, W., Cooper, R. A., Davydov, V., Gibbard, P., Hinnov, L. A., House, M. R., Lourens, L.,
358
J. E. Francis et al.
Luterbacher, H. P., McArthur, J., Melchin, M. J., Robb, L. J., Shergold, J., Villeneuve, M., Wardlaw, B. R., Ali, J., Brinkhuis, H., Hilgen, F. J., Hooker, J., Howarth, R. J., Knoll, A. H., Laskar, J., Monechi, S., Plumb, K. A., Powell, J., Raffi, I., Ro¨hl, U., Sadler, P., Sanfilippo, A., Schmitz, B., Shackleton, N. J., Shields, G. A., Strauss, H., Van Dam, J., van Kolfschoten, T., Veizer, J., & Wilson, D. (2004). A Geologic Time Scale 2004. Cambridge University Press, 589 pp. Grube, R. (2004). Palynologic Studies on Terrestrial Vegetation and Continental Climate in the Southern High Latitudes at the Eocene/Oligocene Boundary Interval. Evolutionary Transformations and Mass Extinctions, Progress Report 2/2004, pp. 70–78. Grube, R., & Mohr, B. (2008). Deterioration and/or cyclicity? The development of vegetation and climate during the Eocene and Oligocene in Antarctica. In: A. K. Cooper, C. R. Raymond, et al. (Eds). Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES X, USGS Open-File Report 2007-1047, 4 pp., Extended Abstract 075. Grube, R., & Mohr, B. A. R. (2004). Vegetation and climate during the Eocene and the Eocene/Oligocene boundary interval in Southern High Latitudes – Preliminary results. In: VII International Organisation of Palaeobotany Conference, Bariloche, Argentina, pp. 49–50, Abstract Volume. Guerstein, G. R., Chiesa, J. O., Guler, M. V., & Camacho, H. H. (2002). Bioestratigrafia Basada en Quistes de Dinoflagelados de la Formacion Cabo Pena (Eoceno terminal-Oligoceno temprano), Tierra del Fuego, Argentina. Revista Espan˜ola de Micropaleontologı´a, 34(1), 105–116. Hambrey, M. J., & Barrett, P. J. (1993). Cenozoic sedimentary and climatic record, Ross Sea region, Antarctica. In: J. P. Kennett, & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change, Part 2. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 60, pp. 91–124. Hambrey, M. J., Barrett, P. J., Ehrmann, E. H., & Larsen, B. (1992). Cenozoic sedimentary processes on the Antarctic continental shelf: The record from deep drilling. Z. Geomorphologie, 86(Suppl.), 73–99. Hambrey, M. J., Barrett, P. J., & Powell, R. D. (2002). Late Oligocene and Early Miocene glacimarine sedimentation in the SW Ross Sea, Antarctica: The record from offshore drilling. In: C. O’Cofaigh, & J. A. Dowdeswell (Eds). GlacierInfluence Sedimentation on High-Latitude Continental Margins, Geological Society of London Special Publication 203, pp. 105–128. Hambrey, M. J., Barrett, P. J., & Robinson, P. H. (1989). Stratigraphy. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. New Zealand DSIR Bulletin, Wellington, 245, pp. 23–48. Hambrey, M. J., Ehrmann, E. H. R., & Larsen, B. (1991). The Cenozoic glacial record of the Prydz Bay continental shelf, East Antarctica. In: J. Barron, B. Larsen, Shipboard Scientific Party. Proceedings of the Ocean Drilling Program,
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
359
Scientific Results, 119, available from http://www-odp.tamu.edu/publications/ 119_SR/119TOC.HTM Hambrey, M. J., & McKelvey, B. C. (2000a). Neogene fjordal sedimentation in the Prince Charles Mountains, East Antarctica. Sedimentology, 47, 577–607. Hambrey, M. J., & McKelvey, B. C. (2000b). Major Neogene fluctuations of the East Antarctic Ice Sheet: Stratigraphic evidence from the Lambert glacier region. Geology, 28(10), 887–891. Hannah, M. J. (1997). Climate controlled dinoflagellate distribution in Late Eocene– Earliest Oligocene strata from CIROS-1 Drillhole, McMurdo Sound, Antarctica. Terra Antartica, 4(2), 73–78. Hannah, M. J., Wilson, G. J., & Wrenn, J. H. (2000). Oligocene and Miocene marine palynomorphs from CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7(4/5), 503–511. Hannah, M. J., Wrenn, J. H., & Wilson, G. J. (1998). Early Miocene and quaternary marine palynomorphs from Cape Roberts Project CRP-1, McMurdo Sound, Antarctica. Terra Antartica, 5(3), 527–538. Haomin, L. (1994). Early Tertiary fossil hill flora from Fildes Peninsula of King George Island, Antarctica. In: S. Yanbin, (Ed.). Stratigraphy and Palaeontology of Fildes Peninsula, King George Island, Antarctica, State Antarctic Committee, Science Press, Beijing, pp. 165–171, Monograph 3. Haskell, T. R., & Wilson, G. J. (1975). Palynology of Sites 280–284, DSDP Leg 29, off Southeastern Australia and Western New Zealand. Initial Reports of the Deep Sea Drilling Program, 29, 723–741. Hathway, B. (2000). Continental rift to back-arc basin; Jurassic-Cretaceous stratigraphical and structural evolution of the Larsen Basin, Antarctic Peninsula. J. Geol. Soc. Lond., 157(2), 417–432. Haywood, A. M., Smellie, J. L., Ashworth, A. C., Cantrill, D. J., Florindo, F., Hambrey, M. J., Hill, D., Hillenbrand, C.-D., Hunter, S. J., Larter, R. D., Lear, C. H., Passchier, S., & van de Wal, R. (2008). Middle Miocene to Pliocene History of Antarctica and the Southern Ocean 405. In: F. Florindo, & M. Siegert (Eds). Antarctic Climate Evolution. Elsevier, Amsterdam, Vol. 8, pp. 401–463. Hiemstra, J. F. (1999). Microscopic evidence of grounded ice in the sediments of the CIROS-1 Core, McMurdo Sound, Antarctica. Terra Antartica, 6, 365–376. Hill, R. S. (1989). Palaeontology-fossil leaf. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. New Zealand DSIR Bulletin, Wellington, 245, pp. 143–144. Holbourn, A., Kuhnt, W., Schulz, M., & Erlrnkeuser, H. (2005). Impacts of orbital forcing and atmospheric carbon dioxide on Miocene ice-sheet expansion. Nature, 438, 483–487. Huber, M., Brinkhuis, H., Stickley, C. E., Do¨o¨s, K., Sluijs, A., Warnaar, J., Schellenberg, S. A., & Williams, G. L. (2004). Eocene circulation of the Southern Ocean: Was Antarctica kept warm by subtropical waters? Paleoceanography, 19(4), PA4026, doi:10.1029/2004PA001014.
360
J. E. Francis et al.
Huber, M., & Sloan, L. C. (2001). Heat transport, deep waters, and thermal gradients: Coupled simulation of an Eocene greenhouse climate. Geophys. Res. Lett., 28, 3481–3484. Hunt, R. (2001). Biodiversity and palaeoecology of Tertiary fossil floras in Antarctica, Ph.D. Thesis, University of Leeds, Leeds, UK. Hunt, R. J., & Poole, I. (2003). Paleogene West Antarctic climate and vegetation history in light of new data from King George Island. In: S. L. Wing, P. D. Gingerich, B. Schmitz, & E. Thomas (Eds). Causes and Consequences of Globally Warm Climates in the Early Paleogene. Geological Society of America, Boulder, CO, Vol. 369, pp. 395–412. Ivany, L. C., Van Simaeys, S., Domack, E. W., & Samson, S. D. (2006). Evidence for an Earliest Oligocene ice sheet on the Antarctic Peninsula. Geology, 34(5), 377–380. Jones, C. M. (2000). The first record of a fossil bird from East Antarctic. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Palaeoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 359–364. Jovane, L., Florindo, F., Coccioni, R., Dinare`s-Turell, J., Marsili, A., Monechi, S., Roberts, A. P., & Sprovieri, M. (2007). The middle Eocene climatic optimum event in the Contessa Highway Section, Umbrian Apennines, Italy. Geol. Soc. Am. Bull., 119, 413–427. Kennett, J. P. (1977). Cenozoic evolution of Antarctic glaciation, the circumAntarctic oceans and their impact on global paleoceanography. J. Geophys. Res., 82, 3843–3859. Kennett, J. P. (1978). The development of planktonic biogeography in the Southern Ocean during the Cenozoic. Mar. Micropaleontol., 3(4), 301–345. Kennett, J. P., Houtz, R. E., Andrews, P. B., Edwards, A. R., Gostin, V. A., Hajo´s, M., Hampton, M. A., Jenkins, D. G., Margolis, S. V., Ovenshine, A. T., & PerchNielsen, K. (1975). Initial Reports of the Deep Sea Drilling Project. U.S. Government Printing Office, Washington, DC, Vol. 29, 1197 pp. Available from http://www.deepseadrilling.org/29/dsdp_toc.htm Kennett, J. P., & Shackleton, N. J. (1976). Oxygen isotopic evidence for the development of the psychryosphere 38 my ago. Nature, 260, 513–515. Kohn, M. J., Josef, J. A., Madden, R., Kay, R., Vucetich, G., & Carlini, A. A. (2004). Climate stability across the Eocene–Oligocene transition, Southern Argentina. Geology, 32, 621–624. Lawver, L. A., & Gahagan, L. M. (1998). Opening of Drake Passage and its impact on Cenozoic ocean circulation. In: T. J. Crowley, & K. C. Burke (Eds). Tectonic Boundary Conditions for Climate Reconstructions. Oxford University Press, New York, pp. 212–223. Lawver, L. A., Gahagan, L. M., & Coffin, M. F. (1992). The development of paleoseaways around Antarctica. In: J. P. Kennett & D. A. Warnke (Eds). The Antarctic Paleoenvironment: a Perspective on Global Change, Part 1. American Geophysical Union, Washington DC, USA, American Geophysical Union Geophysical Monograph, Vol. 56, pp. 7–30.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
361
Lear, C. H., Elderfield, H., & Wilson, P. A. (2000). Cenozoic deep-sea temperatures and global ice volumes from Mg/Ca in benthic foraminiferal calcite. Science, 287, 269–272. Lear, C. H., Rosenthal, Y., Coxall, H. K., & Wilson, P. A. (2004). Late Eocene to Early Miocene ice-sheet dynamics and the global carbon cycle. Paleoceanography, 19(4), PA4015, doi:10.1029/2004PA001039. Levy, R. H., & Harwood, D. M. (2000a). Sedimentary lithofacies of the McMurdo Sound erratics. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Paleoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 39–61. Levy, R. H., & Harwood, D. M. (2000b). Tertiary marine palynomorphs from the McMurdo Sound erratics, Antarctica. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Paleoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 183–242. Li, H. M. (1992). Early Tertiary palaeoclimate of King George Island, Antarctica – Evidence from the fossil hill flora. In: Y. Yoshida, K. Kaminuma, & K. Shiraishi (Eds). Recent Progress in Antarctic Earth Science. Terra Scientific Publishing Company, Tokyo, pp. 371–375. Livermore, R., Eagles, G., Morris, P., & Maldonado, A. (2004). Shackleton Fracture Zone: No barrier to early circumpolar ocean circulation. Geology, 32(9), 797–800. Livermore, R., Nankivell, A., Eagles, G., & Morris, P. (2005). Paleogene opening of Drake Passage. Earth Planet. Sci. Lett., 236(1–2), 459–470. Long, D. H. and Stilwell, J. D. (2000). Fish Remains from the Eocene of Mount Discovery, East Antarctic, in Stilwell, J. D. and Feldmann, R. M. (Eds). Paleobiology and Palaeoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 349–354. Macphail, M. K., & Truswell, E. M. (2004). Palynology of Site 1166, Prydz Bay, East Antarctica. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, 188, available from http:// www-odp.tamu.edu/publications/188_SR/012/012.htm Mao, S., & Mohr, B. A. R. (1995). Middle Eocene dinocysts from Bruce Bank (Scotia Sea, Antarctica) and their paleoenvironmental and paleogeographic implications. Rev. Palaeobot. Palynol., 86, 235–263. Marenssi, S. A. (1995). Sedimentologı´a y Paleoambientes Sedimentarios de la Formacio´n La Meseta, Isla Marambio, Anta´rtida, Ph.D. Thesis, Universidad de Buenos Aires, Buenos Aires, 502 pp. Marenssi, S. A., Net, L. I., & Santillana, S. N. (2002). Provenance, depositional and paleogeographic controls on sandstone composition in an incised valley system: The Eocene La Meseta Formation, Seymour Island, Antarctica. Sedimentary Geol., 150(3–4), 301–321. Marenssi, S. A., Santillana, S., & Rinaldi, C. A. (1998). Stratigraphy of the La Meseta Formation (Eocene), Marambio (Seymour) Island, Antarctica.
362
J. E. Francis et al.
In: S. Casadı´ o (Ed.). Paleo´geno de Ame´rica del Sur y de la Penı´nsula Anta´rtica, Asociacio´n Paleontolo´gica Argentina, Buenos Aires, pp. 137–146, Publicacio´n Especial 5. Marenssi, S. A., Santillana, S. N., & Rinaldi, C. A. (1996). Stratigraphy of La Meseta Formation (Eocene), Marambio Island, Antarctica. In: I Congreso Paleo´geno de Ame´rica del Sur, Santa Rosa, La Pampa, pp. 33–34, Abstracts Volume. McKelvey, B., Hambrey, M. J., Harwood, D. M., Mabin, M. C. G., Webb, P.-N., & Whitehead, J. M. (2001). The Pagodroma Group – A Cenozoic record of the East Antarctic Ice Sheet in the northern Prince Charles Mountains. Antarctic Sci., 13, 455–468. McMinn, A. (1995). Why are there no post-Paleogene dinoflagellate cysts in the Southern Ocean? Micropaleontology, 41, 383–386. Mikolajewicz, U., Maier-Reimer, E., Crowley, T. J., & Kim, K.-Y. (1993). Effect of Drake and Panamanian Gateways on the circulation of an ocean model. Paleoceanography, 8(4), 409–426. Mildenhall, D. C. (1989). Terrestrial palynology. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. New Zealand DSIR Bulletin. Wellington, Vol. 245, pp. 119–127. Miller, K. G., Mountain, G. S., Browning, J. V., Kominz, M., Sugarman, P. J., Christie-Blick, N., Katz, M. E., & Wright, J. D. (1998). Cenozoic global sea level, sequences, and the New Jersey Transect: Results from coastal plain and continental slope drilling. Rev. Geophys., 36, 569–601. Miller, K. G., Wright, J. D., & Browning, J. V. (2005). Visions of ice sheets in a greenhouse world. Mar. Geol., 217, 215–231. Mohr, B. A. R. (1990). Eocene and Oligocene sporomorphs and dinoflagellate cysts from Leg 113 Drill Sites, Weddell Sea, Antarctica. In: N. J. Stewart (Ed.). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 113, pp. 595–612. Mohr, B. A. R. (2001). The development of Antarctic fern floras during the Tertiary, and palaeoclimatic and palaeobiogeographic implications. Palaeontographica B, 259, 167–208. Mohr, B. A. R., & Lazarus, D. B. (1994). Palaeobiogeographic distribution of Kuylisporites and its possible relationship to the extant fern genus Cnemidaria (Cyatheaceae). Ann. Missouri Botanic Gardens, 81, 758–767. Murphy, M. G., & Kennett, J. P. (1975). Development of latitudinal thermal gradients during the Oligocene; oxygen-isotope evidence from the Southwest Pacific. In: J. P. Kennett, C. C. von der Borch, P. A. Baker, C. E. Barton, A. Boersma, J. P. Caulet, W. C. Dudley, Jr., J. V. Gardner, D. Jenkins, D. Graham, W. H. Lohman, E. Martini, R. B. Merrill, R. Morin, C. S. Nelson, C. Robert, M. S. Srinivasan, R. Stein, & A. Takeuchi (Eds). Initial Reports of the Deep Sea Drilling Project, Vol. 90 (2). U.S. Government Printing Office, Washington, DC, pp. 1347–1360, available from http://www.deepseadrilling.org/ 90/dsdp_toc.htm
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
363
Naish, T. R., Woolfe, K. J., Barrett, P. J., et al. (2001). Orbitally induced oscillations in the East Antarctic Ice Sheet at the Oligocene/Miocene boundary. Nature, 413, 719–723. Nong, G. T., Najjar, R. G., Seidov, D., & Peterson, W. (2000). Simulation of ocean temperature change due to the opening of Drake Passage. Geophys. Res. Lett., 27, 2689–2692. Pagani, M., Zachos, J. C., Freeman, K. H., Tipple, B., & Bohaty, S. (2005). Marked decline in atmospheric carbon dioxide concentrations during the Paleogene. Science, 309, 600–603. Palike, H., Shackleton, N. J., & Ro¨hl, U. (2001). Astronomical forcing in Late Eocene marine sediments. Earth Planet. Sci. Lett., 193, 589–602. Pekar, S., & DeConto, R. M. (2006). High-resolution ice-volume estimates for the Early Miocene: Evidence for a dynamic ice sheet in Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 101–109. Pekar, S., DeConto, R. M., & Harwood, D. (2006). Resolving a Late Oligocene conundrum: Deep sea warming and Antarctic glaciation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 29–49. Pirrie, D., Duane, A. M., & Riding, J. B. (1992). Jurassic–Tertiary stratigraphy and palynology of the James Ross Basin: Review and introduction. Antarctic Sci., 4(3), 259–266. Pole, M., Hill, B., & Harwood, D. M. (2000). Eocene plant macrofossils from erratics, McMurdo Sound, Antarctica. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Paleoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 243–251. Pollard, D., & DeConto, R. M. (2005). Hysteresis in Cenozoic Antarctic Ice Sheet variations. Global Planet. Change, 45, 9–21. Pollard, D., DeConto, R. M., & Nyblade, A. (2005). Sensitivity of Cenozoic Antarctic Ice Sheet variations to geothermal heat flux. Global Planet. Change, 49, 63–74. Poole, I., Cantrill, D. J., & Utescher, T. (2005). A Multi-proxy approach to determine Antarctic terrestrial palaeoclimate during the Late Cretaceous and Early Tertiary. Palaeogeogr. Palaeoclimatol. Palaeoecol., 222, 95–121. Poole, I., Hunt, R., & Cantrill, D. (2001). A fossil wood flora from King George Island: Ecological implications for an Antarctic Eocene vegetation. Ann. Bot., 88(1), 33–54. Powell, R. D., Krissek, L. A., & van der Meer, J. J. M. (2000). Preliminary depositional environmental analysis of CRP-2/2A, Victoria Land Basin, Antarctica: Palaeoglaciological and palaeoclimatic inferences. Terra Antartica, 7, 313–322. Prebble, J. G., Raine, J. I., Barrett, P. J., & Hannah, M. J. (2006). Vegetation and climate from two Oligocene glacioeustatic sedimentary cycles (31 and 24 Ma) cored by the Cape Roberts Project, Victoria Land Basin, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 41–57.
364
J. E. Francis et al.
Raine, J. I., & Askin, R. A. (2001). Terrestrial palynology of Cape Roberts Project Drillhole CRP-3, Victoria Land Basin, Antarctica. Terra Antartica, 8(4), 389–400. Reguero, M. A., Marenssi, S. A., & Santillana, S. N. (2002). Antarctic Peninsula and South America (Patagonia) Paleogene Terrestrial environments: Biotic and biogeographic relationships. Palaeogeogr. Palaeoclimatol. Palaeobiol., 179(3–4), 189–210. Riding, J. B., & Crame, J. A. (2002). Aptian to Coniacian (Early-Late Cretaceous) palynostratigraphy of the Gustav Group, James Ross Basin, Antarctica. Cretaceous Res., 23(6), 739–760. Robert, C. M., Exon, N. F., Kennett, J. P., Malone, M. J., Brinkhuis, H., Chaproniere, G. C. H., Ennyu, A., Fothergill, P., Fuller, M. D., Grauert, M., Hill, P. J., Janecek, T. R., Kelly, D. C., Latimer, J. C., McGonigal, R. K., Nees, S., Ninnemann, U. S., Nurnberg, D., Pekar, S. F., Pellaton, C. C., Pfuhl, H. A., Rohl, U., Schellenberg, S. A., Shevenell, A. E., Stickley, C. E., Suzuki, N., Touchard, Y., Wei, W., & White, T. S. (2001). Palaeogene ocean opening south of Tasmania, and palaeoceanographic implications: Preliminary results of clay mineral analyses (ODP Leg 189). C. R. L’Academie Sci. Ser. IIA Earth Planet. Sci., 332(5), 323–329. Roberts, A. P., Wilson, G. S., Harwood, D. M., & Verosub, K. L. (2003). Glaciation across the Oligocene–Miocene boundary in southern McMurdo Sound, Antarctica: New chronology from the CIROS-1 Drill-Hole. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198(1), 113–130. Sadler, P. (1988). Geometry and stratification of uppermost Cretaceous and Paleogene units on Seymour Island, Northern Antarctic Peninsula. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir, Vol. 169, pp. 303–320. Sagnotti, L., Florindo, F., Verosub, K. L., Wilson, G. S., & Roberts, A. P. (1998). Environmental magnetic record of Antarctic palaeoclimate from Eocene/ Oligocene glaciomarine sediments, Victoria Land Basin. Geophys. J. Int., 134, 653–662. Scher, H. D., & Martin, E. E. (2006). Timing and climatic consequences of the opening of Drake Passage. Science, 312, 428–430. Schouten, S., Hopmans, E. C., Schefuss, E., & Damste, J. S. S. (2003). Distributional variations in marine crenarchaeotal membrane lipids: A new tool for reconstructing ancient sea water temperatures? Earth Planet. Sci. Lett., 204, 265–274. Shafik, S., & Idnurm, M. (1997). Calcareous microplankton and polarity reversal stratigraphies of the Upper Eocene Browns Creek clay in the Otway Basin, Southeast Australia: Matching the evidence. Aust. J. Earth Sci., 44, 77–86. Shen, Y. (1994). Subdivision and correlation of Cretaceous to Paleogene volcanosedimentary sequence from Fildes Peninsula, King George Island, Antarctica. In: Y. Shen (Ed.). Stratigraphy and Palaeontology of Fildes Peninsula, King George Island, Antarctica, State Antarctic Committee, Science Press, Beijing, pp. 1–36, Monograph 3.
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
365
Shipboard Scientific Party (2001a). Leg 188 Summary: Prydz Bay-Co-Operation Sea, Antarctica. In: P. E. O’Brien, A. K. Cooper, C. Richter, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports, Vol. 188, available from http://www-odp.tamu.edu/publications/188_IR/188TOC.HTM Shipboard Scientific Party (2001b). Leg summary. In: N. F. Exon, J. P. Kennett, M. J. Malone, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports, Vol. 189, available from http://www-odp.tamu.edu/publications/ 189_IR/chap_01/chap_01.htm Shipboard Scientific Party (2001c). Site 1170. In: N. F. Exon, J. P. Kennett, M. J. Malone, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports, Vol. 189, available from http://www-odp.tamu.edu/publications/ 189_IR/chap_05/chap_05.htm Shipboard Scientific Party (2001d). Site 1172. In: N. F. Exon, J. P. Kennett, M. J. Malone, et al. (Eds). Proceedings of the Ocean Drilling Program, Initial Reports, Vol. 189, available from http://www-odp.tamu.edu/publications/ 189_IR/chap_06/chap_06.htm Sloan, L. C., & Huber, M. (2001). Eocene oceanic responses to orbital forcing on precessional timescales. Paleoceanography, 16(1), 101–111. Sluijs, A., Brinkhuis, H., Stickley, C. E., Warnaar, J., Williams, G. L., & Fuller, M. (2003). Dinoflagellate cysts from the Eocene–Oligocene Transition in the Southern Ocean: Results from ODP Leg 189. In: N. F. Exon, J. P. Kennett, & M. J. Malone (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 189, available from http://www-odp.tamu.edu/publications/189_SR/104/ 104.htm Smellie, J. L., Pankhurst, R. J., Thomson, M. R. A., & Davies, R. E. S. (1984). The geology of the South Shetland Islands: VI. Stratigraphy, geochemistry and evolution, British Antarctic Survey Scientific Report 87, pp. 1–85. Stickley, C. E., Brinkhuis, H., Schellenberg, S. A., Sluijs, A., Ro¨hl, U., Fuller, M., Grauert, M., Huber, M., Warnaar, J., & Williams, G. L. (2004). Timing and nature of the deepening of the Tasmanian Gateway. Paleoceanography, 19(4), PA4027, doi:10.1029/2004PA001022. Stilwell, J., & Zinsmeister, W. (1992). Molluscan systematics and biostratigraphy, Lower Tertiary La Meseta Formation, Seymour Island, Antarctic Peninsula. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 55, 192 pp. Stilwell, J. D., & Feldmann, R. M. (Eds) (2000). Paleobiology and Paleoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, 372 pp. Stover, L. E. (1975). Observations on some Australian Eocene Dinoflagellates. Geosci. Man, 11, 35–45. Strong, C. P., Hollis, C. J., & Wilson, G. J. (1995). Foraminiferal, radiolarian, and dinoflagellate biostratigraphy of Late Cretaceous to Middle Eocene pelagic sediments (Muzzle Group), Mead Stream, Marlborough, New Zealand. N. Z. J. Geol. Geophys., 38, 171–212.
366
J. E. Francis et al.
Sugden, D. E., & Denton, G. H. (2004). Cenozoic landscape evolution of the Convoy Range to Mackay Glacier area, Transantarctic Mountains: Onshore to offshore synthesis. Geol. Soc. Am. Bull., 116, 840–847. Tambussi, C. P., Acosta Hospitaleche, C. I., Reguero, M. A., & Marenssi, S. A. (2006). Late Eocene penguins from West Antarctica; systematics and biostratigraphy. In: J. E. Francis, D. Pirrie, & J. A. Crame (Eds). Cretaceous-Tertiary high-latitude palaeoenvironments; James Ross Basin, Antarctica. Geological Society Special Publications 258, pp. 145–161. Tatur, A., Gazdzicki, A., & del Valle, R. A. (2006). Might Cd/Ca ratio in bivalve shells mark a change in the Southern Ocean circulation at the Eocene/Oligocene boundary? SCAR 2nd. Open Science Conference, SCAR XXIX/COMNAP XVIII – 2006. CD, Poster 0588, 221 pp. Taylor, J., Siegert, M. J., Payne, A. J., Hambrey, M. J., O’Brien, P., Cooper, A. K., & Leitchenkov, G. (2004). Topographic controls on post-Oligocene changes in ice-sheet dynamics, Prydz Bay region, East Antarctica. Geology, 32, 197–200. Thorn, V. C., & DeConto, R. (2006). Antarctic climate at the Eocene–Oligocene boundary – Climate model sensitivity to high latitude vegetation type and comparisons with the Palaeobotanical record. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 134–157. Toggweiler, J. R., & Bjornsson, H. (2000). Drake Passage and paleoclimate. J. Quaternary Sci., 15, 319–328. Torres, T., Marenssi, S. A., & Santillana, S. (1994). Maderas Fo´siles de la isla Seymour, Formacio´n La Meseta, Anta´rtica. Serie Cientı´fica del INACH, Santiago de Chile, 44, 17–38. Troedson, A., & Smellie, J. (2002). The Polonez Cove Formation of King George Island, Antarctica: Stratigraphy, facies and implications for mid-Cenozoic cryosphere development. Sedimentology, 49, 277–301. Truswell, E. M. (1997). Palynomorph assemblages from marine Eocene sediments on the West Tasmanian continental margin and the South Tasman rise. Aust. J. Earth Sci., 4, 633–654. Warnaar, J. (2006). Climatological implications of Australian–Antarctic separation, Doctoral Thesis, Utrecht University, 144 pp. Whitehead, J. M., Quilty, P. G., McKelvey, B. C., & O’Brien, P. E. (2006). A review of the Cenozoic stratigraphy and glacial history of the Lambert Graben – Prydz Bay Region, East Antarctica. Antarctic Sci., 18, 83–99. Wilch, T. I., & McIntosh, W. C. (2000). Eocene and Oligocene volcanism at Mount Petras, Marie Byrd Land: Implications for Middle Cenozoic ice sheet reconstructions in West Antarctica. Antarctic Sci., 12, 477–491. Williams, G. L., Brinkhuis, H., Pearce, M. A., Fensome, R. A., & Weegink, J. W. (2004). Southern Ocean and global dinoflagellate cyst events compared: Index events for the Late Cretaceous—Neogene. In: N. F. Exon, J. P. Kennett, & M. J. Malone (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 189, available from http://www-odp.tamu.edu/publications/189_SR/107/107.htm
From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica
367
Willis, P. M. A., & Stilwell, J. D. (2000). A possible piscivorous crocodile from Eocene deposits of McMurdo Sound, Antarctic. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Palaeoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 355–358. Willumsen, P. S. (2000). Late Cretaceous to Early Paleocene palynological changes in midlatitude Southern Hemisphere, New Zealand. GFF (Geol. Soc. Sweden), 122, 180–181. Wilpshaar, M., Santarelli, A., Brinkhuis, H., & Visscher, H. (1996). Dinoflagellate cysts and Mid-Oligocene chronostratigraphy in the Central Mediterranean region. J. Geol. Soc. Lond., 153, 553–561. Wilson, G. J. (1978). The dinoflagellate species Isabelidinium druggii (Stover) and I. seelandica (Lange); their association in the Teurian of Woodside Creek, Marlborough, New Zealand. N. Z. J. Geol. Geophys., 21, 75–80. Wilson, G. J. (1984). Some new dinoflagellate species from the New Zealand Haumurian and Piripauan Stages (Santonian-Maastrichtian, Late Cretaceous). N. Z. J. Bot., 22, 549–556. Wilson, G. J. (1985). Dinoflagellate biostratigraphy of the Eocene Hampden Section North Otago, New Zealand. N. Z. Geol. Surv. Rec., 8, 93–101. Wilson, G. J. (1988). Paleocene and Eocene dinoflagellate cysts from Waipawa, Hawkes Bay, New Zealand. N. Z. Geol. Surv. Palaeontol. Bull., 57, 1–96. Wilson, G. J. (1989). Marine palynology. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. New Zealand DSIR Bulletin. Wellington, Vol. 245, pp. 129–133. Wilson, G. S. (2000). Glacial geology and origin of fossiliferous-erratic-bearing moraines, southern McMurdo Sound, Antarctica – An alternative ice sheet hypothesis. In: J. D. Stilwell, & R. M. Feldmann (Eds). Paleobiology and Paleoenvironments of Eocene Rocks, McMurdo Sound, East Antarctica. Antarctic Research Series. American Geophysical Union, Washington, DC, Vol. 76, pp. 19–37. Wilson, G. S., Harwood, D. M., Askin, R. A., & Levy, R. H. (1998b). Late Neogene Sirius Group strata in Reedy Valley, Antarctica: A multiple-resolution record of climate, ice-sheet and sea-level events. J. Glaciol., 44, 437–448. Wilson, G. S., Naish, T. R., Aitken, A. R. A., Johnston, L. J., Damaske, D., Timms, C. J., Clifford, A. E., Cooper, A. F., Gorman, A. R., Henrys, S., Horgan, H., Jordan, T. A., & Stuart, G. (2006). Basin development beneath the southern McMurdo Sound ice shelf (SMIS) – Combined geophysical and glacial geological evidence for a potential Paleogene drilling target for ANDRILL. In: Open Science Conference, XXIX SCAR/COMNAP XVII, Hobart, Australia, Abstract. Wilson, G. S., Roberts, A. P., Verosub, K. L., Florindo, F., & Sagnotti, L. (1998a). Magnetobiostratigraphic chronology of the Eocene–Oligocene transition in the CIROS-1 core, Victoria Land Margin, Antarctica: Implications for Antarctic glacial history. Geol. Soc. Am. Bull., 110, 35–47.
368
J. E. Francis et al.
Wilson, T. J. (1999). Cenozoic structural segmentation of the Transantarctic Mountains rift flank in southern Victoria Land, Antarctica. In: F. M. Van der Wateren, & S. Cloetingh (Eds). Lithosphere dynamics and environmental change of the Cenozoic West Antarctic rift system. Global Planet. Change, 23, 105–127. Wrenn, J. H., & Beckmann, S. W. (1982). Maceral, total organic carbon, and palynological analyses of Ross Ice Shelf Project Site J-9 Cores. Science, 216, 187–189. Wrenn, J. H., & Hart, G. F. (1988). Paleogene dinoflagellate cyst biostratigraphy of Seymour Island, Antarctica. In: R. M. Feldmann, & M. O. Woodburne (Eds). Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir., Vol. 169, pp. 321–447. Zachos, J. C., Breza, J. R., & Wise, S. W. (1992). Early Oligocene ice-sheet expansion on Antarctica – Stable isotope and sedimentological evidence from Kerguelen Plateau, Southern Indian Ocean. Geology, 20, 569–573. Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693. Zachos, J. C., Quinn, T. M., & Salamy, K. A. (1996). High-resolution (104 years) deep-sea foraminiferal stable isotope records of the Eocene–Oligocene climate transition. Paleoceanography, 11, 251–266. Zastawaniak, E., Wrona, R., Gazdzicki, A. J., & Birkenmajer, K. (1985). Plant remains from the top part of the Point Hennequin Group (Upper Oligocene), King George Island (South Shetland Islands, Antarctica). Studia Geologica Pol., 81, 143–164.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00009-8
Chapter 9
The Oligocene–Miocene Boundary – Antarctic Climate Response to Orbital Forcing G. S. Wilson1,, S. F. Pekar2, T. R. Naish3, S. Passchier4 and R. DeConto5 1
Department of Geology, University of Otago, P.O. Box 56, Dunedin, New Zealand School of Earth and Environmental Sciences, Queens College, CUNY, 65-30 Kissena Blvd., Flushing, NY 11367, USA and Lamont Doherty Earth Observatory of Columbia University, Palisades, NY 10964, USA 3 Antarctic Research Centre, Victoria University, P.O. Box 600, Wellington, New Zealand and GNS Science, P.O. Box 30368, Lower Hutt, New Zealand 4 Department of Earth and Environmental Studies, Montclair State University, 252 Mallory Hall, 1 Normal Avenue, Montclair, NJ 07403, USA 5 Department of Geosciences, University of Massachusetts, 233 Morril Science Center, 611 North Pleasant Street, Amherst, MA 01003-9297, USA 2
ABSTRACT Recent high-resolution Oligocene–Miocene oxygen isotopic records revealed a relatively transient, ca. 2 myr period, 1 m amplitude cyclicity in isotopic values (Oi and Mi events, respectively). Intriguingly, it has been suggested that these isotopic excursions in oceanic d18O were linked to ephemeral growth and decay in Antarctic Ice Sheets. A great deal of effort in the palaeoceanography community has been focused on developing techniques and gathering additional records to determine if the Antarctic Ice Sheet has behaved in such a transient manner in the past and indeed what factors might have led to the rapid growth and decay of ice sheets. Deciphering between temperature and ice-volume influences in the deep-sea isotopic record has proven somewhat difficult. Approaches have included the sampling of sediment from beneath different water masses, development of an independent palaeothermometer using magnesium/calcium ratios and improving the resolution and accuracy of coastal sea-level records. Despite these advances, it Corresponding author. Tel.: þ64 3 4797509; Fax: þ64 3 4797527;
E-mail:
[email protected] (G.S. Wilson).
370
G. S. Wilson et al.
is only through the recovery of Antarctic drill core records that we have been able to test the resulting hypotheses. Combined with numerical climate models, ice-volume estimates are also available. The Antarctic Oligocene–Miocene record is most complete in the Victoria Land Basin as recovered in the CIROS-1 and CRP-2A drill cores. The strata recovered in both drill cores are cyclic in nature and interpreted to represent periodic advance and retreat of ice across the Antarctic margin in the western Ross Sea concomitant with sea-level fall and rise, respectively. Environmental data suggest a significant Antarctic climate threshold across the Oligocene–Miocene boundary with cooler temperatures implied in the early Miocene with ice volume and palaeo-sea-level estimates suggesting a significant but transient growth in the Antarctic Ice Sheet to B25% larger than present. The Antarctic data are entirely consistent with the predictions from deepsea records including the suggestion that the glacial advance was relatively short lived and interglacial conditions were re-established within a few hundred thousand years. The duration and transience of the Mi1 glacial expansion and swift recovery in Antarctica likely resulted from the limited polar summer warmth from the coincidence of low eccentricity and low-amplitude variability in obliquity of the Earth’s orbit at the Oligocene–Miocene boundary. This was followed by warmer polar summers and increased melt from increased eccentricity and high-amplitude variability in obliquity in the early Miocene, allowing the recovery of vegetation on the craton. Atmospheric CO2 concentrations remained below a 2 pre-industrial threshold, which promoted sensitivity of the climate system to orbital forcing. While climate and ice-sheet modelling support the fundamental role of greenhouse gas forcing as a likely cause of events like Mi1, the models underestimate the range of orbitally paced ice-sheet variability recognised in early Miocene isotope and sea-level records unless accompanied by significant fluctuations in greenhouse gas concentrations. While tectonic influences may have been secondary, they may well have contributed to oceanic cooling recorded at the Cape Roberts Project site in the South Western Ross Sea.
9.1. Introduction The paucity of Cenozoic outcrop on the Antarctic craton has led to the reliance on proxy records (isotopic signatures in microfossils, deep-sea erosion events and former sea levels on distal continental margins) to help unravel the history of climate and ice sheets on Antarctica (Kennett and Shackleton, 1976; Kennett, 1977; Wright and Miller, 1993; Miller and Mountain, 1996; Zachos et al., 2001a; Miller et al., 2005). Much attention has been focused on the search for Antarctic data sets to ‘ground truth’ significant climate trends, events and thresholds observed in these proxy
The Oligocene–Miocene Boundary
371
records. Drilling around the Antarctic margin by seven legs of the Deep Sea Drilling Project (DSDP; Kennett et al., 1974; Hayes and Frakes, 1975) and Ocean Drilling Programme (ODP; Barker et al., 1988a, b; Ciesielski et al., 1988; Barron et al., 1989; Barker et al., 1999; O’Brien et al., 2001), and several sea-ice-based drilling projects (Barrett, 1986, 1989; Cape Roberts Science Team, 1998, 1999, 2000; Fig. 9.1) has recovered Cenozoic sequences, which have allowed the testing of interpretations of Antarctic Glacial history from proxy records and climate models. While the DSDP and ODP core recovery has been between 14 and 40%, riser drilling from sea ice in the South Western Ross Sea has enabled recovered of some high-quality intervals of the Cenozoic (95–98%) recovering prima facie documentation of climate and cryospheric changes in Antarctica. One interval that is particularly well-sampled and well-dated in several drill cores is the Oligocene–Miocene boundary, permitting an accurate comparison to deepsea high-resolution isotopic records from lower latitudes (Naish et al., 2001, 2008; Wilson et al., 2002; Roberts et al., 2003). This chapter reviews recent evidence for a glacial expansion in Antarctica coincident with the Oligocene–Miocene boundary and the Mi1 deep-sea oxygen isotope excursion. The climatic significance of the boundary has only recently become apparent from recalibration of the Oligocene–Miocene time scale using astrochronology (Zachos et al., 2001b). Consequently, age data and chronstratigraphy of the Oligocene–Miocene boundary and the Antarctic strata that contain the boundary are also reviewed. Data sets considered to indicate climate and ice-sheet variability across the boundary include benthic and planktic oxygen isotope (d18O) records (Kennett and Shackleton, 1976; Miller et al., 1991; Wright and Miller, 1992; Paul et al., 2000; Zachos et al., 2001b; Billups et al., 2002) and microfossil geochemistry (Billups and Schrag, 2002; Lear et al., 2004), sequence stratigraphic analyses of Antarctic (Fielding et al., 1997; Naish et al., 2001, 2008) and mid-latitude (Kominz and Pekar, 2001; Pekar et al., 2002; Miller et al., 2005; Pekar and DeConto, 2006; Pekar and Christie-Blick, 2008) continental margin strata, Antarctic palaeobotany and palynology (Askin and Raine, 2000; Barrett, 2007) and physical properties of Antarctic drill core strata including lithology, clay mineralogy, mudrock geochemistry and magnetic mineralogy (Verosub et al., 2000; Ehrmann et al., 2005; Passchier and Krissek, 2008). Finally, the cause of the Mi1 glaciation is considered. 9.1.1. Identification of the Oligocene–Miocene Boundary Early definition of the Oligocene–Miocene boundary relied on the identification of the last occurrence of the calcareous nannofossil Dictyococcites
372
G. S. Wilson et al.
° -30
929 CR
-60 °
522 0°
SO
UT
HA
ME
OCEAN
TI C AN L AT
RI
SOU AFR TH ICA
30
°
1090
CA
° 60
689 690 MR
BS
693 WS AP
KP PB
748
747
WAIS
50
°
°
°
0°
WL
270 RS
0°
C
°
°
12
FI CI PA
10
°
80
˚
30°
20
°
80
40
60
70
CIROS-1
OC E
90°
EAIS
OCEAN
DP SSI
-90°
I AN IND
KGI
CRP-2a VLB
AN
LIA RA ST AU
-12
RS
70
1218
°
0°
15
0°
-15
DSDP/ODP/Antarctic Drillsites Mid Ocean Spreading Ridge 180°
Figure 9.1: Polar stereographic projection showing the location of key Oligocene–Miocene boundary sections discussed in the text. Abbreviations of locations: numbers are DSDP/ODP drill sites; AP, Antarctic Peninsula; BS, Bransfield Strait; CIROS, Cenozoic Investigations in the Ross Sea Drilling Project; CR, Ceara Rise; CRP, Cape Roberts Drilling Project; DP, Drake Passage; EAIS, East Antarctic Ice Sheet; KGI, King George Island; KP, Kerguelen Plateau; MR, Maud Rise; PB, Prydz Bay; RS, Ross Sea; SSI, South Shetland Islands; VLB, Victoria Land Basin; WAIS, West Antarctic Ice Sheet; WS, Weddell Sea.
25 C6Cn.3n
24
25 C6Cn.2r
23.74 Ma C6Cn.1r
C6Cn.2n
24 C6Cn.3n
24.6 Ma
C7n.1n
25 23
24
C6Cn.2r
C6Br
23
23.8 Ma
C7n.1n
OLIGOCENE
C6Cr
C6Cn.1r C6Cn.1n
C6Br
C6Br C6Bn
Ness et al. (1980)
MIOCENE
C6Cn.1n
C6Br
Billups et al. (2004)
Age (Ma)
Cande & Kent (1992, 1995)
Age (Ma)
Berggren et al. (1985)
Age (Ma)
C6Cn.1r C6Cn.1n
23 C6Cn.1r C6Cn.2n 23.033 Ma
C6Cn.3n
Figure 9.2: History of the calibration of the Geomagnetic Polarity Time Scale in the vicinity of the Oligocene– Miocene boundary. Boundary position in Ness et al. (1980) and Berggren et al. (1985) is fixed to the last occurrence of the nannofossil Reticulofenestra bisectus and in Cande and Kent (1992, 1995), Gradstein et al. (2004) and Billups et al. (2004) is fixed to the base of magnetic polarity subchron C6Cn.2n. Ness et al. (1980), Berggren et al. (1985), Cande and Kent (1992, 1995) are calibrated by various mid-ocean ridge spreading rate models. Billups et al. (2004) and Gradstein et al. (2004) are calibrated by astronomical tuning at ODP Site 1090.
C6Cr
24 C6Cn.1n
C6Cn.2n
C6Cn.3n C6Cn.2r C6Cn.2n
C6Cn.2r
23
C6Cr
C6Cr
Age (Ma)
The Oligocene–Miocene Boundary 373
374
G. S. Wilson et al.
bisectus (23.7 Ma; Berggren et al., 1985). However, this has proved problematic in the colder waters, coarser sediments and hiatus prone strata of the Antarctic and Southern Ocean. The reassignment of the boundary by Cande and Kent (1992, 1995) to the slightly older base of Magnetic Polarity subchron C6Cn.2n (23.8 Ma; Fig. 9.2) has made its identification more straightforward in Antarctica and the Southern Ocean but only in relatively complete and continuous stratigraphic successions (Wilson et al., 2002; Roberts et al., 2003). More recently, the recognition of astronomically influenced cyclical physical properties and d18O records in continuously deposited deep successions has enabled astronomical calibration of late Oligocene through early Miocene time. The astronomical calibration suggested that, while still coincident with the base of subchron C6Cn.2n, the boundary was in fact nearly a million years younger (22.970.1 Ma, Shackleton et al., 2000, Pa¨like et al., 2004; 23.03 Ma, Billups et al., 2004, Gradstein et al., 2004; Fig. 9.2). The climatic significance of this was outlined by Zachos et al. (2001b) and Pa¨like et al. (2006) who recognised the coincidence of the Oligocene–Miocene boundary and the Mi1 isotope excursion with an unusual coincidence of low eccentricity and low-amplitude variability in obliquity of the Earth’s orbit (Fig. 9.3). This would have placed the Earth in a sustained period of unusually low seasonality (cold summers), which Zachos et al. (2001b) claimed would have limited polar summer warmth and encouraged ice growth at the poles. Equally, within a few hundred thousand years, the coincidence of increased eccentricity and highamplitude variability in obliquity would have resulted in warmer polar summers and increased summer melt.
9.2. Proxy Records 9.2.1. The Isotopic Record Oxygen isotope ratios (d18O) in foraminiferal tests from deep-sea sedimentary records have long been recognised to represent the Cenozoic climatic (temperature, sea level and ice volume) history of the Earth (e.g. Shackleton et al., 1977 and references therein). However, deciphering the climatic history of Antarctica from d18O values alone in deep-sea records has always proven difficult due to the ambiguity of influence on the signal from the volume of ice on land versus isotopic fractionation, which is related to the water temperature during the precipitation of calcite (Miller et al., 1991). Early studies attempted to separate the two influences by focusing their analyses on
The Oligocene–Miocene Boundary
Eccentricity 0
20
0.02
0.04
375
Mg Temperature (˚C) 1
0.06
2
3
4
5
6
7
21
21
22
22
23
23
Age (Ma)
Age (Ma)
Mi-1
24
24
25
25
26
26
27 0.38
0.39
0.4
0.41
0.42
Obliquity Target
0.43
2.5
2
1.5
1
18
O, PDB (˚/ oo)
Figure 9.3: Oxygen isotope and Magnesium temperature data across the Oligocene–Miocene boundary interval (data from Lear et al., 2004). Obliquity and eccentricity orbital target data from Laskar et al. (2004). All data are plotted against the astronomical time scale presented by Billups et al. (2004).
376
G. S. Wilson et al.
foraminiferal species, such as benthic forms, known to live in water masses thought to be relatively stable in their temperature history (Shackleton and Kennett, 1975; Kennett, 1977), hence deducing that any shorter-term fluctuation was due to ice volume rather than temperature fractionation. While these studies recognised major threshold changes in the climatic deterioration of Antarctica, assuming no Northern Hemisphere ice sheets at the time, their resolution was limited and the record incomplete across the Oligocene–Miocene boundary. In a higher-resolution compilation of Oligocene–Miocene d18O records, Miller et al. (1991) recognised a relatively transient ca. 2 myr period and 1 m amplitude cyclicity in Miocene isotope values (isotope events Mi1-Mi7). The 1 m shifts, they suggested were of the same order as the threshold shifts identified by earlier studies (Kennett and Shackleton, 1976) and represent similar volumes of ice accumulation on the Antarctic craton. The most significant of these shorter order isotopic events, Mi1, was coincident with the Oligocene–Miocene boundary (Fig. 9.3). Originally defined by Miller et al. (1991) from DSDP Site 522 (Figs. 9.1 and 9.3), the event has subsequently been confirmed at numerous locations and the timing, duration and magnitude refined (Zachos et al., 1997; Paul et al., 2000; Zachos et al., 2001a, b; Billups et al., 2002, 2004, Billups and Schrag, 2002). For at least 2 myr prior to the Oligocene–Miocene boundary, d18O values were relatively stable and of low-amplitude variability (o0.5 m; Paul et al., 2000). In contrast, the Mi1 event represents a dramatic B1 m increase in d18O, over a 250 ky period immediately prior to the Oligocene–Miocene boundary peaking coincident with the boundary (Billups et al., 2004; Fig. 9.3). Peak values persisted for only B20 ky before returning to similarly low amplitude but slightly increased late Oligocene mean d18O values over the first B120 ky of the early Miocene (Paul et al., 2000). The covariance of the isotope signal in both benthic and planktic species in the late Oligocene at Equatorial Atlantic ODP Site 929 led Paul et al. (2000) to conclude that the variability was primarily ice volume driven. However, the Mi1 event, itself, is of relatively lower amplitude in planktic records which led Paul et al. (2000) to suggest that only 0.5 m is likely due to ice-volume effects, which, using the late Pleistocene calibration, represents growth of an ice sheet in Antarctica of similar proportion to the present-day East Antarctic Ice Sheet (EAIS). The remaining 0.5 m, they concluded was due to a 21C cooling of bottom waters at ODP Site 929 in the western Equatorial Atlantic. Another approach to deciphering temperature versus ice-volume components of the deep-sea d18O signal was employed by Lear et al. (2004) who determined palaeotemperature independently from Mg/Ca ratios in foraminifera tests across the Oligocene–Miocene boundary at ODP Site 1218 in
The Oligocene–Miocene Boundary
377
the eastern Equatorial Pacific (Figs. 9.1 and 9.3). The 2–31C of cooling predicted in the equatorial Atlantic by Paul et al. (2000) immediately prior to Oligocene–Miocene boundary was confirmed in the Equatorial Pacific by Lear et al. (2004). However, peak cooling preceded peak d18O values at ODP Site 1218 and a slight (11C) warming in bottom waters of the Equatorial Pacific is recorded immediately prior to and during peak Mi1 d18O values followed by another 21C of warming post-Mi1. A similar temperature variation was observed B800 ky prior to the Oligocene–Miocene boundary. Lear et al. (2004) concluded that the Mi1 event did indeed represent a significant increase in continental ice volume and, given the warming influence predicted by Mg/Ca ratios, perhaps a larger ice-volume growth than that predicted by Paul et al. (2000). Billups and Schrag (2002) also suggested that the d18O record from ODP Site 747 represented an ice-volume signal because paired Mg/Ca measurements suggested little change in ocean temperature through the early Miocene. Recent work, however, suggests caution when interpreting stable intervals in Mg/Ca ratios from deep-water sites due to potential saturation of carbonate, which might affect the partitioning of Magnesium into benthic foraminifera (Elderfield et al., 2006; Lear et al., 2008). 9.2.2. Palaeo Sea Levels Sequence and seismic stratigraphy has provided a means of relating the geologic record of continental margins to global sea-level changes that are often related to ice-volume changes at high latitudes (Vail et al., 1977; Haq et al., 1987). Sea-level history is deduced by the recognition of unconformitybounded units (i.e. depositional sequences) deposited in response to a cycle of falling and rising sea level. However, determining the relative role of tectonic subsidence and uplift versus rising and falling global sea-level and hence ice-volume fluctuations (glacioeustasy) is still unresolved, particularly in pre-Pleistocene records (e.g. Macdonald, 1991). Vail et al. (1977) and Haq et al. (1987) attempted to extract the glacioeustatic signal from the comparison of records from several continental margins (Fig. 9.4). However, the large amplitudes of sea level/ice volume, the limited resolution of the resulting record and the use of proprietary data to create the sea-level records, has spurred the scientific community to collect independent data to test the records of Vail et al. (1977) and Haq et al. (1987). The Oligocene– Miocene interval of these records has sparked particular interest. Both Vail et al. (1977) and Haq et al. (1987) predicted sea-level rises and falls of between B50 and 100 m in the late Oligocene. Vail et al. (1977), however,
378
G. S. Wilson et al.
Miller et al. (2005) Sea Level (m) -50 0 +50
Age (Ma)
20
MIOCENE OLIGOCENE
25
0
50 100 150 Sea Level (m) Haq et al. (1987)
Figure 9.4: Eustatic sea-level curves derived from coastal onlap patterns for the Oligocene–Miocene boundary interval from Haq et al. (1987) and Miller et al. (2005). Calibration of Miller et al. (2005) curve is considered more realistic. Curves are adjusted to astronomical time scale of Billups et al. (2004). predicted a fall of some 60 m across the Oligocene–Miocene boundary, whereas Haq et al. (1987) predicted a rise of B100 m across the Oligocene– Miocene boundary. The proprietary nature of much of the data has precluded resolving this conundrum from the same data set. An alternative passive margin stratigraphic data set is available from the New Jersey/New York Bight region of North America. Sea-level changes predicted from sequence stratigraphic analysis have recently been calibrated from coring as part of the Ocean Drilling Programme Legs 150X and 174AX (Miller and Mountain, 1996; Miller et al., 1997a, b). The glacioeustatic contribution to sea-level changes in the Oligocene and earliest Miocene was estimated by combining two-dimensional palaeoslope modelling of the foraminiferal biofacies and lithofacies with two-dimensional flexural backstripping of the margin (Kominz and Pekar, 2001; Pekar and Kominz, 2001). The depth ranges of foraminiferal biofacies were determined from a combination of standard factor analysis techniques and the backstripped geometries. The geometry of the margin through time was determined
The Oligocene–Miocene Boundary
379
using two-dimensional flexural backstripping. Foraminiferal biofacies and lithofacies were then used to constrain the depths of the Oligocene margin profiles obtained from backstripping. A eustatic fall of B40715 or 56 m apparent sea level (apparent sea level is eustasy plus the effects of water loading on the margin, Pekar et al., 2002) was estimated across the Oligocene–Miocene boundary (Fig. 9.4), which is similar to the glacioeustasy predictions from oxygen isotopic and trace metal geochemical data (Paul et al., 2000; Lear et al., 2004). Pekar et al. (2006) provided estimates of Antarctic ice volume and the resulting changes in global sea-level for the late Oligocene by applying d18Oto-sea-level calibrations to deep-sea d18O records from a number of ODP Sites. Their results indicate that the size of the Antarctic Ice Sheet increased from approximately 50% of the present-day EAIS during the latest Oligocene to as much as 25% larger than the present-day EAIS at the Oligocene–Miocene boundary. Ice volume returned to near late Oligocene size in the early Miocene (Pekar and DeConto, 2006).
9.3. Records from the Antarctic Margin The Oligocene–Miocene boundary interval was first sampled in Antarctica at DSDP Site 270 in the Eastern Ross Sea (Fig. 9.1). Drilling recovered a succession of silty mudstones including glaciomarine sediment, which spans the Oligocene–Miocene boundary (Hayes and Frakes, 1975; Leckie and Webb, 1983). Although an abrupt lithological change with a potential hiatus at the Oligocene–Miocene boundary is noted by Leckie and Webb (1983), poor chronological resolution prevents unambiguous correlation with the Mi1 event and the earliest Miocene. The only exposed Oligocene–Miocene boundary strata reported from Antarctica crop out on King George Island (Fig. 9.1) and include the Destruction Bay Formation (Latest Oligocene) and Cape Melville Formation (earliest Miocene; Birkenmajer et al., 1985; Birkenmajer, 1987). In a recent summary of stratigraphy and facies of the succession, Troedson and Riding (2002) concluded that a significant glacial advance occurred at the boundary and that chronological control was good enough to suggest a correlation with the Mi1 event. The facies indicate a significant regional ice grounding event across Bransfield Strait and beyond the South Shetland Islands (Troedson and Riding, 2002). Unfortunately, drilling on the shelf and slope south of the South Shetland Islands (ODP Leg 178; Barker and Carmerlenghi, 2002) did not yield any more definitive records of the Oligocene–Miocene boundary.
380
G. S. Wilson et al.
Late Oligocene/early Miocene strata reported from the East Antarctic margin include Maud Rise (ODP Leg 113 sites 689 and 690; Barker et al., 1988a), the Weddell Sea margin (ODP Leg 113 Site 693; Barker et al., 1988a) and Kerguelen Plateau (ODP Leg 120 sites 747 and 748; Schlich and Wise, 1992; Fig. 9.1). The record at Maud Rise is relatively thin and comprises exclusively siliceous and carbonate ooze, although rare glacial drop stones are reported in strata from Site 689 (Barker et al., 1988a). At the Weddell Sea margin Oligocene–Miocene sediments are also fine grained and include diatom mud, clay and ooze (Barker & Carmerlenghi, 2002). Oligocene–Miocene boundary sediments at Kerguelen Plateau are also carbonate ooze (Schlich and Wise, 1992), however, foraminifera preservation and age resolution were good enough at ODP Site 747 to yield a benthic oxygen isotope stratigraphy across the boundary at that site (Wright and Miller, 1992; Billups and Schrag, 2002; Fig. 9.5) and the amplitude of the Mi1 event was much reduced compared to equatorial values, with a d18O shift of only 0.3 m across the Oligocene–Miocene boundary. A strontium isotope stratigraphy has also been assembled using planktic foraminifera from ODP Site 747 (Oslick et al., 1994). Oslick et al. (1994) reported significant increases in 87Sr/86Sr following the early Miocene Mi isotope events with a B1 myr lag. They suggested that this increase and similar subsequent stepwise increases in early–middle Miocene oceanic 87Sr resulted from changes in the glacial state of East Antarctica. Drilling in Prydz Bay (ODP Legs 119 and 188; Hambrey et al., 1991; Cooper and O’Brien, 2004) did not yield any Oligocene–Miocene age strata δ18O (per mil)
Mg/Ca (mmol/mol) 1
2
3 2.5
2.0
1.5
Age (Ma)
20 Mi1a
Mi1
25
Figure 9.5: Oxygen isotope and Mg/Ca data from the Oligocene–Miocene boundary interval in ODP Site 747 from Kerguelen Plateau. Data from Billups and Schrag (2002) adjusted to time scale of Billups et al. (2004).
The Oligocene–Miocene Boundary
381
and Hambrey et al. (1991) concluded that this was due to erosion beneath an expanded middle-Miocene ice sheet. 9.3.1. McMurdo Sound, South Western Ross Sea The Antarctic Oligocene–Miocene record is most complete in the Victoria Land Basin as recovered in the CIROS-1 and CRP-2A drill holes (Barrett, 1989, Cape Roberts Science Team, 1999; Figs. 9.1 and 9.6). As with Prydz Bay, much of the Oligocene record of the Victoria Land Basin is marked by significant hiatuses (Wilson et al., 1998, 2000a, b; Florindo et al., 2005), however, the latest Oligocene–early Miocene is preserved in both records (Naish et al., 2001; Wilson et al., 2002; Roberts et al., 2003). On the basis of radiometric, biostratigraphic and magnetostratigraphic data, Wilson et al. (2002) placed the Oligocene–Miocene boundary at 183.7 m in the CRP2A core at the base of a normal polarity interval correlated with Polarity Subchron C6Cn.2n using Berggren et al.’s (1995) time scale. However, following the astronomical revision of the late Oligocene through early Miocene time scale (Billups et al., 2004; Gradstein et al, 2004; Pa¨like et al., 2004), Naish et al. (2008) placed the boundary at 130.27 m in an unconformity in the CRP-2A core and revised the age of strata underlying the unconformity to encompass Polarity Chron C7n. Roberts et al. (2003) placed the Oligocene–Miocene boundary at 247 m in the CIROS-1 core B35 km south of CRP-2A. However, following the revision of Naish et al. (2008), the Oligocene–Miocene boundary in the CIROS-1 core more likely occurs in an unconformity at 92 m (Fig. 9.7). The boundary immediately overlies an unconformity at 248.71 m, which might represent as much as 1 myr following the age revision of Antarctic shelf diatom zones (Scherer et al., 2000) implied by Naish et al. (2008). The strata recovered in both the CRP-2A and CIROS-1 drill holes are cyclic in nature and interpreted to represent periodic advance and retreat of ice across the Antarctic margin concomitant with sea-level fall and rise, respectively (Fielding et al., 1997; Naish et al., 2001). Each sequence is organised into a vertical succession, which begins with an erosion surface and is followed by a diamictite and sandstone, which gives way to sparsely fossiliferous bioturbated mudstone representing a cycle of glacial advance and retreat followed by open water conditions across the site of deposition (Naish et al., 2001) in concert with changes in relative sea-level (Dunbar et al., 2008; Fig. 9.6). Naish et al. (2008) estimated the glacioeustatic influence on relative water depth changes by deconvolving the tectonic, isostatic and palaeobathymetric components of water depth. These results are consistent with the d18O
Figure 9.6: Environmental proxy data for the upper part of the CRP-2A drill core. Grain size and clast data, and sequence stratigraphic, palaeobathymetric, depositional environment and ice margin interpretations are from Cape Roberts Science Team (1999). CIA, chemical index of alteration (data from Passchier and Krissek, 2008). ARM, anhysteretic remanent magnetization (data from Verosub et al., 2000). Temperature and meltwater indicators are discussed in Barrett (2007). Nothofagus leaf in CIROS-1 was identified at 215.5 mbsf immediately underlying the Oligocene–Miocene boundary as identified from the revised age model presented in this paper (Fig. 9.7).
382 G. S. Wilson et al.
Polarity
Antarctic shelf diatom zones
C6Cn.1r
C6Cn.1n
22.8
22.6
10
11
87Sr/ 86Sr
9
C6Cn.3n
23.4
C6Cn.2r
23.2
LOL. ornata
24.0 C7n.1n C7n.1r
24.2 C7n.2n
23.8
40Ar/ 39Ar
87Sr/ 86Sr
40Ar/ 39Ar
87Sr / 86Sr
87Sr/ 86Sr
Oligocene-Miocene boundary
350
340
330
320
310
300
290
280
270
260
250
240
230
220
210
200
190
180
170
160
150
140
130
120
110
100
90
80
Depth Lithology (mbsf)
Polarity
CIROS-1 Antarctic shelf diatom zones
C6Cn.2r
C6Cn.1r
C6Cn.1n
22.8
22.6 Nothofagus Leaf (Hill, 1989)
24.2
23.8
23.4
Age (Ma)
?
Oligocene-Miocene boundary
C7n.2n
23.0 C6Cn.2n
?
23.6 C6Cr
24.0 C7n.1n C7n.1r
23.0
C6Cn.2n
LOD. bisectus
23.6 C6Cr
Age (Ma)
Uppe r Ol i gocene
C6Cn.3n
23.2
Upper O l i g o c e ne
Figure 9.7: Revised age models for the CRP-2A and CIROS-1 drill cores from McMurdo Sound using the astronomical time scale of Billups et al. (2004) following the arguments in Naish et al. (2008). Magnetic polarity and diatom biostratigraphy data are from Wilson et al. (2000a, b, 2002) and Roberts et al. (2003). CIROS-1 age model is constrained by diatom biostratigraphic zones with revised ages from Naish et al. (2008).
300
290
280
270
260
250
240
230
220
210
200
190
180
170
160
150
140
130
Depth Lithology (mbsf)
Kisseleviella sp.A
Pterotheca reticulata
Hemiaulus sp.A
? Kisseleviella sp.A P. ret. Hemiaulus sp.A
CRP-2
The Oligocene–Miocene Boundary 383
384
G. S. Wilson et al.
sea-level calibration of Pekar et al. (2002) from the New Jersey margin. Each cycle is interpreted to represent between 10 and 40 m of eustatic variation in the late Oligocene with perhaps 50 m of sea-level fall concomitant with an ice sheet some 20% larger than present coincident with the unconformity, which is correlated with the Oligocene–Miocene boundary by Naish et al. (2008). Grounding of ice across the site at the Oligocene–Miocene boundary is also confirmed by macro- and micro-structures indicative of glacio-tectonic deformation (Passchier, 2000; van der Meer, 2000). The Oligocene–Miocene boundary is marked by some significant changes in physical properties in the CRP-2A core. Late Oligocene sedimentary cycles underlying the boundary are 55–60 m thick and relatively complete, whereas early Miocene sedimentary cycles are 10–20 m thick and truncated (Cape Roberts Science Team, 1999; Fig. 9.6). The clay mineralogy of the strata across the Oligocene–Miocene boundary in CRP-2A records stable physical weathering conditions (Ehrmann et al., 2005). Major element ratios derived from mudrock geochemistry for the same strata show significant shifts in the chemical index of alteration (CIA) across the Oligocene– Miocene boundary, which indicate periods of increased physical weathering and mechanical erosion associated with glacial advance (Passchier and Krissek, 2008). The CIA data reported by Passchier and Krissek (2008; Fig. 9.6) was corrected for the presence of primary volcanic detritus in order to reflect the palaeoclimatic record within the mudrock geochemistry. Short-lived glacial events at B23, B21 and B19 Ma indicated by the CIA data are correlated with the Mi events by Passchier and Krissek (2008) and interpreted to represent significant climatic and ice-sheet events in East Antarctica. Magnetic properties also show a marked change across the Oligocene–Miocene boundary at 130.27 m in the CRP-2A core (Verosub et al., 2000). An earlier change in magnetic properties at 270 m (late Oligocene) is attributed to inception of the McMurdo Volcanic Province (Verosub et al., 2000). Despite these changes in physical properties across the Oligocene– Miocene boundary in the CRP-2A core, palynological data although sparse due to low concentrations of organic matter, indicate a partially open landscape dominated by small Nothofagus (Southern Beech) stands or sparse tundra vegetation persisting through the late Oligocene–early Miocene (Askin and Raine, 2000). Oligocene–Miocene strata in the CIROS-1 core also contain similar amounts of pollen (Mildenhall, 1989) and a Nothofagus leaf fossil was preserved in latest Oligocene strata of the CIROS-1 Core (Hill, 1989; Figure 9.7). Mean summer temperature records derived from the (KþNa)/Al ratios of the CRP cores (Passchier and Krissek, 2008) indicate relatively constant mean summer
The Oligocene–Miocene Boundary
385
temperatures of B101C in the latest Oligocene dropping to B61C in the early Miocene. Marine palynomorphs, however, suggest a more significant change following the Oligocene–Miocene boundary with a significant reduction in the occurrence of prasinophyte algae, which is taken to indicate a reduction in offshore meltwater influence and hence cooler climates (Barrett, 2007).
9.4. Possible Drivers of Change Across the Oligocene–Miocene Boundary 9.4.1. Atmospheric Carbon Dioxide While there has been no specific simulation of the influence of atmospheric CO2 on ice-sheet growth at the Oligocene–Miocene boundary, DeConto and Pollard (2003a) demonstrated the potential link through simulations across the Eocene–Oligocene boundary. Their modelling demonstrated that icesheet inception occurred below a threshold of 3 pre-industrial atmospheric CO2 levels. Model results also demonstrated a strong response of ice volume to orbital forcing as atmospheric CO2 approaches the glaciation threshold, and decreasing orbital variability of an established ice sheet as CO2 approaches pre-industrial levels. Despite a predicted rewarming to preOligocene–Miocene boundary levels in the early Miocene and a major Antarctic glaciation in the middle Miocene (Zachos et al., 2001a), this was not matched by a parallel changes in levels of atmospheric CO2 as determined by geochemical proxies (Pagani et al., 1999, 2005; Pearson and Palmer, 2000; Fig. 9.8). The apparent decoupling between Miocene temperatures and atmospheric CO2 levels led Pagani et al. (1999, 2005) to conclude that, despite a significant decrease in atmospheric CO2 from B500 ppmv in the late Oligocene to new near-modern values at the Oligocene–Miocene boundary, changing atmospheric CO2 levels may have been secondary in driving Miocene Antarctic climatic and ice-sheet evolution.
9.4.2. Ocean Circulation/Tectonic Isolation The progressive opening of oceanic gateways (Fig. 9.1) and progressive tectonic isolation during the Cenozoic stages of Gondwana breakup have been indicated as critical threshold events in the climatic deterioration and inception of ice sheets since the first deep-sea oxygen isotopic records were
386
G. S. Wilson et al.
pCO2 (ppmv)
Eccentricity 0
20
0.02
0.04
500
0
0.06
1000
21
Miocene
22
23
Oligocene
Age (Ma)
Mi-1
24
25
26
0.38 0.39
0.4
0.41 0.42 0.43
Obliquity Target
3
2.5
2
1.5
1
18O, SMOW (°/ )
°°
Figure 9.8: Potential controls on climate and ice sheet behaviour across the Oligocene–Miocene boundary and origin of the Mi-1 d18O event. Obliquity and eccentricity orbital target data from Laskar et al. (2004). d18O data are from Zachos et al. (2001b) and pCO2 data are from Pagani et al. (2005). All data are plotted against the astronomical time scale presented by Billups et al. (2004). See text for discussion.
The Oligocene–Miocene Boundary
387
recovered (Shackleton and Kennett, 1975; Kennett, 1977). Testing of this hypothesis has proven particularly difficult due to uncertainties in the timing of gateway opening and inception of deep-water circulation. While estimates for the timing of the opening and deepening of the Tasmanian Gateway between Australia and Antarctica are reasonably well constrained to the Eocene–Oligocene boundary (Stickley et al., 2004), estimates for the opening of a deep Drake Passage range from the middle Eocene (B50 Ma) to the late Miocene (B6 Ma) (Barker and Burrell, 1977; Barker et al., 2007; Livermore et al., 2007). However, some estimates suggest that deep-water circulation through the Drake Passage may well have been coincident with the Oligocene– Miocene boundary (Barker and Burrell, 1977, 1982). Pekar et al. (2006) and Pekar and Christie-Blick (2008) suggested, however, that Southern Ocean water masses were still relatively poorly mixed in the late Oligocene–early Miocene and that individual records used to compile the composite Cenozoic oxygen isotope curve by Zachos et al. (2001a) were drawn from water masses with different temperature and salinity histories. Hence, introducing an artifact from the splice of different isotopic records (Pekar and Christie-Blick, 2008) previously interpreted to represent significant oceanic warming in the latest Oligocene (Zachos et al., 2001a). A number of numerical ocean modelling studies (Mikolajewicz et al., 1993; Nong et al., 2000; Toggweiler and Bjornsson, 2000; Sijp and England, 2004) have shown that the opening of a deep circum-Antarctic passage can cool the Southern Ocean by 1–31C. While the amount of cooling in these studies is somewhat dependent on modelling details associated with the treatment of the atmosphere (Huber and Nof, 2006), the effects of this range of cooling on continental climate and ice-sheet mass balance have been shown to be small relative to the effects of the falling Cenozoic CO2 concentrations (DeConto and Pollard, 2003a; Huber et al., 2004). For example, recent model simulations testing the importance of sea ice feedback on Antarctic Ice Sheets show that the continental interior is relatively insensitive to changes in Southern Ocean sea surface temperatures, and the effect of even large changes in ocean heat transport and sea ice is generally limited to the continental margins (DeConto et al., 2007). Conversely, the expansion of the EAIS, as presumed to have occurred at Mi1, has a dramatic effect on simulated Southern Ocean sea surface temperatures and sea ice distributions via the ice sheet’s influence on regional temperatures and low-level winds (DeConto et al., 2007; Fig. 9.9). As these simulations clearly show, a growing Mi1 ice sheet would have cooled Southern Ocean sea surface temperatures by several degrees, pushing the 0 1C isotherm equatorwards and increasing the area, thickness, and fractional cover of seasonal and perennial sea ice (DeConto et al., 2007). Furthermore, as the katabatic wind field increased in
a
c
b
°C
d
m
e
°C
f
m
Figure 9.9: South polar seasonal temperatures, sea ice and winds in response to a growing ice sheet as simulated by a GCM (DeConto et al., 2007). (a–c) Climatic conditions in a pre-Mi1 world with isolated ice caps in the continental interior (Top). (d–f) Climatic conditions with a fully glaciated east antarctica as presumed to have existed at the time of Mi1 (bottom). With the exception of ice sheet geometry, boundary conditions are identical in both simulations including the same late Palaeogene palaeogeography, 2 pre-industrial CO2 (560 ppmv), and a relatively cold austral summer orbit conducive to Antarctic Ice Sheet growth. Ice sheet geometries are taken from prior GCM-ice sheet simulations of Antarctic glaciation (DeConto and Pollard, 2003a). Austral Summer (December, January, February) and winter (June, July, August) seasonal climatologies are shown on the top and bottom of (a–f), respectively. (a, d) Seasonal surface (2 m) air temperature, (b, e) seasonal sea ice extent and thickness in metres and (c, f) lowest level (sigma level ¼ 0.189) GCM winds with vector scale length equivalent to 21C per m1 of wind velocity.
The Oligocene–Miocene Boundary
389
intensity, the enhanced polar easterlies and westerlies would have increased ocean frontal divergence and upwelling, with possible implications for the marine carbon cycle and CO2 drawdown (DeConto et al., 2007). Such mechanisms have been implicated as important contributors to the dynamics of Quaternary glacial cycles (Stephens and Keeling, 2000; Archer et al., 2003), but they have yet to be considered in a Miocene context. While discussions of tectonic influences on Antarctic climate evolution usually focus on Southern Ocean gateways, Miocene ice sheets could have also been sensitive to changes in tropical climate associated with Tethyan tectonics. As the Southern Ocean gateways were widening and deepening, the eastern Tethys was closing. Ocean modelling studies have shown that the progressive closure of the Tethys affected the location of deep-water formation and the thermohaline component of the meridional overturning circulation, ocean heat transport, and both tropical and high-latitude sea surface temperatures (Hotinski and Toggweiler, 2003; von der Heydt and Dijkstra, 2006). While the Antarctic interior appears to have been relatively insensitive to changes in the Southern Ocean, the modern Antarctic interior receives much of its moisture from the low mid-latitudes and significant changes in the tropics and associated teleconnections to polar latitudes could be important. Considering the timing of Tethyan closure relative to Antarctic Ice Sheet expansion in the Miocene, the sensitivity of ice-sheet evolution to low-latitude versus circum-Antarctic sea surface temperatures should be tested in future modelling studies. The perspective provided by numerical climate modelling suggests falling greenhouse gas concentrations around the time of the Oligocene–Miocene boundary (Pagani et al., 2005) had a greater impact on Antarctic climate than the direct, physical effects of ocean gateways. However, the indirect effects of the gateways, including their influence on the marine carbon cycle and atmospheric CO2 should also be considered. These indirect effects may be found to be more important to Cenozoic climate events like Mi1 than the direct influence of the gateways on ocean circulation and heat transport (Mikolajewicz et al., 1993; DeConto and Pollard, 2003a; Huber et al., 2004). 9.4.3. Orbital Parameters If varying orbital parameters were to result in ice growth on the Antarctic craton, this was most likely to have occurred when summer insolation is minimised either by the seasonal timing of aphelion (precession) during periods of relatively high eccentricity, periods of low eccentricity which minimised the
390
G. S. Wilson et al.
effects of precession, even when perihelion occurs during austral summer, or during periods of reduced amplitude in obliquity. All of these orbital configurations result in relatively cool summers and hence reduce the potential for summer melt. Zachos et al. (2001b) measured oxygen isotope ratios from Ocean Drilling Programme Equatorial Atlantic (Ceara Rise) Site 929 for a B5 myr interval spanning the Oligocene–Miocene boundary. The Mi1 event was found to correlate with both a minima in the low frequency (400 ky) eccentricity cycle and a prolonged minima in the amplitude of obliquity, a co-occurrence with a reoccurrence interval of 2.4 myr or longer (Pa¨like et al., 2004, 2006). This sustained period of unusually low seasonality (cold summers) was, however, relatively transient with warmer summers returning within a few hundred thousand years from the coincidence of increased eccentricity and high-amplitude variability in obliquity which would have resulted in warmer polar summers and increased summer melt (Zachos et al., 2001b). 9.4.4. Ice-Sheet Hysteresis Coupled climate–ice sheet models have been reasonably successful in simulating sudden Cenozoic glaciation events such as Oi1 and Mi1 (DeConto and Pollard, 2003a, b). For example, beginning with an ice-free continent and assuming gradually declining greenhouse gas concentrations and accounting for orbital forcing, DeConto and Pollard (2003a) simulated the sudden stepwise glaciation of East Antarctica within a 200-ky interval. The simulated ice sheet was comparable in volume to the modern EAIS, but significantly smaller than the volume of Mi1 ice reconstructed from the proxy isotope and sea-level records discussed above. Subsequent modelling work, including a representation of ice shelves not included in their earlier simulations (Pollard and DeConto, 2007), have shown that an Antarctic Ice Sheet B20–25% bigger than today would have required a glaciated West Antarctica, and ice grounding lines extending close to the continental shelf break around much of the margin. An Mi1 ice sheet of this size would have been similar in geometry to the ice sheet that existed at the Last Glacial Maximum (Huybrechts, 2002). However, the presumably warmer ocean at Mi1 might have been unconducive to the seaward migration of grounding lines, so this scenario maybe difficult to reconcile from a modelling perspective. Furthermore, the cold south polar conditions implied by such an Ice Sheet also implies global temperatures low enough to have allowed significant glaciation in the Northern Hemisphere, especially during orbital periods which produced cold boreal summers. While Greenland may have
The Oligocene–Miocene Boundary
391
contained some glacial ice as early as the Eocene (Eldrett et al., 2007), the Oligocene–Miocene boundary is B20 myr before the onset of the fist significant Northern Hemisphere glacial cycles. Clearly, some important problems remain in terms of reconciling the magnitude of the Mi1 event. While the rapid growth of Antarctic ice at Mi1 can be explained through a combination of decreasing greenhouse gas concentrations and orbital forcing (with other possible influences from mountain uplift and/or ocean circulation), the ephemeral nature of the event and subsequent variability of ice volume are also problematic from a modelling perspective (Pollard and DeConto, 2005). As shown in both simple and sophisticated numerical icesheet models, (Weertman, 1961; Huybrechts, 1994; Pollard and DeConto, 2005), the high albedo and elevation of large polar-centred ice sheets produce considerable hysteresis. In a scenario of cooling climate, a polar ice sheet can grow suddenly, once the snow line intersects sufficient land area in mountains and high plateau. The non-linear jump in ice volume is facilitated by height–mass balance and albedo feedbacks, as the ice sheet spreads horizontally (albedo feedback) and more of the parabolic ice surface rises above the snow line and out of the ablation zones around its margins (height–mass balance feedback) (Abe-Ouchi and Blatter, 1993; DeConto and Pollard, 2003a). The high elevation and albedo of the ice sheet inhibit the ice sheet from disappearing during subsequent warming interval, unless temperatures (snow lines) rise far above their initial values (elevation) at the time of glacial onset (Huybrechts, 1994). Pollard and DeConto (2005) studied this hysteresis effect in a coupled GCM–ice sheet model and in a simple flowline model with parameterised mass-balance forcing. They concluded that the hysteresis effect is strong enough to preclude orbital forcing from driving the range of Cenozoic ice-volume variability seen in the oxygen isotope and sea-level records described above, unless the orbital forcing is accompanied by significant changes in greenhouse concentrations. During favourable (cold austral summer) orbital periods, the atmospheric CO2-glaciation threshold for Antarctica is B2 pre-industrial levels, while CO2 must approach B4 pre-industrial levels during a warm austral summer orbital period to trigger the collapse of the interior EAIS. If the sensitivity of the models to orbital and greenhouse gas forcing is reasonable, the short duration of the peak Mi1 event would require a significant perturbation to the carbon cycle, producing significant global warming soon after the peak glacial interval. Greenhouse gas variability of this magnitude is not evident in existing proxy reconstructions of early Miocene CO2 (e.g. Pagani et al., 2005), however, higher-resolution records will be required to resolve this type of CO2 variability across key climatic events like the Oligocene–Miocene boundary.
392
G. S. Wilson et al.
9.5. Summary and Conclusions Strata recovered from the Antarctic margin indicate a significant glacial advance at the Oligocene–Miocene boundary reaching the south Shetland Islands on the Antarctic Peninsula (Troedson and Riding, 2002) and grounding in Prydz Bay and the South Western Ross Sea as indicated by hiatuses in drill cores (Hambrey et al., 1991; Roberts et al., 2003; Naish et al., 2008). Ice rafted as far north as Maud Rise (Barker et al., 1988a, b) and the central Ross Sea (Leckie and Webb, 1983) but did not appear to reach the Kerguelen Plateau (Schlich and Wise, 1992). Pre-Oligocene–Miocene boundary strata indicate a late Oligocene Antarctic Ice Sheet (Cape Roberts Science Team, 1999), which expanded to an ice volume of the order of 20% greater than the present ice sheet at the Oligocene–Miocene boundary (Naish et al., 2008). The glacial expansion, however, although significant in extent and volume, must have been relatively transient and neither cold nor extensive enough to extinguish Nothofagus tundra vegetation (Askin and Raine, 2000; Roberts et al., 2003), which persisted across the boundary despite a slight drop in temperature (Passchier and Krissek, 2008). Marine palynomorphs, however, indicate that coastal temperatures did not return to the warmth of the late Oligocene with a much reduced freshwater melt input to coastal regions (Barrett, 2007). Data from the Antarctic continent are entirely consistent with the shortlived (200 ky) ice-volume increase from 40% of present Antarctic ice volume to 25% greater than present Antarctic ice volume across the Oligocene– Miocene boundary with concomitant oceanic deep-water cooling implied by the Mi1 isotopic excursion recognised in equatorial and Southern Hemisphere deep-sea sedimentary records (Paul et al., 2000). Ice-volume estimates are confirmed by the backstripped stratigraphic records from the New Jersey Margin (Kominz and Pekar, 2001; Pekar et al., 2002), however, accommodating this much ice on Antarctica when global temperatures were presumably warmer than today may prove difficult from a modelling perspective. Warm summer mean temperatures were re-established soon after the Oligocene–Miocene boundary, although a few degrees cooler than pre-Miocene summer mean temperatures. The duration and transience of the Mi1 glacial expansion and swift recovery in Antarctica likely resulted from the limited polar summer warmth from coincidence of low eccentricity and low-amplitude variability in obliquity of the Earth’s orbit at the Oligocene– Miocene boundary (Zachos et al., 2001b). This was followed by warmer polar summers and increased melt from increased eccentricity and highamplitude variability in obliquity in the early Miocene, allowing the recovery of vegetation on the Antarctic craton. Atmospheric CO2 concentrations
The Oligocene–Miocene Boundary
393
remained below the 2 pre-industrial threshold, which promoted sensitivity of the climate system to orbital forcing during cold Austral summers. While climate and ice-sheet modelling supports the fundamental role of greenhouse gas forcing punctuated by orbital forcing as a likely cause of events like Mi1 (DeConto and Pollard, 2003a; Huber et al., 2004; DeConto et al., 2007; Pa¨like et al., 2006), the models underestimate the range of orbitally paced ice-sheet variability recognised in early Miocene isotope and sea-level records unless accompanied by significant fluctuations in greenhouse gas concentrations (Pollard and DeConto, 2005). While, tectonic influence may have been secondary, they may well have contributed to oceanic cooling recorded at the Cape Roberts Project site in the South Western Ross Sea (Barrett, 2007).
REFERENCES Abe-Ouchi, A., & Blatter, H. (1993). On the initiation of ice sheets. Ann. Glaciaol., 18, 203–207. Archer, D. E., Martin, P. A., Milovitch, J., Broykin, V., Plattner, G. K., & Ashendel, C. (2003). Model sensitivity in the effect of Antarctic sea ice and stratification on atmospheric pCO2. Paleoceanography, 18(1), 1012, doi:10.1029/ 2002/PA000760. Askin, R. A., & Raine, J. I. (2000). Oligocene and early Miocene terrestrial palynology of the Cape Roberts Drillhole CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 493–501. Barker, P. F., & Burrell, J. (1977). The opening of the Drake Passage. Mar. Geol., 25, 15–34. Barker, P. F., & Burrell, J. (1982). The influence upon Southern Ocean circulation, sedimentation and climate change of the opening of Drake Passage. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, pp. 377–385. Barker, P. F., & Carmerlenghi, A. (2002). Glacial history of the Antarctic Peninsula from Pacific Margin sediments. Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 178, doi:10.2973/odp.proc.sr.178.238.2002. Barker, P. F., Carmerlenghi, A., Acton, G. D., et al. (1999). Proceedings of the Ocean Drilling Program, Initial Reports. College Station, TX (Ocean Drilling Program), Vol. 178, [online]. Barker, P. F., Filippelli, G. M., Florindo, F., Martin, E. E., & Scher, H. D. (2007). Onset and role of the Antarctic Circumpolar Current. Deep Sea Res. II, 54, 2388–2398. Barker, P. F., Kennett, J. P., O’Connell, S., et al. (1988a). Proceedings of the Ocean Drilling Program, Initial Reports. College Station, TX (Ocean Drilling Program), Vol. 113, 774 pp.
394
G. S. Wilson et al.
Barker, P. F., Kennett, J. P., & Scientific Party (1988b). Weddell sea palaeoceanography: Preliminary results of ODP Leg 113. Palaeogeogr. Palaeoclimatol. Palaeoecol., 67, 75–102. Barrett, P. J. (Ed.) (1986). Antarctic Cenozoic history from the MSSTS-1 drillhole, McMurdo Sound. DSIR Bulletin. DSIR Publishing, Wellington, Vol. 237, 174 pp. Barrett, P. J. (Ed.) (1989). Antarctic Cenozoic history from the CIROS-1 drillhole, McMurdo Sound. DSIR Bulletin. DSIR Publishing, Wellington, Vol. 245, 254 pp. Barrett, P. J. (2007). Cenozoic climate and sea level history from glacimarine strata off the Victoria Land coast, Cape Roberts Project, Antarctica. In: M. J. Hambrey, P. Christoffersen, N. Glasser, & B. Hubbard (Eds). Glacial Sedimentary Processes and Products. International Association of Sedimentologists, Special Publication. Blackwell Scientific, Oxford, UK, Vol. 39 (in press). Barron, J., Larsen, B., Baldauf, J., et al. (1989). Proceedings of the Ocean Drilling Programme, Initial Reports. College Station, TX (Ocean Drilling Program), Vol. 119, 942 pp. Berggren, W. A., Kent, D. V., Flynn, J., & Van Couvering, J. A. (1985). Cenozoic geochronology. Geol. Soc. Am. Bull., 96, 1407–1418. Berggren, W. A., Kent, D. V., Swisher, C. C. III, & Aubrey, M. P. (1995). A revised Cenozoic geochronology and chronostratigraphy. Soc. Sediment. Geol. (Special Publication), 54, 129–212. Billups, K., Channell, J. E. T., & Zachos, J. C. (2002). Late Oligocene to early Miocene geochronology and paleoceanography from the Subantarctic South Atlantic. Paleoceanography, 17, doi:10.1029/2000PA000568. Billups, K., Pa¨like, H., Channell, J. E. T., Zachos, J. C., & Shackleton, N. J. (2004). Astonomic calibration of the late Oligocene through early Miocene geomagnetic polarity time scale. Earth Planet. Sci. Lett., 224, 33–44. Billups, K., & Schrag, D. P. (2002). Paleotemperatures and ice volume of the past 27 myr revisted with paired Mg/Ca and 18O/16O measurements on benthic foraminifera. Paleoceanography, 17, doi:10.1029/2000PA000567. Birkenmajer, K. (1987). Oligocene-Miocene glacio-marine sequences of King George Island (South Shetland Islands), Antarctica. Paleontol. Pol., 49, 113–122. Birkenmajer, K., Gazdzicki, A., Kreuzer, H., & Muller, P. (1985). K-Ar dating of the melville glaciation (early Miocene) in West Antarctica. Polish Acad. Sci. Bull. Earth Sci., 33, 15–23. Cande, S. C., & Kent, D. V. (1995). Revised calibration of the geomagnetic polarity timescale for the late Cretaceous and Cenozoic. J. Geophys. Res., 100, 6093–6095. Cande, S. C., & Kent, D. V. (1992). A new geomagnetic polarity timescale for the late Cretaceous and Cenozoic. J. Geophys. Res., 97, 13917–13951. Cape Roberts Science Team (1998). Studies from the Cape Roberts Project, Ross Sea, Antarctica – Initial report on CRP-1. Terra Antartica, 5, 1–187. Cape Roberts Science Team (1999). Studies from the Cape Roberts Project, Ross Sea, Antarctica – Initial report on CRP-2/2A. Terra Antartica, 6, 1–173.
The Oligocene–Miocene Boundary
395
Cape Roberts Science Team (2000). Studies from the Cape Roberts Project, Ross Sea, Antarctica – Initial report on CRP-3. Terra Antartica, 7, 1–209. Ciesielski, P. F., Kristoffersen, Y., Clement, B., et al. (1988). Proceedings of the Ocean Drilling Program, Initial Reports. College Station, TX (Ocean Drilling Program), Vol. 114, 815 pp. Cooper, A. K., & O’Brien, P. E. (2004). Leg 188 synthesis: Transitions in the glacial history of the Prydz Bay region, East Antarctica, from ODP drilling. Proceedings of the Ocean Drilling Programme, Scientific Results, Vol. 188, doi:10.2973/ odp.proc.sr.188.001.2004. DeConto, R. M., & Pollard, D. (2003a). Rapid Cenozoic glaciation of Antarctica induced By declining atmospheric CO2. Nature, 421, 245–249. DeConto, R. M., & Pollard, D. (2003b). A coupled climate-ice sheet modeling approach to the early Cenozoic history of the Antarctic Ice Sheet. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 39–53. DeConto, R. M., Pollard, D., & Harwood, D. (2007). Sea ice feedback and Cenozoic evolution of Antarctic climate and ice sheets. Paleoceanography, 22, PA3214, doi:10.1029/2006PA001350. Dunbar, G. B., Naish, T., Powell, R., & Barrett, P. J. (2008). Contraining the amplitude of late Oligocene bathymetric changes in western Ross Sea during orbitally-induced oscillations in the East Antarctic Ice Sheet: (1) Implications for glacimarine sequence stratigraphic models. Palaeogeogr. Palaeoclimatol. Palaeoecol., 258, 50–65. Ehrmann, W. U., Setti, M., & Marinoni, L. (2005). Clay minerals in Cenozoic sediments off Cape Roberts (McMurdo Sound, Antarctica) reveal palaeoclimatic history. Palaeogeogr. Palaeoclimatol. Palaeoecol., 229, 187–211. Elderfield, H., Yu, J., Anand, P., Kiefer, T., & Nyland, B. (2006). Calibrations for benthic foraminiferal Mg/Ca paleothermometry and the carbonate ion hypothesis. Earth Planet. Sci. Lett., 250, 633–649. Eldrett, J. S., Harding, I. C., Wilson, P. A., Butler, E., & Roberts, A. P. (2007). Continental ice in Greenland during the Eocene and Oligocene. Nature, 466, 176–179. Fielding, C. R., Woolfe, K. J., Purdon, R. G., Lavell, M., & Howe, J. (1997). Sedimentological and stratigraphical re-evaluation of the CIROS-1 Core, McMurdo Sound, Antarctica. Terra Antartica, 4, 149–160. Florindo, F., Wilson, G. S., Roberts, A. P., Sagnotti, L., & Verosub, K. L. (2005). Magnetostratigraphic chronology of a late Eocene to early Miocene glacimarine succession from the Victoria Land Basin, Ross Sea, Antarctica. Global Planetary Change, 45, 207–236. Gradstein, F. M., Ogg, J. G., & Smith, A.G. (Eds) (2004). A geological time scale. Cambridge University Press, Cambridge, 589 pp. Hambrey, M. J., Ehrmann, W. U., & Larsen, B. (1991). Cenozoic glacial record of the Prydz Bay Continental Shelf, East Antarctica. Proceedings of the Ocean Drilling Programme, Scientific Results, Vol. 120, pp. 77–132. Haq, B. U., Hardenbol, J., & Vail, P. R. (1987). Chronology of fluctuating sea levels since the Triassic. Science, 235, 1156–1167.
396
G. S. Wilson et al.
Hayes, D. E., Frakes, L. A., Barret, P. J., et al. (1975). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, 1017 pp. Hill, R. S. (1989). Fossil Leaf. In: P. J. Barrett (Ed.). Antarctic Cenozonic History from the CIROS-1Drillhole, McMurdo Sound. DSIR Bull. DSIR Publishing, Wellington, Vol. 245, pp. 143–144. Hotinski, R. M., & Toggweiler, J. R. (2003). Impact of a Tethyan Circumglobal Passage on ocean heat transport and ‘‘equable’’ climates. Paleoceanography, 18(1), 1007, doi:10.1029/2001PA000730. Huber, M., Brinkhuis, H., Stickley, C. E., Doos, K., Sluijs, A., Warner, J., Schellenberg, S. A., & Williams, G. L. (2004). Eocene circulation of the Southern Ocean: Was Antarctica kept warm by subtropical waters? Paleoceanography, 19(PA4026), doi:10.1029/2004PA001014. Huber, M., & Nof, D. (2006). The ocean circulation in the Southern Hemisphere and its climatic impacts in the Eocene. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 9–28. Huybrechts, P. (1994). Formation and disintegration of the Antarctic Ice Sheet. Ann. Glaciol., 20, 336–340. Huybrechts, P. (2002). Sea-level changes at the LGM from ice-dynamic reconstructions of the Greenland and Antarctic Ice Sheets during the glacial cycles. Quaternary Sci. Rev., 21, 203–231. Kennett, J. P. (1977). Cenozoic evolution of Antarctic glaciation, the circumAntarctic ocean, and their impact on global paleoceanography. J. Geophys. Res., 82, 3843–3860. Kennett, J. P., Houtz, R. E., Andrews, P. B., et al. (1974). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 29, 1197 pp. Kennett, J. P., & Shackleton, N. J. (1976). Oxygen isotopic evidence for the development of the psychrospheric 38 myr ago. Nature, 260, 513–515. Kominz, M. A., & Pekar, S. F. (2001). Oligocene eustasy from two-dimensional sequence stratigraphic backstripping. Geol. Soc. Am. Bull., 113, 291–304. Laskar, J., Gastineau, M., Joutel, F., Robutel, P., Levrard, B., & Correia, A. (2004). Long term evolution and chaotic diffusion of the insolation quantities of Mars. Icarus, 170, 343–364 . Lear, C. H., Bailey, T. R., Pearson, P. N., Coxall, H. K., & Rosenthal, Y. (2008). Cooling and ice growth across the Eocene-Oligocene transition. Geology, 36, 251–254. Lear, C. H., Rosenthal, Y., Coxall, H. K., & Wilson, P. A. (2004). Late Eocene to early Miocene ice sheet dynamics and the global carbon cycle. Paleoceanography, 19, doi:10.1029/2004PA001039. Leckie, R. M., & Webb, P.-N. (1983). Late Oligocene–early Miocene glacial record of the Ross Sea, Antarctica: Evidence from DSDP site 270. Geology, 11, 578–582. Livermore, R., C-D., H., Meredith, M., & Eagles, G. (2007). Drake Passage and Cenozoic climate: An open and shut case? Geochem. Geophys. Geosyst., 8, Q01005, doi:10.1029/2005GC001224.
The Oligocene–Miocene Boundary
397
MacDonald, D. I. M. (Ed.) (1991). Sedimentation, tectonics and eustasy – Sea-level changes at active margins. Special Publications of the International Association of Sedimentologists. Blackwell Scientific Publications, Oxford, UK, Vol. 12, 518 pp. Mikolajewicz, U., Maier-Reimer, E., Crowley, T. J., & Kim, K.-Y. (1993). Effect of Drake and Panamanian Gateways on the circulation of an ocean model. Paleoceanography, 8, 409–426. Mildenhall, D. (1989). Terrestrial palynology. In: P. J. Barrett (Ed.). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound. DSIR Bull. DSIR Publishing, Wellington, Vol. 245, pp. 119–127. Miller, K. G., Aubrey, M. P., Browning, J. V., Bukry, D., Burckle, L. D., Feigenson, M. D., Goss, M., Gwynn, D., Heibel, T., Kent, D. V., Liu, C., Mullikin, L., Pekar, S. F., Powars, D. S., Queen, D., Sugarman, P. J., & van Fossen, M. C. (1997a). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 150X, 388 pp. Miller, K. G., Mountain, G. S., & the Leg 150 Shipboard Party, and Members of the New Jersey Coastal Plain Drilling Project. Drilling and Dating New Jersey Oligocene-Miocene Sequences (1996). Ice volume, global sea level, and exxon records. Science, 271, 1092–1095. Miller, K. G., Sugarman, P. J., Browning, J. V., Aubrey, M. P., Brenner, G. J., Bukry, D., Burckle, L. D., Cramer, B. S., Curran, J., Dalton, R. F., Feigenson, M. D., Lawrence, R. P., Metzger, J. M., Olsson, R. K., Pekar, S. F., Reilly, T. J., Stewart, M., Uptegrove, J., & Wright, J. D. (1997b). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 174AX, 43 pp. Miller, K. G., Wright, J. D., & Browning, J. V. (2005). Visions of ice sheets in a greenhouse world. Mar. Geol., 217, 215–231. Miller, K. G., Wright, J. D., & Fairbanks, R. G. (1991). Unlocking the icehouse: Oligocene-Miocene oxygen isotope, eustasy, and margin erosion. J. Geophys. Res., 96, 6829–6848. Naish, T. R., Wilson, G. S., Dunbar, G., & Barrett, P. J. (2008). Constraining the amplitude of late Oligocene bathymetric changes in western Ross Sea during orbitally-induced oscillations in the East Antarctic Ice Sheet: (2) Implications for global sea-level changes. Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 66–76. Naish, T. R., Woolfe, K. J., Barrett, P. J., Wilson, G. S., Atkins, C., Bohaty, S., Bucker, C. J., Claps, M., Davey, F. J., Dunbar, G. B., Dunn, A. G., Fielding, C. R., Florindo, F., Hannah, M. J., Harwood, D. M., Henrys, S. A., Krissek, L., Lavelle, M., van der Meer, J., McIntosh, M. C., Niessen, F., Passchier, S., Powell, R. D., Roberts, A. P., Sagnotti, L., Scherer, R. P., Strong, C. P., Talarico, F., Verosub, K. L., Villa, G., Watkins, D. K., Webb, P. N., & Wonik, T. (2001). Orbitally induced oscillations in the East Antarctic Ice Sheet at the Oligocene/Miocene boundary. Nature, 413, 719–723. Ness, G., Levi, S., & Counch, R. (1980). Marine magnetic anomaly timescales for the Cenozoic and late Cretaceous: A pre´cis, critique and synthesis. Rev. Geophys. Space Phys., 18, 753–770.
398
G. S. Wilson et al.
Nong, G. T., Najjar, R. G., Seidov, D., & Peterson, W. (2000). Simulation of ocean temperature change due to the opening of Drake Passage. Geophys. Res. Lett., 27, 2689–2692. O’Brien, P. E., Cooper, A. K., Richter, C., et al. (2001). Proceedings of the Ocean Drilling Program, Initial Reports. College Station, TX (Ocean Drilling Program), Vol. 188, [Online]. Oslick, J., Miller, K. G., Feigenson, M. D., & Wright, J. D. (1994). OligoceneMiocene strontium isotopes: Stratigraphic revisions and correlations to an inferred glacioeustatic record. Paleoceanography, 9, 427–443. Pagani, M., Arthur, M. A., & Freeman, K. H. (1999). Micoene evolution of atmospheric carbon dioxide. Paleoceanography, 14, 273–292. Pagani, M., Zachos, J. C., Freeman, K. H., Tipple, B., & Bohaty, S. (2005). Marked decline in atmospheric carbon dioxide concentrations during the Paleogene. Science, 309, 600–603. Pa¨like, H., Laskar, J., & Shackleton, N. J. (2004). Geologic constraints on the chaotic diffusion of the solar system. Geology, 32, 929–932. Pa¨like, H., Norris, R. D., Herrle, J. O., Wilson, P. A., Coxall, H. K., Lear, C. H., Shackleton, N. J., Tripati, A. K., & Wade, B. S. (2006). The heartbeat of the Oligocene climate system. Science, 314, 1894–1898. Passchier, S. (2000). Soft-sediment deformation features in core from CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 401–412. Passchier, S., & Krissek, L. A. (2008). Oligocene-Miocene Antarctic continental weathering record and paleoclimatic implications, Cape Roberts Drilling Project, Ross Sea, Antarctica. Palaeogeogr. Palaeoclmatol. Palaeoecol., 260, 30–40. Paul, H. A., Zachos, J. C., Flower, B. P., & Triparti, A. (2000). Orbitally induced climate and geochemical variability across the Oligocene/Miocene boundary. Paleoceanography, 15, 471–485. Pearson, P. N., & Palmer, M. R. (2000). Atmospheric carbon dioxide over the past 60 million years. Nature, 406, 695–699. Pekar, S. F., & Christie-Blick, N. (2008). Resolving apparent conflicts between oceanographic and Antarctic climate records and evidence for a decrease in pCO2 during the Oligocene through early Miocene (34–16 Ma). Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 41–49. Pekar, S. F., Christie-Blick, N., Kominz, M. A., & Miller, K. G. (2002). Calibrating eustasy to oxygen isotopes for the early icehouse world of the Oligocene. Geology, 30, 903–906. Pekar, S. F., & DeConto, R. M. (2006). High-resolution ice-volume estimates for the early Miocene: Evidence for a dynamic ice sheet in Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 101–109. Pekar, S. F., DeConto, R. M., & Harwood, D. M. (2006). Resolving a late Oligocene conundrum: Deep-sea warming versus Antarctic glaciation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 29–40.
The Oligocene–Miocene Boundary
399
Pekar, S. F., & Kominz, M. A. (2001). Two-dimensional paleoslope modeling: A new method for estimating water depths for benthic foraminiferal biofacies and paleo shelf margins. J. Sediment. Res., 71, 608–620. Pollard, D., & DeConto, R. M. (2005). Hysteresis in Cenozoic Antarctic Ice Sheet variations. Glob. Planet. Change, 45, 9–21. Pollard, D., & DeConto, R. M. (2007). A coupled ice-sheet/ice-shelf/sediment model applied to a marine-margin flowline: Forced and unforced variations. In: M. Hambrey, P. Christoffersen, N. Glasser, & B. Hubbard (Eds). Glacial Sedimentary Processes and Products. International Association of Sedimentologists Special Publication. Blackwell Scientific, Oxford, UK, pp. 35–53. Roberts, A. P., Wilson, G. S., Harwood, D. M., & Verosub, K. L. (2003). Glaciation across the Oligocene-Miocene boundary in Southern McMurdo Sound, Antarctica: New chronology from the CIROS-1 drill hole. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 113–130. Scherer, R. P., Bohaty, S. M., & Harwood, D. M. (2000). Oligocene and lower Miocene siliceous microfossil biostratigraphy of Cape Roberts Project Core CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 417–442. Schlich, R., Wise, S. W. Jr., Palmer Julson, A. A., et al. (1992). The geologic and tectonic evolution of the Kerguelen Plateau: An introduction to the scientific results of Leg 120. Proceedings of Ocean Drilling Program, Scientific Results, Vol. 120, pp. 5–30 Shackleton, N. J., Hall, M. A., Raffi, I., Tauxe, L., & Zachos, J. (2000). Astronomical calibration for the Oligocene-Miocene boundary. Geology, 28, 447–450. Shackleton, N. J., & Kennett, J. P. (1975). Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: Oxygen and carbon isotope analyses in DSDP sites 277, 279 and 281. Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 29, 743 pp. Shackleton, N. J., Lamb, H. H., Worssam, B. C., Hodgson, J. M., Lord, A. R., Shotton, F. W., Schove, D. J., & Cooper, L. H. N. (1977). The oxygen isotope stratigraphic record of the late pleistocene. Philos. Trans. R. Soc. Lond., 280B, 169–182. Sijp, W. P., & England, M. H. (2004). Effect of the Drake Passage through flow on global climate. J. Phys. Oceanogr., 34, 1254–1266. Stephens, B. B., & Keeling, R. F. (2000). The influence of Antarctic sea ice on glacial-interglacial CO2 variations. Nature, 404, 171–174. Stickley, C. E., et al., (2004). Timing and nature of the deepening of the Tasmanian Gateway. Paleoceanography, 19(PA4027), doi:10.1029/2004PA001022. Toggweiler, J. R., & Bjornsson, H. (2000). Drake Passage and paleoclimate. J. Quaternary Sci., 15, 319–328. Troedson, A. L., & Riding, J. B. (2002). Upper Oligocene to lowermost Miocene strata of King George Island, South Shetland Islands, Antarctica: Stratigraphy,
400
G. S. Wilson et al.
facies analysis, and implications for the glacial history of the Antarctic Peninsula. J. Sediment. Res., 72, 510–523. Vail, P. R., Mitchum, R. M. Jr., & Thompson, S. (1977). Seismic stratigraphy and global changes of sea level, part 4: Global cycles of relative changes of sea level. Am. Assoc. Pet. Geol. Mem., 26, 83–97. Van der Meer, J. J. M. (2000). Microscopic observations on the upper 300 metres of CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 339–348. Verosub, K. L., Florindo, F., Sagnotti, L., Roberts, A. P., & Wilson, G. S. (2000). Environmental magnetism of Oligocene-Miocene glaciomarine strata from CRP2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 599–608. von der Heydt, A., & Dijkstra, H. A. (2006). Effect of ocean gateways on the global ocean circulation in the late Oligocene and early Miocene. Paleoceanography, 21, PA10111, doi:10.1029/2005PA001149. Weertman, H. (1961). Stability of ice-age ice sheets. J. Geophys. Res., 66, 3783–3792. Wilson, G. S., Bohaty, S., Fielding, C. R., Florindo, F., Hannah, M. J., Harwood, D. M., McIntosh, W. C., Naish, T. R., Roberts, A. P., Sagnotti, L., Scherer, R. P., Strong, C. P., Verosub, K. L., Villa, G., Watkins, D. K., Webb, P. N., & Woolfe, K. J. (2000a). Chronostratigraphy of CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 647–654. Wilson, G. S., Florindo, F., Sagnotti, L., Verosub, K. L., & Roberts, A. P. (2000b). Magnetostratigraphy of Oligocene-Miocene glaciomarine strata from the CRP2/ 2A core, Victoria Land Basin. Terra Antartica, 7, 631–646. Wilson, G. S., Lavelle, M., McIntosh, W. C., Roberts, A. P., Harwood, D. M., Watkins, D. K., Villa, G., Bohaty, S., Florindo, F., Sagnotti, L., Naish, T. R., Scherer, R. P., & Verosub, K. L. (2002). Integrated chronostratigraphic calibration of the Oligocene-Miocene boundary at 24.070.1 Ma from the CRP2A drill core, Ross Sea, Antarctica. Geology, 30, 1043–1046. Wilson, G. S., Roberts, A. P., Verosub, K. L., Florindo, F., & Sagnotti, L. (1998). Magnetobiostratigraphic chronology of the Eocene-Oligocene transition in the CIROS-1 core, Victoria Land margin, Antarctica: Implications for Antarctic glacial history. Geol. Soc. Am. Bull., 110, 35–47. Wright, J. D. & Miller, K. G. (1992). Miocene stable isotope stratigrpahy, Site 747, Kerguelen Plateau. Proceedings of the Ocean Drilling Programme, Scientific Results, Vol. 120, pp. 855–866. Wright, J. D., & Miller, K. G. (1993). Southern ocean influence on late Eocene to Miocene deepwater circulation. Antarctic Res. Series, 60, 1–25. Zachos, J. C., Flowers, B. P., & Paul, H. A. (1997). A high resolution chronology of orbitally paced climate oscillations across the Oligocene/Miocene boundary. Nature, 388, 567–570. Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K. (2001a). Trends rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693. Zachos, J. C., Shackleton, N. J., Revenaugh, J., Pa¨like, H., & Flower, B. P. (2001b). Climate response to orbital forcing across the Oligocene-Miocene boundary. Science, 292, 274–278.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00010-4
Chapter 10
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean Alan M. Haywood1,, John L. Smellie2, Allan C. Ashworth3, David J. Cantrill4, Fabio Florindo5, Michael J. Hambrey6, Daniel Hill2, Claus-Dieter Hillenbrand2, Stephen J. Hunter2,1, Robert D. Larter2, Caroline H. Lear7, Sandra Passchier8 and Roderick van de Wal9 1
School of Earth & Environment, University of Leeds, Leeds LS2 9JT, UK Geological Sciences Division, British Antarctic Survey, High Cross, Madingley Road, Cambridge CB3 0ET, UK 3 Department of Geosciences, North Dakota State University, Fargo, ND 58105-5517, USA 4 Royal Botanic Gardens Melbourne, Private Bag 2000, Birdwood Avenue, South Yarra, Victoria 3141, Australia 5 Istituto Nazionale di Geofisica e Vulcanologia, Via di Vigna Murata 605, 00143 Rome, Italy 6 Institute of Geography & Earth Sciences, University of Wales, Aberystwyth, Ceredigion SY23 3DB, UK 7 School of Earth, Ocean and Planetary Sciences, Cardiff University, Main Building, Park Place, Cardiff CF10 3YE, UK 8 Department of Earth and Environmental Studies, Mallory Hall 252, Montclair State University, Montclair, NJ 07043, USA 9 Institute for Marine and Atmospheric Research Utrecht, Utrecht University, Princetonplein 5, 3584 Utrecht, The Netherlands 2
ABSTRACT This chapter explores the Middle Miocene to Pliocene terrestrial and marine records of Antarctica and the Southern Ocean. The structure of the chapter makes a clear distinction between terrestrial and marine records as well as proximal (on or around Antarctica) and more distal records (Southern Ocean). Particular geographical regions are identified that reflect the areas for which the majority of palaeoenvironmental and palaeoclimatic information exist. Specifically, Corresponding author. Tel.: þ44(0)113 343 8657; Fax: þ44(0)113 343 6716;
E-mail:
[email protected] (A.M. Haywood).
402
A. M. Haywood et al.
the chapter addresses the terrestrial sedimentary and fjordal environments of the Transantarctic Mountains and Lambert Glacier region, the terrestrial fossil record of Antarctic climate, terrestrial environments of West Antarctica, and the marine records of the East Antarctic Ice Sheet (EAIS), the West Antarctic Ice Sheet (WAIS) and the Antarctic Peninsula Ice Sheet (APIS), as well as the marine record of the Southern Ocean. Previous and current studies focusing on modelling Middle Miocene to Pliocene climate, environments and ice sheets are discussed.
10.1. Introduction The Middle-to-Late Miocene interval is believed to represent a time of significant ice-sheet expansion on Antarctica (e.g. Miller et al., 1991, 2005; Lear et al., 2000; Turco et al., 2001; Billups and Schrag, 2002; Shevenell et al., 2004). The stable isotope record of the deep sea demonstrates that a midMiocene ‘‘climatic optimum’’, at B15 Ma, was followed by strong enrichment in oceanic d18O and climatic cooling over the next 6 Ma (e.g. Zachos et al., 2001). During this interval, the East Antarctic Ice Sheet (EAIS) is thought to have been a major and permanent ice sheet, although fluctuations in the size of EAIS may still have occurred (e.g. Westerhold et al., 2005). Denton et al. (1984) proposed that during this time, the EAIS overrode the Transantarctic Mountains (Fig. 10.1). Recent studies from the western Dry Valleys indicate that the atmosphere cooled by as much as B201C prior to an erosional event that is linked to the EAIS overtopping the mountains (Lewis et al., 2007). If such a sequence is repeated elsewhere (i.e. cooling and development of cold-based alpine glaciers preceded ice-sheet overriding), then it would suggest that the initial rise in deep-sea benthic d18O (Zachos et al., 2001) reflects deep-water cooling, followed later by ice-sheet expansion. Furthermore, Lewis et al. (2006) propose that immense freshwater floods to the Southern Ocean from large subglacial lakes beneath the expanded EAIS occurred between 14.4 and 12.4 Ma. The discharges are not only considered to have been the erosive force that formed such prominent Dry Valleys features as the Labyrinth, but are also considered through their impact on oceanic circulation to be a cause of mid-Miocene climatic change. A different explanation for the enrichment in oceanic d18O is that it represents the onset of significant glaciation on West Antarctica (Mercer, 1978; Ciesielski et al., 1982). This is supported by the first occurrence of ice rafted debris (IRD) at Deep Sea Drilling Project (DSDP) Site 325 in the Bellingshausen Sea during the Early to Middle Miocene
Figure 10.1: Map showing geographical locations discussed in the text. Antarctic bed topography also shown from the BEDMAP dataset (Lythe et al., 2000).
404
A. M. Haywood et al.
(Hollister et al., 1976). However, the relative role that the EAIS and West Antarctic Ice Sheet (WAIS) played in the Middle Miocene Climate Transition (MMCT) has recently been questioned by new seismicstratigraphic data from the Ross Sea revealing at least five major intervals of grounded ice advance and retreat in the Middle Miocene (Bart, 2003; Chow and Bart, 2003). Much of this ice was sourced from West Antarctica, suggesting the presence of a large and dynamic WAIS prior to the MMCT (Bart, 2003; Chow and Bart, 2003). The stability of Antarctic Ice Sheets during the Late Miocene and Pliocene has been the subject of almost continuous debate for more than 20 years. The key questions in this argument are when did the EAIS switch from being polythermal and dynamic to predominantly cold based and stable, and could relatively short-lived climatic warm intervals have been sufficient to influence the Antarctic Ice Sheet? A long-standing view is that the EAIS became stable during the Middle Miocene, evidence for which is primarily derived from apparent landscape stability and well-dated surfaces and ash deposits in the Dry Valleys region along the western border of the Ross Sea (e.g. Denton et al., 1993; Marchant et al., 1993a; Sugden, 1996). An alternate view is that terrestrial glacial deposits, known as the Sirius Group, scattered throughout the Transantarctic Mountains, indicate dynamic ice-sheet conditions even during the Pliocene. This conclusion is based on the occurrence of Pliocene (and older) diatoms reworked into glacial deposits (Harwood, 1983, 1986; Webb et al., 1984, 1996; Harwood and Webb, 1986; Wilson, 1995; Wilson et al., 1998). The dynamic nature of the Pliocene Antarctic Ice Sheet is supported by the Pagodroma Group along the flanks of the Lambert Glacier (e.g. Hambrey and McKelvey, 2000a,b). Each view is internally consistent and scientists have been presented with a significant challenge in reconciling the different views. When considering the size and diverse landscapes of Antarctica, we should not be surprised to see a degree of heterogeneity in the climate and environmental response. Yet the current state of knowledge is so contradictory that the scientific community has become polarized into two camps (‘‘stabilists and dynamicists’’) over the issue of Middle Miocene to Pliocene conditions on Antarctica (e.g. Harwood and Webb, 1998; Stroeven et al., 1998a,b).
10.2. East Antarctic Terrestrial Environments Two main pre-Quaternary sedimentary sequences recording glacial events are found on land in East Antarctica, the Sirius Group in the Transantarctic
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
405
Mountains and the Pagodroma Group in the Prince Charles Mountains flanking the Lambert Glacier (Fig. 10.2). The two sequences are quite different in character. The Sirius Group is a mainly terrestrial glacigenic succession whose age has proved highly controversial, with Pliocene, midMiocene or even older ages suggested. The Pagodroma Group is an iceproximal fjordal sequence up to several hundred metres thick, and is well dated on the basis of in situ marine fossils. All these well-exposed deposits are important in unravelling the history of the EAIS, as they are the nearest
Figure 10.2: (a) Map of Antarctica showing key sites of investigations of on-shore pre-quaternary glacigenic sediments. (b) The distribution of known Sirius Group outcrops. (c) Location of the Pagodroma Group sites along the Lambert Graben (redrawn from Hambrey and McKelvey, 2000a; Hambrey et al., 2003).
406
A. M. Haywood et al.
deposits to the heart of the ice sheet, and yield valuable information concerning palaeogeography and palaeoclimate. 10.2.1. Transantarctic Mountains Outcrops of glacigenic sedimentary rocks of the Sirius Group occur as erosional remnants scattered throughout the Transantarctic Mountains as far south as 861S (Fig. 10.3). The deposits are exposed in two typical settings: (1) as thin erosional remnants at high elevation in palaeovalleys or on flat mountain summits, or (2) as sequences more than 100 m thick along the walls of broad trunk valleys occupied by large outlet glaciers draining the modern EAIS. The main sedimentary facies are massive diamicts, massive boulder gravels, stratified pebbly sands and muds, and laminated sand/mud couplets with dropstones (Fig. 10.3). Diamicts overlie spectacular grooved and striated pavements in some areas, notably at Roberts Massif. In the Dominion Range, bordering Beardmore Glacier, well-preserved Nothofagus leaves, stems and roots, plus mosses and insects are interbedded with the diamict (Francis and Hill, 1996; Ashworth et al., 1997). The diamicts are dominated by subangular to subrounded striated and facetted clasts. The environment of deposition is interpreted as largely subglacial to proglacial, with extensive wet-based, probably polythermal glaciers being the agent of sediment transport, although a marine influence is evident through marine fossil remains in part of the Beardmore Glacier succession (e.g. Mercer, 1972; Mayewski and Goldthwait, 1985; McKelvey et al., 1991; Webb et al., 1996; Stroeven and Prentice, 1997; Wilson et al., 1998, 2002; Hambrey et al., 2003). The lithostratigraphy and depositional setting of the Sirius Group at two of the southernmost locations is summarized in Table 10.1. A debate was ignited when Webb et al. (1984), based on recycled marine diatoms, linked the Sirius Group to glacial expansion after early Pliocene deglaciation and marine incursion of East Antarctica. In contrast, Denton et al. (1984) attributed the Sirius Group to mid-Miocene overriding of the Transantarctic Mountains based on geomorphological evidence from the Dry Valleys. A dynamic ice-sheet hypothesis and a stable ice-sheet hypothesis developed, which represented contrasting views of Neogene Antarctic climate and glacial dynamics. Pivotal to this debate is our understanding of the uplift and erosional, hence unroofing, history of the Transantarctic Mountains (e.g. Webb et al., 1984; Behrendt and Cooper, 1991; Sugden et al., 1995; Kerr and Huybrechts, 1999). The Transantarctic Mountains are divided into several crustal blocks which vary in size on a range of scales and which probably experienced uplift at different rates and
Figure 10.3: Glacial erosional features and sedimentary characteristics of the Sirius Group, Transantarctic Mountains. (A) Aerial view of sub-Sirius Group glacially abraded surface on Roberts Massif, showing patch of diamict, postdepositional faults, and a superimposed recessional quaternary moraine system. (B) Multiple beds of diamict comprising the W100 m thick Shackleton Glacier formation at the type locality alongside upper Shackleton Glacier. (C) Large-Scale Glacial Grooves on Jurassic Dolerite, predating deposition of the Sirius Group. (D) Striated dolerite surface with overlying Sirius Group diamict, Roberts Massif. (E) Typical massive diamict facies of the Sirius Group, showing alignment of clasts, interpreted as basal till, Roberts Massif. (F) Laminated silt and thin diamicts with large dropstone, interpreted as an ice-contact lake deposit in the Shackleton Glacier formation, Bennett Platform (redrawn from Hambrey et al., 2003).
Battye Glacier (mid-Miocene)
Bardin Bluffs (Pliocene–?Pleistocene)
Mount Johnston (?Oligocene–early Miocene)
Fisher Bench (mid-Miocene)
Fisher Massif
Formations
Amery Oasis
Massive and stratified diamictite; sandy breccia/conglomerate; partially silicified wood fragments; well lithified
Massive and weakly stratified diamict, massive boulder gravel, stratified sandstone, breccia; sand/ mud laminate with dropstones at Shackleton Glacier. Well preserved Nothofagus flora at Beardmore Glacier. All facies well indurated
Principal facies
Massive boulder gravel; massive diamict; gravel; laminate with dropstones. Rich diatom microflora. All weakly to well indurated
Principal facies
Shackleton erosion surface/dominion erosion surface
Cloudmaker (Pliocene–?Miocene)
Shackleton Glacier
Eroded remnants as erratics
Meyer Desert (Pliocene)
Beardmore Glacier
Bennett Platform
Shackleton Glacier
Formations
Ice-proximal fjordal, grounding-line fan complexes predominant, to distal glaciomarine
Paleoenvironment
Subglacial with extensive mass-movement and fluvial reworking; welldeveloped flora
Subglacial; some supraglacial; proglacial glaciofluvial; ice contact lake
Paleoenvironment
Ages of formations are given where determined using diatom biostratigraphy. No correlation is implied in the Sirius Group. Data from Webb et al. (1996) for the Sirius Group at Beardmore Glacier; Hambrey et al. (2003) for the Sirius Group at Shackleton Glacier; Hambrey and McKelvey (2000a) for the Pagodroma Group.
Pago-droma
Group
Sirius
Group
Table 10.1: Representative stratigraphy of the Sirius and Pagodroma Groups with principal lithofacies and interpretation of palaeoenvironment. 408 A. M. Haywood et al.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
409
during different periods, principally in the late Cretaceous and early Tertiary (prior to c. 40 Ma; Stump and Fitzgerald, 1992; Fitzgerald, 1994; Bussetti et al., 1999; van der Wateren et al., 1999). The uplift histories of adjacent blocks might also be different. The younger uplift and erosional history is only relatively well known for the Dry Valleys block, as a result of detailed detrital studies by the Cape Roberts Project (CRP). Those studies showed that at the deepest levels cored, equivalent to c. 34 Ma, the kilometre-thick Kirkpatrick Basalts (Jurassic) had already been almost completely removed and the sand modes were dominantly derived from sandstones of the Victoria Group, then the Taylor Group (Devonian–Triassic Beacon Supergroup; Smellie, 2000a,b, 2001a,b, unpublished; Talarico et al., 2000; Sandroni and Talarico, 2001). By c. 33 Ma, the 2 km-thick Beacon Supergroup had been cut through to expose outcrops of basement rock (early Palaeozoic and Precambrian granitoids and metamorphic rocks), which then began to contribute significant detritus. Between that time and c. 29 Ma, tectonic stability and, presumably, little uplift-related erosion are inferred from the essentially unchanging detrital modes. Further changes in the detrital modes between 24 and 16 Ma suggest another phase of instability and possibly uplift/unroofing, after which the record is not preserved. Although the detrital record is interpreted here in terms of simple uplift and unroofing, an alternative explanation is that the changes observed might reflect varied phases of climate-related erosion independent of uplift (e.g. enhanced downcutting by glaciers; cf. Kerr and Huybrechts, 1999). This is an ambiguity that has yet to be resolved. According to the dynamic ice-sheet hypothesis, the diatom assemblages incorporated in the Sirius Group record periods when East Antarctic basins were ice-free and became inundated by the sea (Harwood, 1986; Harwood and Webb, 1998). The time intervals of which diatoms are lacking record the stages when either ice was covering the inland basins or the floors of the basins were exposed. According to stabilists, however, the basic assumption derived from ice-sheet modelling is that deglaciation requires considerable climatic warming (Huybrechts, 1993). A substantial body of internally consistent evidence (including ash and cosmogenic dating) for a pre-middle Miocene landscape, subsequently unmodified by ice, has been published in numerous papers (e.g. Marchant et al., 1993a,b,c, 1996; Sugden et al., 1995; Sugden, 1996; Sugden and Denton, 2004). For example, isotopic ages of W14 Ma for the in situ, unweathered, ash deposits from the higher elevated regions of the Dry Valleys (Marchant et al., 1993b) are in disagreement with the concept of warming causing major deglaciation in the early–midPliocene. The diatom evidence for a Pliocene age of the Sirius Group has been disputed (Barrett, 1996; Gersonde et al., 1997; Stroeven et al., 1998a,b).
410
A. M. Haywood et al.
Field studies and studies of aerial photographs provided evidence of rejuvenation of faulting during and after deposition of the Sirius Group (Hambrey et al., 2003). Moreover, based on morphostratigraphic constraints and provenance analyses, several authors pointed out that the Sirius Group comprises deposits of multiple glaciations within different wet-based icesheet drainage systems, which operated during consecutive stages of glacial denudation in the Transantarctic Mountains (e.g. Mercer, 1968, 1972; Brady and McKelvey, 1979, 1983; McKelvey et al., 1991; Stroeven, 1997; Van der Wateren et al., 1999; Passchier, 2001, 2004; Hambrey et al., 2003). Unfortunately, the absolute ages of the individual formations within the Sirius Group remain inconclusive and lithostratigraphic correlations are hampered by the complex tectonic framework of the Transantarctic Mountains. Given the considerable linear geographical spread of the Sirius Group over a distance of c. 1,500 km and spanning nearly 101 of latitude, the grouping of these strata under a single group name has exacerbated the arguments concerning the age. Resolution of the age question remains a key challenge for Antarctic geologists, but it is likely that, although the Sirius Group may contain disputed Pliocene or Miocene elements, it could also extend back to Oligocene time in view of the presence of glaciomarine sediments of this age offshore (Barrett, 1996; Hambrey et al., 2002; Francis et al., 2008). 10.2.2. Lambert Glacier Region The Pagodroma Group is a succession of massive diamicts and boulder gravels with minor stratified diamicts, laminites, sand and gravel, cropping out in the Prince Charles Mountains along the western margin of the Lambert Graben (Figs. 10.2 and 10.4). The best known outcrops are at Amery Oasis and Fisher Massif (Bardin, 1982; Hambrey and McKelvey, 2000a; McKelvey et al., 2001; Whitehead et al., 2003, 2006a), but equivalent exposures are known from Mount Menzies, the last exposed rock in this region heading polewards. The total outcrop distance of known glacigenic strata along the flank of the graben is about 800 km. The graben is occupied by the Lambert Glacier, a south–north-flowing outlet glacier, whose total ice-drainage area represents 13 per cent or 1 million km, of the EAIS. It has acted as a conduit for ice flowing into Prydz Bay since its initial formation at the Eocene–Oligocene transition (Barron et al., 1991; Strand et al., 2003). Although the Pagodroma Group is regarded as the East Antarctic equivalent of the Sirius Group, there are major differences in lithofacies, facies thickness and geometry, fossil content and palaeoenvironment. For example, these
Figure 10.4: Glacial erosional features and sedimentary characteristics of the Pagodroma Group, Prince Charles Mountains. (A) Uplifted segment of a palaeofjord bottom and sidewall (arrowed), filled with nearly 300 m of iceproximal fjordal sediment of the Miocene Fisher Bench formation, Fisher Massif. (B) Striated pavement on Amery Group Sandstone, Radok Lake, Amery Oasis; this surface is coeval with the lower part of the Pliocene Bardin Bluffs formation. (C) Typical massive to crudely stratified boulder gravel of the Miocene Fisher Bench formation, Fisher Massif, interpreted as icecontact deposits associated with a grounding-line fan. (D) Weakly stratified diamict and boulder gravel of the Pliocene Bardin Bluffs formation at the type locality, interpreted as ice-proximal deposits; the cliff face here is about 60 m high. (E) Laminated silt and thin diamictites representing deposition adjacent to a grounding-line fan as quiet water cyclopels and cyclopsams, interrupted by gravity-flow deposition, Miocene Battye Glacier formation, Dragons Teeth, Amery Oasis (redrawn from Hambrey and McKelvey, 2000a).
412
A. M. Haywood et al.
deposits contain in situ diatom assemblages that can be used to date the formations directly (Whitehead et al., 2003, 2004). The four formally defined formations within the Pagodroma Group (Table 10.1) have age ranges from early Miocene (or possibly Oligocene) to Pliocene–Pleistocene (Hambrey and McKelvey, 2000b; Whitehead et al., 2003, 2006a). The strata at Mount Menzies (Whitehead and McKelvey, 2002) have not yet yielded any datable fossils. The massive diamicts and boulder gravels indicate deposition in an ice-proximal environment near a grounding line, whereas the stratified facies represent more distal iceberg deposition. The diamicts contain striated and facetted clasts within a matrix that is finer grained than modern Antarctic tills. Dropstone structures occur in thin laminated beds and resemble glacigenic marine rhythmites (cyclopels and cyclopsams) that commonly occur to the sides and front of a groundingline fan (Fig. 10.5). The depositional environments are considered to be analogous to the modern fjords of East Greenland with fast-flowing polythermal tidewater glaciers (Hambrey and McKelvey, 2000b). The Battye Glacier Formation contains diatom-bearing beds (opal contents up to 15 wt.%) with in situ mollusc assemblages, which are assumed to have been deposited in ice-distal settings (Whitehead et al., 2006a). The Pagodroma Group has a cumulative thickness of W800 m and provides evidence for major shifts in the position of the grounding line of the Lambert Glacier throughout the Neogene (Hambrey and McKelvey, 2000a). The Pagodroma Group represents remnants of a much more extensive fjord-fill sequence that has been uplifted above sea level (a.s.l.) (Fig. 10.4). Palaeofjord floors have been uplifted to different levels at Amery Oasis and Fisher Massif, the highest being nearly 1,500 m on the latter. In contrast, the main Lambert trough has been excavated to 2,000 m below sea level, the sediment from which has been delivered into Prydz Bay since the Eocene/ Oligocene transition, where it forms the prograded continental shelf, including the huge late Neogene Prydz Trough-Mouth Fan that was drilled during Ocean Drilling Program (ODP) Leg 188 (Passchier et al., 2003; O’Brien et al., 2004). The Prydz Bay continental margin has been drilled by ODP Legs 119 and 188 (see Barron et al., 1991; Cooper et al., 2004; Whitehead et al., 2006b). Numerical modelling of ice-sheet expansion and recession in the Lambert Glacier–Prydz Bay region has demonstrated that when major phases of erosion and sediment delivery took place, the ice dynamics were strongly controlled by the changing bathymetry of the iceeroded graben (Taylor et al., 2004). The clay mineralogy of the Pagodroma Group indicates a dominance of illite and chlorite in the Middle Miocene or older Mount Johnston and Fisher Bench Formations, indicative of physical weathering and erosion of
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
413
Figure 10.5: Distribution of Neogene (15–2 Ma) terrestrial volcanic outcrops in West Antarctica. local metavolcanic and gneissic source rock terrains under glacial conditions (Ehrmann et al., 2003). In contrast, the Middle–Late Miocene Battye Glacier Formation contains significant contributions of smectite and kaolinite (Whitehead et al., 2006a), and the Pliocene–Pleistocene Bardin Bluffs Formation has the highest kaolinite concentrations in the absence of smectite. Formation of kaolinite and smectite in source rock terrains generally requires prolonged exposure of source rocks to chemical
414
A. M. Haywood et al.
weathering. The chemically altered nature of the strata is confirmed by bulk chemical and rock magnetic studies (Bloemendal et al., 2003; Passchier and Whitehead, 2006). Recycling of weathered materials from local sedimentary source rocks cannot be ruled out (Ehrmann et al., 2003). The origin and significance of the chemically weathered materials in the Middle Miocene and younger Pagodroma Group formations remains inconclusive at this time and further field and laboratory studies are necessary to resolve this issue. The two regions described above represent the best studied examples of pre-Quaternary onshore glacigenic successions, but there are others elsewhere, e.g. the Antarctic Peninsula (Hambrey et al., 2008) and the Grove Mountains (Fang et al., 2004). When a better-dated Sirius Group can be linked to the offshore record in McMurdo Sound, and the data from the Prince Charles Mountains fully integrated with the emerging Prydz Bay record, we will have a good understanding of the behaviour of the EAIS since its inception in these segments of the Antarctic margin. 10.2.3. The Terrestrial Fossil Record of East Antarctic Climate The onset of Antarctic glaciation during the Oligocene dramatically altered terrestrial environments ultimately resulting in widespread extinction of plant and animal taxa (Cantrill, 2001). Tundra environments, based on pollen from the Cape Roberts borehole, are known to have existed in the Ross Sea region through the Oligocene until at least the early Miocene (Askin and Raine, 2000; Raine and Askin, 2001). The final extinction of plants and animals with more complex life histories is assumed to be associated with the shift to the polar desert climate which prevails today. Whether or not this immediately followed the mid-Miocene warm interval or was as late as the mid-Pliocene is the most outstanding problem in Neogene climatic history in need of resolution. Pre-Pleistocene glacigenic sequences (Sirius and Pagodroma Groups) are patchily exposed throughout the Transantarctic and Prince Charles mountains. They are dominated by glaciomarine fjordal sediments, lodgement tills and outwash deposits (see previous section for more details). At the Oliver Bluffs on the upper Beardmore Glacier, fossiliferous sequences occur within the Meyer Desert Formation. The discovery of wood (Carlquist, 1987) and later Nothofagus leaves (Hill and Truswell, 1993; Webb and Harwood, 1993; Francis and Hill, 1996; Hill et al., 1996) from a location about 500 km from the South Pole makes this arguably the most important fossil site in Antarctica. The fossils provide an unusual window
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
415
into Neogene environments and biogeography but until questions surrounding their age can be better resolved, their value as palaeoclimatic proxies is lessened. The deposits have been assigned a Pliocene age using a biostratigraphic argument based on reworked marine diatoms (Harwood, 1986; Webb et al., 1996). Recently, however, cosmogenic dates obtained on boulders from moraines that post-date the Meyer Desert Formation indicate ages of greater than 5 Ma (Ackert and Kurz, 2004). The authors propose various age models based on different rates of weathering and discuss different amounts of uplift, but based on their discussion, a Miocene age for the Meyer Desert Formation is preferred over a Pliocene age (Ackert and Kurz, 2004). If the older age for the Meyer Desert Formation prevails, then it is tempting to speculate that the degree of melting of the ice sheet that resulted in fjords opening up into the Transantarctic Mountains was associated with the midMiocene climatic optimum. That interpretation would be more acceptable to researchers in the Dry Valleys who present evidence that the shift from wetto cold-based glaciation occurred during the Middle Miocene (Denton et al., 1993; Marchant et al., 1993a; Sugden, 1996). The fossil Nothofagus leaves from the Meyer Desert Formation were described as a new species, N. beardmorensis, but with morphological similarities to extant species in South America and Tasmania (Hill et al., 1996). Originally, a krummholz growth form (scrubby, stunted growth form) was assigned to the Nothofagus (Webb and Harwood, 1993) but later studies of the growth ring anatomy (Francis and Hill, 1996) suggested a low prostrate and spreading habitus, more similar to ground-hugging shrubs growing today along exposed parts of treeline on Isla Navarino at the southern tip of South America (Ashworth and Kuschell, 2003, p. 193). The prostrate growth forms implied colder growing conditions than estimated using the krummholz interpretation, and considerably lower than the comparison to the cold-temperate rain forest of southern South America made on the basis of a pollen study (Askin and Markgraf, 1986; Mercer, 1986). Nearest living relative comparisons, along with physiological experiments into cold tolerance of Nothofagus, yielded theoretical estimates of mean summer month temperatures of at least 51C for 3 months and low winter month temperatures of 15 to 221C. Mean annual temperature (MAT) was estimated to have been in the range of 8 to 121C (Francis and Hill, 1996; Ashworth and Kuschel, 2003) compared to an interpolated sealevel MAT at latitude 851S today of 261C (Ashworth and Kuschel, 2003). Although much of the research to date has focused on Nothofagus, the vegetation was more complex consisting of cushion plants, grasses, ranunculids (buttercups) and mosses (Ashworth and Cantrill, 2004).
416
A. M. Haywood et al.
It should be emphasized that to date, only one site has been reported from the Beardmore Glacier and leaves and wood of Nothofagus have not been reported from other Sirius Group sequences although they have been discovered recently in pre-Pleistocene glacial sequences of the western Dry Valleys (Ashworth et al., unpublished). The Meyer Desert Formation also contains fossils of beetles, including two weevil species, a higher fly, single species each of a freshwater gastropod and bivalve, an ostracod and a fish (Ashworth and Kuschel, 2003; Ashworth and Preece, 2003; Ashworth and Thompson, 2003; Ashworth et al., 1997). Using overlapping autecological requirements for Nothofagus, listroderine weevils and freshwater molluscs, mean summer temperatures are estimated to have been 4–51C for at least two summer months and the MAT was estimated to be 81C, similar to the temperatures estimated from the physiological requirements of the Nothofagus fossils (Ashworth and Kuschel, 2003; Ashworth and Preece, 2003). The palaeosols, which are stratigraphically higher than the fossiliferous beds, indicate polar conditions (Retallack et al., 2001), and their development may signal the end of warmer and wetter climates in Antarctica. Regardless of age, there is a question regarding the likelihood of a tundra biome re-establishing itself in the interior of Antarctica during a warm phase if it had been extirpated during an earlier cold phase. The question is important because it sets limits on how cold the climate of Antarctica could have been before the Meyer Desert Formation biota colonized the interior. Did the species disperse from refugia around the margins of Antarctica, in which case the climate had to have been continually warmer around the margins of the continent through at least the Early Miocene? Or, if they became extinct prior to the mid-Miocene climatic optimum or mid-Pliocene warm interval (B3 Ma), could they have reinvaded the continent? For Nothofagus, seed germination is considered to be critical to answering this question. Studies suggest that Nothofagus seeds cannot survive prolonged periods of immersion in salt water (Hill et al., 1996) and consequently cannot be easily dispersed across oceanic barriers. The implications of this are that components of the vegetation survived on Antarctica even during earlier glacial phases. There is a scenario, however, based on phylogenetic and palynological studies, that allows for the possibility that Nothofagus may have dispersed from Australia to New Zealand after the Tasman Sea formed (McGlone et al., 1996). Even if transTasman dispersal occurred, it would have been from land areas with similar climatic conditions that would favour successful colonization. For organisms to disperse from South America or New Zealand to Antarctica would require not only much greater distances of dispersal but for organisms to re-establish
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
417
themselves in complex communities in a different climatic regime. This makes the long-distance dispersal scenario for recolonization of Antarctica even less probable. We note that since deglaciation at the end of the Pleistocene, climatic conditions on many of the sub-Antarctic islands have been favourable for colonization by Nothofagus, listroderine weevils and the types of freshwater molluscs found in the Meyer Desert Formation, but it has not occurred. This strongly implies that refugia were present on the margins of Antarctica, from which plants, insects and freshwater molluscs could migrate to the continental interior during a time or times of significant climatic warming (cf. Convey et al., in press).
10.3. West Antarctic Terrestrial Environments West Antarctica contains numerous outcrops that preserve, or potentially preserve, information on Miocene and Pliocene terrestrial palaeoenvironments. All are volcanic, although several contain associated, usually minor, sedimentary strata. Because of the long time gaps between eruptions in any volcanic centre, the environmental record they preserve is typically of a coarse resolution, and they have often been ignored as a potentially important source of palaeoenvironmental information. The outcrops are often remote or inaccessible, and many have been rarely visited. Summary information on all those outcrops of Middle Miocene and Pliocene age is given in Table 10.2 (see also Fig. 10.5). The eruptive environment is relatively well understood for a few but remains uncertain for many.
10.3.1. Antarctic Peninsula Most of the outcrops in the Antarctic Peninsula have been relatively well investigated in recent years. They form two main types: outcrops dominated by a major stratovolcano (James Ross Island Volcanic Group) and outcrops formed by multiple monogenetic volcanic fields and small isolated centres (Bellingshausen Sea Volcanic Group; Smellie, 1999). James Ross Island, situated at the northern end of the Antarctic Peninsula, comprises a single large stratovolcano 1.5 km high and about 60 km in diameter, together with numerous small satellite centres. About 50 individual eruptive episodes have been identified, with most retaining information on environmental conditions between 6.25 Ma and today. All but a few show features diagnostic of glacial conditions (Smellie, 2006). The associated ice sheets had a provenance
Ames Range
Marie Byrd Land Mount Flint
Hudson Mountains
Jones Mountains
14.7–10.0; 10.3–10.0; 7.6–5.7
14.1
10–7? (imprecise) 8.5–3.7
Undated
o6.5
James Ross Island region
Ellsworth Land Snow Nunataks
6
7.5–5.4; 2.5– 2.7
Age (Ma)
Merrick Mountains
Antarctic Peninsula Alexander Island
Outcrop
K–Ar
K–Ar
K–Ar
K–Ar
40/39Ar
K–Ar
K–Ar
Dating method
Unknown
Unknown
Subaerial, subaqueous and/or glacial
Glacial (2?)
Glacial (1)
Three overlapping stratovolcanoes
Stratovolcano
Multiple small centres
Multiple monogenetic centres
Tiny degraded outcrop remnants Dissected stratovolcano and multiple satellite centres
Subaerialc; glacial? Glacial (1) and glaciomarine (interglacial)
Multiple monogenetic centres
Comments
Glacial (1 and 2)b; subaerial
Environmental characteristicsa
LeMasurier and Thomson (1990) LeMasurier and Thomson (1990)
LeMasurier and Thomson (1990)
LeMasurier and Thomson (1990)
Nelson (1975), Skilling (1994)
Smellie and Hole (1997), Smellie (1999) Smellie (1999)
Principal sources
Table 10.2: Characteristics of volcanic outcrops of Neogene Age (15–2 Ma) in West Antarctica.
418 A. M. Haywood et al.
10.1–9.1
9.56–3.19
9.39–8.22
Toney Mountain
Hobbs Coast
Mount Murphy
5.7–4.2
10.4–8.54; 6.4
K–Ar
8.5–6.4; 7.86– 7.57
40/39Ar
K–Ar
K–Ar
K–Ar
40/39Ar
K–Ar
K–Ar
11.7–8.6
10.4–10.0; 3.0
K–Ar
13.7
Mount Cumming, Executive Committee Range Mount Sidley, Executive Committee Range Mount Bursey
Whitney Peak, Executive Committee Range Mount Hampton, Executive Committee Range Mount Hartigan, Executive Committee Range Two poorly exposed volcanoes: Lavris Peak and Tusing Peak Undissected stratovolcano Three coalesced stratovolcanoes
Subaerial?c
Subaerialc and snow contact
Glacial (1 and (mainly) 2) and subaerialc
Glacial
Unknown
Unknown
Two overlapping undissected stratovolcanoes Plateau lava basal succession Numerous small dissected centres Dissected stratovolcano
Undissected stratovolcano
Subaerial?c
Subaerial?c
Undissected stratovolcano
Subaerial?c
LeMasurier and Thomson (1990) LeMasurier and Thomson (1990) Smellie (2000a,b), LeMasurier (2002), Wilch and McIntosh (2002)
LeMasurier and Thomson (1990)
LeMasurier and Thomson (1990)
LeMasurier and Thomson (1990)
LeMasurier and Thomson (1989)
LeMasurier and Thomson (1988) Middle Miocene to Pliocene History of Antarctica and the Southern Ocean 419
2.58–2.32
4.5–3.4
Mount Berlin
Fosdick Mountains
K–Ar
K–Ar
K–Ar
40/39Ar
40/39Ar
Dating method
Three volcanoes: Mt Rees, Mt Frakes and Mt Steere Two overlapping undissected stratovolcanoes Two coalesced undissected stratovolcanoes
Glacial (2); subaerialc
Unknown
Subaerialc and glacial?
Unknown
Small monogenetic centres
Comments
Glacial (1)
Environmental characteristicsa
LeMasurier and Thomson (1990)
LeMasurier and Thomson (1990)
LeMasurier and Thomson (1990)
Smellie (2001a,b), Wilch and McIntosh (2002) Wilch and McIntosh (2002)
Principal sources
b
Ignores late-stage subaerial parasitic cones (Marie Byrd Land), often very much younger (by m.y.) than the main stratocone successions, which they overlie. Glacial 1: ‘‘thick ice’’ conditions (few hundred metres); Glacial 2:‘‘thin ice’’ conditions (o150 m); after Smellie (2000a,b). c Environmentally undiagnostic since subaerial volcanics may be supraglacial; evidence for (lower elevation) glacial conditions not exposed.
a
9.13–6.82; 8.46–5.74; 4.25–1.81 4.9–4.7
Crary Mountains
Mount Moulton
6.8–5.95
Age (Ma)
Mt. Murphy satellite centres
Outcrop
Table 10.2: (Continued ).
420 A. M. Haywood et al.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
421
rooted both in the Antarctic Peninsula and in a local ice cap centred on the volcano itself (Hambrey and Smellie, 2006). The ice was wet based and erosive throughout, although there are few thermal regime data for units younger than 2 Ma. Conversely, a few of the volcanic units showed evidence for eruptions in marine conditions, and marine fossils in interbedded sedimentary deposits also suggest that warmer ice-poor conditions (interglacials) occurred during at least three broad periods: 6.5–5.9, 5.03–4.22 and o0.88 Ma (Smellie et al., 2006a,b). The James Ross Island Volcanic Group is now known to contain numerous sedimentary interbeds, mainly varieties of diamicts with a glacial association that are only beginning to be investigated fully (Hambrey and Smellie, 2006; Hambrey et al., 2008). By contrast, the Bellingshausen Sea Volcanic Group contains multiple small volcanic centres with ages ranging between 7.5 and 2.5 Ma (Smellie, 1999). The glacial environment was wet-based, temperate or subpolar, and varied between thin and thick ice conditions (o70 and W500 m, respectively), although not all outcrops yielded the full range of information (Table 10.2; Smellie et al., 1993; Smellie and Hole, 1997). There also exists a small outcrop of 2.5 Ma tephra (Hornpipe Heights Formation, northern Alexander Island) erupted under essentially dry subaerial conditions, presumably when any associated glacial cover was below the present outcrop (i.e. below 750 m a.s.l.). Scant evidence from elsewhere on Alexander Island suggests that glacial conditions may have been present at 2.5 Ma, although few Pliocene samples are dated and the age resolution is poor (Smellie, 1999). 10.3.2. Ellsworth Land Three outcrops occur in this region: Snow Nunataks, Jones Mountains and Hudson Mountains (Table 10.2). The Hudson Mountains are particularly poorly known. Conversely, Snow Nunataks comprises several small undated nunataks that have been divided into the Mount Benkert and Mount McCann formations, formed of subaqueous stratified hyalotuff and subaerial lava, respectively (Smellie, 1999). The formations are cogenetic and resemble basaltic glaciovolcanic centres known as tuyas (Smellie, 2000a,b) erupted probably under ice thicknesses that locally exceeded 350 m. However, a marine environment, while probably less likely, cannot be excluded since so little is known about the area. A 700 m-thick Cenozoic volcanic succession resting on a subhorizontal polished and striated glacial unconformity crops out at Jones Mountains (Rutford et al., 1972; Hole et al., 1994). Diamictite lenses interpreted as tillite overlie the unconformity and the
422
A. M. Haywood et al.
overlying sequence has been divided into at least two units of pillow lava, hyaloclastite breccia and hyalotuff. Striated and facetted erratic boulders are also scattered throughout the lower volcanic unit. The descriptions most closely resemble volcanic sequences formed under a thin glacial cover. However, since individual ‘‘thin ice sequences’’ are theoretically limited to thicknesses r150 m (Smellie, 2000a,b, 2001a,b), the much greater total thickness of volcanic rock present suggests that more than two volcanic sequences are present. This is also suggested by the presence of numerous reworked hyalotuff lenses dispersed within the sequence, and possibly by the wide range in the published isotopic ages (mainly 7–10 Ma; Table 10.2), which might suggest an outcrop constructed from multiple eruptions that occurred over a long period. 10.3.3. Marie Byrd Land Marie Byrd Land contains the largest Cenozoic volcanic field in Antarctica (LeMasurier and Thomson, 1990). It is also the most remote region and is among the least visited and least well known of all Antarctic volcanoes. Numerous reports published in the 1970s and 1980s suggested a subglacial origin for many centres (summarized by LeMasurier and Rocchi, 2005). While agreeing generally that the eruptive environment was probably glacial, the published criteria used most often (e.g. mainly presence of hyaloclastite; less often interbedded tillite and/or glacially scoured surfaces) are too simplistic and permit only limited interpretation of the eruptive environment, except in rare outcrops where more modern detailed studies have been carried out. In addition, most Marie Byrd Land volcanoes oc. 12 Ma in age have undergone only minimal dissection, probably as a result of a change to a polar thermal regime (LeMasurier and Rocchi, 2005; cf. Armienti and Baroni, 1999), so only the latest and topographically highest erupted products are available for examination. As those units were usually erupted subaerially, they provide little of environmental use except a limiting elevation for any coeval ice sheet. However, caldera and sector collapses have led to a few inland volcanoes being well exposed (e.g. Crary Mountains, Mount Sidley: Wilch and McIntosh, 2002), and many of the coastal volcanoes are deeply dissected (e.g. Mount Murphy: LeMasurier et al., 1994; Smellie, 2000a,b, 2001a,b; LeMasurier, 2002; Wilch and McIntosh, 2002; LeMasurier and Rocchi, 2005). Only the better exposed and described examples are considered here. Mount Murphy is situated on the coast in eastern Marie Byrd Land. It is a complex polygenetic edifice constructed from at least three centres, of which
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
423
only the westernmost is well exposed. The basal section, up to c. 1,400 m a.s.l., is basaltic and it has a thinner capping sequence of mainly trachyte lavas. The basaltic section is dominated by multiple units individually several tens of metres thick and composed of stratified hyalotuff, hyaloclastite breccia and sheet or pillow lava. Each unit is separated by a glacial unconformity that is often draped by a thin diamict (tillite), and a few units are capped by strombolian scoria. Although LeMasurier (2002) interpreted the Mount Murphy succession in terms of lava-fed deltas (and therefore, by implication, eruptions beneath ‘‘thick ice’’), the sequences are much more similar to those associated with a thin glacial cover, i.e. composed mainly of snow, firm and/or fractured ice that was never more than tens of metres thick (Smellie, 2000a,b, unpublished). The sequences extend down to below 400 m a.s.l., and they reflect the local dynamics of a small ice cap probably similar in appearance to that centred on Mount Murphy today. Conversely, several satellite centres at Icefall Nunatak, Turtle Rock and Hedin Nunatak have younger ages of 6.8–5.95 Ma (Table 10.2). They were associated with a low-elevation (o600 m a.s.l.) ice sheet composed of relatively thick ice (at least several hundred metres). That ice was thicker than today and had a surface elevation that fluctuated through 200 m during the period (Smellie, 2000a,b, 2001a,b). A wet-based thermal regime (temperate or subpolar ice) characterized the glacial cover associated with the main (8–9 Ma) Mount Murphy edifice, but its nature is unknown for the younger satellite centres. Finally, Mount Murphy is overlain by a thin laminated mudstone interpreted as a glacial lake sequence related to a significant but undated overriding event (o3.5 Ma) by an extraordinarily thick former ice sheet that was at least 1,500 m thicker than at present (LeMasurier et al., 1994). Crary Mountains (9.13–1.81 Ma; Table 10.2) are situated 250 km inland of Mount Murphy, where the surrounding WAIS surface elevation is 1,600 m a.s.l. Only the two older northern centres (Mount Rees and Mount Steere) are well exposed and have ages in the range between 8 and 9 Ma, i.e. coeval with Mount Murphy. The mafic to intermediate successions are individually o100 m thick and composed of alternating ‘‘dry’’ lavadominated and ‘‘wet’’ hyaloclastite-rich lithofacies, interpreted as lavas interacting with local slope ice and snow (Wilch and McIntosh, 2002). Thus, the coeval Crary Mountains and Mount Murphy successions are broadly similar, being products of interaction with a thin glacial cover. They differ, however, in that the former lack the interbedded tillites and glacial surfaces that are so prominent at Mount Murphy, a difference that was tentatively interpreted to be due to the Crary Mountains ‘‘ice’’ being cold based.
424
A. M. Haywood et al.
Finally, sector collapse at Mount Sidley has exposed a magnificent 1,300 m-high caldera wall section that extends from c. 2,700 to W4,000 m a.s.l. The outer slopes are essentially undissected and expose only subaerially erupted volcanic products, including a widespread ‘‘dry’’ trachyte– phonolite fall deposit. By contrast, the caldera wall section is mainly formed of interbedded phonolite and trachyte lavas, domes and ignimbrite ranging in age from 5.69 to 4.18 Ma. The sequence is essentially subaerial. Evidence for glacial conditions is absent, apart from minor signs of water interaction (stratified hyalotuffs and thin vitroclastic lava bases) in one of the subsequences dated as 5.36–4.81 Ma. The latter was interpreted as due to limited glaciovolcanic interaction, presumably local slope ice/snow. However, the basal 1.5 km of the volcano is obscured by the present-day WAIS and the earlier (and lower elevation) environmental history is therefore unseen.
10.4. The Marine Record of the East Antarctic Ice Sheet Due to the nearly complete ice coverage of the Antarctic continent, a record of past Antarctic climate lies in the layers of sediments eroded over million years from the continent and deposited in sedimentary basins around its margin and in the Southern Ocean. During the past ca. 20 years, nine ODP legs have significantly advanced our understanding of the Cenozoic tectonics and palaeoenvironments of the Antarctic region. These legs include Leg 113 in the Weddell Sea (Barker et al., 1988, 1990), Leg 114 in the sub-Antarctic South Atlantic (Ciesielski et al., 1988, 1991), Leg 119 in Prydz Bay and on Kerguelen Plateau (Barron et al., 1989, 1991), Leg 120 on Kerguelen Plateau (Schlich et al., 1989; Wise et al., 1992), Leg 177 in the southeast Atlantic sector of the Southern Ocean (Gersonde et al., 1999, 2003), Leg 178 on the Antarctic Peninsula (Barker et al., 2002), Leg 181 (Carter et al., 2000), Leg 188 in Prydz Bay (O’Brien et al., 2001; Cooper et al., 2004) and Leg 189 in the Tasmanian region (Exon et al., 2001, 2004). The pioneer of Antarctic drilling was the DSDP in 1973 (Hayes et al., 1975). Following that, four distinct phases of drilling have taken place in the Ross Sea region, in sedimentary basins under the direct influence of East Antarctica, by rigs mounted on the fast-ice that rings the southern part of the Ross Sea just north of the Ross Ice Shelf (see Chapter 3 for further details). These include the Dry Valley Drilling Project (DVDP) in 1970–1975 (McGinnis, 1981), the McMurdo Sound Sediment and Tectonic Studies (MSSTS) in 1979 (Barrett, 1986), the Cenozoic Investigations in the Western
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
425
Ross Sea (CIROS) in 1984 and 1986 (Barrett, 1989; Barrett et al., 1992) and the multinational CRP in 1997–1999 (Cape Roberts Science Team, 1998, 1999, 2000; Hambrey et al., 1998; Barrett et al., 2000, 2001). These efforts are continuing through ANDRILL (Antarctic geological drilling) programme, which demonstrated ability to recover high quality marine and glacimarine sedimentary drill cores from high latitude ice-covered areas (Naish et al., 2007; Florindo et al., 2008; Harwood et al., 2008). To date, in spite of these efforts, important questions and problems remain unresolved and are the subject of a considerable debate within the scientific community. Among these are the causes of the mid-Miocene cooling event at around 14 Ma, and the warming–cooling climatic phases of the Pliocene. 10.4.1. Middle Miocene Cooling Marine deep-sea oxygen isotope records indicate that a period of ice-sheet instability during the Early and Middle Miocene is superimposed on a longterm warming trend that culminated in the mid-Miocene climatic optimum from B17 to 15 Ma (Miller et al., 1987; Zachos et al., 2001). This phase was followed by a gradual cooling with a major cooling event occurring at the Middle–Late Miocene boundary B14 Ma that continued for B1–2 m.y. (MMCT; Flower and Kennett, 1995). The Prydz Bay area was visited during the Leg 119 (Barron et al., 1989, 1991) and Leg 188 (O’Brien et al., 2001; Cooper et al., 2004) drilling campaigns of the ODP. Middle Miocene–Pliocene sediments were drilled at Sites 739, 742, 1165 and 1166. An interesting stratigraphic record across this transition was recovered during ODP Leg 188 when three sites were drilled proximal to the EAIS across the Prydz Bay continental shelf (Site 1166), slope (Site 1167) and rise (Site 1165). Of these, Site 1165 (64122uS; 67113uE) records a history of sedimentation on the continental rise extending back to earliest Miocene times (about 22 Ma). Several long-term changes characterize this record, including an overall trend of decreasing sedimentation rates from the bottom to the top of the hole. There is a progressive decrease in the sedimentation rate above about 308 m b.s.f. (metres below sea floor), which is marked by a transition from predominantly dark-grey fissile claystones with abundant silt laminae to grey diatom-bearing clays with a higher abundance of IRD (Fig. 10.6). At this transition, the total clay content also increases. The chronology of this sequence indicates a Middle Miocene age (ca. 14.3 Ma) for the lithological transition. Correlation to ODP Hole 747A on the Kerguelen Plateau shows that this transition coincides with the base of the Mi-3/3a d18O event (Florindo et al., 2003). Middle Miocene
Figure 10.6: Lithostratigraphic column of Site 1165 and correlation between Middle Miocene intervals to ODP Hole 747A from the Kerguelen Plateau (ODP Leg 120). The lithological transition at about 308 m b.s.f., which is characterized by an increase in ice-rafted debris, sand grains and total clay content, coincides with the base of Mid-Miocene glacial event (Mi-3/3a) (modified from Florindo et al., 2003).
426 A. M. Haywood et al.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
427
changes in sedimentation at Site 1165 may have been caused by the development of a cooler ice sheet characterized by decreased rates of glacial erosion and decreased production of melt water. In such an event, less suspended sediment is delivered to the continental rise, but ice at sea level produces more icebergs and, hence, more ice rafting of debris. 10.4.2. Late Miocene–Early Pliocene Ice-Sheet Fluctuations Long-term variation in the marine oxygen isotopic data (e.g. Hodell and Venz, 1992; Kennett and Hodell, 1995; Zachos et al., 2001) are suggestive of a warming in the earliest Pliocene, culminating in the mid-Pliocene Climatic Optimum at B3 Ma. However, the oxygen isotope ratios do not allow major changes in ice volume on Antarctica. Obliquity-driven oscillations (41 kyr) in d18O of up to 0.6m, which dominate the Pliocene period, while capable of producing B20–30 m sea-level changes (depending on the temperature contribution to the d18O signal), are also considered insufficient to cause large-scale deglaciation of Antarctica without a significant increase in the temperature of the deep ocean. Marine sedimentary deposits exposed on land and on the continental shelf, slope and rise show evidence of dynamic ice-sheet behaviour in the Late Miocene–Early Pliocene and Early Pliocene warming. At Marine Plain in the Vestvold Hills, East Antarctica, a ca. 8 m thick early Pliocene sequence of diatomaceous sands and silts is exposed (Pickard et al., 1988). During deposition of the diatomaceous sediments, the ice margin must have been ca. 50 km further inland and no floating ice was covering the site. The presence of Cetacean skeletons, including dolphins and a right whale, and preliminary isotope measurements led Quilty (1993) to infer warmer conditions during deposition than currently prevail. Water temperatures may have been as high as 51C. A marked increase in sea-ice indicator diatoms and a coarsening upward of the sediments in the upper 3 m of the deposit indicate cooling conditions and glacial expansion. The sediments at ODP Sites 739, 742 and 1166 in Prydz Bay record evidence of repeated advances and retreats of the Lambert ice stream across the shelf in the Late Miocene through Early Pliocene (Hambrey et al., 1991; Passchier et al., 2003; O’Brien et al., 2004). During repeated advances of the Lambert ice stream to the shelf break glacigenic debris flows build up a trough-mouth fan on the continental slope. It is worth mentioning that the IRD signal inferred from the W250-mm size fraction at Site 1165, in a sediment drift on the continental rise off Prydz Bay, suggests warmer conditions in the Early Pliocene
428
A. M. Haywood et al.
(Gru¨tzner et al., 2005). The low sea-ice concentrations derived from diatom assemblages (Whitehead et al., 2005) and the warm-water periods inferred from high abundances of silicoflagellates (Dictyocha) at Site 1165 (Whitehead and Bohaty, 2003) are consistent with isotopic estimates of warmer conditions than present during the Early Pliocene (Hodell and Venz, 1992). In addition, three periods of enhanced accumulations of sediment originating from East Antarctica are recognized in the Late Miocene–Pliocene record of the East Kerguelen Ridge sediment drift, in the Indian Ocean to the north of Prydz Bay (Joseph et al., 2002). The sediment pulses are interpreted as periodically warmer, less stable, ice-sheet conditions during this time. DVDP-10, 11 and 15, MSSTS-1 and CIROS-2 drill-holes in the Ross Sea Region have recovered partial late Neogene records that suggest a dynamic Plio-Pleistocene history of glacial advance and retreat in the Taylor and Ferrar valleys. Of these, a 166 m thick Plio-Pleistocene sequence of diatombearing glaciomarine strata was cored in Ferrar Fiord by the CIROS-2 drilling campaign (163131uE; 77141uS) (Barrett et al., 1992; Barrett and Hambrey, 1992). The Pliocene sequence is B62 m thick and is composed of diamictite dominated by basement lithologies interbedded with thin mudstone intervals, indicating several early Pliocene advances and retreats of ice through the Transantarctic Mountains. Further evidence for Pliocene Antarctic ice volume fluctuation is recorded by glaciomarine strata from the DVDP-10 and 11 drill-holes (163132uE; 77135uS) (McGinnis, 1981; Ishman and Reick, 1992; Wilson, 1993). Combined foraminiferal and magnetostratigraphic analyses indicate highly fluctuating palaeoenvironmental conditions in the Late Miocene–Early Pliocene with less ice-influenced conditions in the early Pliocene (Ishman and Reick, 1992). Late Neogene trough-mouth fans have been identified in seismic data on the Ross Sea and Weddell Sea continental slopes. The Crary Fan on the Weddell Sea margin was formed by glacial sediment supply from Dronning Maud Land and ice drainage from East Antarctica through the Weddell Sea (Bart et al., 1999), whereas the Ross Sea fan was fed by ice draining through the Northern basin (Bart et al., 2000). In contrast, the geometry of Late Neogene sequences on the Wilkes Land margin is characterized as steep prograding foresets and trough-mouth fans that are less developed (Escutia et al., 2005). However, the close proximity to the deep subglacial Wilkes Land basin results in ice grounding below sea level, which may have contributed to ice-sheet behaviour that is different from the other portions of the East Antarctic margin. Interpretations of the seismic data record a dynamic Late Neogene EAIS with large expansions and contractions for the Ross Sea and Weddell Sea portions of the East Antarctic margin (Bart et al., 1999, 2000).
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
429
10.5. The Marine Record of the West Antarctic and Antarctic Peninsula Ice Sheets The late Cenozoic glacial history of West Antarctica is archived in the sedimentary sequences of its continental margins. During the last 20 years, multichannel seismic data were acquired from the margins of the Weddell Sea, the Antarctic Peninsula, the Bellingshausen and Amundsen seas, and the eastern Ross Sea. DSDP Legs 28 (Ross Sea margin) and 35 (Bellingshausen Sea) and ODP Legs 113 (Weddell Sea) and 178 (Pacific margin of the Antarctic Peninsula) recovered sedimentary sequences spanning the Neogene to Quaternary. In the following, we summarize the main results of these studies and their implications for the history of WAIS and Antarctic Peninsula Ice Sheet (APIS) from the Middle Miocene to Pliocene. Seismostratigraphic studies suggest that an early stage of glaciation characterized by presence of local ice caps and tidewater glaciers had affected the Ross Sea region already during the Late Oligocene and the Early Miocene (De Santis et al., 1995, 1999; Anderson and Shipp, 2001). Glacial and glaciomarine sediments interbedded with fluvial to deep-water mudstones were deposited on the innermost part of the shelf (CIROS-1 drill site), while open-marine sedimentation dominated on the outer shelf (Abreu and Anderson, 1998; Anderson and Shipp, 2001). This depositional pattern points to cool temperate interglacial conditions, but gives no direct evidence for presence of the WAIS (Abreu and Anderson, 1998; Anderson and Shipp, 2001). Based on seismic-stratigraphic investigations and sedimentary records drilled at DSDP Sites 270 and 272, De Santis et al. (1999) inferred that a glacial transitional regime developed in the Ross Sea region during the Early and Middle Miocene. Most of the subglacial erosion and deposition still took place on the inner shelf, but occasionally local ice caps expanded onto the outer shelf. Around this time, the first significant amounts of IRD were deposited in the Bellingshausen Sea (DSDP Leg 35, Sites 323 and 325) and the Weddell Sea (ODP Leg 113, Sites 693 and 694) (Abreu and Anderson, 1998). The onset of IRD sedimentation at Site 325 points to an establishment of West Antarctic ice caps outside the Ross Sea region during the Early to Middle Miocene (Hollister et al., 1976; Abreu and Anderson, 1998; Barker et al., 2002). By analysing Mg/Ca ratios of calcareous benthic and planktonic foraminifera tests, Lear et al. (2000) and Shevenell et al. (2004) demonstrated that ice-sheet build-up in Antarctica between 14.2 and 13.8 Ma contributed to the pronounced, global increase in d18O ratios in benthic foraminifera tests observed by Zachos et al. (2001), even though the related eustatic sea-level fall was rather indistinct (Abreu and Anderson, 1998; Miller et al.,
430
A. M. Haywood et al.
2005). While most of the Middle Miocene ice-volume increase may have occurred in East Antarctica and possibly the Arctic (Moran et al., 2006), seismic studies also point to an expansion of ice caps in West Antarctica or their coalescence to a WAIS smaller than present (Abreu and Anderson, 1998; Anderson and Shipp, 2001). In the Ross Sea region, seismic data and sedimentary records reveal that grounded ice advanced at least five times across the shelf (Chow and Bart, 2003). A comparison with foraminiferal oxygen isotope and sea-level data shows that at least two of these grounding events fall into the early part of the Middle Miocene (Bart, 2003). Bart and Anderson (1995) proposed on the basis of seismic profiles from the western Antarctic Peninsula outer shelf that glacial erosional surfaces formed during the Middle to Late Miocene, when huge sediments drifts started to grow on the adjacent continental rise as indicated by seismic investigations (Rebesco et al., 1997, 2002). Bart et al. (2005) combined seismic data from the western Antarctic Peninsula shelf with sedimentological data from ODP Leg 178 Site 1097 drilled on the shelf, and concluded that the earliest grounding events on the Antarctic Peninsula shelf are of Late Miocene age. Seismic studies of the sedimentary development of the Crary Fan in the southeastern Weddell Sea indicate a minimum of 5 long-term and 14 small-scale ice-sheet expansions across the adjacent shelf from the Middle Miocene to the Pleistocene (Anderson and Shipp, 2001, and references therein). However, present-day ice-drainage patterns raise the question of whether or not WAIS expansion made a major contribution to the inferred ice advances (cf. Bart et al., 1999). Despite episodic ice expansion in parts of West Antarctica during the Middle Miocene, benthic and neritic planktonic diatoms, which were reworked from the Antarctic shelf and deposited in the central Weddell Sea (Site 694), point to the absence of a major WAIS or APIS from the Middle Miocene to the early part of the Late Miocene (Kennett and Barker, 1990; cf. Chow and Bart, 2003). Widespread glacial unconformities assumed to be Late Miocene in age are observed in seismic profiles from the eastern Ross Sea margin (Cooper et al., 1991; De Santis et al., 1995; Anderson and Shipp, 2001). The grounded ice advances were accompanied by the deposition of glacial and glaciomarine strata (DSDP Leg 28 Site 270 and CIROS-1), which might exhibit an interior West Antarctic provenance (Anderson and Shipp, 2001) and which are interbedded with meltwater deposits and diatom oozes documenting temperate interglacial episodes (Abreu and Anderson, 1998). De Santis et al. (1999) correlated seismic data with sedimentary records drilled at DSDP Leg 28 sites and proposed an onset of fully glacial conditions in the Ross Sea region during the latest Miocene and Early Pliocene. A large and thick WAIS advanced up to the shelf edge (Accaino et al., 2005) and gave the
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
431
Ross Sea shelf its typical modern seabed morphology, which is characterized by landward dipping, overdeepening and troughs carved by grounded ice streams (De Santis et al., 1999). Shelf progradation observed in seismic profiles west of the Antarctic Peninsula indicates a general expansion of APIS assumed to be Late Miocene to Early Pliocene in age (Larter and Barker, 1991; Larter and Cunningham, 1993; Bart and Anderson, 1995; Larter et al., 1997; Barker et al., 2002; Bart et al., 2005). Similar shelf progradation is reported from seismic profiles elsewhere on the continental margin in the Bellingshausen and Amundsen seas although the age of the prograding sequences in these areas is unconstrained (Nitsche et al., 1997, 2000; Scheuer et al., 2006). Episodic advance and retreat of grounded ice across the eastern and western Antarctic Peninsula shelf throughout the latest Miocene and Early Pliocene are evident from seismic unconformities (Sloan et al., 1995; Larter et al., 1997; Bart and Anderson, 2000; Bart, 2001; Bart et al., 2005) and the onset of shelf overdeepening west of the Antarctic Peninsula (Camerlenghi et al., 2002). Recovery of sediments deposited in subglacial and glaciomarine settings from the western Antarctic Peninsula shelf (ODP Leg 178 Sites 1097 and 1103) confirm episodic APIS grounding events at least since the Latest Miocene to Early Pliocene (Barker et al., 1999a, 2002; Eyles et al., 2001). The general APIS expansion during the Late Miocene to Early Pliocene coincided with a northward shift of the Antarctic Circumpolar Current (ACC) (Herna´ndez-Molina et al., 2006), pronounced growth of large sediment drifts (Rebesco et al., 1997), clay-mineral fluctuations (Hillenbrand and Ehrmann, 2005) and distinct maxima in accumulation rates of terrigenous detritus (Gru¨tzner et al., 2005) on the western Antarctic Peninsula rise. The sedimentary development of the drifts suggests episodic, enhanced supply of glacial-sourced sediments from the adjacent shelf from ca. 9.4 Ma onwards. Considerable growth of APIS and WAIS during the Late Miocene is also indicated by the onset of IRD deposition and sedimentation of a clay mineral assemblage pointing to the dominance of physical weathering in the South Orkney region (Site 696) and of rapid turbidite deposition with a West Antarctic provenance in the central Weddell Sea (Site 694) (Kennett and Barker, 1990). Anderson and Shipp (2001) infer that during the Late Miocene, the WAIS reached its modern size and, at times, was even larger than today. The seismic and core data from the western Antarctic Peninsula margin mentioned above suggest a similar scenario for the APIS. At the Miocene/Pliocene boundary, the deposition of IRD at Site 325 increased (Hollister et al., 1976), while the first significant deposition of IRD started at Site 322 near Drake Passage (Abreu and Anderson, 1998). Kennett and Barker (1990) attributed a shift from pelagic to hemipelagic deposition
432
A. M. Haywood et al.
at Site 696 and from turbiditic to hemipelagic sedimentation at Site 694 during the earliest part of the Pliocene to the development of a stable APIS and WAIS or climatic warming in West Antarctica. Seismic stratigraphy combined with sedimentological data from DSDP Leg 28 drill cores indicates that the eastern Ross Sea shelf reached a late fully glacial stage with an overdeepened topography, which was mainly influenced by glacial processes of erosion and deposition resulting in relatively low sedimentation rates (De Santis et al., 1999). Throughout the Early Pliocene, high-frequency grounding events are observed in seismic profiles from the eastern Ross Sea shelf and the western Antarctic Peninsula shelf (Alonso et al., 1992; Larter and Cunningham, 1993; Larter et al., 1997; Bart and Anderson, 2000; Bart, 2001). Cyclicity in drift sedimentation recorded in drill cores from the Antarctic Peninsula rise confirms that ice streams repeatedly advanced to the shelf break throughout the Pliocene (Barker et al., 2002; Hillenbrand and Ehrmann, 2005; Hepp et al., 2006). During the Early Pliocene, sedimentation rates increased in the deep Bellingshausen Sea (DSDP Leg 35 Sites 322, 323 and 325) and the Weddell Sea (ODP Leg 113 Sites, 693 and 697) (Barker, 1995). On the western Antarctic Peninsula continental rise, down-slope transport exceeded along-slope transport throughout the Pliocene (Herna´ndez-Molina et al., 2006), and the sediment supply from the shelf enhanced and maintained the sediment drifts (Rebesco et al., 1997, 2002). Seismic studies of the continental margin in the western Bellingshausen and Amundsen seas revealed truncations of foreset reflectors and extensive progradation, probably during the Pliocene (Nitsche et al., 1997, 2000; Scheuer et al., 2006), suggesting that grounded ice eroded the shelf and transported the terrigenous detritus across the shelf edge. According to Scheuer et al. (2006), the accumulation rate of terrigenous sediment on the adjacent continental rise reached its maximum during the Pliocene and Quaternary, which may indicate a relatively late establishment of WAIS in the adjacent hinterland (cf. Nitsche et al., 1997). In contrast, sedimentation rates on other parts of the West Antarctic margin reached distinct maxima during the Late Miocene (Site 1095; Barker et al., 2002) or during the Early Pliocene (Barker, 1995). During the Late Pliocene, sedimentation rates decreased at the drifts on the Antarctic Peninsula rise (ODP Leg 178, Sites 1095, 1096 and 1101; Barker et al., 1999a, 2002), in the deep Bellingshausen Sea (DSDP Leg 35, Sites 322, 323 and 325; Barker, 1995) and in the Weddell Sea (ODP Leg 113, Sites 694 and 697; Kennett and Barker, 1990; Barker, 1995). A concomitant transition from progradation to aggradation is observed in seismic profiles from elsewhere on the West Antarctic margin, from the Antarctic Peninsula to the eastern Ross Sea (Larter and Barker, 1991; Larter and Cunningham, 1993;
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
433
De Santis et al., 1995; Larter et al., 1997; Nitsche et al., 1997, 2000). In contrast to these findings, Rebesco et al. (2006) compared changes in the geometries of seismic reflection profiles and concluded that enhanced progradation of the continental shelf and slope started almost all-around Antarctica at ca. 3 Ma. These authors suggest that APIS, WAIS and EAIS switched from a predominantly polythermal to a predominantly cold-based mode at this time of major ice-sheet build-up in the Northern Hemisphere (cf. Barker, 1995).
10.6. Marine Records of the Southern Ocean Today there are four main water masses south of the polar front: Antarctic Surface Water; Warm Deep Water (WDW); Circumpolar Deep Water (CDW); and Antarctic Bottom Water (AABW) (see Chapter 4 for further details). WDW flows into the Southern Ocean from the North Atlantic, Pacific and Indian Oceans. This nutrient-rich water mass upwells at the Antarctic Divergence, replacing sinking Antarctic Surface Water. CDW, the dominant water mass in the Southern Ocean, is a mixture of WDW and Antarctic waters. AABW forms as cold, saline shelf waters sink and mix with CDW prior to flowing northwards out of the Southern Ocean. The Antarctic coastal current flows from east to west around the continent. North of the Antarctic Divergence, prevailing westerlies drive surface waters towards the east, forming the ACC, which in places extends to the seafloor. In part, these oceanographic conditions are a response to climate on the Antarctic continent. For example, formation of sea ice increases the density of shelf waters, aiding the formation of AABW. Conversely, changing ocean– atmosphere circulation patterns influence global climate by determining the transport of heat and moisture onto the polar continent. A commonly cited example is the thermal isolation of Antarctica by the ACC.
10.6.1. The MMCT The MMCT witnessed a major reorganization of global deepwater circulation patterns associated with the expansion of the Antarctic Ice Sheet (e.g. Woodruff and Savin, 1989; Wright et al., 1992; Flower and Kennett, 1995). The widespread presence of hiatuses in Southern Ocean sedimentary cores is one indicator of this Middle Miocene reorganization of ocean circulation patterns. Although these hiatuses present difficulties in
434
A. M. Haywood et al.
reconstructing past oceanographic conditions using marine cores, a broad consensus regarding the history of some of the Southern Ocean water masses has emerged in recent years. One proxy used to reconstruct ocean circulation patterns is the carbon isotopic composition of benthic foraminifera, as water masses generally become enriched in the light isotope of carbon as they age. Interbasinal offsets in benthic foraminiferal carbon isotopes indicate that the Early Miocene Southern Ocean received warm saline deep water sourced from low latitudes in the Tethyan region (Woodruff and Savin, 1989; Wright et al., 1992). The flow of this warm saline water mass likely ceased in the Middle Miocene, and the consequent reduction in meridional heat transport has been called upon as a trigger for the ice-sheet expansion in the Middle Miocene (Woodruff and Savin, 1989; Flower and Kennett, 1995). The ACC connects the Pacific, Atlantic and Indian Oceans and thus plays an important role with respect to global heat transport. Inception of a deep circumpolar ACC possibly occurred near the Oligocene–Miocene boundary, well before the MMCT (e.g. Pfuhl and McCave, 2005; Lyle et al., 2007). However, a variety of different proxies all point to temporal variations in the intensity of the ACC since its inception. A reduction in surface-to-thermocline foraminiferal d18O gradients at DSDP Site 516 indicates reduced thermal stratification in the surface waters of the southwestern South Atlantic around 16 Ma (Pagani et al., 2000). At the same time, the abundance of alkenones at this site decreased, implying a reduced nutrient input to this area. These observations have been interpreted to reflect a reduction in Antarctic Intermediate Water caused by a reduction in the strength of the ACC (Pagani et al., 2000). Re-establishment of a strong surface-to-thermocline temperature gradient in this region points to renewed flow of the ACC coincident with the expansion of the Antarctic Ice Sheet in the Middle Miocene (Pagani et al., 2000). Because the oceanic residence time of neodymium is comparable to the ocean mixing time, the neodymium isotope system can be used as a radiogenic isotope tracer of ocean water masses. A study of the neodymium isotopic composition of sediments from DSDP Site 266 in the Australian– Antarctic basin indicates an increased contribution of neodymium from the Kerguelen volcanic province during the Middle Miocene, suggesting enhancement of the ACC during the climate transition (Vlastelic et al., 2005). Pacific ferromanganese crusts that grew in equatorial Pacific bottom water display a gradual change in their neodymium isotopic composition between
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
435
38 and B20 Ma, which has been interpreted as increasing flow of Southern Ocean waters into the Pacific Ocean via the Deep Western Boundary Current (DWBC) as a result of progressive build-up of the ACC (Van de Flierdt et al., 2004). The isotope trends in the ferromanaganese crusts reversed in the Middle Miocene, likely reflecting increased mixing in the Pacific as a result of increased production and export of AABW associated with the build-up of Antarctic ice (Van de Flierdt et al., 2004). Independent evidence for intensification of the DWBC associated with the Middle Miocene Antarctic Ice Sheet growth is found in the record of sortable silt mean size from ODP Site 1123 situated on the Chatham Rise (Hall et al., 2003). The sortable silt mean size increases with stronger near bottom current flow and selective deposition and winnowing. The ODP Site 1123 record indicates increased flow speeds during the MMCT, interpreted as reflecting increased export of Southern Ocean waters to the Pacific Ocean. Furthermore, the intensity of the DWBC was found to exhibit variability on the timescale of the 41 kyear orbital obliquity cycle (Hall et al., 2003). High-resolution foraminiferal proxy records from the South Tasman Rise (southwest Pacific) give further independent evidence for a strengthening of the ACC associated with the ice-sheet expansion in the Middle Miocene. Shevenell et al. (2004) generated paired oxygen isotope (d18O) and Mg/Ca records using a planktonic foraminifera, G. bulloides, from ODP Site 1171. The d18O record reflects a combination of temperature and the oxygen isotopic composition of seawater, the latter a function of global ice volume in addition to regional salinity. The Mg/Ca record is used as a salinityindependent temperature proxy. The ODP Site 1171 Mg/Ca record displays a B71C cooling (point-to-point) between 14.2 and 13.8 Ma, which suggests that the paired d18O record reflects a concomitant surface water freshening above the South Tasman Rise. These records are interpreted as reflecting an intensification of the ACC between 14.2 and 13.8 Ma in a series of steps paced by orbital cycles (Shevenell et al., 2004), although it has also been suggested that a component of the surface water freshening may represent massive melting pulses of the Antarctic Ice Sheet (Holbourn et al., 2005). 10.6.2. Late Miocene–Early Pliocene The Late Miocene witnessed further high latitude cooling, summarized in Hodell and Kennett (1986). A cooling of Southern Ocean upper circumpolar and intermediate water masses is documented in benthic foraminiferal oxygen isotope records from ODP Site 1088 in the sub-Antarctic South
436
A. M. Haywood et al.
Atlantic, which display a two-step increase between B7.4 and 6.9 Ma (Billups, 2002). Associated with this late Neogene cooling are further changes in Southern Ocean oceanography. A 10–5 Ma hiatus in the sediment package underlying the Bounty Fan at ODP Site 1122 suggests an intensification of the ACC, associated with the Late Miocene climate cooling and expansion of the WAIS (Carter et al., 2004). The strengthening of North Atlantic Deep Water (NADW) in the Late Miocene led to global oceanic circulation patterns approaching the modern configuration (e.g. Wright et al., 1991). Late Miocene strengthening of NADW is supported by benthic foraminiferal carbon isotope records from ODP Site 1088 in the sub-Antarctic South Atlantic, which indicate an influx of a nutrient-depleted water mass into the Southern Ocean at B6.6 Ma (Billups, 2002). The relative flux of NADW into the Southern Ocean reached modern proportions by B6 Ma, and subsequently intensified yet further, supporting speculation that the climatic warmth of the Early Pliocene resulted from enhanced thermohaline overturn (Kwiek and Ravelo, 1999; Billups, 2002). The warm period in the Early Pliocene coincides with an apparent increase in biological productivity as indicated by increased opal deposition rates at ODP Sites 1165, 1095, 1096 and several gravity core sites in the Southern Ocean, which has been interpreted as reflecting a reduction in the extent of sea ice (Hillenbrand and Fu¨tterer, 2002; Gru¨tzner et al., 2005; Hillenbrand and Ehrmann, 2005; Hillenbrand and Cortese, 2006). Surprisingly, this warm interval appears to coincide with glacial advance (indicated by increased terrigenous input from Antarctica), perhaps supporting the ‘‘snowgun hypothesis’’ of Prentice and Matthews (1991) (Gru¨tzner et al., 2005). Large oscillations in the Al/Ti ratio, the lead isotopic composition and the neodymium isotopic composition of sediments at DSDP Site 266 in the Australian–Antarctic basin have been interpreted to reflect either variations in the intensity of the ACC or variations in weathering fluxes resulting from periods of instability of the Antarctic Ice Sheet around 1.9 and 3.3 Ma (Vlastelic et al., 2005). Diatom assemblages at ODP Sites 1165 and 1166 suggest that sea-ice cover remained lower than today through much of the Pliocene (Whitehead et al., 2005), and silicoflagellate assemblages have been used to pinpoint three intervals (3.7, 4.3–4.4 and 4.6–4.8 Ma) within the Pliocene when sea surface temperatures in the Southern Ocean were roughly 51C warmer than today (Whitehead and Bohaty, 2003). Similar events are recorded in ODP Leg 178 Sites, pointing to circum-Antarctic warming that may have been responsible for ice-sheet loss (Escutia et al., 2007).
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
437
10.7. Modelling Antarctic Climates and Ice Sheets Aspects of global Middle Miocene to Pliocene climates and environments have been examined in palaeoclimate modelling studies. For example, the impact that oceanic gateways such as the Panama Gateway (Klocker et al., 2005), the Indonesian Throughflow (Cane and Molnar, 2001) and Tethyan circumglobal passage (Hotinski and Toggweiler, 2003) had on global oceanic circulation patterns and heat/moisture transports has received considerable attention. The role of Tibetan uplift, as well as other regional orographic and palaeogeographic variations, in changing atmospheric circulation patterns and the development of monsoon systems has also been a key area of research (Prell and Kutzbach, 1992; Kutzbach et al., 1993; Fluteau et al., 1999; Bice et al., 2000; Zhisheng et al., 2001; Liu et al., 2003; Kutzbach and Behling, 2004; Ramstein et al., 2005). Reconstructing and predicting the climatic implications of altered distributions of vegetation types (Dutton and Barron, 1996, 1997; Cosgrove et al., 2002; Haywood et al., 2002a; Haywood and Valdes, 2006), sea surface temperature gradients (SST) and latitudinal temperature gradients/heat transports have been studied in detail (Chandler et al., 1994; Sloan et al., 1996; Rind, 1998; Haywood et al., 2000; Barreiro et al., 2006). However, very few palaeoclimate modelling studies have specifically addressed the question of how climates and environments changed on Antarctica during this interval. The majority of global palaeoclimate modelling experiments have used specified Greenland and Antarctic ice volumes based solely on estimates of global sea level. Therefore, modelling studies conducted thus far could be considered as representing a suite of sensitivity experiments that have tested the response of climate models to different initial conditions rather than accurate climate ‘‘retrodictions’’. One example of a palaeoclimate modelling study that did focus on Antarctica is Haywood et al. (2002b). In this study, they attempted to assess the sensitivity of a climate model to different Antarctic Ice Sheet configurations and to address, indirectly at least, the controversy existing over the behaviour of the EAIS during the Pliocene epoch and whether or not it was possible for Nothofagus remains founds in Sirius Group deposits in the Transantarctic Mountains to be Pliocene in age. Three sensitivity experiments were performed for the Pliocene using the UK Meteorological Office’s General Circulation Model (GCM). Each experiment used a different configuration of Antarctic ice conforming to a range of sea-level estimates supported by geological evidence (specifically 35, 25 and 12–15 m higher than modern sea level). Climate outputs from these experiments were used to drive
438
A. M. Haywood et al.
a biome model capable of predicting the equilibrium vegetation state for a given climate forcing. For the experiments using an Antarctic Ice Sheet configuration equivalent to a +12 to 15 and +25 m sea level, the biome model predicted the occurrence of an exclusively tundra type of vegetation on deglaciated regions of Antarctica (Fig. 10.7). In the +35 m experiment, climate ameliorated sufficiently to allow the model to predict deciduous taiga montane forests in areas of Antarctica where fossil Nothofagus material has been recovered from Sirius Group sediments (Fig. 10.7). This work suggested that it is feasible for Nothofagus to have existed in Antarctica during the Pliocene with ice-sheet extents calibrated to global sea-level records. This research had certain limitations relating to the experimental design. The configuration of Antarctic ice was specified within the GCM and the resulting climatologies were only used to drive an offline biome model. A more sophisticated approach would use a coupled climate–ice and vegetation model capable of predicting, rather than being prescribed with, the equilibrium condition for Antarctic ice cover and vegetation with climate. As a step towards this, Francis et al. (2007) dynamically coupled the Topdown Representation of Interactive Flora and Foliage Including Dynamics (TRIFFID) dynamic global vegetation model to the HadAM3 GCM set up for the mid-Pliocene. The results supported the earlier study by Haywood et al. (2002b) since the TRIFFID model predicted tundra vegetation on deglaciated regions of Antarctica. A number of studies have utilized ice-sheet models to predict the response of the EAIS to increasing surface temperatures the results from which, in a few cases, have been related to the debate over how dynamic or stable the Antarctic Ice Sheet was likely to have been during the Middle Miocene to Pliocene (e.g. Barker et al., 1999b). Huybrechts (1993) used a threedimensional ice-sheet model to determine ice-sheet geometries under various kinds of climatic conditions. A surface temperature warming of between 17 and 201C was needed to generate the ice-free corridor over the Pensacola and Wilkes subglacial basins hypothesized by Harwood (1983) and Harwood and Webb (1986). For temperature rises of less than 51C, the model predicted that the EAIS increased in size due to an increase in snowfall. Even with the most favourable model set up for ice loss, a surface temperature rise of 151C was still required to make the Pensacola and Wilkes subglacial basins ice-free which Huybrechts (1993) concluded was unlikely for the interval in question. However, there are a number of limitations inherent in such an approach. First, by assessing the response of the modern EAIS to a certain climatic forcing, an assumption is made that the EAIS during the Middle Miocene and Pliocene was the same as it is today. Other important differences may
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
439
Pliocene control experiment
Biome key:
Pliocene dyanamic experiment
Pliocene stable experiment
Figure 10.7: Predictions of Antarctic vegetation distributions for each of the Pliocene experiments using climatological means derived from the UK Meteorological Office General Circulation Model within a biome model (BIOME 4; Kaplan, 2001).
440
A. M. Haywood et al.
also have existed. For example, the Transantarctic Mountains may have been substantially lower in the past (McKelvey et al., 1991). Kerr and Huybrechts (1999) tested this possibility in an ice-sheet-modelling study and concluded that the surface elevation of individual mountain blocks has only a very local effect on EAIS dynamics. Second, by applying a uniform climate forcing within the model, the possibility for important regional-scale variations in temperature and precipitation, resulting from the differences in Middle Miocene and Pliocene boundary conditions, is ignored. Third, the ice-sheet model used did not include a representation of ice shelves or ice streams. Ice-sheet theory states that the stability of an ice sheet grounded below sea level depends on its surrounding ice shelves that are believed to buffer the ice sheet from oceanic changes (Lingle, 1984). Any weakening of these ice shelves may lead to a runaway process in which the entire marine ice sheet potentially collapses. Support for this hypothesis comes from recent observations of the terrestrial APIS. After the collapse of the major part of the Larsen A ice shelf in 1995, the flow of glaciers, which had fed the former ice shelf, accelerated significantly (De Angelis and Skvarca, 2003). The ice shelves provide a mechanism for linking ocean temperature change with the retreat of ice shelves and grounded ice-sheet instability that may have been important during the Middle Miocene to Pliocene interval. Current research is using an alternative approach to explore the nature and behaviour of the Antarctic Ice Sheet during the Middle Miocene to Pliocene interval. Instead of attempting to reconstruct the actual state of the ice sheet at a specific time, which is extremely difficult given the incomplete geological record and uncertainties inherent in the techniques used for palaeoenvironmental reconstruction, modelling studies are now focusing on developing ensemble predictions which encompass the range of plausible behaviour of the ice sheet for particular periods. For the Pliocene, a suite of palaeoclimate modelling experiments which cover the range of possible boundary conditions, and which have been validated against global Pliocene palaeoenvironmental proxies, are being used to provide three-dimensional thermomechanical ice-sheet models with the required climatological forcings to predict the extent of the both the Greenland Ice Sheet and EAIS. Ranges of different initial conditions are used within ice-sheet models so that icesheet hysteresis can be explored in detail. The dependence of these results on the GCM boundary conditions and the initial state of the ice sheets leads to a range of ice-sheet reconstructions, but it seems likely that the mean state for the Pliocene Greenland Ice Sheet and EAIS were, to some extent, smaller than they are today (Fig. 10.8). For Antarctica, the model results clearly show that the Wilkes and Aurora Basins are the areas of the EAIS that are most susceptible to increased temperatures. Nevertheless, even in the most
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
441
Figure 10.8: Examples of (A) smallest and (B) largest modelled EAIS in equilibrium with reconstructed Pliocene climate. extreme scenario, the predicted EAIS still covers all but the northernmost reaches of the Wilkes subglacial basin, which is suggested to be a source of the diatoms found within the Sirius Group sediments (Hill et al., 2007). Comparison of the model-predicted ice sheet and sea-level mean state to observed Pliocene highstands (e.g. Dowsett and Cronin, 1990) provides a potential match, which may be improved in future modelling exercises that incorporate variability in orbital forcing.
10.8. Summary 10.8.1. Terrestrial and Fjordal Sedimentary Environments of the Transantarctic Mountains and Lambert Glacier Region Two main pre-Quaternary sedimentary sequences recording glacial events are found on land in East Antarctica, the Sirius Group in the Transantarctic Mountains and the Pagodroma Group in the Prince Charles Mountains flanking the Lambert Glacier. The Sirius Group is a mainly terrestrial glacigenic succession, commonly a few tens of metres thick or as a thin drape, but whose age has proved highly controversial, with Pliocene, Middle Miocene or even older ages suggested. The Pagodroma Group is an
442
A. M. Haywood et al.
ice-proximal fjordal sequence up to several hundred metres thick, and is well dated on the basis of in situ marine fossils. All these well-exposed deposits are important in unravelling the history of the EAIS, as they are the nearest deposits to the heart of the ice sheet, and yield valuable information concerning palaeogeography and palaeoclimate. The two regions described above represent the best studied examples of pre-Quaternary onshore glacigenic successions, but there are others elsewhere, e.g. the Antarctic Peninsula and the Grove Mountains (Fang et al., 2004). When a better dated Sirius Group can be linked to the offshore record in McMurdo Sound, and the data from the Prince Charles Mountains fully integrated with the emerging Prydz Bay record, we will have a good understanding of the behaviour of the EAIS since its inception. 10.8.2. The Terrestrial Fossil Record of Antarctic Climate The terrestrial Neogene biota of Antarctica is poorly known with fossils coming only from exposures on the Beardmore Glacier and from boreholes in the Ross Sea region. This will soon change with the recent discoveries of Neogene fossil assemblages in the Dry Valleys (Ashworth, A. C., Lewis, A. R. and Marchant, D. M., personal communication, 2006). What is known so far is that the Neogene biota had a low diversity and was very different than the diverse forested habitats of the early Palaeogene. Evidently, tundra vegetation was present in the Transantarctic Mountains from the Oligocene until at least the Early Miocene and during that time it appears that there was a declining biodiversity (Askin and Raine, 2000; Raine and Askin, 2001). The final extinction of several vascular plants and animals with more complex life histories is assumed to be associated with the shift to the polar desert climate which prevails today in the interior of the continent. Whether or not this immediately followed the mid-Miocene climatic optimum or was as late as the mid-Pliocene is the most outstanding problem in Neogene climatic history in need of resolution. The extinction implies that a climatic threshold was passed in which temperatures and available moisture were significantly reduced after the event than those which supported the tundra biota. The most parsimonious interpretation to explain the occurrence of the Neogene biota is that the organisms were descendents of the Gondwana biota that had existed on the continent for millions of years, long after the fragmentation of the continent. The conclusion that plants and animals continued to inhabit Antarctica for millions of years after the continent became isolated makes their extinction
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
443
an especially significant event and supports the argument that it was associated with a major reorganization of the global climate system. 10.8.3. Terrestrial Environments of West Antarctica Over the past 7 Ma, the APIS has always been relatively thin (few hundred metres). It was also wet based and erosive, at least until about 2 Ma ago. There is also evidence for several ice-poor periods (interglacials) preserved in northern volcanic sequences, particularly in latest Miocene and Early Pliocene. A relatively thin ice cover (probably tens of metres) existed in Ellsworth Land, and it may have locally reached a few hundred metres, but is very poorly dated. Although there are numerous large volcanoes in Marie Byrd Land, each potentially preserving important evidence for past glacial conditions, most are essentially undissected and the only eruptive products available for study are subaerial units. They thus yield minimal useful environmental information other than maximum possible elevations for any former ice sheet(s) that may be hypothesized from other evidence. A glacial cover existed in Marie Byrd Land by 9–8 Ma, at least. It was thin (tens of metres) and formed local icecaps on the large volcanoes. It was wet based at the coast but may have been dry-based (polar) inland. A thicker ice sheet (at least several hundred metres) existed by c. 7–6 Ma, but its thermal regime is unknown. There is also evidence for at least one important episode of overriding by an unusually thick ice sheet in Pliocene or younger time (W1.5 km thick; o3.5 Ma). 10.8.4. The Marine Record of the East Antarctic Ice Sheet Marine deep-sea oxygen isotope records indicate that a period of ice-sheet instability during the Early and Middle Miocene is superimposed on a longterm warming trend that culminated in the late Middle Miocene climatic optimum from B17 to 15 Ma (Miller et al., 1987; Zachos et al., 2001). This phase was followed by a gradual cooling with a major cooling event occurring at the Middle–Late Miocene boundary B14 Ma that continued for B1–2 m.y. (Flower and Kennett, 1995). Long-term variation in the marine oxygen isotopic data (e.g. Hodell and Venz, 1992; Kennett and Hodell, 1995; Zachos et al., 2001) is suggestive of a warming in the earliest Pliocene, culminating in the mid-Pliocene Climatic Optimum at B3 Ma. However, the oxygen isotope ratios do not allow major changes in ice volume on Antarctica. Obliquity-driven oscillations (41 kyr) in
444
A. M. Haywood et al.
d18O of up to 0.6m, which dominate the Pliocene period, while capable of producing B20–30 m sea-level changes (depending on the temperature contribution to the d18O signal), are also considered insufficient to cause large-scale deglaciation of Antarctica without a significant increase in the temperature of the deep ocean. 10.8.5. Marine Evidence for Antarctic Peninsula and West Antarctic Ice Sheets From marine evidence, it is possible to deduce that significant ice-sheet growth in West Antarctica commenced during the Middle Miocene around ca. 14 Ma. In the Ross Sea and Weddell Sea regions, WAIS reached its fully glacial configuration, comparable with today, during the Late Miocene. The APIS established its present size during the later part of the Late Miocene. In the western Bellingshausen and Amundsen seas, the WAIS may not have grown to its full size until the Early Pliocene. Throughout the Late Miocene and Pliocene, WAIS and APIS repeatedly advanced across the adjacent shelves, and during the Early Pliocene, WAIS and APIS may have been more dynamic (Barker, 1995; Bart, 2001; Hepp et al., 2006). Seismic profiles from the West Antarctic continental margin suggest there was no prolonged deglaciation during the late Neogene, but these records cannot be used to determine whether or not the WAIS and APIS decreased drastically in size during short-term interglacial periods. Investigations on sedimentary records drilled at the ODP Leg 178 sites do not corroborate a significant ice volume reduction (Barker et al., 2002; Hillenbrand and Ehrmann, 2005). 10.8.6. Marine Records of the Southern Ocean The widespread presence of hiatuses in Southern Ocean sedimentary cores is one indicator of the reorganization of ocean circulation patterns that occurred during the MMCT, which is believed to have been associated with the expansion of the Antarctic Ice Sheet (e.g. Woodruff and Savin, 1989). Interbasinal offsets in benthic foraminiferal carbon isotopes suggest that the Early Miocene Southern Ocean received warm saline deep water sourced from low latitudes in the Tethyan region (e.g. Woodruff and Savin, 1989). The flow of this warm saline water mass likely ceased in the Middle Miocene, and the consequent reduction in meridional heat transport is a possible trigger for the ice-sheet expansion in the Middle Miocene. Furthermore, a number of Southern Ocean records reconstruct an increase in the strength of
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
445
the ACC which may also have played an important role in the MMCT (e.g. Pagani et al., 2000). During the Late Miocene, Southern Ocean records indicate further highlatitude cooling possibly associated with the establishment and/or growth of the WAIS and an intensification of the ACC (e.g. Kennett and Barker, 1990). By 6 Ma, the relative flux of NADW into the Southern Ocean reached modern proportions, and subsequently intensified yet further, supporting the speculation that the relative climatic warmth of the Early Pliocene was a result of enhanced thermohaline overturn (e.g. Kwiek and Ravelo, 1999). During the Pliocene, biological productivity increased suggesting a reduction in sea-ice extent/coverage at this time (e.g. Gru¨tzner et al., 2005; Hillenbrand and Cortese, 2006). Furthermore, silicoflagellate assemblages have been used to pinpoint three intervals (3.7, 4.3–4.4 and 4.6–4.8 Ma) within the Pliocene when sea surface temperatures in the Southern Ocean were roughly 51 warmer than today (Whitehead and Bohaty, 2003).
10.8.7. Modelling Antarctic Middle Miocene to Pliocene Climates and Antarctic Ice Sheets A number of palaeoclimate modelling studies have been conducted for the Middle Miocene to Pliocene period (e.g. Ramstein et al., 2005). However, only a few of them have focused on reconstructing the climate and environments of Antarctica (see Haywood et al., 2002b; Francis et al., 2007; Hill et al., 2007). A limited number of studies, using a prescribed EAIS within the climate model which is based on global sea-level estimates, suggest that during warm phases of the Pliocene (specifically the mid-Pliocene), climatic conditions of Antarctica could have ameliorated sufficiently to allow tundra forms of vegetation to exist in deglaciated regions. Idealized sensitivity experiments (e.g. Huybrechts, 1993), in which ice-sheet models are forced with different temperature and precipitation regimes, suggest that a surface temperature warming of between 17 and 201C is needed to generate the ice-free corridor over the Pensacola and Wilkes subglacial basins hypothesized by Harwood (1983) and Harwood and Webb (1986) for the Pliocene. For temperature rises of less than 51C, the models predict that the EAIS increased in size due to an increase in snowfall. More recent studies in which climatologies derived from globally evaluated climate modelling simulations for the Pliocene that are used offline to drive an ice-sheet model suggest that the mean state for the Pliocene Greenland Ice Sheet and EAIS were, to some extent, smaller than today. For Antarctica, the model results
446
A. M. Haywood et al.
clearly show that the Wilkes and Aurora Basins are the areas of the EAIS that are most susceptible to increased temperatures. Nevertheless, even in the most extreme scenario, the predicted EAIS still covers all but the northernmost reaches of the Wilkes subglacial basin, which is suggested to be a source of the diatoms found within the Sirius Group sediments (Hill et al., 2007).
ACKNOWLEDGEMENTS This chapter is a contribution to the SCAR programme Antarctic Climate Evolution (ACE) and the BAS core science programme Greenhouse to Icehouse Evolution of the Antarctic Cryosphere and Palaeoenvironment (GEACEP). Acknowledgements for funding and logistical and technical support with the field campaigns and drilling projects addressing Middle Miocene and Pliocene are made to UK Natural Environment Research Council, the US Office of Polar Programs (NSF-OPP), Antarctica New Zealand, the Australian Antarctic Division, the Italian National Antarctic Program (PNRA), the Alfred Wegener Institute for Polar and Marine Research (AWI), Deutsche Forschungsgemeinschaft (DFG) and the Ocean Drilling Program.
REFERENCES Abreu, V. S., & Anderson, J. B. (1998). Glacial eustasy during the Cenozoic: Sequence stratigraphic implications. Am. Assoc. Petroleum Geol. Bull., 82, 1385–1400. Accaino, F., Bo¨hm, G., & Brancolini, G. (2005). Analysis of Antarctic glaciations by seismic reflection and refraction tomography. Mar. Geol., 216, 145–154. Ackert, R. P., & Kurz, M. D. (2004). Age and uplift rates of Sirius Group sediments in the Dominion Range, Antarctica, from surface exposure dating and geomorphology. Global Planet. Change, 42, 207–225. Alonso, B., Anderson, J. B., Diaz, J. I., & Bartek, L. R. (1992). Pliocene–Pleistocene seismic stratigraphy of the Ross Sea: Evidence for multiple ice sheet grounding episodes. In: D. H. Elliot (Ed.). Contributions to Antarctic Research III. Antarctic Research Series. AGU, Washington, D.C., Vol. 56, pp. 93–103. Anderson, J. B., & Shipp, S. S. (2001). Evolution of the West Antarctic Ice Sheet. In: R. A. Bindschadler (Ed.). The West Antarctic Ice Sheet: Behaviour and Environment. Antarctic Research Series, Vol. 77. AGU, Washington, D.C., pp. 45–57. Armienti, P., & Baroni, C. (1999). Cenozoic climatic change in Antarctica recorded by volcanic activity and landscape evolution. Geology, 27, 617–620.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
447
Ashworth, A. C., & Cantrill, D. J. (2004). Neogene vegetation of the Meyer Desert Formation (Sirius Group) Transantarctic Mountains, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 213, 65–82. Ashworth, A. C., Harwood, D. M., Webb, P. N., & Mabin, M. C. G. (1997). A weevil from the heart of Antarctica. Quaternary Proc., 5, 15–22. Ashworth, A. C., & Kuschel, G. (2003). Fossil weevils (Coleptera: Curculionidae) from latitude 85 degrees south, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 191, 191–202. Ashworth, A. C., & Preece, R. C. (2003). The first freshwater molluscs from Antarctica. J. Molluscan Stud., 69, 89–92. Ashworth, A. C., & Thompson, C. (2003). Palaeontology: A fly in the biogeographic ointment. Nature, 423, 135–136. Askin, R. A., & Markgraf, V. (1986). Palynomorphs from the Sirius Formation, Dominion Range, Antarctica. Antarctic J. U.S., 21, 34–35. Askin, R. A., & Raine, J. L. (2000). Oligocene and Early Miocene terrestrial palynology of the Cape Roberts Drillhole CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antartica, 7, 493–501. Bardin, V. I. (1982). Composition of east Antarctic moraines and some problems of Cenozoic history. In: C. Craddock (Ed.). Third Symposium on Antarctic Geology and Geophysics. The University of Wisconsin Press, Madison, WI, USA, pp. 1069–1076. Barker, P. F. (1995). The proximal marine sediment record of Antarctic climate since the Late Miocene. In: A. K. Cooper, P. F. Barker, & G. Brancolini, G. (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. Antarctic Research Series, Vol. 68. AGU, Washington, D.C., pp. 25–57. Barker, P. F., Barrett, P. J., Cooper, A. K., & Huybrechts, P. (1999b). Antarctic glacial history from numerical models and continental margin sediments. Palaeogeogr. Palaeoclimatol. Palaeoecol., 150, 247–267. Barker, P. F., Camerlenghi, A., Acton, G. D., et al. (1999a). Proceedings of the Ocean Drilling Program, Initial Reports, 178 (CD-ROM), available from: Ocean Drilling Program, Texas A&M University, College Station, TX 77845-9547, USA. Barker, P. F., Camerlenghi, A., Acton, G. D., & Ramsay, A. T. S. (Eds) (2002). Proceedings of the Ocean Drilling Program, Scientific Results, 178 (CD-ROM), available from Ocean Drilling Program, Texas A&M University, College Station, TX 77845-9547, USA. Barker, P. F., Kennett, J. P., et al. (1988). Proceedings of the ODP, Initial Reports, 113, Ocean Drilling Program, College Station, TX. Barker, P. F., Kennet, J. P., et al. (1990). Proceedings of the ODP, Scientific Results, 113, Ocean Drilling Program, College Station, TX. Barreiro, M., Philander, G., Pacanowski, R., & Fedorov, A. (2006). Simulations of warm tropical conditions with application to middle Pliocene atmospheres. Climate Dyn., 26, 349–365.
448
A. M. Haywood et al.
Barrett, P. J. (Ed.) (1986). Antarctic Cenozoic History from the MSSTS-1 Drillhole, McMurdo Sound, Antarctica: Wellington, New Zealand, Department of Scientific and Industrial Research Bulletin, Vol. 237. DSIR Publishing, 174 pp. Barrett, P. J. (Ed.) (1989). Antarctic Cenozoic History from the CIROS-1 Drillhole, McMurdo Sound, Antarctica, Department of Scientific and Industrial Research Bulletin, 245. DSIR Publishing, Wellington, 254 pp. Barrett, P. J. (1996). Antarctic palaeoenvironments through Cenozoic times – A review. Terra Antartica, 3, 103–119. Barrett, P. J., Adams, C. J., McIntosh, W. C., Swisher, C. C., & Wilson, G. S. (1992). Geochronological evidence supporting Antarctic deglaciation around 3 million years ago. Nature, 359, 816–818. Barrett, P. J., et al. (Eds) (2000). Studies from the Cape Roberts Project, Ross Sea, Antarctica, scientific results of CRP-2/2A, parts I & II. Terra Antarctica, 7, 1–665. Barrett, P. J., et al. (Eds) (2001). Studies from the Cape Roberts Project, Ross Sea, Antarctica, scientific results of CRP-3, parts I & II. Terra Antarctica, 8, 1–621. Barrett, P. J., & Hambrey, M. J. (1992). Plio-Pleistocene sedimentation in Ferrar Fiord, Antarctica. Sedimentology, 39(1), 109–123. Barron, J., Larsen, B., et al. (1989). Proceedings of the ODP, Initial Reports, 119, Ocean Drilling Program, College Station, TX. Barron, J., Larsen, B., Baldauf, J., & Shipboard Scientific Party (1991). Leg 119, Kerguelen Plateau and Prydz Bay, Antarctica. In: Proceedings of the Ocean Drilling Program, Scientific Results, 119, 1003 pp. Bart, P. J. (2001). Did the Antarctic Ice Sheets expand during the Early Pliocene? Geology, 29, 67–70. Bart, P. J. (2003). Were West Antarctic Ice Sheet grounding events in the Ross Sea a consequence of East Antarctic Ice Sheet expansion during the Middle Miocene? Earth Planet. Sci. Lett., 216(1–2), 93–107. Bart, P. J., & Anderson, J. B. (1995). Seismic record of glacial events affecting the Pacific margin of the northwestern Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. Antarctic Research Series, Vol. 68. AGU, Washington, D.C., pp. 75–95. Bart, P. J., & Anderson, J. B. (2000). Relative temporal stability of the Antarctic Ice Sheets during the late Neogene based on the minimum frequency of outer shelf grounding events. Earth Planet. Sci. Lett., 182, 259–272. Bart, P. J., DeBatist, M., & Jokat, W. (1999). Interglacial collapse off Crary trough mouth fan, Weddell Sea, Antarctica: Implications for Antarctic glacial history. J. Sedimentary Res., 69, 1276–1289. Bart, P. J., Anderson, J. B., Trincardi, F., & Shipp, S. S. (2000). Seismic data from the Northern basin, Ross Sea, record extreme expansions of the East Antarctic Ice Sheet during the late Neogene. Mar. Geol., 166, 31–50. Bart, P. J., Egan, D., & Warny, S. A. (2005). Direct constraints on Antarctic Peninsula ice sheet grounding events between 5.12 and 7.94 Ma. J. Geophys. Res., 110, F04008, doi:10.1029/2004JF000254.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
449
Behrendt, J. C., & Cooper, A. (1991). Evidence of rapid Cenozoic uplift of the shoulder escarpment of the Cenozoic West Antarctic rift system and a speculation on paleoclimate forcing. Geology, 19, 315–319. Bice, K. L., Scotese, C. R., Seidov, D., & Barron, E. J. (2000). Quantifying the role of geographic change in Cenozoic ocean heat transport using uncoupled atmosphere and ocean models. Palaeogeogr. Palaeoclimatol. Palaeoecol., 161, 295–310. Billups, K. (2002). Late Miocene through Early Pliocene deep water circulation and climate change viewed from the sub-Antarctic South Atlantic. Palaeogeogr. Palaeoclimatol. Palaeoecol., 185, 287–307. Billups, K., & Schrag, D. P. (2002). Paleotemperatures and ice volume of the past 27 Myr revisited with paired Mg/Ca and O-18/O-16 measurements on benthic foraminifera. Paleoceanography, 17(1), Art. No. 1003. Bloemendal, J., Ehrmann, W., Hambrey, M. J., McKelvey, B. C., Matthews, R., & Whitehead, J. M. (2003). Geochemical and rock magnetic records from sediments of the Cenozoic Pagodroma Group, Prince Charles Mountains, East Antarctica: Implications for provenance and weathering. Antarctic Sci., 15(3), 365–378. Brady, H., & McKelvey, B. (1979). The interpretation of a tertiary tillite at Mount Feather, Southern Victoria Land, Antarctica. J. Glaciol., 22(86), 189–193. Brady, H., & McKelvey, B. (1983). Some aspects of the Cenozoic glaciation of Southern Victoria Land, Antarctica. J. Glaciol., 29(102), 343–349. Bussetti, M., Spadini, G., van der Wateren, F. M., Cloetingh, S., & Zanolla, C. (1999). Kinematic modelling of the West Antarctic rift system, Ross Sea, Antarctica. Global Planet. Change, 23, 79–104. Camerlenghi, A., Rebesco, M., De Santis, L., Volpi, V., & DeRossi, A. (2002). The Antarctic Peninsula Pacific margin: Modelling flexure and decompaction with constraints from ODP Leg 178 initial results. In: J. A. Gamble, D. N. B. Skinner, & S. Henrys (Eds). Antarctica at the Close of a Millennium. Proceedings of the VIII International Symposium on Antarctic Earth Sciences, Wellington, New Zealand, July 1999, Royal Society of New Zealand, Wellington. R. Soc. N. Z. Bull., 35, 261–267. Cane, M. A., & Molnar, P. (2001). Closing of the Indonesian seaway as a precursor to east African aridification around 3 to 4 million years ago. Nature, 411, 157–162. Cantrill, D. (2001). Nothofagus from the Early Oligocene of Antarctica: Implications for the vegetation history. Terra Antarctica, 8, 1–6. Cape Roberts Science Team (1998). Initial report from CRP-1, Cape Roberts Project, Antarctica. Terra Antarctica, 5, 1–187. Cape Roberts Science Team (1999). Studies from the Cape Roberts Project, Ross Sea, Antarctica. Initial report on CRP-2/2A. Terra Antarctica, 6, 1–173. with supplement 245 pp. Cape Roberts Science Team (2000). Studies from the Cape Roberts Project, Ross Sea, Antarctica. Initial report on CRP-3. Terra Antarctica, 7, 1–209. with supplement 305 pp.
450
A. M. Haywood et al.
Carlquist, S. (1987). Pliocene Nothofagus wood from the Transantarctic Mountains. Aliso, 11, 571–583. Carter, L., Carter, R. M., & McCave, I. N. (2004). Evolution of the sedimentary system beneath the deep Pacific inflow off eastern New Zealand. Mar. Geol., 205, 9–27. Carter, R. M., McCave, I. N., Richter, C., et al. (2000). Proceedings of the ODP, Initial Reports, 181, available from http://www-odp.tamu.edu/publications/ 181_IR/181ir.htm Chandler, M., Rind, D., & Thompson, R. (1994). Joint investigations of the middle Pliocene climate II: GISS GCM northern hemisphere results. Global Planet. Change, 9, 197–219. Chow, J. M., & Bart, P. J. (2003). West Antarctic Ice Sheet grounding events on the Ross Sea outer continental shelf during the middle Miocene. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198(1–2), 169–186. Ciesielski, P. F., et al. (1988). Leg 114. In: Proceedings of the ODP, Initial Reports, 114, 815 pp. Ciesielski, P. F., et al. (1991). Leg 114. In: Proceedings of the ODP, Scientific Results, 114, 826 pp. Ciesielski, P. F., Ledbetter, M. T., & Ellwood, B. B. (1982). The development of Antarctic glaciation and the Neogene palaeoenvironment of Maurice Ewing Bank. Mar. Geol., 46(1–2), 1–51. Convey, P., Gibson, J. A. E., Hillenbrand, C.-D., Hodgson, D. A., Pugh, P. J. A., Smellie, J. L., & Stevens, M. I. (2008). Antarctic terrestrial life – Challenging the history of the frozen continent. Biol. Rev., 83(2), 103–117, doi:10.1111/j.1469185X.2008.00034.x. Cooper, A. K., Barrett, P. J., Hinz, K., Traube, V., Leitchenkov, G., & Stagg, H. M. J. (1991). Cenozoic prograding sequences of the Antarctic continental margin: A record of glacio-eustatic and tectonic events. Mar. Geol., 102, 175–213. Cooper, A. K., O’Brien, P. E., & Richter, C. (Eds) (2004). Proceedings of the ODP, Scientific Results, 188, available from http://www-odp.tamu.edu/publications/ 188_SR/188sr.htm. Cosgrove, B. A., Barron, E. J., & Pollard, D. (2002). A simple interactive vegetation model coupled to the GENESIS GCM. Global Planet. Change, 32, 253–278. De Angelis, H., & Skvarca, P. (2003). Glacier surge after ice shelf collapse. Science, 299, 1560–1562. De Santis, L., Anderson, J. B., Brancolini, G., & Zayatz, I. (1995). Seismic record of late Oligocene through Miocene glaciation on the central and eastern continental shelf of the Ross Sea. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. Antarctic Research Series, Vol. 68. AGU, Washington, D.C., pp. 235–260. De Santis, L., Prato, S., Brancolini, G., Lovo, M., & Torelli, L. (1999). The eastern Ross Sea continental shelf during the Cenozoic: Implications for the West Antarctic Ice Sheet development. Global Planet. Change, 23, 173–196.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
451
Denton, G. H., Prentice, M. L., Kellogg, D. E., & Kellogg, T. B. (1984). Late tertiary history of the Antarctic ice sheet: Evidence from the Dry Valleys. Geology, 12, 263–267. Denton, G. H., Sugden, D. E., Marchant, D. R., Hall, B. L., & Wilch, T. I. (1993). East Antarctic Ice Sheet sensitivity to Pliocene climatic change from a Dry Valleys perspective. Geografiska Annaler, 75A, 155–204. Dowsett, H. J., & Cronin, T. M. (1990). High eustatic sea level during the Middle Pliocene: Evidence from the southeastern US Atlantic coastal plain. Geology, 18, 435–438. Dutton, J. F., & Barron, E. J. (1996). GENESIS sensitivity to changes in past vegetation. Palaeoclimates: Data Model., 1, 325–354. Dutton, J. F., & Barron, E. J. (1997). Miocene to present vegetation changes: A possible piece of the Cenozoic puzzle. Geology, 25, 39–41. Ehrmann, W., Bloemendal, J., Hambrey, M., McKelvey, B., & Whitehead, J. (2003). Variations in the composition of the clay fraction of the Cenozoic Pagodroma Group, East Antarctica: Implications for determining provenance. Sedimentary Geol., 161(1–2), 131–152. Escutia, C., Ba´rcena, M. A., Lucchi, R., Romero, O., & Ballegeer, M. (2007). Early Pliocene circum-Antarctic warming events between 3.5 and 3.7 Ma recorded in sediments from ODP Sites 1165 (Prydz Bay) and 1095 and 1096 (Antarctic Peninsula). In: A.Cooper, C. Raymond, and the ISAES Editorial Team (Eds). Antarctica: A Keystone in a Changing World-Online Proceedings for the Tenth International Symposium on Antarctic Earth Sciences. USGS Open-File Rep. 2007–1047:104. (Extended Abstract) Escutia, C., De Santis, L., Donda, F., Dunbar, R. B., Cooper, A. K., Brancolini, G., & Eittreim, S. L. (2005). Cenozoic ice sheet history from East Antarctic Wilkes land continental margin sediments. Global Planet. Change, 45, 51–81. Exon, N. F., Kennett, J. P., & Malone, M. J. (2004). Leg 189 synthesis: Cretaceous– Holocene history of the Tasmanian Gateway. In: N. F. Exon, J. P. Kennett, & M. J. Malone (Eds). Proceedings of the ODP, Scientific Results, 189, available from http://www-odp.tamu.edu/publications/189_SR/synth/synth.htm Exon, N. F., Kennett, J. P., Malone, M. J., et al. (2001). Proceedings of the ODP, Initial Reports, 189, available from http://www.odp.tamu.edu/publications/ 189_IR/189ir.htm Eyles, N., Daniels, J., Osterman, L. E., & Januszcak, N. (2001). Ocean Drilling Program Leg 178 (Antarctic Peninsula): Sedimentology of glacially influenced continental margin topsets and foresets. Mar. Geol., 178, 135–156. Fang, A., Liu, X., Li, X., & Huang, F. (2004). Sedimentary environments of the Cenozoic sedimentary debris found in the moraines of the Grove Mountains, East Antarctica, and its climatic implications. Prog. Nat. Sci., 14, 223–234. Fitzgerald, P. G. (1994). Thermochronologic constraints on post-Paleozoic tectonic evolution of the central Transantarctic Mountains, Antarctica. Tectonics, 13, 818–836.
452
A. M. Haywood et al.
Florindo, F., Bohaty, S. M., Erwin, P. S., Richter, C., Roberts, A. P., Whalen, P. A., & Whitehead, J. M. (2003). Magnetobiostratigraphic chronology and palaeoenvironmental history of Cenozoic sequences from ODP sites 1165 and 1166, Prydz Bay, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 69–100. Florindo, F., Harwood, D., Levy, R., and SMS Project Science Team (2008). ANDRILL’S success during the 4th International Polar Year. Scientific Drilling, 6, 29–31, doi:10.2204/iodp.sd.6.03.2008. Flower, B. P., & Kennett, J. P. (1995). Middle Miocene deepwater paleoceanography in the southwest Pacific: Relations with East Antarctic Ice Sheet development. Paleoceanography, 10, 1095–1112. Fluteau, F., Ramstein, G., & Besse, J. (1999). Simulating the evolution of the Asian and African monsoons during the past 30 Myr using an atmospheric General Circulation Model. J. Geophys. Res. Atmospheres, 104(D10), 11995–12018. Francis, J. E., Marenssi, S., Levy, R., Hambrey, M., Thorn, V. T., Mohr, B., Brinkhuis, H., Warnaar, J., Zachos, J., Bohaty, S., & DeConto, R. (2008). From Greenhouse to Icehouse – The Eocene/Oligocene in Antarctica 311. In: F. Florindo, & M. Siegert (Eds). Antarctic Climate Evolution. Elsevier, Amsterdam, Vol. 8, pp. 303–362. Francis, J. E., Haywood, A. M., Ashworth, A., & Valdes, P. J. (2007). Tundra environments of the Neogene Sirius Group, Antarctica: Evidence from the geological record and coupled atmosphere–vegetation models. J. Geol. Soc. Lond., 164, 17–322, doi:10.1144/0016-76492005-191. Francis, J. E., & Hill, R. S. (1996). Fossil plants from the Pliocene Sirius Group, Transantarctic Mountains: Evidence for climate from growth rings and fossil leaves. Palaios, 11, 389–396. Gersonde, R., Hodell, D. A., & Blum, P. (Eds) (2003). Proceedings of the ODP, Scientific Results, 177, available from http://www-odp.tamu.edu/publications/ 177_SR/177sr.htm Gersonde, R., Hodell, D. A., Blum, P., et al. (1999). Proceedings of the ODP, Initial Reports, 177, available from http://www-odp.tamu.edu/publications/177_IR/ 177TOC.HTM Gersonde, R., Kyte, F. T., Bleil, U., Diekmann, B., Flores, J. A., Gohl, K., Grahl, G., Hagen, R., Kuhn, G., Sierro, F. J., Voelker, D., Abelmann, A., & Bostwick, J. A. (1997). Geological record and reconstruction of the Late Pliocene impact of the Eltanin asteroid in the Southern Ocean. Nature, 390, 357–363. Gru¨tzner, J., Hillenbrand, C.-D., & Rebesco, M. (2005). Terrigenous flux and biogenic silica deposition at the Antarctic continental rise during the late Miocene to early Pliocene: Implications for ice sheet stability and sea ice coverage. Global Planet. Change, 45, 131–149. Hall, I. R., McCave, I. N., Zahn, R., Carter, L., Knutz, P. C., & Weedon, G. P. (2003). Paleocurrent reconstruction of the deep Pacific inflow during the Middle Miocene: Reflections of East Antarctic Ice Sheet growth. Paleoceanography, 18, 1040, doi: 10.1029/2002PA000817.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
453
Hambrey, M. J., Barrett, P. J., & Powell, R. D. (2002). Late Oligocene and early Miocene glacimarine sedimentation in the SW Ross Sea, Antarctica: The record from offshore drilling. In: C. O’Cofaigh, & J. A. Dowdeswell (Eds). Glacier-Influence Sedimentation on High-Latitude Continental Margins. Special Publication of the Geological Society 203, pp. 105–128. Hambrey, M. J., Ehrmann, W. U., & Larsen, B. (1991). Cenozoic glacial record of the Prydz Bay continental shelf, East Antarctica. In: J. Barron, B. Larsen et al. (Eds). Proc. ODP, Sci. Results, Vol. 119: College Station, TX (Ocean Drilling Program), pp. 77–132. Hambrey, M. J., & McKelvey, B. (2000a). Neogene fjordal sedimentation on the Western margin of the Lambert Graben, East Antarctica. Sedimentology, 47, 577–607. Hambrey, M. J., & McKelvey, B. (2000b). Major Neogene fluctuations of the East Antarctic Ice Sheet: Stratigraphic evidence from the Lambert Glacier region. Geology, 28(10), 887–890. Hambrey, M. J., & Smellie, J. L. (2006). Distribution, Lithofacies and Environmental Context of Neogene Glacial Sequences on James Ross and Vega Islands, Antarctic Peninsula. Geological Society, London, Special Publications 258, pp. 187–200. Hambrey, M. J., Smellie, J. L., Nelson, A. A., & Johnson, J. S. (2008). Late Cenozoic glacier–volcano interaction on James Ross Island and adjacent areas, Antarctic Peninsula region. Geol. Soc. Am. Bull., 120(5–6), 709–731. Hambrey, M. J., Webb, P. N., Harwood, D. M., & Krissek, L. A. (2003). Neogene Glacial record from the Sirius Group of the Shackleton Glacier Region, Central Transantarctic Mountains, Antarctica. GSA Bull., 115(8), 994–1015. Hambrey, M. J., Wise, S. W. Jr., Barrett, P. J., Davey, F. J., Ehrmann, W. U., Smellie, J. L., Villa, G. & Woolfe, K. J. (Eds) (1998). Scientific report on CRP-1, Cape Roberts Project, Antarctica. Terra Antartica, 5, 713 pp. Harwood, D. M. (1983). Diatoms from the Sirius Formation, Transantarctic Mountains. Antarctic J. U.S., 18(5), 98–100. Harwood, D. M. (1986). Diatom Biostratigraphy and Paleoecology with a Cenozoic History of Antarctic Ice Sheets, Ph.D. Thesis, Ohio State University, Columbus, OH. Harwood, D. M., & Webb, P. N. (1986). Recycled marine microfossils from basal debris-ice in ice-free valleys of Southern Victoria Land. Antarctic J. U.S., 21(5), 87–88. Harwood, D. M., & Webb, P. N. (1998). Glacial transport of diatoms in the Antarctic Sirius Group: Pliocene refrigerator. GSA Today, 8(4), 1–8. Harwood, D. M., Florindo, F., Talarico, F., & Levy, R. (2008). Studies from the ANDRILL, Southern McMurdo Sound Project, Antarctica – Initial Science Report on AND-2A. Terra Antartica. Hayes, D. E., Frakes, L. A., et al. (Eds) (1975). Initial Reports of the Deep Sea Drilling Project: Washington, D.C., Vol. 28. US Government Printing Office, 1017 pp.
454
A. M. Haywood et al.
Haywood, A. M., & Valdes, P. J. (2006). Vegetation in a warmer world simulated using a dynamic global vegetation model for the mid Pliocene. Palaeogeogr. Palaeoclimatol. Palaeoecol., 237(2–4), 412–427. Haywood, A. M., Valdes, P. J., Francis, J. E., & Sellwood, B. W. (2002a). Global middle Pliocene biome reconstruction: A data/model synthesis. Geochem. Geophys. Geosyst., 3(1), 1072, doi: 10.1029/2002GC000358. Haywood, A. M., Valdes, P. J., & Sellwood, B. W. (2000). Global scale palaeoclimate reconstruction of the middle Pliocene climate using the UKMO GCM: Initial results. Global Planet. Change, 25(3/4), 239–256. Haywood, A. M., Valdes, P. J., Sellwood, B. W., & Kaplan, J. O. (2002b). Antarctic climate during the middle Pliocene: Mode l sensitivity to ice sheet variation. Palaeogeogr. Palaeoclimatol. Palaeoecol., 182(1–2), 93–115. Hepp, D. A., Moerz, T., & Gru¨tzner, J. (2006). Pliocene glacial cyclicity in a deepsea sediment drift (Antarctic Peninsula Pacific margin). Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 181–198. Herna´ndez-Molina, F. J., Larter, R. D., Rebesco, M., & Maldonado, A. (2006). Miocene reversal of bottom water flow along the Pacific margin of the Antarctic Peninsula: Stratigraphic evidence from a contourite sedimentary tail. Mar. Geol., 228, 93–116. Hill, D., Haywood, A. M., Hindmarsh, R. C. A., & Valdes, P. J. (2007). Characterising Pliocene ice-sheets: Evidence from data and models. In: M. Williams, A. Haywood, J. Gregory, & D. Schmidt (Eds). Deep Time Perspectives on Climate Change: Marrying Biological Proxies & Climate Models. The Micropalaeontological Society Special Publication, The Geological Society, London, pp. 517–538. Hill, R. S., Harwood, D. M., & Webb, P. N. (1996). Nothofagus beardmorensis (Nothofagaceae), a new species based on leaves from the Pliocene Sirius Group, Transantarctic Mountains, Antarctica. Rev. Palaeobot. Palynol., 94, 11–24. Hill, R. S., & Truswell, E. M. (1993). Nothofagus fossils in the Sirius Group, Transantarctic Mountains: Leaves and pollen and their climatic significance. Antarctic Res. Ser., 60, 67–73. Hillenbrand, C. D., & Cortese, G. (2006). Polar stratification – A critical view from the Southern Ocean. Palaeogeogr. Palaeoclimatol. Palaeoecol., 242, 240–252. Hillenbrand, C. D., & Ehrmann, W. (2005). Late Neogene to Quaternary environmental changes in the Antarctic Peninsula region: Evidence from drift sediments. Global Planet. Change, 45, 165–191. Hillenbrand, C.-D., & Fu¨tterer, D. K. (2002). Neogene to Quaternary deposition of opal on the continental rise west of the Antarctic Peninsula, ODP Leg 178, Sites 1095, 1096 and 1101. In: P. F. Barker, A. Camerlenghi, G. D. Acton, & A. T. S. Ramsay (Eds). Proceedings of the ODP, Scientific Results, 178, pp. 1–33 (CDROM), available from Ocean Drilling Program, Texas A&M University, College Station, TX 77845-9547, USA. Hodell, D. A., & Kennett, J. P. (1986). Late Miocene–Early Pliocene stratigraphy and paleoceanography of the South Atlantic and Southwest Pacific Ocean; a synthesis. Paleoceanography, 1, 285–311.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
455
Hodell, D. A., & Venz, K. A. (1992). Toward a high-resolution stable isotopic record of the Southern Ocean during the Plio-Pleistocene (4.8 to 0.8 Ma). In: J. P. Kennett & D. A. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change, Part I. Antarctica Research Series 56. American Geophysical Union, Washington, DC, pp. 265–310. Holbourn, A., Kuhnt, W., Schulz, M., & Erlenkeuser, H. (2005). Impacts of orbital forcing and atmospheric carbon dioxide on Miocene ice-sheet expansion. Nature, 438, 483–487. Hole, M. J., Storey, B. C., & LeMasurier, W. E. (1994). Tectonic setting and geochemistry of Miocene alkalic basalts from the Jones Mountains, West Antarctica. Antarctic Sci., 6, 85–92. Hollister, C. D., Craddock, C., et al. (1976). Initial Reports of the Deep Sea Drilling Project 35, US Government Printing Office, Washington, DC, 929 pp. Hotinski, R. M., & Toggweiler, J. R. (2003). Impact of a Tethyan circumglobal passage on ocean heat transport and ‘‘equable’’ climates. Paleoceanography, 18(1), Art. No. 1007. Huybrechts, P. (1993). Glaciological modelling of the Late Cenozoic East Antarctic Ice Sheet: Stability or dynamism? Geograf. Ann., 75, 221–238. Ishman, S. E., & Reick, H. J. (1992). A Late Neogene Antarctic glacioeustatic record, Victoria Land Basin margin, Antarctica. Antarctic Res. Ser., 56, 327–348. Joseph, L. H., Rea, D. K., Van der Pluijm, B. A., & Gleason, J. D. (2002). Antarctic environmental variability since the Late Miocene: ODP Site 745, the East Kerguelen sediment drift. Earth Planet. Sci. Lett., 201, 127–142. Kaplan, J. O. (2001). Geophysical Applications of Vegetation Modeling, Ph.D. Thesis, Lund University, Lund. Kennett, J. P., & Barker, P. F. (1990). Latest Cretaceous to Cenozoic climate and oceanographic developments in the Weddell Sea, Antarctica: An ocean-drilling perspective. In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, 113, Ocean Drilling Program, College Station, TX, pp. 937–960. Kennett, J. P., & Hodell, D. A. (1995). Stability or instability of Antarctic Ice Sheets during warm climates of the Pliocene? GSA Today, 5(1), 9–22. Kerr, A., & Huybrechts, P. (1999). The response of the East Antarctic Ice Sheet to the evolving tectonic configuration of the Transantarctic Mountains. Global Planet. Change, 23, 213–299. Klocker, A., Prange, M., & Schulz, M. (2005). Testing the influence of the Central American seaway on orbitally forced northern hemisphere glaciation. Geophys. Res. Lett., 32(3), Art. No. L03703. Kutzbach, J. E., & Behling, P. (2004). Comparison of simulated changes of climate in Asia for two scenarios: Early Miocene to present, and present to future enhanced greenhouse. Global Planet. Change, 41(3–4), 157–165. Kutzbach, J. E., Prell, W. L., & Ruddiman, W. F. (1993). Sensitivity of Eurasian climate to surface uplift of the Tibetan Plateau. J. Geol., 101(2), 177–190.
456
A. M. Haywood et al.
Kwiek, P. B., & Ravelo, A. C. (1999). Pacific Ocean intermediate and deep water circulation during the Pliocene. Palaeogeogr. Palaeoclimatol. Palaeoecol., 154, 191–217. Larter, R. D., & Barker, P. F. (1991). Neogene interaction of tectonic and glacial processes at the Pacific margin of the Antarctic Peninsula. In: D. I. M. Macdonald (Ed.). Sedimentation, Tectonics and Eustasy. Special Publication No. 12 of the International Association of Sedimentologists, pp. 165–186. Larter, R. D., & Cunningham, A. P. (1993). The depositional pattern and distribution of glacial–interglacial sequences on the Antarctic Peninsula Pacific margin. Mar. Geol., 109, 203–219. Larter, R. D., Rebesco, M., Vanneste, L. E., Gamboa, L. A. P., & Barker, P. F. (1997). Cenozoic tectonic, sedimentary and glacial history of the continental shelf west of Graham Land, Antarctic Peninsula. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, Part 2. Antarctic Research Series, Vol. 71. AGU, Washington, D.C., pp. 1–27. Lear, C. H., Elderfield, H., & Wilson, P. A. (2000). Cenozoic deep-sea temperatures and global ice volumes from Mg/Ca in benthic foraminiferal calcite. Science, 287, 269–272. LeMasurier, W. E. (2002). Architecture and evolution of hydrovolcanic deltas in Marie Byrd Land, Antarctica. In: J. L. Smellie & M. G. Chapman (Eds). Volcano–Ice Interaction on Earth and Mars. Geological Society of London, Special Publications 202, pp. 115–148. LeMasurier, W. E., & Thomson, J. W. (Eds) (1990). Volcanoes of the Antarctic plate and southern oceans. Antarctic Res. Ser., 48, 1487. LeMasurier, W. E., & Rocchi, S. (2005). Terrestrial record of post-Eocene climate history in Marie Byrd Land, West Antarctica. Geografiska Annaler, 87A, 51–66. LeMasurier, W. E., Harwood, D. M., & Rex, D. C. (1994). Geology of Mount Murphy volcano: An 8 m.y. history of interaction between a rift volcano and the West Antarctic Ice Sheet. Geol. Soc. Am. Bull., 106, 265–280. Lewis, A. R., Marchant, D. R., Kowalewski, D. E., Baldwin, S. L., & Webb, L. E. (2006). The age and origin of the Labyrinth, western dry valleys, Antarctica: Evidence for extensive middle Miocene subglacial floods and freshwater discharge to the Southern Ocean. Geology, 34, 513–516. Lewis, A. R., Marchant, D. R., Ashworth, A. C., Hemming, S. R., & Machlus, M. L. (2007). Major middle Miocene global climate change: Evidence from East Antarctica and the Transantarctic Mountains. Geol. Soc. Am. Bull., 119, 1449–1461. Lingle, C. S. (1984). A numerical model of interactions between a polar ice stream and the ocean: Application to ice stream E, West Antarctica. J. Geophys. Res., 89, 3524–3549. Liu, X. D., Kutzbach, J. E., Liu, Z. Y., et al. (2003). The Tibetan Plateau as amplifier of orbital-scale variability of the East Asian monsoon. Geophys. Res. Lett., 30(16), Art. No. 1839.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
457
Lyle, M., Gibbs, S., Moore, T. C., & Rea, D. K. (2007). Late Oligocene initiation of the Antarctic Circumpolar Current: Evidence from the South Pacific. Geology, 35, 691–694, doi: 10.1130/G23806A.1. Lythe, M. B., Vaughan, D. G., & the BEDMAP Consortium (2000). BEDMAP – Bed Topography of the Antarctic, 1:10 000000 Scale Map. BAS (MISC) 9. British Antarctic Survey, Cambridge. Marchant, D. R., Denton, G. H., Sugden, D. E., & Swisher, C. C. S. III. (1993b). Miocene glacial stratigraphy and landscape evolution of the western Asgard Range, Antarctica. Geografiska Annaler, 75A(4), 303–330. Marchant, D. R., Denton, G. H., & Swisher, C. C. III (1993c). Miocene–Pliocene– Pleistocene glacial history of Arena Valley, Quartermain Mountains, Antarctica. Geografiska Annaler, 75A, 269–302. Marchant, D. R., Denton, G. H., Swisher, C. C. III, & Potter, N. Jr. (1996). Late Cenozoic Antarctic palaeoclimate reconstructed from volcanic ashes in the dry valleys. Geol. Soc. Am. Bull., 108, 181–194. Marchant, D. R., Swisher, C. C. III, Lux, D. R., West, D. P. Jr., & Denton, G. H. (1993a). Pliocene paleoclimate and East Antarctic ice-sheet history from surficial ash deposits. Science, 260, 667–670. Mayewski, P. A., & Goldthwait, R. P. (1985). Glacial Events in the Transantarctic Mountains: A Record of the East Antarctic Ice Sheet Antarctic Research Series, Vol. 36. American Geophysical Union, pp. 275–324. McGinnis, L. D. (Ed.) (1981). Dry Valley Drilling Project: Antarctic Research Series, Vol. 33. American Geophysical Union, Washington, D.C., 465 pp. McGlone, M. S., Mildenhall, D. C., & Pole, M. S. (1996). History and paleoecology of New Zealand Nothofagus forests. In: T. T. Veblen, R. S. Hill, & J. Read (Eds). The Ecology and Biogeography of Nothofagus Forests. Yale University Press, Yale, USA, pp. 83–130. McKelvey, B. C., Harwood, M. J., Harwood, D. M., Mabin, M. C. G., Webb, P. N., & Whitehead, J. M. (2001). The Pagodroma Group – A Cenozoic record of the East Antarctic Ice Sheet in the northern Prince Charles Mountains. Antarctic Sci., 13(4), 455–468. McKelvey, B. C., Webb, P. N., Harwood, D. M., & Mabin, M. C. G. (1991). The Dominion Range Sirius Group: A record of the Late Pliocene–Early Pleistocene Beardmore Glacier. In: M. R. A. Thompson, J. A. Crame, & J. W. Thompson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, UK, pp. 675–682. Mercer, J. H. (1968). Glacial geology of the Reedy Glacier area, Antarctica. Geol. Soc. Am. Bull., 79, 471–486. Mercer, J. H. (1972). Some observations on the glacial geology of the Beardmore Glacier area. In: R. J. Adie (Ed.). Antarctic Geology and Geophysics. Universitetsforlaget, Oslo, Norway, pp. 427–433. Mercer, J. H. (1978). West Antarctic Ice Sheet and CO2 greenhouse effect: A threat of disaster. Nature, 271, 321–325.
458
A. M. Haywood et al.
Mercer, J. H. (1986). Southernmost Chile: A modern analogue of the southern shores of the Ross embayment during Pliocene warm intervals. J. Antarctic States Am., 21, 103–105. Miller, K. G., Fairbanks, R. G., & Mountain, G. S. (1987). Tertiary oxygen isotope synthesis, sea-level history, and continental margin erosion. Paleoceanography, 2, 1–19. Miller, K. G., Wright, J. D., & Fairbanks, R. G. (1991). Unlocking the ice house – Oligocene–Miocene oxygen isotopes, eustasy, and margin erosion. J. Geophys. Res. Solid Earth Planets, 96(B4), 6829–6848. Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E., Sugarman, P. J., Cramer, B. S., Christie-Blick, N., & Pekar, S. F. (2005). The Phanerozoic record of global sea-level change. Science, 310, 1293–1298. Moran, K., Backman, J., Brinkhuis, H., et al. (2006). The Cenozoic palaeoenvironment of the Arctic Ocean. Nature, 441, 601–605, doi: 10.10138/nature04800. Naish, T. R., Powell, R. D., & Levy, R. H. (2007). Initial science results from AND-1B, ANDRILL McMurdo Ice Shelf Project, Antarctica. Terra Antartica, 14(3), 111–328. Nelson, P. H. H. (1975). The James Ross Island Volcanic Group of north-east Graham Land. Br. Antarct. Surv. Sci. Rep., 54, 1–62. Nitsche, F. O., Cunningham, A. P., Larter, R. D., & Gohl, K. (2000). Geometry and development of glacial continental margin depositional systems in the Bellingshausen Sea. Mar. Geol., 162, 277–302. Nitsche, F. O., Gohl, K., Vanneste, K., & Miller, H. (1997). Seismic expression of glacially deposited sequences in the Bellingshausen and Amundsen Seas, West Antarctica. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, Part 2. Antarctic Research Series, Vol. 71. AGU, Washington, D.C., pp. 95–108. O’Brien, P. E., Cooper, A. K., Florindo, F., Handwerger, D., Lavelle, M., Passchier, S., Pospichal, J. J., Quilty, P. G., Richter, C., Theissen, K. M., & Whitehead, J. M. (2004). Prydz Channel fan and the history of extreme ice advances in Prydz Bay. In: A. K. Cooper, P. E. O’Brien, & C. Richter (Eds). Proceedings of the ODP, Scientific Results, 188, available from ohttp://www-odp.tamu.edu/publications/188_SR/016/ 016.htmW. O’Brien, P. E., Cooper, A. K., Richter, C., et al. (2001). Proceedings of the ODP, Initial Reports, 188, available from http://www-odp.tamu.edu/publications/ 188_IR/188ir.htm Pagani, M., Arthur, M. A., & Freeman, K. H. (2000). Variations in Miocene phytoplankton growth rates in the southwest Atlantic: Evidence for changes in ocean circulation. Paleoceanography, 15, 486–496. Passchier, S. (2001). Provenance of the Sirius Group and related upper Cenozoic glacial deposits from the Transantarctic Mountains, Antarctica: Relation to landscape evolution and ice-sheet drainage. Sedimentary Geol., 144, 263–290.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
459
Passchier, S. (2004). Variability in geochemical provenance and weathering history of Sirius Group strata, Transantarctic Mountains: Implications for Antarctic glacial history. J. Sedimentary Res., 74(5), 607–619. Passchier, S., O’Brien, P. E., Damuth, J. E., Januszczack, N., Handwerger, D. A., & Whitehead, J. M. (2003). Pliocene–Pleistocene glaciomarine sedimentation in eastern Prydz Bay and development of the Prydz trough-mouth fan, ODP Sites 1166 and 1167, East Antarctica. Mar. Geol., 199, 179–305. Passchier, S., & Whitehead, J. M. (2006). Anomalous geochemical provenance and weathering history of Plio-Pleistocene glaciomarine fjord strata, Bardin Bluffs formation, East Antarctica. Sedimentology, 53, 929–942. Pfuhl, H. A., & McCave, I. N. (2005). Evidence for late Oligocene establishment of the Antarctic circumpolar current. Earth Planet. Sci. Lett., 235, 715–728. Pickard, J., Adamson, D. A., Harwood, D. M., Miller, G. H., Quilty, P. G., & Dell, R. K. (1988). Early Pliocene marine sediments, coastline and climate of East Antarctica. Geology, 16, 158–161. Prell, W. L., & Kutzbach, J. E. (1992). Sensitivity of the Indian monsoon to forcing parameters and implications for its evolution. Nature, 360(6405), 647–652. Prentice, M. L., & Matthews, R. K. (1991). Tertiary ice sheet dynamics: The snow gun hypothesis. J. Geophys. Res., B, 96, 6811–6827. Quilty, P. G. (1993). Coastal east Antarctic Neogene sections and their contribution to the ice sheet evolution debate. In: J. P. Kennett & D. Warnke (Eds). The Antarctic Paleoenvironment: A Perspective on Global Change. Antarctic Research Series 60, 251–264. Raine, J. I., & Askin, R. A. (2001). Terrestrial palynology of Cape Roberts Project drillhole CRP-3, Victoria Land Basin, Antarctica. Terra Antartica, 8, 389–400. Ramstein, G., Khodri, M., Donnadieu, Y., et al. (2005). Impact of the hydrological cycle on past climate changes: Three illustrations at different time scales. C. R. Geosci., 337(1–2), 125–137. Rebesco, M., Camerlenghi, A., Geletti, R., & Canals, M. (2006). Margin architecture reveals the transition to the modern Antarctic Ice Sheet ca. 3 Ma. Geology, 34, 301–304, doi: 10.1130/G22000.1. Rebesco, M., Larter, R. D., Barker, P. F., Camerlenghi, A., & Vanneste, L. E. (1997). The history of sedimentation on the continental rise west of the Antarctic Peninsula. In: P. F. Barker, & A. K. Cooper (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin, Part 2. Antarctic Research Series, Vol. 71. AGU, Washington, D.C., pp. 29–49. Rebesco, M., Pudsey, C. J., Canals, M., Camerlenghi, A., Barker, P. F., Estrada, F., Lucchi, R., & Giorgetti, A. (2002). Sediment drift and deep-sea channel systems, Antarctic Peninsula Pacific margin. In: D. A. V. Stow, J.-C Faugeres, J. Howe, C. Pudsey, & A. Viana (Eds). Deep-Water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics, Geological Society of London Memoirs, 22, 353–371.
460
A. M. Haywood et al.
Retallack, G. J., Krull, E. S., & Bockheim, J. G. (2001). New grounds for reassessing palaeoclimate of the Sirius Group, Antarctica. J. Geol. Soc. Lond., 158, 925–935. Rind, D. (1998). Latitudinal temperature gradients and climate change. J. Geophys. Res., 103(D6), 5943–5971. Rutford, R. H., Craddock, C., White, C. M., & Armstrong, R. L. (1972). Tertiary glaciation in the Jones Mountains. In: R. J. Adie (Ed.). Antarctic Geology and Geophysics. Universitetsforlaget, Oslo, pp. 239–243. Sandroni, S., & Talarico, F. (2001). Petrography and provenance of basement clasts and clast variability in CRP-3 drillcore (Victoria Land Basin, Antarctica). Terra Antartica, 8, 449–467. Scheuer, C., Gohl, K., Larter, R. D., Rebesco, M., & Udintsev, G. (2006). Variability in Cenozoic sedimentation along the continental rise of the Bellingshausen Sea, West Antarctica. Mar. Geol., 227, 279–298. Schlich, R., Wise, S. W. Jr., et al. (1989). Proceedings of the ODP, Initial Reports, 120, Ocean Drilling Program, College Station, TX. Shevenell, A. E., Kennett, J. P., & Lea, D. W. (2004). Middle Miocene Southern Ocean cooling and Antarctic cryosphere expansion. Science, 305, 1766–1770. Skilling, I. P. (1994). Evolution of an englacial volcano: Brown Bluff, Antarctica. Bull. Volcanol., 56, 573–591. Sloan, B. J., Lawver, L. A., & Anderson, J. B. (1995). Seismic stratigraphy of the Larsen Basin, eastern Antarctic Peninsula. In: A. K. Cooper, P. F. Barker, & G. Brancolini (Eds). Geology and Seismic Stratigraphy of the Antarctic Margin. Antarctic Research Series, Vol. 68. AGU, Washington, D.C., pp. 59–74. Sloan, L. C., Crowley, T. J., & Pollard, D. (1996). Modeling of middle Pliocene climate with the NCAR GENESIS general circulation model. Mar. Micropaleontol., 27(1/4), 51–61. Smellie, J. L. (1999). Lithostratigraphy of Miocene – Recent, alkaline volcanic fields in the Antarctic Peninsula and eastern Ellsworth Land. Antarctic Sci., 11, 362–378. Smellie, J. L. (2000a). Erosional history of the Transantarctic Mountains deduced from sand grain detrital modes in CRP-2/2A, Ross Sea, Antarctica. Terra Antartica, 7, 545–552. Smellie, J. L. (2000b). Subglacial eruptions. In: H. Sigurdsson (Ed.). Encyclopedia of Volcanoes. Academic Press, San Diego, pp. 403–418. Smellie, J. L. (2001a). History of Oligocene erosion, uplift and unroofing of the Transantarctic Mountains deduced from sandstone detrital modes in CRP-3 drillcore, Victoria Land Basin, Antarctica. Terra Antartica, 8, 481–489. Smellie, J. L. (2001b). Lithofacies architecture and construction of volcanoes erupted in englacial lakes: Icefall Nunatak, Mount Murphy, eastern Marie Byrd Land, Antarctica. In: N. Riggs, & J. D. L. White (Eds). Volcaniclastic Sedimentation in Lacustrine Settings, International Association of Sedimentologists Special Publication 3, pp. 73–98. Smellie, J. L. (2006). The relative importance of supraglacial versus subglacial meltwater escape in basaltic subglacial tuya eruptions: An important unresolved conundrum. Earth Sci. Rev., 74, 241–268.
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
461
Smellie, J. L., & Hole, M. J. (1997). Products and processes in Pliocene – Recent, subaqueous to emergent volcanism in the Antarctic Peninsula: examples of englacial Surtseyan Volcano construction. Bull. Volcanol., 58, 628–646. Smellie, J. L., Hole, M. J., & Nell, P. A. R. (1993). Late Miocene valley-confined subglacial volcanism in Northern Alexander Island, Antarctic Peninsula. Bull. Volcanol., 55, 273–288. Smellie, J. L., McArthur, J. M., McIntosh, W. C., & Esser, R. (2006a). Late Neogene interglacial events in the James Ross Island region, northern Antarctic Peninsula, dated by Ar/Ar and Sr-isotope stratigraphy. Palaeogeogr. Palaeoclimatol. Palaeoecol., 242, 169–187. Smellie, J. L., McIntosh, W. C., & Esser, R. (2006b). Eruptive environment of volcanism on Brabant Island: Evidence for thin wet-based ice in northern Antarctic Peninsula during the late Quaternary. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 233–252. Strand, K., Passchier, S., & Na¨si, J. (2003). Implications of quartz grain microtextures for onset Eocene/Oligocene glaciation in Prydz Bay, ODP Site 1166, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 101–111. Stroeven, A., Burckle, L. H., Kleman, J., & Prentice, M. L. (1998a). Atmospheric transport of diatoms in the Antarctic Sirius Group. Pliocene deep freeze. GSA Today, 8, 1–5. Stroeven, A., Burckle, L. H., Kleman, J., & Prentice, M. L. (1998b). Atmospheric transport of diatoms in the Antarctic Sirius Group: Pliocene deep freeze. GSA Today, 8(4), 1–5. Stroeven, A. P. (1997). The Sirius Group of Antarctica: Age and environments. In: C. A. Ricci (Ed.). VII International Symposium on Antarctic Earth Sciences, Siena, Italy. The Antarctic Region: Geological Evolution and Processes, Terra Antartica Publication, pp. 747–761. Stroeven, A. P., & Prentice, M. L. (1997). A case for Sirius Group alpine glaciation at Mount Fleming, South Victoria Land, Antarctica: A case against Pliocene East Antarctic Ice Sheet reduction. Geol. Soc. Am. Bull., 109(7), 825–840. Stump, E., & Fitzgerald, P. G. (1992). Episodic uplift of the Transantarctic Mountains. Geology, 20, 161–164. Sugden, D., & Denton, G. (2004). Cenozoic landscape evolution of the Convoy Range to Mackay Glacier area, Transantarctic Mountains: Onshore to offshore synthesis. Geol. Soc. Am. Bull., 116(7–8), 840–857. Sugden, D. E. (1996). The East Antarctic Ice Sheet: Unstable ice or unstable ideas? Trans. Inst. Br. Geographers Nr., 21, 443–454. Sugden, D. E., Denton, G. H., & Marchant, D. R. (1995). Landscape evolution of the Dry Valleys, Transantarctic Mountains: Tectonic implications. J. Geophys. Res., 100, 9949–9967. Talarico, F., Sandroni, S., Fielding, C. R., & Atkins, C. (2000). Variability, petrography and provenance of basement clasts from CRP-2/2A drillcore (Victoria Land Basin, Ross Sea, Antarctica). Terra Antartica, 7, 529–544.
462
A. M. Haywood et al.
Taylor, J., Siegert, M. J., Payne, A. J., Hambrey, M. J., O’Brien, P. E., Cooper, A. K., & Leitchenkov, G. (2004). Topographic controls on post-Oligocene changes in icesheet dynamics, Prydz Bay region, East Antarctica. Geology, 32, 197–200. Turco, E., Hilgen, F. J., Lourens, L. J., et al. (2001). Punctuated evolution of global climate cooling during the late Middle to early Late Miocene: High-resolution planktonic foraminiferal and oxygen isotope records from the Mediterranean. Paleoceanography, 16(4), 405–423. Van de Flierdt, T., Frank, M., Halliday, A. N., Hein, J. R., Hattendorf, B., Gu¨nther, D., & Kubik, P. W. (2004). Deep and bottom water export from the Southern Ocean to the Pacific over the past 38 million years. Paleoceanography, 19, PA1020, doi:10.1029/2003PA000923. Van der Wateren, F. M., Dunai, T. J., VanBalen, R. T., Klas, W., Verbers, A. L. L. M., Passchier, S., & Herpers, U. (1999). Contrasting Neogene denudation histories of different structural regions in the Transantarctic Mountains rift flank constrained by cosmogenic isotope measurements. Global Planet. Change, 23, 145–172. Vlastelic, I., Carpentier, M., & Lewin, E. (2005). Miocene climate change recorded in the chemical and isotopic (Pb, Nd, Hf) signature of Southern Ocean sediments. Geochem. Geophys. Geosyst., 6, Q03003, doi: 10.1029/2004GC000819. Webb, P.-N., & Harwood, D. M. (1993). Pliocene Nothofagus (southern beech) from Antarctica: Phytogeography, dispersal strategies, and survival in high latitude glacial deglacial environments. In: J. Alden, J. C. Mastrantonio, & S. Odum (Eds). Forest Development in Cold Climates. Plenum, New York, pp. 135–165. Webb, P.-N., Harwood, D. M., Mabin, M. G. C., & McKelvey, B. C. (1996). A marine and terrestrial Sirius Group succession, middle Beardmore Glacier-Queen Alexandra Range, Transantarctic Mountains, Antarctica. Mar. Micropaleontol., 27(1/4), 273–297. Webb, P. N., Harwood, D. M., McKelvey, B. C., Mercer, J. H., & Stott, L. D. (1984). Cenozoic marine sedimentation and ice-volume variation on the East Antarctic craton. Geology, 12, 287–291. Westerhold, T., Bickert, T., & Rohl, U. (2005). Middle to late Miocene oxygen isotope stratigraphy of ODP site 1085 (SE Atlantic): New constrains on Miocene climate variability and sea-level fluctuations. Palaeogeogr. Palaeoclimatol. Palaeoecol., 217(3–4), 205–222. Whitehead, J. M., & Bohaty, S. M. (2003). Pliocene summer sea-surface temperature reconstruction using silicoflagellates from Southern Ocean ODP Site 1165. Paleoceanography, 18, 1075–1086. Whitehead, J. M., Ehrmann, W., Harwood, D. M., Hillenbrand, C.-D., Quilty, P. G., Hart, C., Taviani, M., Thorn, V., & McMinn, A. (2006a). Late Miocene paleoenvironment of the Lambert Graben embayment, East Antarctica, evident from: Mollusc paleontology, sedimentology and geochemistry. Global Planet. Change, 50, 127–147. Whitehead, J. M., Harwood, D. M., McKelvey, B. C., Hambrey, M. J., & McMinn, A. (2004). Diatom biostratigraphy of the Cenozoic fjordal Pagodroma Group,
Middle Miocene to Pliocene History of Antarctica and the Southern Ocean
463
Northern Prince Charles Mountains, East Antarctica. Aust. J. Earth Sci., 51(4), 521–547. Whitehead, J. M., Harwood, D. M., & McMinn, A. (2003). Distal glacimarine facies of the Battye Glacier Formation, Amery Oasis, northern Prince Charles Mountains, Antarctica. Sedimentology, 50, 531–552. Whitehead, J. M., & McKelvey, B. C. (2002). Cenozoic glacigene sedimentation and erosion at the Menzies Range, southern Prince Charles Mountains, Antarctica. J. Glaciol., 48, 226–236. Whitehead, J. M., Quilty, P. G., McKelvey, B. C., & O’Brien, P. E. (2006b). A review of the Cenozoic stratigraphy and glacial history of the Lambery GrabenPrydz Bay region, East Antarctica. Antarctic Sci., 18, 83–99. Whitehead, J. M., Wotherspoon, S., & Bohaty, S. M. (2005). Minimal Antarctic sea ice during the Pliocene. Geology, 33, 137–140. Wilch, T. I., & McIntosh, W. C. (2002). Lithofacies analysis and 40Ar/39Ar geochronology of ice–volcano interactions at Mt. Murphy and the Crary Mountains, Marie Byrd Land, Antarctica. In J. L. Smellie, & M. G. Chapman (Eds). Volcano–Ice Interaction on Earth and Mars, Geological Society, London, Special Publications 202, pp. 237–253. Wilson, G. S. (1993). Ice Induced Sea Level Change in the Late Neogene, Unpublished Ph.D. Thesis lodged in the Victoria University of Wellington Library, 452 pp. Wilson, G. S. (1995). The Neogene Antarctic Ice Sheet. A dynamic or stable feature? Quaternary Sci. Rev., 14, 101–123. Wilson, G. S., Barron, J. A., et al. (2002). The Mount Feather diamicton of the Sirius Group: An accumulation of indicators of Neogene Antarctic glacial and climatic history. Palaeogeogr. Palaeoclimatol. Palaeoecol., 182, 117–131. Wilson, G. S., Harwood, D. M., Askin, R. A., & Levy, R. H. (1998). Late Neogene Sirius Group strata in Reedy Valley, Antarctica: A multiple-resolution record of climate, ice-sheet and sea-level events. J. Glaciol., 44, 437–447. Wise, S. W. Jr., Schlich, R., et al. (1992). Proceedings of the ODP, Scientific Results, 120, Ocean Drilling Program, College Station, TX. Woodruff, F., & Savin, S. M. (1989). Miocene deepwater oceanography. Paleoceanography, 4, 87–140. Wright, J. D., Miller, K. G., & Fairbanks, R. G. (1991). Evolution of modern deepwater circulation: Evidence from the late Miocene Southern Ocean. Paleoceanography, 6, 275–290. Wright, J. D., Miller, K. G., & Fairbanks, R. G. (1992). Early and Middle Miocene stable isotopes: Implications for deepwater circulation and climate. Paleoceanography, 7, 357–389. Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–693. Zhisheng, A., Kutzbach, J. E., Prell, W. L., et al. (2001). Evolution of Asian monsoons and phased uplift of the Himalayan Tibetan plateau since Late Miocene times. Nature, 6833, 62–66.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00011-6
Chapter 11
Late Pliocene–Pleistocene Antarctic Climate Variability at Orbital and Suborbital Scale: Ice Sheet, Ocean and Atmospheric Interactions Tim Naish1,2,, Lionel Carter1, Eric Wolff3, David Pollard4 and Ross Powell5 1
Antarctic Research Centre, Victoria University of Wellington, P.O. Box 600, Wellington, New Zealand 2 GNS Science, P.O. Box 30368, Lower Hutt, New Zealand 3 British Antarctic Survey, High Cross, Madingley Road, Cambridge CB3 OET, UK 4 Earth and Environmental Systems Institute, Pennsylvania State University, University Park, PA 16802, USA 5 Department of Geology and Environmental Geosciences, Northern Illinois University, DeKalb, IL 60115-2854, USA
ABSTRACT Continental margin drill core and seismic data indicate that between 3.0 and 2.5 Ma, high-latitude climate cooling drove both the West and East Antarctic Ice Sheets towards their present expanded cold polar state. Ice margins developed permanent marine termini with ice shelves. Direct physical sedimentary records of Antarctic Ice Sheet variability (e.g. ice-rafted debris, proximal glacimarine cycles) and more distal ocean records of sea-ice distribution (e.g. diatom palaeoecology), thermohaline circulation (e.g. sortable silt), ocean temperatures (e.g. d18O), frontal dynamics and surface circulation (e.g. palaeoecological assemblages and sea-surface temperature (SST) reconstructions) all show a strong covariance with the 41 kyr cycle in Earth’s obliquity between 3 and 1 Ma. Glacial periods are characterised by northward expansion of seasonal sea ice, SSTs up to 61C colder than now, equatorward migration of frontal zones by B5–101
Corresponding author. Tel.: þ64 4 4636197; Fax: þ64 4 4635186;
E-mail:
[email protected] (T. Naish).
466
T. Naish et al.
latitude, intensification of zonal westerly winds, invigorated surface circulation (e.g. Antarctic Circumpolar Current) and intensified abyssal currents. These processes lead the d18O ice volume maximum by 3–7 kyr, at which time Antarctic Ice Sheets were fully extended onto the continental shelf. Antarctic ice volume changes were in part controlled by the effect of Northern Hemisphere glacioeustasy on its marine margin, and this mechanism accounts for much of the orbital variability of the last 2.6 myr. An enigmatic interval of foraminiferal ooze and coccolith-bearing sediments in Weddell Sea and Prydz Bay sediment cores, together with a bioclastic limestone in the Ross Sea at B1 Ma, imply a significant warming and change in ocean chemistry around the periphery of Antarctica. This event, which occurs within the short normal polarity Jaramillo Subchron in Ross Sea cores, is correlated with warm Marine Isotope Stage 31. The anomalous warming implies an increase of 4–61C in SST, possible incursion of Subantarctic Surface Waters and depression of the lysocline – an event that is apparently unique in the last 3 myr. The proximal geological record implies a second cooling step occurred at the Mid-Pleistocene Climate Transition (ca. 900 ka). Glacialinterglacial cycles, spanning the last 1 myr, recovered in drill cores from the Ross Sea and Prydz Bay are dominated by coarse-grained, ice-proximal diamicts of cold, polar subglacial affinity. Open-ocean pelagites and hemipelagites are rare on the shelf. The atmospheric temperature and greenhouse gas records from Antarctic ice cores show a pronounced increase in amplitude at the 100 ka periodicity, that is coherent and in phase with marine temperature records, but lags ice volume signal. The records of greenhouse gases from the Dome C ice core show very strong congruence with many features of the temperature record, and are consistent with CO2 in particular playing a significant role in temperature amplification. This also suggests that the Southern Ocean plays a leading role in controlling the atmospheric concentration of CO2 on glacial-interglacial time scales. Higher dust flux to Antarctica during glacial maxima compared with warm periods is considered to have radiative affects over Antarctica by providing nutrients (e.g. Fe) to the Southern Ocean promoting higher algal productivity and atmospheric CO2 drawdown. Most of the variation in total grounded ice volume in Antarctic glacial-cycle ice sheet models is due to expansion and contraction of grounding lines across continental shelves, mostly in the West Antarctic Ice Sheet (WAIS) Ross and Weddell sectors, and is equivalent to B15–20 m of sea level. The East Antarctic Ice Sheet interior responds in the opposite sense, contracting slightly at Northern Hemispheric glacial maxima due to lower model snowfall rates. In general millennial-scale cycles occur in high-resolution Antarctic ice cores, particularly in methane and temperature records. However, they appear lower amplitude and the timing may differ from their Northern Hemisphere counterparts (Dansgaard–Oeschger or D-O events). Although uncertainties between the age of the gas (methane) and the ice enclosing it in ice cores makes evaluation of the precise timing of inter-hemispheric, millennial-scale cycles difficult, some D-O cycles appear to be synchronous. The pattern of millennial-scale variability superimposed on G-I climate cycles
Late Pliocene–Pleistocene Antarctic Climate Variability
467
in Antarctic ice cores occurs both during and prior to the last 100 kyr cycle. Millennial-scale Antarctic warm periods (A1–A7) are well expressed in SST, sortable silt and d18O records from the southern mid-latitude ocean. One of these events, the Antarctic Cold Reversal, at B14.2–12.4 ka, is associated with an abrupt cooling of B21C in the Southern Ocean, an expansion of ice shelves and sea ice, modest intensification of winds and intensification in deep abyssal inflow along eastern New Zealand through the Pacific gateway.
11.1. Introduction Two major scientific breakthroughs occurred in the late twentieth century that revolutionised our understanding of Earth’s Quaternary climate system and provided new insight into its pattern of behaviour at both orbital and suborbital time scales. (1) In the mid-1970s, the first oxygen isotope records, developed from deepocean sedimentary archives (e.g. Shackleton and Opdyke, 1976), demonstrated that numerous glaciations of the Northern Hemisphere (NH) continents, and associated fluctuations in global sea level, had occurred in response to variations in Earth’s orbital geometry (Hays et al., 1976). Canonical Milankovitch theory (Milankovitch, 1941) had predicted that long-term variations in NH summer insolation influenced by precession (20 kyr cycle) and obliquity (40 kyr cycle) controlled the variability of the NH ice sheets. Subsequent studies indicated that these orbital influences produced globally synchronous responses in: (i) atmospheric circulation (e.g. Petit et al., 1999), (ii) ocean circulation (e.g. Rahmstorf, 2002), (iii) terrestrial climate (e.g. Leroy and Dupont, 1994; Ding et al., 2002), (iv) sea level (e.g. Chappell and Shackleton, 1986; Naish et al., 1998) and (iv) Antarctic ice volume (e.g. Denton et al., 1986). (2) In the late 1980s and early 1990s, the first deep ice core records from Greenland identified a pervasive B1,500 yr long climate cycle manifested as a series of warm phases termed Dansgaard–Oeschger (D-O) interstadials that punctuated the otherwise cold conditions of the glacials (Dansgaard et al., 1993; Hammer et al., 1997). These D-O climate oscillations were also identified in ocean temperature and ice-rafted debris records from NH high-latitude marine sediment cores (e.g. Bond et al., 1997), and indicated a major reorganisation of the NH climate involving switching of the oceanatmospheric system between two principal modes, warm and cold (e.g. Broecker et al., 1990), over decades to just a few years (Severinghaus et al., 1998).
468
T. Naish et al.
Millennial-scale climate variability has been associated with far-reaching, but not necessarily globally synchronous influences. These include: (i) abrupt changes in meridional oceanic overturning (e.g. Broecker, 2000; Van Kreveld et al., 2000; Weaver et al, 2003), fluctuating NH ice volumes (Alley & MacAyeal, 1994) associated with metre-scale changes in sea level (Chappell et al., 1996), atmospheric circulation as recorded by Antarctic ice cores (Blunier et al., 1998; EPICA Community Members, 2006) and the response of Southern Hemisphere (SH) glaciers (e.g. Denton and Hendy, 1994). This burgeoning Quaternary climate dataset, developed largely from NH records, highlighted the dearth of information from Antarctica and the Southern Ocean. A number of important hypotheses and questions began to form concerning the role of Antarctic Ice Sheets and sea ice on ocean circulation and sea level in the Late Pliocene–Pleistocene bipolar glacial world. These included: 1. What is the fundamental response of the Antarctic Ice Sheets to orbital variations? 2. What is the influence of local insolation forcing, and what is the precise phase relationship between NH and SH climatic processes? 3. Did the dramatic millennial-scale, D-O climate cycles of the NH manifest themselves in the SH? 4. If 3 were true, what was the amplitude of the response, and was the SH in-phase or out-of-phase with the D-O cycles? The short duration of these cycles and the inherent uncertainties with age models of the climate proxy records (marine and ice core) being compared is problematic. 5. To what extent did Antarctic ice volume variability at both orbital and suborbital scale modulate or even drive global climate? This question is especially relevant in terms of its influence on seasonal sea-ice distribution, the production of deep water and thermohaline circulation. 6. What has been the contribution of Antarctic ice to global ice volume and sea-level fluctuations through the Quaternary? 7. How have the Antarctic Ice Sheets contributed, and evolved, through major Late Neogene climate thresholds such as: (i) the global cooling B3.0– 2.5 myr associated with the initiation of NH glaciation, and (ii) the onset of the 100 ka climate cycle during the Mid-Pleistocene Transition (MPT). Owing to Antarctica’s remoteness, its extensive sea-ice apron and its extended ice sheet cover for the last 34 million years, such seemingly fundamental questions are only now being addressed, through the recovery of well-dated, deep marine sedimentary and ice core archives that span many glacial–interglacial (G-I) cycles. These records largely stem from three major
Late Pliocene–Pleistocene Antarctic Climate Variability
469
international drilling initiatives: (1) the European Programme for Ice Coring in Antarctica (EPICA) and its predecessor project at Vostok station, (2) the Integrated Ocean Drilling Programme (IODP) and its predecessor Ocean Drilling Programme (ODP), and (3) continental margin drilling programmes from floating ice platforms such as the Dry Valleys Drilling Project (DVDP), Cenozoic Investigations in Ross Sea programme (CIROS), the Cape Roberts Project and the ANDRILL Program. In this chapter we summarise the present state of knowledge of Late Pliocene–Pleistocene Antarctic-Southern Ocean climate behaviour using data and interpretations from Antarctic ice cores (e.g. Vostok and EPICA), Southern Ocean sediment cores (ODP–IODP) and Antarctic continental margin geological drill cores (CIROS-2, Cape Roberts Project, ANDRILL). We also review the advances provided from numerical ice sheet and global climate models, with special emphasis on the potential of the new datasets to be integrated with a new generation of numerical models.
11.2. Glacial Variability from the Continental Margin Geological Record Continuous sedimentary records of Late Pliocene–Pleistocene (last B2.5 myr) glacial variability are generally lacking on the Antarctic continental shelf (defined as the glacially, over-deepened sea floor above the 1,000 m isobath, Fig. 11.1A). This is because: (1) ice grounding during Quaternary ice sheet expansions onto the shelf has produced significant hiatuses, and (2) the ‘riser-less’ rotary coring systems employed by the ODP, and its predecessors, usually result in poor core recovery (o50%) of unconsolidated coarse-grained glacimarine sediments. While, the finergrained pelagic and hemipelagic sedimentary records recovered from deepwater sediment drifts on the continental rise have yielded more continuous records, these do not provide direct physical evidence of past glacial fluctuations. The history of Antarctic margin Cenozoic geological drilling is summarised by Barrett (this volume, Chapter 3). Here we review in more detail the Late Cenozoic intervals of those records. Since the mid-1970s, ship-based continental margin drilling of Late Pliocene–Pleistocene sediments has been focused in three main regions (Fig. 11.1): (1) Antarctic Peninsula (ODP Leg 178; Barker et al., 1999) and Weddell Sea (ODP Leg 113; Kennett and Barker, 1990); (2) Prydz Bay (ODP Leg 119; Barron et al., 1989; ODP Leg 188; O’Brien et al., 2001); (3) Ross Sea (DSDP Leg 28; Hayes and Frakes et al., 1975; Hays et al., 1976). Average core recovery on the continental shelf has ranged from 8%
470
T. Naish et al.
Late Pliocene–Pleistocene Antarctic Climate Variability
471
(Leg 178 sites 1097, 1100, 1102, 1103) to 30% (Leg 119, sites 739–743) (see Fig. 11.1 and this volume, Chapter 3). The continental margin off Wilkes Land is scheduled for drilling by the IODP in January 2009 (IODP expedition 318, Escutia et al., 2006). In parallel with ocean drilling activities, a different drilling approach based on a closed-circulation, continuous wireline coring system used by the minerals industry, was being applied from land and sea-ice platforms in the McMurdo Sound region. The first glacimarine Pleistocene records were recovered by the DVDP (McGinnis, 1981) from the deglaciated mouth of the Taylor Valley with significantly better core recovery (B90%) than ocean drilling. The first attempts at drilling from sea ice were made in the late 1970s with limited success and core recoveries of 50–60% (DVDP-15; Barrett and Treves, 1981; MSSTS-1; Barrett, 1986). During the 1980s and 1990s progressively more sophisticated sea-riser technology was employed by the CIROS (Barrett, 1989) and the Cape Roberts Project (CRP; CRST, 1999) that enabled the first deep Cenozoic records to be recovered (95–98%). Nevertheless, Pliocene–Pleistocene strata remained poorly sampled due to glacial erosion and non-deposition at these sites. A region that had experienced high rates of Pliocene–Pleistocene basin subsidence and sedimentation was identified beneath the McMurdo Ice Shelf in the centre of the Victoria Land Basin (Naish et al., 2006), and was drilled by the ANDRILL Programme’s McMurdo Ice Shelf (MIS) Project in the austral summer of 2006–2007 (Naish et al., 2007). The 1,285 m long AND-1B drill core recovered by the MIS Project provides the most complete (98% core recovery) and continuous sedimentary record of Pliocene–Pleistocene climate and glacial history from the Antarctic continental margin to date (Naish et al., 2008) and is reviewed in more detail below.
11.2.1. Weddell Sea Pliocene–Pleistocene sediments were recovered in the Weddell Sea during ODP Leg 113 from three distinctly different deep-water environments Figure 11.1: (A) Location of ice cores from the Antarctic interior and ocean drill cores (DSDP, IODP, ODP) from the continental margin and rise discussed in text. (B) Location of sea-ice and ice-shelf based geological drill core records from the Victoria Land Basin (VLB) in McMurdo Sound. (C) Geological cross section (A-A’, Fig. 11.1B) across the VLB shows the stratigraphic relationships and age of strata recovered by drill cores in Southern McMurdo Sound.
472
T. Naish et al.
(Fig. 11.1): hemipelagic and biogenic sedimentation on the South Orkney microcontinent influenced by the WAIS (Site 697), a turbiditic sequence in the deep Weddell Basin influenced by both WAIS and the East Antarctic Ice Sheet (EAIS) (Site 694) and hemipelagic and terrigenous sedimentation on the east Antarctic continental margin (site 693). From these sites, Kennett and Barker (1990) were the first to identify a cooling step during the Late Pliocene that they associated with development of the present cold polar thermal regime of the Antarctic Ice Sheets. This is marked between 3 and 2.5 Ma, by a regional reduction in sedimentation rates and diatom abundances, a deterioration in microfossil preservation and a dominance of the sea-ice diatom Eucampia antarctica. At this time cooling led to expansion of WAIS and EAIS to the edge of the continental shelf coincident with the development of the NH continental ice sheets. Leg 113 cores also showed that Quaternary expansions of the Antarctic Ice Sheet displayed pronounced orbital-scale, G-I sedimentary cyclicity characterised by relatively low siliceous biogenic components and higher average particle size during glacial deposition, alternating with larger silt-sized grains and higher diatom abundances during interglacials. Thus, glacial periods were characterised by increased sea-ice cover, lower productivity and stronger bottom water flows (Pudsey et al., 1988; Pudesy, 1990). The converse was true for interglacial periods. With the exception of the G-I cycles in the upper 20 m of Site 697 core, which were correlated with Marine Isotope Stages 10-1, the chronostratigraphy did not allow accurate dating of the Early and Middle Pleistocene, G-I cycles. An enigmatic, 2.5 m thick interval of foraminiferal ooze, bearing coccolith assemblages occurred in a number of Leg 113 cores, and implied an interval of warming and depression of the lysocline. While the age of this calcareous unit remains poorly constrained in the Weddell Sea region, it may correlate with a similar unit reported from Prydz Bay and Ross Sea sediment cores and dated at B1 Ma (see below). 11.2.2. Antarctic Peninsula Poor core recovery from shelf sites during Leg 178 drilling on the Pacific margin of the Antarctic Peninsula resulted in virtually no sedimentary record (Barker et al., 1999) of the progradational wedges that are inferred to have built the continental shelf during Quaternary oscillations of the WAIS (Larter and Barker, 1989; Larter and Cunningham, 1993). Notwithstanding this, the characteristics of the Pleistocene diamicts and glacimarine sediments that were recovered from the shelf sites were consistent with sedimentation
Late Pliocene–Pleistocene Antarctic Climate Variability
473
on the margin since the Last Glacial Maximum (LGM). This observation led Barker et al. (1999) to propose that the topset deposits known as Antarctic Margin seismic sequence unit S1 were deposited under the influence of an expanded and periodically grounded WAIS. Rebesco et al. (2006) correlated a grid of seismic reflection data collected on the Pacific shelf margin after Leg 178 into the more complete and betterdated, continental rise Site 1101. They showed a transition across the seismic sequence unit S2-S1 boundary marked by a change from deep erosion and debris flows to the progradation of relatively stable wedges on the continental slope occurred B3.0 Ma. The timing is coincident with a significant reduction in sedimentation rate over the distal margin area consistent with observations on the other side of the Antarctic Peninsula in the Weddell Sea (Kennett and Barker, 1990). Both studies argued that the change in depositional style corresponds to major ice sheet expansion onto the continental shelf associated with reduced sediment supply and meltwater volume as the Antarctic Ice Sheets attained their present cold polar state. Drilling of continental rise sediment drifts at sites 1095 and 1096 produced a more continuous record of Pleistocene G-I variability expressed as lithologic alternations between interglacial biosiliceous silts and glacial nonfossiliferous, terrigenous laminated clays and silts. Intriguingly, time-series analysis of the properties of G-I cycles at sites 1095 and 1096 shows no periodicity in the orbital frequency bands. However, iceberg-rafted debris (IRD) mass accumulation rates in the Pliocene–Pleistocene drift record at Site 1101 show strong covariance with the B40 kyr, obliquity frequency between 3 and 1 Ma, after which 100 kyr frequencies dominate (Cowan, 2002). Two large spikes in IRD abundance at 2.8 and 0.88 Ma have been interpreted as corresponding to Late Pliocene cooling and ice sheet expansion onto the shelf, and a second phase of cooling and ice expansion across the MPT (see below). The occurrence of orbitally influenced ice rafting in records around the Antarctic Peninsula from B2.8 Ma has led a number of workers to suggest a NH eustatic influence on the Quaternary stability of the Antarctic Ice Sheet margin (e.g. Barker et al., 1999; Cowan, 2002; Hillenbrand and Ehrmann, 2005). 11.2.3. Prydz Bay Drilling was undertaken at five sites (739–743, Fig. 11.1) during ODP Leg 119 on a transect across the continental shelf of Prydz Bay to elucidate the long-term glacial history and evolution of the EAIS (Barron et al., 1989). Although the late Pliocene to Pleistocene interval is limited by poor core
474
T. Naish et al.
recovery and inadequate age control, a cyclic record of tills and ice proximal, glacimarine sediments implies a succession of advances and retreats by the EAIS across the region (Hambrey et al., 1991). During Leg 188, two sites sampled Late Pliocene–Pleistocene strata (O’Brien et al., 2001). Site 1167 was the first hole to be drilled in an Antarctic margin trough mouth fan, and the core provides an expanded view of glacier fluctuations in the Prydz Bay region during the last 2 Ma. The sedimentary record is dominated by several metres thick diamictons deposited as debrites separated by thin mud units (O’Brien et al., 2007). The debrites represent slumping and remobilisation of glacial outwash deposits and tills when the EAIS grounding line is near the shelf edge. Intervening interglacials are represented by glacimarine muds deposited by the retreating ice sheet. While the diamicton-mud cycles maybe of orbital duration, hiatuses within the section and insufficient age control preclude the duration of cycles from being established. O’Brien et al. (2007) also pointed out that the trend to expanded ice in the Prydz Bay region during the Late Pliocene maybe associated with the movement of interior precipitation to the coast, either from changes in storm tracks or the increase in ice around the EAIS margin. Moreover, they argue that this process could have caused the change to more laterally confined ice streams since the early Pliocene. A significant change in sedimentation rate and sediment composition suggests a colder and less erosive ice sheet developed following the MPT at about 0.78 Ma. Three of the sedimentary cycles above the Bruhnes– Matuyama (B-M) magnetic reversal boundary (0.78 Ma) correspond to three cycles in a d18O record from Neogloboquadrina pachyderma at Site 1167 that are tentatively correlated with Marine Isotope Stages 21-16 and indicate orbital control on ice sheet fluctuations (Thiessen et al., 2003). Site 1165 sampled sediment drift deposits on the continental rise and comprises a strongly cyclic Pleistocene interval in its upper 50 m in particle size, grey-scale and clay mineral records (O’Brien et al., 2007). A recent revision of the Pleistocene bio- and magnetostratigraphy of the sites show an incursion of calcareous nannofossils in the Site 1165 core at around 1 Ma (Villa et al., 2007), coincident with a decrease in d18O values in records from both 1167 and 1165 (Thiessen et al., 2003). Late Pliocene–Early Pleistocene glacimarine deposits are described on land from the uplifted sequences of the Bardin Bluffs Formation (uppermost formation of the Pagodroma Group) in the Prince Charles Mountains (e.g. Whitehead and McKelvey, 2001). While the formations within this group are geographically separated they have been correlated into a composite stratigraphic section on the basis of diatom biostratigraphy (e.g. McKelvey et al., 2001; Whitehead et al., 2004). Bardin Bluffs Formation
Late Pliocene–Pleistocene Antarctic Climate Variability
475
sediments mark a dramatic change from older warmer intervals of diatombearing glacimarine sediments to coarse-grained, ice-proximal diamicts overlying a striated erosion surface above the Amery Group. 11.2.4. Ross Sea The pioneering coring expedition of DSDP Leg 28 in the Ross Sea drilled four sites on the continental shelf: Sites 270–272 in the south eastern quadrant of the Ross continental shelf and Site 273 in the west-central portion of the shelf (Hayes and Frakes et al., 1975) (Fig. 11.1). While core recovery was low (B30%), a composite Pliocene–Pleistocene record of W200 m was sampled at sites 270–272 comprising terrigenous silts and biosiliceous-bearing muds and diatomaceous ooze with ice-rafted debris throughout. A cyclostratigraphic interpretation was not reported for these sites, but a major glacial unconformity of inferred Late Pliocene age identified in Ross Sea seismic data, and correlated into Site 270, is interpreted as widespread westward expansion of the WAIS ice streams into the Ross Sea (Bart, 2004). Rebesco et al. (2006) have correlated this event with evidence elsewhere for continentwide development of a marinebased ice margin on the continental shelf, associated with development of a thicker, cold ice sheet similar to the present-day thermal regime. Immediately prior to the coring of Leg 28 the Dry Valley Drilling Project (DVDP) recovered the first Pliocene–Pleistocene glacimarine sediment cores from the continent (McGinnis, 1981). A B200 m thick composite section based on DVDP Sites 10 and 11 in the Lower Taylor Valley, comprised a succession of ice-proximal Late Pliocene glacimarine diamictons, sands and gravels passing up-section into fluvio-delatic and beach sands of Late Pleistocene age (McKelvey, 1981; Powell, 1981). Lateral facies relationships and correlations between the two sites indicate that the down valley Site 10 was most proximal to the grounding line. This relationship together with the occurrence of McMurdo volcanic clasts in diamictons implied an eastern ice source during glacial advances associated with expansions of the WAIS (Porter and Beget, 1981). McMurdo volcanic-bearing diamictons occur above an unconformity at 203 m depth and 146 m depth in DVDP 11 and 10, respectively, corresponding to the proximity of the Gauss–Gilbert polarity transition (2.6 Ma) (Elston and Bressler, 1981). Thus, although some uncertainties surround the chronology of these cores remain, expansion of WAIS-sourced ice into the Dry Valleys is consistent with timing of Late Pliocene ice expansion elsewhere. Barrett and Hambrey (1992) report a Late Pliocene–Pleistocene succession of McMurdo volcanic-bearing tills in
476
T. Naish et al.
the upper 100 m of the CIROS-2 core from Ferrar Fiord. They interpret the succession to represent the onset of ice advance from the Ross/WAIS Ice Sheet at around 3–2.5 Ma. 11.2.5. The ANDRILL Programme: McMurdo Ice Shelf Record of Late Pliocene to Pleistocene Climate Variability During the austral summer of 2006–2007, a new Antarctic geological drilling programme, ANDRILL, successfully recovered a 1,285 m long record of strata (AND-1B) from beneath the McMurdo Ice Shelf sector of the Ross Ice Shelf. This record spans the last 14 myr (Fig. 11.2; Naish et al., 2007).
11.2.5.1. Orbital-scale, glacial–interglacial sedimentary cycles Forty unconformity-bound glacimarine cycles in the upper 600 m of the core record orbitally influenced fluctuations in the areal extent of the WAIS margin in the Ross Embayment, and the evolution of the whole Antarctic Ice Sheet from a period of relative global warmth in the Early Pliocene through a profound cooling step in deep-sea oxygen isotope records B3–2.5 Ma, to the development of the present cold polar Antarctic thermal regime during the last B1 million years (Naish et al., 2007, 2008). The rocks were interpreted in terms of lithofacies associations – sediments representing specific environments of deposition. These ranged from open marine diatomites, mudstones and turbidites deposited during interglacials to ice-proximal massive and stratified diamictites, conglomerates and sandstones representing glacial periods. During glacial periods the ice sheet had a laterally extensive marine terminus extending well out into the Ross Sea beyond the drill site. In interglacials the drill site was covered by either an ice shelf (similar to present day), or lay in open water when the ice sheet margin had retreated onto the continent, with local deposition of marine diatoms, terrigenous mud and occasional debris from floating ice.
11.2.5.2. Chronology A preliminary age model for the upper 700 m of drill core constructed from diatom biostratigraphy and radiometric ages on volcanic material allows a unique correlation between B70% of the magnetic polarity stratigraphy and the Geomagnetic Polarity Time Scale (GPTS) (Wilson et al., 2007).
Figure 11.2: Lithological column of the upper 600 m of the AND-1B drill core recovered by the ANDRILL MIS Project. Correlation of biostratigraphically constrained palaeomagnetic reversals (Wilson et al., 2007) with the oxygen-isotope time scale (Lisiecki and Raymo, 2005) are shown. Glacial surfaces of erosion mark boundaries of orbital-scale, glacimarine sedimentary cycles (Naish et al., 2008). Density, clast abundance and glacial proximity curves highlight the glacial and sedimentary cyclicity of G-I cycles. I, Ice Contact; P, Proximal Glacimarine; D, Distal Glacimarine; M, Open Marine.
478
T. Naish et al.
The age model provides several well-constrained intervals displaying relatively rapid (o1 m/1,000 yr) and continuous accumulation of sediment punctuated by several 0.5–1 myr stratal hiatuses representing more than half of the last 7 myr. Thus, the AND-1B record provides several highly resolved ‘windows’ into the Late Cenozoic development of the Antarctic Ice Sheets.
11.2.5.3. Implications for late Cenozoic ice sheet history The Pliocene period (5–2 Ma) is characterised by a dynamic ice margin with interglacials displaying pelagic diatomite, implying high phytoplankton productivity in locally open water. A W80 m thick interval of diatomite between 370 and 460 m below sea floor (mbsf) represents an extended period of open water in the Ross Embayment and high phytoplankton productivity. Two erosional unconformities of greater than 100 kyrs dated at B3.3 and 2.7 Ma enclosed by diamictites occur at B285 and 250 mbsf, and are interpreted to represent stepwise cooling, expansion of the WAIS and glacial erosion in Western Ross Embayment. The timing leads and is coeval with the development of the NH ice sheets and evidence for ice expansion on other regions of the Antarctic continental shelf already discussed. (e.g. Rebesco et al., 2006). Stratigraphic analysis and a new high-resolution integrated chronostratigraphy (Wilson et al., 2007) for the overlying Late Pliocene AND-1B cycles, particularly the abrupt transitions from diamictites to biosiliceous deposits, implies rapid oscillations between glacial and marine environments with significant volume changes in the WAIS at a frequency of 41 kyr. Intriguingly, the facies also imply a cooler style of ice sheet compared with the Early Pliocene cycles, indicating reduced amounts of subglacial meltwater and local terrigenous sediment input from the McMurdo Sound area (McKay et al., accepted). Naish et al (submitted) propose that between 3.5 and 2.5 Ma, high-latitude climate cooling drove both the WAIS and EAIS towards their present expanded cold polar state. Relatively warm, often terrestrially based ice margins were replaced by more permanent marine termini and the development of ice shelves. Mass balance changes affecting these colder ice sheets may have occurred primarily through calving processes at the marine margins, rather than melting. From B2.6 Ma polar Antarctic ice volume changes may have became significantly influenced by NH, glacio-eustasy and this mechanism may account for much of the orbital variability of the WAIS since B2.6 Ma.
Late Pliocene–Pleistocene Antarctic Climate Variability
479
A dramatic change in sedimentary cycle architecture occurs above an erosion surface at 82.74 mbsf in the AND-1B core, that is more or less coincident with the B-M polarity transition, 0.78 ka. Eight fluctuations in the proximity of WAIS margin, expressed as, polar-style sedimentary cycles, occur within the interval above the B-M boundary and correspond to an interval of eight, 100 kyr duration d18O oscillations between Marine Isotope Stages 20 and 2. These diamictite-dominated cycles indicate a predominance of a cold polar ice sheet over the drill site for most of the Late Pleistocene, which retreated to form an ice shelf during interglacial minima. The erosional surface beneath these cycles removes about B100 kyr of the upper part of the Matuyama Chron above a normal polarity interval of biosiliceous-bearing mudstone assigned to the Jaramillo Subchron (Wilson et al., 2007). An anorthoclase-bearing tuff within the mudstone has yielded a 40Ar/39Ar age of 1.01470.004 Ma (Wilson et al., 2007) and indicates that this period of open marine pelagic sedimentation in western Ross Sea occurred during the warm interglacial of Marine Isotope Stage 31 providing the last evidence for an open Ross Embayment in the AND-1B record. The calving line of the Ross Ice Shelf appears not to have retreated south of its present interglacial position during subsequent ‘warmer-than-Holocene super-interglacials’ (e.g. Marine Isotope Stages 11, 9 and 5; Jouzel et al., 2007). These observations of Late Pleistocene relative stability of the Ross Ice Shelf seem incompatible with the geological evidence for Late Pleistocene collapse of the WAIS (e.g. Scherer et al., 1998), and remnant shorelines 20 m above present sea level during these interglacials (Hearty et al., 1999).
11.3. Atmospheric Variability from Ice Cores So far ice cores reach only through the last third (800 ka) of the Pleistocene epoch (Jouzel et al., 2007). However, for that period they provide iconic datasets that can help to unlock some aspects of Antarctic climate evolution. This is because they supply a record of Antarctic climate (temperature and precipitation) that drives the evolution of the ice sheet, and they also supply the records of the most important forcings (especially greenhouse gases, but less directly other forcings); these have in turn presumably controlled the climate. One thing ice cores cannot constrain so well is the response of the ice sheets to climate change, and that is the value of the continental shelf and deep marine geological records, and the integration of both.
480
T. Naish et al.
11.3.1. The Nature of the Ice Core Record Ice cores are a sedimentary record like many others, but comprise snow and ice rather than terrigenous particles, with layers deposited sequentially, although unlike other records the layers thin with depth due to ice flow. They have the benefit that they record the atmosphere (including sampling precipitation, aerosol and trace gases) more directly than any other archive, without biological mediation. Highly resolved records, potentially datable by counting annual layers, are recovered at sites with high snow accumulation rates. The longest records extend back hundreds of thousands of years. However, at these sites annual layers cannot be resolved and far more difficult exercises are needed in order to derive a reliable age scale (e.g. Parrenin et al., 2007a). A particular strength of ice cores is that so many aspects of climate forcing and response are recorded in one core, while a weakness (less obvious for this book that concentrates on Antarctic climate) is that most ice core records come from the polar regions (particularly Greenland and Antarctica). The ice core record is held in three distinct forms. Firstly, the water molecules themselves, through their isotopic content, contain a record of the 16 temperature at the time the snow fell. The ratio of H18 2 O/H2 O, or of HDO/ 2 H2O (where D is deuterium, H) is generally considered to be a good proxy for site temperature, and the primary record of most ice core studies. This arises from the fact that, after water vapour has been evaporated from a warm ocean, it steadily loses water as it cools on the way to the poles, with the heavier isotopes having a lower vapour pressure and, therefore, condensing more readily, which leaves an airmass depleted in heavy isotopes to continue towards the poles. Although the real relationship is of course more complicated (Jouzel et al., 2003), in essence this leads to more negative isotopic values (expressed as per mil deviations from the composition of ocean water) in colder climates. As an aside, it is worth noting that it is the buildup of ice with negative isotope values in the ice sheets that leads to positive isotopic excursions in the oceans during colder (more glaciated times). When snow falls, it takes with it aerosol particles and some gases that have an affinity for water. More are added just after deposition. It is in this part of the record, of material trapped in surface snow, that we find for example the spikes of sulphuric acid that are found in snow for a few years after major volcanic eruptions, as well as components of sea salt and terrestrial dust. Finally, snow turns to solid ice in the cold polar regions only under the pressure of overlying ice that sinters the snow grains together. This is the process that seals samples of atmospheric air in air bubbles, typically at
Late Pliocene–Pleistocene Antarctic Climate Variability
481
60–100 m depth, and it is in these bubbles that all the stable trace gases, including carbon dioxide and methane, are found. As already discussed, most ice cores come from the polar regions. In Greenland, the oldest continuous ice core so far is from North GRIP (North Greenland Ice Core Project Members, 2004), at 123 kyr, reaching just into the last interglacial. Several other cores, spanning the latitudes of Greenland, cover a slightly shorter period continuously. In Antarctica (Fig. 11.1), the oldest ice cores, spanning several G-I cycles, come from the East Antarctic plateau sites with very low snow accumulation rates. For many years, Vostok, spanning four glacial cycles and 420 kyr (Petit et al., 1999), was the oldest ice core, and it remains the longest in terms of depth. A slightly shorter period has so far been published from Dome Fuji (Watanabe et al., 2003), although a new core has recently extended the age span towards that of Dome C. The European EPICA ice core at Dome C now contains the longest (in age) continuous sequence, at just over 800 kyr (EPICA Community Members, 2004; Jouzel et al., 2007). A number of other cores have been drilled in the coastal regions of Antarctica that reach beyond the LGM (e.g. Law Dome, Siple Dome, Berkner Island); these are important because they hold the potential to derive records of climate and ice sheet size in the outlet regions of the continent. Finally, central West Antarctica is so far represented only by the Byrd core, but a new effort from the US Antarctic Programme (WAIS Divide Project) promises a new high-quality core in the next few years.
11.3.2. Antarctic Climate and Forcing at Glacial–Interglacial Time scales Rather than give a historical account of increasing knowledge in this area (which would focus strongly on Vostok), we will here use the longest (in age) record from Dome C to describe what is known, and then discuss what the similarities and differences of other records indicate. Figure 11.3 shows the record (using water isotopes) of Antarctic temperature at Dome C over the last 800 ka (Jouzel et al., 2007). The age scale has been derived using an ice flow and snow accumulation model constrained, and in some parts tuned, to independent age markers of known age (Parrenin et al., 2007a). The maximum uncertainty of the absolute ages is estimated to be better than 6 kyr throughout the core. Through most of its span it shows good coherence with signals recorded in the benthic marine oxygen isotope stack (Fig. 11.3, representing some combination of global ice
482
T. Naish et al.
δ18O / ‰
2.5
5.5
15.1
19.3
3.0 3.5 4.0 LGM
-360 δD / ‰
11.3
-390 -420 -450 4
ΔT / °C
0 -4 -8 -12
0
200
400
600
800
Age / EDC3/LR04 ka bp (1950)
Figure 11.3: Top panel: marine benthic oxygen isotope stack on LR04 age scale (Lisiecki and Raymo, 2005). Lower panels: 800 kyr record, on EDC3 age scale of Deuterium, and temperature (difference from the last millennium, derived using a correction for seawater isotopic content and modelled ice-sheet altitude) (Jouzel et al., 2007). Numbers above the marine curve represent selected Marine Isotope Stages. volume and deep-water temperature) whose time scale was derived in a completely different way (Lisiecki and Raymo, 2005). The first obvious feature of the record is that a generally cold climate (up to 91C colder than the late Holocene in this part of Antarctica) was interrupted approximately every 100 kyr by warmings (such as the Holocene interglacial) lasting approximately 10–30 kyr. The last four interglacials (back to Marine Isotope Stage 11) had periods that were 2–4.51C warmer than the last millennium. Before Marine Isotope Stage 11, a different pattern is seen: although 100 kyr still emerges as the dominant period of the record (Jouzel et al., 2007), the earlier interglacials are significantly cooler than the later ones, and the system seems to spend a higher proportion of time in the warmer phase. There is not yet any convincing explanation for this change in amplitude of the signal (discussed below). The next question that arises is whether the Dome C record can be considered representative for a larger part of the Antarctic. Comparison with
Late Pliocene–Pleistocene Antarctic Climate Variability
483
Vostok (around 500 km away) shows excellent agreement, but more importantly a very similar signal (albeit with small differences in the age scale) is seen at Dome Fuji, on the opposite side of the East Antarctic plateau (Watanabe et al., 2003; Kawamura et al., 2007). This similarity suggests that the Dome C record can stand as a signal of climate across much of the polar plateau. The 100 kyr periodicity seen in the ice core record is of course already well-known from other climate records, and assumed to derive in some way from orbital forcing amplified by internal feedbacks. One of the most important amplifiers is the greenhouse gas concentration of the atmosphere. The records of greenhouse gases from Dome C (completed by those from Vostok in as yet unmeasured sections of Dome C) extend so far to 650 ka (and are currently being continued to 800 ka). They show very strong congruence with many features of the temperature record, and are consistent with CO2 in particular playing a significant role in such amplification (Figs. 11.3 and 11.4). Concentrations of CO2 are typically around 180–200 ppmv in the coldest parts of each cycle, and reached 280–300 ppmv during the warm periods back to Marine Isotope Stage 11 (Siegenthaler et al., 2005). They were however considerably lower (B250 ppmv) during the warm periods before Marine Isotope Stage 11, scaling with Antarctic temperature (Fig. 11.4). The very strong similarity of CO2 concentrations and Antarctic temperature strongly suggests that the Southern Ocean plays a leading role in controlling the atmospheric concentration of CO2 on these time scales, as already implied by more detailed records of the last deglaciation (Monnin et al., 2001). Concentration of CO2 in the atmosphere today (ca. 380 ppmv) is nearly 30% higher than the highest value seen in the previous 650 kyr. Methane (CH4) concentrations (Spahni et al., 2005) show strong similarities to the Antarctic temperature record (Fig. 11.4), with high concentrations in warm periods, low concentrations in glacial maxima and lower concentrations in ‘weaker’ interglacials compared to ‘stronger’ ones. However in this case, even at the low resolution shown, it is apparent that there are higher-frequency millennial-scale variations superimposed on the orbital-scale trends. In the most recent climatic cycle (Blunier and Brook, 2001) these correspond to the D-O events that will be discussed later. Methane concentrations are most likely mainly controlled by changing wetland extents in the northern high latitudes and tropics, along with changes in sinks (Valdes et al., 2005), and thus are relevant to Antarctic climate only as a relatively minor amplifier. Ice cores have provided other data that are important for understanding Antarctic climate and ice sheets. The snow accumulation rate is clearly
484
T. Naish et al.
Age / EDC3 ka
nss Ca flux / μg m-2 a-1
δD / ‰
-390 -420 1000
100
CH4 / ppbv
10 700 600 500 400
CO2 / ppmv
300 270 240 210 180 0
200
400
600
Age / EDC2 ka (gases)
Figure 11.4: Records of Deuterium, non-sea-salt Calcium flux (Wolff et al., 2006), CO2 (Siegenthaler et al., 2005) and CH4 (Spahni et al., 2005) measured at Dome C. Note that there maybe small timing mismatches because the gases are shown on the EDC2 age scale, while the other records are on EDC3. important for the mass balance of the ice sheet. In central Antarctica, it has generally been treated by a modelling approach in which the accumulation rate is thermodynamically related to the local temperature. Snow accumulation scales non-linearly with temperature, and was around half its present value in central East Antarctica during the coldest periods of the last 800 kyr than it is at present. Ice sheet elevation in central Antarctica is assumed to have varied little over the last 800 kyr: simple modelling exercises (Parrenin et al., 2007b) used as input to age models suggest variations of less than 200 m at both Dome C and Dome Fuji. Measurements of air volume (which has atmospheric pressure as one among its controlling variables) provide only a limited constraint, confirming that altitude changes of more than a few
Late Pliocene–Pleistocene Antarctic Climate Variability
485
hundred metres have not occurred. Of course, it is possible for the areal extent of an ice sheet to change considerably without any great change in the altitude of its central parts. The dust flux to Antarctica (represented in Fig. 11.4 by non-sea-salt calcium) was much higher during glacial maxima than during warm periods. This is believed to be a result of changes in climate in southern South America (Wolff et al., 2006), although the exact nature of the relevant changes is not yet clear. In any case the result is a change in aerosol content of the atmosphere over Antarctica (with possible radiative impacts) and a change in the amount of dust deposited in the Southern Ocean which may also play a role in stimulating phytoplankton productivity, and thus carbon dioxide drawdown from the atmosphere. Finally in this section, it has been suggested that the sea salt content of Antarctic ice cores might yield a record of past sea ice extent (Wolff et al., 2006). Sea salt, as expected under this idea, is at higher concentrations during cold periods, consistent with the findings of marine sediments (Gersonde et al., 2005) that sea ice was significantly extended at such times. 11.3.3. Millennial-Scale Changes in Antarctic Climate The climate of the last glacial period is strongly dominated by the occurrence of rapid millennial-scale climate changes known as D-O events (e.g. North Greenland Ice Core Project Members, 2004). These occur in Greenland as rapid (occurring in o40 years) warmings by more than 101C (Huber et al., 2006). Counterparts to them are clearly seen in numerous other NH records. What concerns us here is what was happening in Antarctica in relation to those events, and what it tells us about the causes. The longest D-O events have very clear counterparts, initially labelled as ‘A-events’ in Antarctic climate records. Using the high time-resolution available from the Byrd ice core, Blunier and Brook (2001) showed that while Greenland was cold before a D-O warming, Antarctica experienced a steady warming (by up to B31C). As soon as Greenland temperature jumped, Antarctic temperature started to cool again. It was suggested that this was the pattern expected from simple models involving changes in ocean heat transport (Stocker and Johnsen, 2003), with Antarctica and the Southern Ocean able to accumulate heat when ocean heat transport to the north was weak. Further support for this model has come from the finding (using the EPICA cores at both Dome C and Dronning Maud Land (Kohnen station)) that there is a subdued southern counterpart to every D-O event (EPICA
486
T. Naish et al.
Community Members, 2006). It has not yet been possible to show whether the timing of the weaker, shorter events is consistent with the model. However, central Antarctic climate of the last glacial appears to have been dominated by slow warmings and coolings of between 0.5 and 31C that are most likely associated with changes in meridional ocean heat transport, and thus with rapid warmings in the north. The southern warm events have been named as ‘Antarctic Isotopic Maxima (AIM)’ (Blunier and Brook, 2001), with the A-events representing the largest AIM. Very similar temperature variability was found in the Dome C record for previous glacial periods as for the most recent one, strongly suggesting that the same mode of millennial variability held sway for the last 800 kyr (Jouzel et al., 2007).
11.3.4. Climate of the Antarctic Periphery Only a small number of ice core records from the Antarctic Ice Sheet periphery extend beyond the LGM. The 1,178 m deep ice core at Law Dome appears to reach into the last interglacial period (Morgan et al., 1997), although the whole of the last glacial period is held within the lowest 100 m, and no attempt has been made to provide an age scale for the oldest part. At least during the last deglaciation, the timing of the signal from this core appears at least similar to that from other Antarctic cores. Of most interest for the current paper is the finding that no isotopic values typical of inland ice are found anywhere in the core. This implies that, although the ice at Law Dome may have thickened during some parts of the glacial by up to 300 m (Delmotte et al., 1999), it was never overridden by inland ice during the last 100 ka. Working clockwise around the continent, Taylor Dome (TD), adjacent to many of the marine core records, has yielded a 554 m deep core to bedrock (Steig et al., 1998). The age scale has been tentatively extended to 300 ka (Grootes et al., 2001); however, Termination II (130 ka) is only 23 m above bedrock, so the record is most useful only for one climatic cycle. Comparison with Vostok during the last climatic cycle shows a very similar pattern of variations, but with a larger amplitude for the equivalents of the A-events (large AIM). The assumptions made to create the time scale for the glacial part of the record preclude any discussion about the phasing of these changes in the TD record relative to those of central Antarctica. There has of course been a suggestion (Steig et al., 1998) that, at least during the last termination, TD may have shown a different pattern of change from other Antarctic sites, more in line with those of Greenland. This assertion has been strongly
Late Pliocene–Pleistocene Antarctic Climate Variability
487
challenged (Mulvaney et al., 2000) because of issues about the TD time scale in this section, and must be considered doubtful. The 1,004 m deep bedrock core at Siple Dome (SD), on the West Antarctic side of the Ross Sea, also covers much of the last glacial period. Careful synchronisation of this core to that of Byrd and to Greenland cores (Brook et al., 2005) shows a similar pattern and timing of millennial-scale change to that of Byrd, with recognisable A-events, and with SD (like Byrd) warming while Greenland remained cold, and apparently starting to cool when Greenland warmed. An abrupt climate change, not observed in other Antarctic cores, is seen at SD at 22 ka (Taylor et al., 2004); there is as yet no clear explanation for this apparently very local climate shift. Two other near-coastal cores have already penetrated beyond the LGM, at Talos Dome and Berkner Island, although publications from them are still awaited. Early indications from Berkner Island (facing the Weddell Sea) suggest that it was also never overridden by inland ice during the last climatic cycle (Mulvaney et al., 2007) as some models suggest. Taking all the records, from central East Antarctica, coastal cores and Byrd in central West Antarctica, it appears likely (despite the questions raised by the TD core) that the continent sees the main millennial- and orbital-scale changes more or less synchronously and with a similar pattern. Further work to separate changes in ice sheet altitude from changes in climate will elucidate the causes of somewhat different amplitudes for millennial-scale change at different sites. Finally in this section, we note the potential for coastal cores such as Law Dome and Berkner Island to provide critical constraints on the extent of the Antarctic Ice Sheet during at least the last glacial cycle. 11.3.5. Future Challenges for Ice Coring The ice core community has prepared a strategy for the next decade or more that includes two tasks of particular importance for this chapter (Brook and Wolff, 2006). Firstly, the community has agreed on the value of further cores around Antarctica (and the Arctic) reaching at least 40 ka back in time, to further understand the spatial and temporal pattern of millennial-scale changes. Secondly, sites with continuous ice sequences reaching well beyond 800 ka should be targeted for drilling. While a major geophysical survey will be required to locate such ice, the goal of reaching back as far as 1.5 Ma, into the period when 40 kyr cyclicity is well-established in climate records, is of great importance. It offers the possibility for example to test ideas about the role of changing CO2 concentrations, and the relative roles of
488
T. Naish et al.
northern and southern ice (Raymo et al., 2006), in the transition from the 40–100 kyr world.
11.4. Oceanic Variability from Southern Ocean Sediment Cores As Antarctica changes, so does the ocean. The long-term evolution of the ice sheets and the shorter-term oscillations superimposed upon them, have produced far-reaching effects on the circulation and water masses of both hemispheres. In this section, we focus on the oceans’ responses to G-I cycles over the last million years with emphasis on the last few cycles simply because they are generally the best documented. This younger part of the record is also assessed for the effects of millennial-scale change as driven from Antarctica but also from the NH. Here we review briefly the wealth of palaeoceanographic research and the reader is referred to the reference list as a window into an extensive literature.
11.4.1. Glacial–Interglacial Cycles Past influences of Antarctica on the Southern Ocean are best gauged from well-dated records of environmental proxies, which are interpreted against a history of Antarctic ice behaviour preserved in ice cores and sediment cores from high-latitude locations. While we focus on Antarctica, it is made with full acknowledgement of NH forcing especially as the dominant driver of eustatic sea-level oscillations and as a co-driver of thermohaline circulation.
11.4.1.1. Thermohaline circulation Essentially, the sinking of upper ocean waters in northern high latitudes forms North Atlantic Deep Water (NADW) that migrates south to mix with southern-sourced waters comprised of (i) dense bottom waters generated around the Antarctic margin and (ii) deep waters recycled in the Indian and Pacific oceans especially within the Antarctic Circumpolar Current (see Carter et al., this volume Chapter 4). The resultant Circumpolar Deep Water is the most voluminous in the Southern Ocean and dominates the deep inflows into the Indian and Pacific basins (Fig. 11.5). Yet despite its hybrid character, it retains the signatures of its source waters in particular NADW, which is identified by the salinity maximum and, from a palaeoceanographic
Late Pliocene–Pleistocene Antarctic Climate Variability
489
Figure 11.5: Generalised outline of the modern Southern Ocean with the westerly wind-driven Antarctic Circumpolar Current defined by its zonal jets or fronts that include the Subantarctic Front (SAF), Polar Front (PF), Southern Front (SF) and Southern Boundary (SB) after Orsi et al. (1995). ODP Sites 1090 and 1123 containing W1 myr records of ocean change are also located.
490
T. Naish et al.
perspective, by its elevated d13C content (Charles and Fairbanks, 1992). Changes in this circulation over G-I cycles are subject to a continuing debate, much of which focuses on the input of NADW to the Southern Ocean. Charles and Fairbanks (1992), Rahmstorf (2002), Ninnemann and Charles (2002) amongst others favour a reduction in NADW input to the Southern Ocean in glacial times. Hall et al. (2001) support this decline and suggest that it was compensated by an increase in the supply of Antarctic-sourced deep water, at least in the New Zealand gateway for the Pacific Deep Western Boundary Current (DWBC). Glacial conditions potentially favoured bottom water production through increased windiness plus an expansion of sea ice and associated brine rejection. Certainly Hall et al. (2001) record an increase in the speed of the Pacific DWBC in glacial times as reflected by an increase in the size of ‘sortable silt’, a grain-size proxy for relative current strength (see also Venuti et al., 2007). For time frames longer than individual G-I cycles, the sortable silt data point to three phases of Pacific DWBC vigour during the last 1.2 myr. These three phases cover the transition from a world dominated by 41 kyr climatic cycles to one where 100 kyr cycles prevailed and include: Phase 1 (1.2–0.87 Ma) of moderately strong flow; Phase 2 or the MPT (0.87–0.40 Ma) of weaker flow, and Phase 3 (0.40 Ma to present) when strongest flows prevailed. The alternative hypothesis that little or no decline in NADW inflow occurred in glacial times is supported by Moy et al. (2006), Rosenthal et al. (1997) and others on the basis of various palaeoceanographic proxies including the distribution of d13C between ocean basins over G-I cycles. As NADW flows to the Southern Ocean, its d13C content reduces due to remineralisation of organic carbon and mixing with southern-sourced waters with lower d13C. Depletion continues into the Indian and Pacific oceans, the latter having the lowest d13C (Kroopnick, 1985). By comparing d13C profiles measured on benthic foraminifers from the main ocean basins bathed by NADW-sourced waters, it is possible to identify any variability of those waters. Moy et al. (2006) compared benthic d13C for cores spanning the last 180 kyr from the equatorial Atlantic, the Indian–Pacific boundary at South Tasman Rise and the equatorial Pacific. Equatorial Atlantic d13C was higher than at South Tasman Rise, but followed the same G-I cyclicity and maintained a gradient suggesting maintenance of northern-source deep water for the last 160 kyr. In addition, d13C data for the Tasman Rise and equatorial Pacific were very similar highlighting a close tracking of Southern Ocean and central Pacific waters. While the debate continues, most recent data (e.g. Moy et al., 2006; I. N. McCave, submitted and personal communication 2007) tend to favour a more subdued input of NADW and no major reorganisation of Southern Ocean water masses during glacial periods.
Late Pliocene–Pleistocene Antarctic Climate Variability
491
11.4.1.2. Intermediate waters The dispersal of southern-sourced, intermediate depth waters is a key process in distributing heat, salt, nutrients and gases to equatorial and northern latitudes (e.g. Sarmiento et al., 2004). The northward transport of Subantarctic Mode Water (SAMW) and Antarctic Intermediate Water (AAIW) largely balances the southward export of NADW (Gordon, 1986). Both SAMW and AAIW form in the vicinity of the Subantarctic Front through buoyancy loss related to heat exchange, freshwater input and a northward Ekman transport although changes in wind stress and lateral advection may also play a role (e.g. McCartney, 1977; Rintoul et al., 2001; Toggweiler et al., 2006). While the mode(s) of intermediate water formation is not well understood, observations show that SAMW and underlying AAIW descend to B500 and B500–1,400 m depths, respectively. SAMW formation is fairly widespread, whereas AAIW generation is favoured in the SE Pacific and SW Atlantic. Given their source at the wind-forced ocean surface, intermediate waters typically have elevated d13C contents as revealed by hydrographic data and intermediate depth benthic foraminifers, assuming other factors have not come into play (e.g. Lynch-Stieglitz et al., 1994; Pahnke and Zahn, 2005). In contrast, glacial benthic records show a marked reduction in d13C that is interpreted as a diminished supply of intermediate water. As outlined in more detail in the next section, the glacial Southern Ocean was (i) substantially cooler with sea-surface temperatures (SSTs) up to 61C colder than now (e.g. Mashiotta et al., 1999; EPICA Community Members, 2004; Barrows et al., 2007), (ii) more extensive following an equatorward migration of southern waters by B5–101 latitude (Howard and Prell, 1992; Gersonde et al., 2005), (iii) windier following an equatorward displacement and intensification of zonal westerly winds (Shulmeister et al., 2004; Toggweiler et al., 2006) and (iv) supported a more extensive cover of sea ice B100% more than present (Gersonde et al., 2005). Pahnke and Zahn (2005) have argued that meltwater from the expanded sea ice, concomitant with its northward transport within an invigorated Ekman layer, reduced the surface density and the buoyancy-driven sinking of SAMW and AAIW. Yet despite this decline, intermediate waters remained a thermal conduit that linked Antarctica and the equatorial ocean judging by the strong coherence of their respective temperature records (Lamy et al., 2004; Kiefer et al., 2006). This linkage has been invoked by Weaver et al. (2003) as a potential control of meridional overturning in the North Atlantic. Pulses of meltwater from Antarctica could freshen AAIW, which upon arrival in the North Atlantic could potentially increase the density in contrast with surface waters to encourage sinking and the formation of NADW.
492
T. Naish et al.
11.4.1.3. Surface waters G-I cycles have a marked influence on the upper ocean by virtue of its direct interaction with the atmosphere. Reconstructions of SSTs for the last 25–400 kyr – the approximate range of records recovered by piston cores – highlight marked glacial cooling at orbital frequencies in the Atlantic sector of the Southern Ocean (e.g. Mortyn et al., 2002; Gersonde et al., 2003, 2005; Bianchi and Gersonde, 2004; Pahnke and Sachs, 2006), the Indian sector (e.g. Howard and Prell, 1992; Gersonde et al., 2005; Barrows et al., 2007), the SW Pacific (Fig. 11.5; Weaver et al., 1998; Barrows et al., 2000, 2007; Pahnke et al., 2003; Neil et al., 2004; Pahnke and Zahn, 2005) and the SE Pacific (e.g. Mashiotta et al., 1999; Lamy et al., 2004). A wide range of geochemical, isotopic and microfossil proxies is used to derive SSTs (see summary in Wefer et al., 1999), which can vary according to the proxy (Barrows et al., 2007). The quest for reliable comparisons between ocean basins is further confounded by some palaeoceanographic observations that record localised effects such as upwelling, rather than regional signals (e.g. Pahnke and Sachs, 2006). With such limitations in mind, the aforementioned studies reveal that ocean cooling was most pronounced around 40–461S during major glaciations (e.g. Marine Isotope Stages 2, 6, 8, 10, 12) when SSTs were 4–61C lower than present (Fig. 11.6). This change was recorded in most sectors of the Southern Ocean with the possible exception of the central Pacific where a sparse database suggests that cooling was less severe (Gersonde et al., 2005). Certainly the western and eastern margins of the Pacific Basin were subject to 4–61C coolings during the LGM (Lamy et al., 2004; Barrows et al., 2007). By comparison, the peak interglacials of Marine Isotope Stages 5e and 11 record SSTs up to 31CS warmer than present. The amplitude of SST swings reduces southwards (Howard and Prell, 1992; Gersonde et al., 2003) and once south of the Antarctic Polar Front, LGM SSTs, for example were B31C cooler than present at B501S and r11C cooler south of B521S, at least in the Atlantic-SE Indian sectors. SST records extending over a million years or more from the Southern Ocean are restricted mainly to DSDP and ODP sites, which because of their sparse coverage prevent identification of regional patterns (Fig. 11.7). Nonetheless, they highlight temporal trends especially over the transition from 41 kyr-paced to 100 kyr-paced G-I cycles. Of note are SST records from (i) ODP 1090 located between the modern Subtropical and Subantarctic fronts in the SE Atlantic (Becquey and Gersonde, 2002; Hodell et al., 2002), and (ii) ODP 1123 positioned at the northern limit of the Southern Ocean, within the Subtropical Front in the SW Pacific (Fig. 11.5; Crundwell et al., 2008). They were chosen because they are both based on foraminifer transfer
Late Pliocene–Pleistocene Antarctic Climate Variability
493
Figure 11.6: (B) Stacked SST, normalised to a modern value of 01C, for the Southern Ocean, compiled by Barrows et al. (2007). It highlights the marked cooling of the LGM, when temperatures were up to 51C cooler than now and the abrupt swings in SST that are near-synchronous with warm phases in Antarctica (C) but mainly leading prominent D-O events in the Northern Hemisphere (A). functions and come from open-ocean sites beyond local complexities imposed by bathymetry. Both records exhibit the same three phases of SST variability (Fig. 11.7), but the type and degree of variability at times differ in response to the regional oceanography. Phase 1: 41 kyr cycles (1.2 Ma and earlier to 0.87 Ma). SW Pacific temperatures exhibit high amplitude fluctuations of 4–71C within a G-I range of 9–181C. These fluctuations are superimposed on an overall cooling trend that is especially evident with interglacial optima, which decline progressively from 18 to 141C. Contemporaneous SE Atlantic SSTs have more restricted amplitudes of 1–31C and G-I range of 3–51C. A cooling trend is less clear,
Figure 11.7: Composite plot of Antarctic and Southern Ocean climate proxies spanning the last 1.2 myr. The data allow the coupled G-I variability within the atmospheric, oceanic and cryospheric systems to be evaluated in the context of longer-term changes extending from the 41 kyr climate system (Phase 1) through the MPT (Phase 2) and into the 100 kyr climate system (Phase 3).
494 T. Naish et al.
Late Pliocene–Pleistocene Antarctic Climate Variability
495
but an irregular decline is evident from B1.06 to 0.87 Ma when interglacial optima went from 7 to 41C. The general cooling at both sites reflects an expansion of the Antarctic cryosphere accompanied by a northward migration of the cold surface waters, while at depth there may have been a general increase in the thermohaline circulation at least into the Pacific Ocean (Hall et al., 2001). Thus, at ODP 1123, Subtropical Water was displaced by Subantarctic Water with the former returning at the last part of the terminations and early interglacials. In contrast, the more southerly sited ODP 1090 was mainly affected by circumpolar surface waters through much of the G-I cycles thus accounting for the colder temperatures with subdued amplitudes and G-I ranges. Phase 1 terminated at both sites at the prominent interglacial of Marine Isotope Stage 21. Phase 2: MPT (B0.87–0.4 Ma). This fundamental transition from a 41–100 kyr dominant world was marked at both sites by persistently warmer and more prolonged interglacial periods that were interrupted by some of the coldest glacial periods in the Pleistocene. G-I cyclicity was most marked in the SE Atlantic, which came under the influence of Subantarctic Water and even Subtropical Water during Marine Isotope Stage 15. Not surprisingly, that southward migration of Subtropical Water was particularly well shown at ODP 1123 where it dominated SSTs from Marine Isotope Stages 15–13 with the intervening glacial, Marine Isotope Stage 14, bringing only a modest cooling (o21C). The general warming and southward migration of warmer waters during the MPT is consistent with a reduced Antarctic influence, which is supported by a less vigorous thermohaline circulation into the Pacific as suggested by Hall et al. (2001). The marked G-I variability is interpreted by Becquey and Gersonde (2002) to reflect the increasing influence of 100 kyr cycles, which in turn are driven primarily by changes in NH ice sheets. However, the impact of these drivers appears to be blunted by subtropical influences as attested by ODP 1123 record especially over Marine Isotope Stages 15–13. Phase 3: 100 kyr cycles (B0.4 Ma to present). In the SE Atlantic the onset of the 100 kyr world was accompanied by highly variable SSTs with G-I contrasts of up to 81C. These major fluctuations occurred against a background of mainly warmer temperatures at ODP 1090, which lay mainly within Subantarctic Water. While also exhibiting 100 kyr-paced SSTs, the G-I differences at ODP 1123 are less marked with all but one (Marine Isotope Stages 6-5) less than 51C. Such a muted response may reflect the influence of the south Pacific Subtropical Inflow (e.g. Carter et al., 2008). Despite differences in SST amplitudes, the SE Atlantic and SW Pacific data reveal a progressive reduction of B31C in interglacial optima over Phase 3.
496
T. Naish et al.
11.4.1.4. Ocean fronts From the preceding discussion it is clear that the variability of SSTs over G-I cycles is intimately linked to the meridional displacement of polar waters, which is affected by the state of the Antarctic cryosphere and zonal winds. Glaciations witnessed a major expansion of winter sea ice. In the LGM, sea ice grew by B100% compared to now, extending 3–81 latitude in the Atlantic, 7–101 in the Indian and possibly B2–51 in the Pacific, bearing in mind the last region has sparse data coverage (Gersonde et al., 2005). This expansion was accompanied by an intensification and equatorward displacement of zonal winds (Thiede, 1979; Shulmeister et al., 2004) along with a northward migration of cold water and ocean fronts (e.g. Howard and Prell, 1992; Becquey and Gersonde, 2002; Gersonde et al., 2003; Crundwell et al., 2008). The extent of the migrations depend upon the strength and duration of the climatic drivers, as well as the morphology of the Southern Ocean floor, for example any G-I displacement of fronts off eastern New Zealand, is limited by extensive submarine elevations such as the Chatham Rise where the STF is restricted to 1–21 latitude along the Rise crest (Weaver et al., 1998; Sikes et al., 2002). In contrast, frontal systems and associated water masses in the open Southern Ocean are free to migrate further afield (Fig. 11.5). However, it is an open question as to how SSTs represent shifts in ocean fronts. As pointed out by Gersonde et al. (2005), control sites are too scattered to delineate the temperature and salinity gradients that define frontal systems. Accordingly, the following outline of frontal migrations is indicative only. The data of Becquey and Gersonde (2002), Gersonde et al. (2005) and Howard and Prell (1992) indicate northward migrations of 41, 5– 101 and 2–31 for Antarctic Polar Front in the Atlantic, Indian and Pacific sectors, respectively, during the LGM. At the same time, the SAF shifted 4–51 and 5–101 in the Atlantic and Indian, respectively. SSTs from the easternmost (Lamy et al., 2004) and westernmost Pacific (Crundwell et al., 2008) suggest similar northward displacements. In contrast, migrations of the STF were restricted to only ca. 2–31 and 51 latitude in the Atlantic and Indian sectors suggesting an intensification of oceanographic gradients at the northern limit of the Southern Ocean (Gersonde et al., 2005). The reasons for such limited migration are not clear but it maybe reflect a counteraction by the subtropical gyres in the major basins. In addition to the G-I cycles, long-term records from the ODP sites (Fig. 11.7) identify broad changes in the fronts. At ODP 1090, the PF prevailed, representing a possible shift of B71 latitude north of its present position (Becquey and Gersonde, 2002). During the MPT the site was dominated by the SAF, but with interruptions by glacial excursions of the
Late Pliocene–Pleistocene Antarctic Climate Variability
497
PF. In Phase 3, the SAF again dominated but was interspersed by interglacial incursions of the STF. In the SW Pacific, the more northerly located ODP 1123 spent most of the last 1.2 myr within or north of the STF (Crundwell et al., 2008) except during peak glacial periods. Taking the mean annual SST of the northward edge of the modern SAF to be 111C (Uddstrom and Oien, 1999), then the mean annual palaeo-SST profiles reveal that (i) Phase 1 was a time of frequent glacial incursions of the SAF and associated waters representing a northward migration of B51 latitude; (ii) Phase 2 MPT witnessed only infrequent SAF migrations, which ceased in the latter half when the STF prevailed and (iii) the SAF was located well south reaching Site 1123 only in the major glaciation of Marine Isotope Stage 6.
11.4.1.5. Surface ocean currents As noted in the previous section, SST data indicate northward migrations of the PF and SAF during glacial periods. As these fronts accommodate most of the flow within the ACC (e.g. Whitworth, 1988; Rintoul et al., 2001) the inference is that this major current system also moved north, possibly by B51 or more of latitude. However, in light of (i) limitations of SST data to define past frontal positions, (ii) the marked topographic steering of the modern ACC by mid-ocean ridges (e.g. Gordon et al., 1978; Orsi et al., 1995; Moore et al., 1999) and (iii) the strong latitudinal variability of modern ACC fronts such as the PF, which ranges over 5–71 of latitude (Moore et al., 1999), it is unclear how the ACC behaved. Uncertainty also arises as to whether the ACC strengthened under the stronger wind regimes of glacial periods. Certainly the ACC is mainly wind-driven, with the extent of this forcing dependent upon the position of mid-latitude westerly winds (Toggweiler et al., 2006). Wind strengthens the ACC by direct shear but also enhances the northward Ekman transport, which in turn is compensated by the southward transport and eventual upwelling of deep water at the Antarctic margin. These effects are most pronounced when winds are aligned with the ACC as is the case today. Although glacial period westerly winds (Moreno et al., 1999; Sigman and Boyle, 2000) and the main ACC fronts were located further north than today, it is not known if winds and current were aligned. However, where migration of the ACC is restricted by the seabed morphology as in the Scotia Sea (Pudsey and Howe, 1998), or off eastern New Zealand (Neil et al., 2004), the ACC appears to have strengthened, although it is unclear if this was a response to increased wind-forcing, compression of fronts against the bathymetry or both these processes.
498
T. Naish et al.
11.4.1.6. Iceberg discharge and migration A proxy for G-I behaviour of Antarctic ice is found in the IRD records of Southern Ocean sediments. ODP Site 1011, off the Antarctic Peninsula provides a generalised IRD history spanning 3 Ma (Cowan, 2002). Features are phases of pronounced IRD deposition around 2.8, 1.9 and 0.85 Ma, interleaved by more subdued deposition that followed G-I cycles with a frequency of 41 kyr from 2.2 to 1.0 Ma and 100 kyr frequency after 0.4 Ma. In the latter cycles, the late glacials and following interglacials were the main times of ice rafting. A more detailed record for the last 300 kyr in the Weddell Sea confirms the strong IRD signal at G-I transitions, but shows high deposition of IRD continuing into the early part of the following glacial period (Fig. 11.8; Grobe and Mackensen, 1992). The Weddell Sea record also reveals a consistent decline in IRD production since 200 kyr
Figure 11.8: Normalised stacked profiles for IRD generated at source in the Weddell Sea (Grobe and Mackensen, 1992) and deposited at a distant depocentre – the Campbell Plateau, New Zealand (Carter et al., 2002). Both profiles show similar broad outlines but the key difference is that the times of maximum IRD production are not faithfully reproduced off New Zealand especially during a G-I transition when a warming ocean and reducing windiness would be less favourable for iceberg preservation.
Late Pliocene–Pleistocene Antarctic Climate Variability
499
(Marine Isotope Stage 7). Superimposed on the G-I cycles are IRD pulses of millennial-scale frequency which, given uncertainties of age models, are traceable from the SE Atlantic to SW Pacific, at least for the past 70 kyr (Labeyrie et al., 1986; Kanfoush et al., 2000; Carter et al., 2002). The Kanfoush et al. (2000) IRD events coincide with periods of NADW production and warm interstadials in the North Atlantic. This led to the proposition that the increased inflow of relatively warm NADW and/or a rise in sea level destabilised ice sheets to promote iceberg discharge. Once calved, icebergs from the Weddell to Ross Sea regions were most likely captured within the Weddell, Ross and un-named coastal gyres judging by modern iceberg trajectories (Keys, 1990). Nevertheless, some were eventually entrained by the ACC to move east as identified from modern iceberg paths (Tchernia and Jeannin, 1983), last glacial IRD dispersal patterns (Cooke and Hays, 1982), and a meltwater signal generated about 35–17 kyr (Labeyrie et al., 1986). Not surprisingly, the Weddell Sea IRD profiles (Grobe and Mackensen, 1992) broadly correlate with those off eastern New Zealand (Fig. 11.8; Carter et al., 2002). However, there are also differences in that the interglacial records are more subdued off New Zealand, a feature Carter et al. (2002) attributed to iceberg melting en route from Antarctica. In contrast, glacial records are better correlated indicating better preservation of travelling icebergs on account of colder glacial seas and possibly an intensified ACC (Pudsey and Howe, 1998; Neil et al., 2004). 11.4.2. Millennial-Scale Cycles Antarctic ice cores record frequent abrupt changes in temperature at millennial scales that are difficult to explain by variations in orbital parameters. Over the last 90 kyr, Antarctica has been subject to seven prominent warm periods designated A1–A7 (Blunier and Brook, 2001, and discussed above). With the possible exception of A5 and A6, the others are well expressed in SST records from southern mid-latitudes (Fig. 11.6; Barrows et al., 2007). But perhaps the best-documented abrupt change in terms of identifying oceanic responses is the Antarctic Cold Reversal (ACR), which is timed at around B14.1–12.4 ka according to the EPICA deuterium record (Blunier et al., 1997; EPICA Community Members, 2004). An abrupt cooling of B21C appears to have been accompanied by an expansion of sea ice (Shemesh et al., 2002) and a modest intensification of winds as identified from sea salt and dust proxies in ice cores (Stenni et al., 2001; Ro¨thlisberger et al., 2002). As noted previously, an expanded Antarctic cryosphere can invigorate the THC (e.g. Hall et al., 2001), and judging by the modern oceanography,
500
T. Naish et al.
such a change would be translated rapidly through the Southern Ocean (e.g. Haine et al., 1998). Thus, off eastern New Zealand, the Pacific gateway for the THC, benthic d18O becomes heavier directly in phase with the ACR (Carter et al., 2008). In contrast, change in the surface ocean about the time of the ACR is variable. South of the Subtropical Front, SSTs cool in phase with the ACR, as recorded in the SW Pacific (Pahnke et al., 2003), SE Pacific (Lamy et al., 2004) and the south Atlantic (Kanfoush et al., 2000; Sachs et al., 2001; Shemesh et al., 2002; Gersonde et al., 2003). These sites have direct atmospheric and oceanic links to Antarctica, especially during glacial periods when polar effects are enhanced through the northward migration of westerly winds and surface ocean waters. By comparison, mid-latitude records have a more delayed and muted response to the ACR. Off eastern New Zealand, for example, there was no obvious response until B13.5 ka when the ACR was at its coldest (Carter et al., 2008). Then, SSTs became cooler, marine fertility dropped and the uppermost ocean either became more mixed or the thermocline more shallow. Onshore, pollen and speleothem data reveal cooler, windier conditions that were accompanied by an expansion of glaciers and, at the coast, by a likely stillstand in sea level (Carter et al., 1986; Turney et al., 2003; Williams et al., 2004). The delayed reaction to the ACR is likely connected to a contemporaneous re-establishment of the Subtropical Inflow as it migrated south during the deglacial phase (e.g. Martinez, 1994). Like the southwest Pacific, the south Indian Ocean also cooled out of phase. SSTs reduced by 0.81C between 13.2 and 12 ka, B1,000 years after commencement of the ACR demonstrating the regionality of abrupt climate change effects (Stenni et al., 2001). As core chronologies improve (see Steig, 2001), it is becoming apparent that some millennial-scale variability in the Southern Ocean is linked with abrupt changes documented in the NH. Barrows et al. (2007) draw attention to the timing of Southern Ocean warm phases, A1–A4, with warm phases of D-O cycles, 8, 12, 14 and 17 identified in the Greenland ice core, GISP 2. Likewise, Sachs and Anderson (2005) observe phases of marked algal productivity off eastern New Zealand that occurred within 1–2 ka of massive iceberg influxes associated with Heinrich Events 1–6 in the North Atlantic. Another northern cold event that has received much attention in the Southern Ocean is the Younger Dryas (YD) of 13–11.5 kyr (e.g. Morigi et al., 2003; Turney et al., 2003; Bianchi and Gersonde, 2004; Sachs and Anderson, 2005; amongst others). Because of the YD’s partial overlap with the ACR (14.1–12.4 kyr) and the Oceanic Cold Reversal (13.2–12 kyr) in the Indian Ocean (Stenni et al., 2001) conclusive evidence for a YD effect is still under debate, the resolution of which hinges on the quality of palaeoenvironmental chronologies. Once resolved, the mechanisms that translate YD and other
Late Pliocene–Pleistocene Antarctic Climate Variability
501
abrupt NH perturbations to the Southern Ocean (and vice versa) can be better evaluated to allow a more informed assessment of potential change in the future.
11.5. Modelling of Pleistocene Ice Volume Variations 11.5.1. Northern Hemisphere, the Last Million Years Variations in global ice volume during the Pleistocene, and probably the Pliocene, have been dominated by growth and decay of Northern Hemispheric ice, and most ice-sheet modelling of this period has been concerned with the Laurentide, Cordilleran, Greenland and Eurasian ice sheets. Weertman (1976) was the first to include ice dynamics in model simulations of long-term ice-sheet evolution. He used a perfectly plastic icesheet profile representing a north–south cross section of the Laurentide Ice Sheet, and forced it from 400 ka to the present with a prescribed pattern of net annual accumulation minus ablation versus latitude and elevation (termed a snowline pattern below), shifted vertically in proportion to summer insolation at 651N to represent actual Milankovitch orbital variations. This study was the first to capture the several thousand-year lag of ice volume behind climate forcing due to the mass inertia of the ice sheet. It produced reasonable B23–41 kyr cycles in direct response to precessional and obliquity orbital variations, but did not realistically simulate the dominant observed 100 kyr ice cycles. Subsequent work extended this study to (i) use energy-balance climate models (EBMs) instead of a prescribed snowline pattern (Pollard, 1978; Berger et al., 1990; Deblonde and Peltier, 1993) and (ii) use ice-sheet flow models instead of plastic profiles, allowing bedrock topography and lagged depression below the ice (Birchfield et al., 1981; Pollard, 1982; Hyde and Peltier, 1987). For the most part, the addition of climate models did not change Weertmans’ results significantly, and just showed that the smooth summer-temperature variations in EBMs have much the same effect as a shifted snowline pattern. The additions of high-latitude topography and lagged bedrock depression and rebound were found to be more important: high-latitude plateaus allow nascent ice caps to initiate more easily following interglacials, and deep bedrock depressions at glacial maxima amplify subsequent retreat due to warmer air temperatures at lower elevations. Although these mechanisms produced more non-linear response, they were still not enough to produce realistic 100 kyr response and full G-I cycles.
502
T. Naish et al.
Retreats are triggered in these models by hot northern summer orbital configurations that are coeval with the start of major observed terminations, but the modelled retreats are not complete, allowing substantial ice cover to persist through interglacials. It was found that some additional physics were needed that come into play during the major deglaciations to terminate each 100 kyr cycle, and a large variety of mechanisms have been proposed. These range from cyclic freezing versus melting and surging at the ice-sheet base (Oerlemans, 1982b; MacAyeal, 1993), calving by proglacial lakes (Pollard, 1982), prescribed climatic effects of atmospheric CO2 and/or North Atlantic heat transport and thermohaline circulation (Tarasov and Peltier, 1997), and atmospheric dust concentrations (Peltier and Marshall, 1995). These mechanisms involve or respond to some other long-term component (besides ice volume) with intrinsic time scales of thousands to tens of thousands of years, which is logically necessary because mechanisms that only involve fast-time-scale components (atmosphere, snow, sea ice, upper ocean) would have been triggered not just during major deglaciations but at other times with similar orbits and mid-range ice extents. However, there are two exceptions: snow aging (Gallee et al., 1992) and sea-ice switching (Gildor and Tziperman, 2001). Other aspects of the 100 kyr cycles have been studied recently, such as phase locking by eccentricity or obliquity/precession (e.g. Tziperman et al., 2006) and whether these cycles are deterministic at all (Wunsch, 2003). Although many of the models mentioned above produce realistic Pleistocene ice volume time series including B23, 41, and 100 kyr cycles, there is still no consensus on what processes are responsible for the rapid inceptions and complete deglaciations that are key to the dominant 100 kyr cycles of the last million years. An entirely different approach is to use zero-dimensional ‘box’ models comprised of a few coupled non-linear ordinary differential equations for quantities such as global ice volume, mean ocean temperature and atmospheric CO2 level (Matteucci, 1989; Saltzman and Verbitsky, 1993; Paillard, 1998). These can be tuned to yield fairly realistic ‘ice volume’ time series forced by Pleistocene orbital variations. Hargreaves and Annan (2002) used Saltzman’s model and Monte Carlo Markov Chain (MCMC) techniques to find the optimal parameter set and probability density functions yielding the most realistic results; similar approaches have used three-dimensional ice sheet models (Tarasov and Peltier, 2004). Paillard’s zero-dimensional results basically support the general conclusion from more physical models that nonlinear thresholds are needed to successfully produce 100 kyr cycles from the higher-frequency orbital forcing. However, the highly conceptual formulation of the zero-dimensional models make it hard to translate their results into specific action items for improvements in more physically explicit models.
Late Pliocene–Pleistocene Antarctic Climate Variability
503
A unique approach was taken by Bintanja et al. (2005) to address the ambiguity between ice volume and local sea-water temperature inherent in deep-sea core d18O records. They used a NH ice-sheet flowline model, driven through multiple G-I cycles over the last million years with simplified climate–orbital linkages as in many previous studies. But instead of simply running the model forward in time, (i) a standard empirical formula was used to predict foraminiferal d18O at any time from the model’s current deep-sea temperature and ice volume, and (ii) at each time step the climate temperature was determined not from a physical model, but from the requirement that, when the ice model was stepped forward, the computed foraminiferal d18O yielded the observed deep-sea-core value for the next time step. This produced a self-consistent estimate of the relative contributions of Northern Hemispheric ice versus deep-ocean temperature in d18O core records, with the main conclusion that the ice-sheet contribution varied from B10% in the beginning of glacial cycles to B60% at glacial maxima; the latter is consistent with relict pore water measurements (Schragg et al., 1996). Some attempts have been made to use General Circulation Models (GCMs) in conjunction with long-term ice-sheet variations. Computer time is prohibitive for direct GCM integrations longer than a thousand years or so, and several strategies have evolved to circumvent this limitation. Several studies have driven three-dimensional ice sheet models through the last glacial cycle (last B125 kyr), with climates obtained by interpolating between two stored GCM snapshots at the LGM (21 ka) and modern. The weighting of the two GCM climates is proportional to an observed ice core record (usually Greenland GRIP) representing NH climate of the last 125 kyr (Marshall and Clarke, 1999b; Zweck and Huybrechts, 2005; Charbit et al., 2007). However, this procedure heavily constrains the predicted variations of ice volume (on millennial, orbital and 105-year time scales) to those in the ice core record, so the approach is useful mainly for diagnostic assessment of processes in the ice model. For instance, one striking diagnostic result is the greatly expanded extent of basal melting under the Laurentide as LGM is approached, which is consistent with the basal freeze-thaw-surge hypothesis for amplification of 100 kyr cycles (Marshall and Clark, 2002; cf. Johnson and Fastook, 2002). To couple relatively coarse-grid GCM climates (B100’s km) to the finer ice-sheet model grids and topographies (B10’s km) in these experiments, straightforward downscaling techniques have been developed (e.g. Ramstein et al., 1997; Thompson and Pollard, 1997; Marshall and Clarke, 1999a). First, monthly mean air temperatures and precipitation are interpolated both horizontally and vertically to the ice surface locations. Degree–day parameterisations are used to convert monthly temperatures to snow or ice melt, which are able to capture subtle effects of orbital changes in seasonal
504
T. Naish et al.
insolation (e.g. Huybers, 2006). Some calculations explicitly account for refreezing of meltwater. Also, ‘anomaly’ techniques are often used to reduce biases in GCM climate, by only using GCM predicted changes in climate from the present, combined with modern observed climatology. This is usually done additively for temperature, and by ratio for precipitation to avoid negative precipitation values. A few studies have used asynchronous coupling between GCMs and ice sheets, whereby the GCM is run for a few decades every few thousand years and its climate is used to drive the ice sheet model alone for the next few thousand years. Then the new ice-sheet extent is used to update the GCM’s surface conditions, and the sequence is repeated. For the Pleistocene, this has been done only over the last 21 kyr through the last deglaciation (Charbit et al., 2002). More recently, Earth Models of Intermediate Complexity (EMICs) have made it possible to run coupled climate–ice sheet models through long periods (Charbit et al., 2005). Realistic Pleistocene cycles have been obtained in this way (Calov and Ganopolski, 2007), but, as found in some of the earlier studies above, prescribed variations of atmospheric CO2 and North Atlantic ocean heat transport were needed to obtain complete 100 kyr cycles. A new approach using climate parameterisations based on past GCM snapshots has recently been taken by Abe-Ouchi et al. (2007). Many Global Climate Model (GCMs) and Regional Climate Model (RCMs) have performed ‘snapshot’ simulations of Pleistocene climates at individual times, usually with prescribed ice sheets. One of the first such applications was Gates (1976) at the LGM, using prescribed CLIMAP SSTs and ice sheet reconstructions; recently, fully coupled A/OGCMs have been used for LGM (Hewitt et al., 2003; Kim et al., 2003), and GCMs and EMICs have been applied to other times such as the end of the last interglacial, sometimes driving nascent ice sheets (deNoblet et al., 1996; Vettoretti and Peltier, 2004; Kubatzki et al., 2006). A few studies have used RCMs or zoomed-region GCMs to simulate the mass balance over NH ice sheets at LGM and other times (Hostetler et al., 2000; Krinner et al., 2004). A detailed survey of these simulations is outside the scope of this chapter, but two general points emerge concerning ice sheets: (i) there is wide scatter between GCMderived mass balances of ice sheets, due to the strong sensitivity of summer melt to air temperatures over the ablation zones (Pollard and PMIP, 2000), and (ii) although reasonable mass balance patterns on major NH ice sheets can be attained, several characteristics consistently disagree with geologic data, such as too much ice buildup over Alaska and Siberia (Charbit et al., 2007). It is relatively easy to adjust the downscaling of climate or other parameters to achieve a reasonable mass balance for a given GCM, time and ice sheet, but
Late Pliocene–Pleistocene Antarctic Climate Variability
505
much more challenging to achieve good results for other times and other ice sheets using exactly the same model and parameter values. 11.5.2. Mid-Pleistocene Transition Before the MPT at around 900 ka BP, the 100 kyr cycle was essentially absent in d18O deep-sea core records, and most of the spectral power was at the orbital obliquity period of B41 kyr, with little at the precessional period B23 kyr (Mudelsee and Schulz, 1997). A number of models that explicitly involve NH ice sheets have attempted to understand this abrupt change. Most rely on a special mechanism that allows 100 kyr cycles in the model only when the maximum ice sheet size is large, and not before B1 Ma when the ice sheets never reached the required size. This size evolution is accomplished in turn by an assumed long-term trend in some external model parameter such as declining atmospheric CO2 (Deblonde and Peltier, 1991; Mudelsee and Schulz, 1997; Berger et al., 1999). Clark and Pollard (1998) achieved the same result not by a prescribed long-term trend, but by adding a deforming sediment component underneath the ice sheet. Starting with a 50 m regolith at 3 Ma, the early ice sheets were kept thin by the greater basal sliding due to the sediment. After the sediment is eroded by repeated glacial growth and retreat, by about 1 Ma enough hard bedrock is exposed to support the greater basal shear stresses of thick ice sheets, allowing the 100 kyr mechanism in that model (proglacial calving) to operate. However, the true cause(s) of the MPT remain unknown. Other proposed mechanisms are noted below in section 11.6.4. None of these models address the dominance of B41 kyr spectral power (driven by obliquity) compared to B23 kyr (driven by precession) before the MPT. One recent theory (Raymo et al., 2006) hypothesises that before the MPT, the EAIS was somewhat smaller than at present and much more variable, with terrestrial margins and ablation (summer-melting) zones that responded significantly to orbitally driven summer-temperature variations. If so, precession-driven ice volume variations would have been out of phase between the hemispheres, and obliquity-driven variations in phase, resulting in a cancelling between northern and southern ice volume variations at 23 kyr, and dominant 41 kyr power. This hypothesis is testable by establishing the character and variability of the Antarctic Ice Sheet during the Pliocene up to the MPT (see below). Another recent hypothesis notes that total melt season insolation is affected more strongly by obliquity than by precession (Huybers, 2006); this is testable by models with seasonal cycles and positive-degree-day parametrisations of ice melt.
506
T. Naish et al.
11.5.3. Millennial-Scale Variability Most climate and ice sheet modelling that addresses Pleistocene millennialscale variability has been in the form of snapshots for individual times or modes, and not in long-term time sequences. This is true for mechanisms involving see-saw connections, ocean response to meltwater discharge, and thermohaline interactions between southern and northern high latitudes (reviewed in Rahmstorf, 2002; Clark et al., 2007). One exception is the modelling of the ice-sheet binge-purge mechanism for Heinrich Events (HEs) by Calov et al. (2002) who used a three-dimensional Laurentide ice sheet coupled to an EMIC. They integrated the coupled system through 200,000 years and obtained quasi-periodic HE-like surges due to basal melting and refreezing over Hudson Bay and Strait. However, the true causes of HEs and whether they are internal ice sheet oscillations or manifestations of oceanic sea-saw behaviour are active research questions (Clark et al., 2007).
11.5.4. Antarctica Little data are available that directly records Antarctic Ice Sheet variations during the last few million years. Most climate and ice sheet modelling of Pleistocene Antarctica has been directed at individual times: LGM, modern and future (e.g. Oerlemans, 1982a; Warner and Budd, 1998; Huybrechts et al., 2004). Throughout the Pliocene and Pleistocene the EAIS is usually thought to have been stable with few deeply submerged marine margins, narrow continental shelf areas limiting expansion (Anderson et al., 2002) and marginal summer temperatures that are too cold to allow significant surface melt even with favourable orbital configurations. One regional exception is the Prydz Bay–Lambert Graben drainage system, where geologic and seismic studies have found significant regional variations in glacial extent through the later Cenozoic (McKelvey et al., 2001; Whitehead et al., 2004; O’Brien et al., 2007). On continental scales, though, the EAIS has been insensitive to the B120 m sea-level variations caused by Northern Hemispheric ice cycles, and by orbitally and CO2-induced local air temperature changes. Contrary views have been proposed by Raymo et al. (2006) involving smaller and more variable EAISs prior to the MPT (B1 Ma, see above), and by several researchers (e.g. Webb and Harwood, 1991) who suggest that Sirius Group deposits in the Transantarctic mountains imply drastic EAIS retreat in the Wilkes Bay sector during the Pliocene (see Miller and Mabin, 1998).
Late Pliocene–Pleistocene Antarctic Climate Variability
507
In contrast, the WAIS is grounded well below sea level with extensive deep grounding lines and wide continental shelves in the Weddell and Ross embayments. Consequently it is much more sensitive than the EAIS to variations in sea level and ocean temperatures via ice shelf basal melt. The WAIS is thought to have repeatedly expanded to the continental shelf limits and back during the Pleistocene, driven primarily by sea-level variations of B120 m caused by NH ice growth and decay. A few three-dimensional Antarctic ice models have been run in this way through the last 400 kyr (Ritz et al., 2001; Huybrechts, 2002), forced by prescribed temperature and sea-level variations from ice core and deep sea core records. These threedimensional ice models have the capacity to some degree to simulate floating ice and advancing and retreating marine grounding lines. They capture what is known of the WAIS and EAIS behaviour since the LGM reasonably well, including the retreat of WAIS grounding lines over the Ross and Weddell embayments (Conway et al., 1999; Anderson et al., 2002), as illustrated in Fig. 11.9. However, the models’ treatments of basal hydrology and grounding line migration are simplified and open to question, especially in the light of recent theoretical developments in grounding line treatment that suggests coarse-grid models are not capable of realistic migration (Schoof, 2007), and fuller-stress models are needed to capture drawdown in ice stream regions (Payne et al., 2004). Most of the variation in total grounded ice volume in the Huybrechts and Ritz glacial-cycle models is due to expansion and contraction of grounding lines across continental shelves, mostly in the WAIS Ross and Weddell sectors, but also to the west of the Antarctic Peninsula and other areas. The modelled G-I (BLGM vs. modern) change in Antarctic ice volume is equivalent to B15–20 m of sea level. The EAIS interior responds in the opposite sense, contracting slightly at Northern Hemispheric glacial maxima due to lower model snowfall rates (which are reduced as temperatures fall). At these times, modelled surface elevations in the central EAIS plateau decrease by B100–150 m compared to the present, with minor volume changes equivalent to just a few metres of sea level. Just as for the NH, more sophisticated regional climate, ocean and ice models have been applied to Antarctica but only for particular ‘snapshot’ times, mainly the present and LGM. These include RCMs (Bailey and Lynch, 2000; Bromwich et al., 2004), ice-sheet models including basal sediment and hydrology (Vogel et al., 2003), longitudinal stresses (Payne et al., 2004), subglacial lakes (Siegert, 2005; Pattyn and Siegert, 2007) and regional ocean models capable of resolving sub-ice shelf circulation (Jenkins and Holland, 2002; Holland et al., 2003). The combination of additional physical processes and higher model resolution could well be important for
Figure 11.9: Snapshots of Antarctic Ice Sheet Evolution during the last G-I cycle, simulated by a three-dimensional ice-sheet model driven by parameterised forcing, from Huybrechts (2002; his Fig. 3). Shown is surface ice elevation relative to present sea level. Contour interval is 250 m; thick lines are for every 1,000 m; the lowest contour is for 250 m and generally close to the grounding line (s.l.e., sea-level equivalent). The x- and y-axis-coordinates are arbitrary spatial values used by the model.
508 T. Naish et al.
Late Pliocene–Pleistocene Antarctic Climate Variability
509
the long-term evolution of Antarctic ice, but their use in time-evolving simulations awaits future work.
11.6. Synthesis: Antarctic Climate Evolution Since B3 Ma 11.6.1. Late Pliocene Cooling, Ice Sheet Expansion and the Development of a Marine Ice Sheet Margin Around Antarctica In the Weddell Sea, Kennett and Barker (1990) were the first to identify a cooling step marked by a reduction in sedimentation rate, lower diatom abundance and the presence of sea-ice diatoms during the Late Pliocene. Along the Pacific margin of the Antarctic Peninsula, seismic data correlated into ODP Leg 178 drill cores, documented a change in depositional style associated with reduced sediment supply and meltwater volume as the Antarctic Ice Sheets expanded and began cooling towards their present state from around 3–2.5 Ma (Rebesco et al., 2006). A large spike in IRD abundance at 2.8 Ma recorded in Antarctic Peninsula continental rise Site 1101 core also supports the development of a marine ice terminus along the edge of the WAIS (Cowan, 2002). In the Ross Sea, Bart (2004) interpreted a major glacial unconformity in seismic reflection data of inferred Late Pliocene age as evidence for widespread westward expansion of the WAIS ice streams into the Ross Sea. Although some uncertainties surround the chronology of DVDP cores in the mouth of Taylor Valley, and CIROS-2 cores in Ferrar Fiord, expansion of WAIS-sourced ice into the Dry Valleys also appears to have occurred about this time (McKelvey, 1981; Powell, 1981). Based on the above and new evidence from the ANDRILL AND-1B record (Naish et al., submitted) it appears that between 3.0 and 2.5 Ma, highlatitude climate cooling drove both the WAIS and EAIS towards their present expanded cold polar state. Relatively warm and often terrestrially based ice margins were replaced by more permanent marine termini and the development of ice shelves.
11.6.2. Late Pliocene–Early Pleistocene G-I Climate Cycles and the Role of Northern Hemisphere Glacio-Eustasy Direct physical sedimentary records of Antarctic Ice Sheet variability (e.g. IRD, proximal glacimarine cycles) and more distal ocean records of sea-ice
510
T. Naish et al.
distribution (e.g. diatom palaeoecology), thermohaline circulation (e.g. grain size), ocean temperatures (e.g. d18O), frontal dynamics and surface circulation (e.g. faunally derived assemblages and SSTs) all show a strong covariance with the 41 kyr cycle in Earth’s obliquity prior to the last 800 kyr. It has been proposed that long-term variations in the duration of summer insolation controlled by orbital obliquity (Huybers, 2006), generally dominates over the influence of the intensity of NH insolation (e.g. Raymo et al., 2006) prior to the Mid-Pleistocene. Marine Isotope Stage 31 may be the exception (Scherer et al., 2008). Strong coupling between the different elements of the SH ice-sheet–ocean-climate system, supports a primary obliquity influence on G-I variability between 3 and 1 Ma. Glacial periods result in the northward expansion of seasonal sea ice, SSTs up to 61C colder than now, equatorward migration of frontal zones by B5–101 latitude (Howard and Prell, 1992; Gersonde et al., 2005), equatorward displacement and intensification of zonal westerly winds (Shulmeister et al., 2004; Toggweiler et al., 2006), invigorated surface circulation (ACC) and intensified abyssal currents (e.g. Hall et al., 2001). All of these processes appear to have occurred 3–7 kyr before the d18O ice volume maximum (e.g. Crundwell et al., 2008) when Antarctic Ice Sheets are fully extended onto the continental shelf. The converse is true for interglacials. Stratigraphic analysis of 41 kyr duration, Late Pliocene glacimarine cycles in the AND-1B core straddling the Gauss–Matuyama (G-M) polarity transition (ca. 2.6 Ma) display abrupt transitions between ice-proximal diamictites and open marine biosiliceous deposits, implying rapid oscillations between glacial and marine environments with significant volume changes in the WAIS (Naish et al., 2008). Intriguingly, the facies also imply a cooler style of ice sheet compared with the Early Pliocene cycles, indicating reduced amounts of subglacial meltwater and terrigenous sediment (McKay et al., accepted). From B2.6 Ma polar Antarctic ice volume changes may have been controlled by the effect of NH glacio-eustasy on its marine margin, and that this mechanism accounts for much of the orbital variability since 2.6 Ma (cf. Raymo et al., 2006).
11.6.3. Marine Isotope Stage 31: Anomalous Continental-Scale Warmth An enigmatic interval of foraminiferal ooze and coccolith-bearing assemblages in the Weddell Sea and Prydz Bay cores together with a bioclastic limestone in the Ross Sea at B1 Ma, imply a significant warming and change in ocean chemistry around the periphery of Antarctica.
Late Pliocene–Pleistocene Antarctic Climate Variability
511
This event that occurs within the short normal polarity Jaramillo Subchron is correlated with warm Marine Isotope Stage 31 (Villa et al., 2007; Scherer et al., 2008). The anomalous warming implies an increase of 4–61C in SST, possible incursion of Subantarctic Surface Waters and depression of the lysocline – an event that is apparently unique in the last 3 Ma. The warming may occur during a unique configuration in orbital parameters that may have produced an extended interval (20 kyr) of unusually warm (e.g. Raymo et al., 2006; Scherer et al., 2008), or alternatively unusually long (Denton and Huybers, 2008) SH summers. Planned future recovery of a 1–1.5 Ma ice core from Antarctica should shed valuable light on the greenhouse gas and temperature conditions at this time.
11.6.4. Mid-Pleistocene Climate Transition and Antarctica G-I variability in the 100 Kyr Glacial World The change in frequency of the G-I climate cycle between 900–700 kyr, from the 40 kyr duration, that dominated much of the glacial Cenozoic (last B35 Ma), to B100 kyr cycles remains one of the most poorly understood events in palaeoclimatology. Explanations have involved amplification of the weak orbital influence by non-linear feedbacks in the carbon cycle (Shackleton, 2000), the internal dynamics of large continental ice sheets (Clark and Pollard, 1998), precessional-forcing (Raymo, 1997) and obliquity-beat skipping (Huybers and Wunsch, 2005). Depositional characteristics of the Late Pleistocene AND-1B G-I sedimentary cycles such as: the predominance of subglacial diamictites, thin interglacial ice shelf mudstones and no evidence for meltwater suggests a further cooling and stabilisation of the margins of the Antarctic Ice Sheets after the B-M polarity transition B800 kyr. A similar pattern occurs in ODP Site 1167 core from the Prydz Bay trough mouth fan which shows a significant reduction in sedimentation rate and change in sediment provenance attributed to a less erosive, dry-based ice sheet from B1 Ma (O’Brien et al., 2007). The calving line of the Ross Ice Shelf appears not to have retreated south of its present interglacial position during subsequent ‘warmer-thanHolocene, super-interglacials’ (e.g. Marine Isotope Stages 11, 9 and 5; e.g. Jouzel et al., 2007). This evidence is in contradiction to hypothesised Late Quaternary collapse of the WAIS (Scherer et al., 1998) and far-field evidence for sea-levels at þ20 m above present assigned to Marine Isotope Stage 11 (Hearty et al., 1999).
512
T. Naish et al.
Ocean records (e.g. ODP 1123 and 1090) imply northward expansion of seasonal sea ice, intensified thermohaline circulation and northward displacement of wind-driven, gyral circulation and subantarctic water masses from B700 kyr to present. The atmospheric temperature and greenhouse gas records from Antarctic ice cores show a pronounced 100 kyr periodicity that is coherent and in phase with marine temperature records (e.g. SSTs, Crundwell et al., 2008), but leads the deconvolved ice volume signal (e.g. d18O, Shackleton, 2000). The records of greenhouse gases from Dome C show very strong congruence with many features of the temperature record, and are consistent with CO2 in particular playing a significant role in temperature amplification, and also suggest that the Southern Ocean plays a leading role in controlling the atmospheric concentration of CO2 on G-I time scales. Most of the variation in total grounded ice volume in Antarctic glacial-cycle ice sheet models (e.g. Ritz et al., 2001; Huybrechts, 2002) is due to expansion and contraction of grounding lines across continental shelves, mostly in the WAIS Ross and Weddell sectors, but also to the west of the Antarctic Peninsula and other areas and is equivalent to B15–20 m of sea level. The EAIS interior responds in the opposite sense, contracting slightly at Northern Hemispheric glacial maxima due to lower model snowfall rates (which are reduced as temperatures fall). Glacial surface elevations based on ice core temperature modelling (Parrenin et al., 2007b) suggest central EAIS plateau decreased by B100–150 m compared to the present, with minor volume changes equivalent to just a few metres of sea level. In contrast ice build up on coastal Antarctica was of the order of several hundred metres during the LGM (e.g. Law Dome, Delmotte et al., 1999). In situ accumulation, rather than flow, is indicated from the ice composition which precludes an interior Antarctic source. Higher dust flux to Antarctica during glacial maxima compared with warm periods is considered to have radiative affects over Antarctica and provide nutrients (e.g. Fe) to the Southern Ocean promoting higher algal productivity and atmospheric CO2 drawdown (Wolff et al., 2006).
11.6.5. Millennial-Scale Climate Variability In general millennial-scale cycles occur in high-resolution Antarctic ice cores, particularly in methane and temperature records (e.g. Byrd, EPICA Dome C). However, they appear lower amplitude and their timing may differ from their NH counterparts (D-O events).
Late Pliocene–Pleistocene Antarctic Climate Variability
513
The pattern of millennial-scale variability superimposed on G-I climate cycles in Antarctic ice cores occurs both during and prior to the last 100 kyr. The millennial-scale climate cycles labelled as ‘A-events’ in Antarctic temperature records suggest that Antarctica and the Southern Ocean may be out-of-phase with Greenland warming events (Blunier and Brook, 2001). It has been suggested that this was the pattern expected from simple models involving changes in ocean heat transport (Stocker and Johnsen, 2003), with Antarctica and the Southern Ocean able to accumulate heat when ocean heat transport to the north was weak. Although uncertainties between the age of the gas (methane) and the ice enclosing it in Antarctic ice cores makes evaluation of the precise timing of inter-hemispheric millennial-scale cycles difficult, in some D-O cycles the temperature maxima in the north and south appear to be synchronous (e.g. EPICA Community Members, 2006). In this case, methane concentrations are most likely controlled by changing wetland extents in the northern high latitudes and tropics, along with changes in sinks (Valdes et al., 2005), and thus are relevant to Antarctic climate only as a relatively minor amplifier. However, a Pacific Basin ocean clathrate source for the methane has also been proposed (Kennett et al., 2000). Millennial-scale Antarctic warm periods (A1–A7; Blunier and Brook, 2001) are well expressed in SST, thermohaline circulation (sortable silt) and d18O records from the southern mid-latitude ocean (Barrows et al., 2007; Carter et al., 2008). One of these events, ACR, at B14.2–12.4 ka, is associated with an abrupt cooling of B21C in the Southern Ocean and an expansion of ice shelves and sea ice (Shemesh et al., 2002), modest intensification of winds (Stenni et al., 2001; Ro¨thlisberger et al., 2002) and intensification in deep abyssal inflow along eastern New Zealand through the Pacific gateway (Carter et al., 2008). Thus, millennial-scale climate variability in Antarctica and Southern Ocean appears to display many similar characteristics and processes (perhaps with reduced amplitude) to those observed at the orbital scale (e.g. SSTs, sea ice, water masses and circulation).
ACKNOWLEDGEMENTS Referees comments by Phil O’Brien and Mike Bentley helped improve the text and sharpen some of the conclusions. Comments and edits on a draft manuscript by Robert McKay are greatly appreciated.
514
T. Naish et al.
REFERENCES Abe-Ouchi, A., Segawa, T., & Saito, F. (2007). Climatic conditions for modeling the Northern Hemispheric ice sheets through the ice age cycle. Climates Past, 3, 423–438. Alley, R. B., & MacAyeal, D. R. (1994). Ice-rafted debris associated with binge/ purge oscillations of the Laurentide ice sheet. Paleoceanography, 9, 503–511. Anderson, J. B., Shipp, S. S., Lowe, A. L., et al. (2002). The Antarctic Ice Sheet during the last glacial maximum and its subsequent retreat history: A review. Quaternary Sci. Rev., 21(1–3), 49–70. Bailey, D. A., & Lynch, A. H. (2000). Development of an Antarctic regional climate system model. Part II: Station validation and surface energy balance. J. Climate, 13(8), 1351–1361. Barker, P. F., Camerlenghi, A., Acton, G. D., et al. (1999). Proceedings of the Ocean Drilling Program, Initial Reports, 178 pp. [Online]. Barrett, P. J. (Ed.) (1986). Antarctic Cenozoic history from the MSSTS-1 drillhole, McMurdo Sound, Antarctica, NZ DSIR Bulletin. Science Information Publishing Centre, Wellington, Vol. 237, 174 pp. Barrett, P. J. (Ed.) (1989). Antarctic Cenozoic history from the CIROS-1 drillhole, McMurdo Sound, Antarctica, NZ DSIR Bulletin. Science Information Publishing Centre, Wellington, Vol. 245, 254 pp. Barrett, P. (2008). A history of Antarctic Cenozoic Glaciation – view from the Continent. In: F. Florindo, & M. Siegert (Eds). Antarctic Climate Evolution. Elsevier, Amsterdam, Vol. 8, pp. 33–83. Barrett, P. J., & Hambrey, M. (1992). Plio-Pleistocene sedimentation in Ferrar Fiord, Antarctica. Sedimentology, 39, 109–123. Barrett, P. J., & Treves, S. B. (1981). Sedimentology and petrology of core from DVDP 15, Western McMurdo Sound. In: L. D. McGinnis (Ed.). Dry Valley Drilling Project. Antarctic Research Series 81. AGU, Washington, DC, pp. 281–314. Barron, J., Larsen, B., et al. (1989). Proceedings of the Ocean Drilling Program. Initial Reports, Vol. 119, pp. 15–44. Barrows, T. T., Juggins, S., de Deckker, P., Calvo, E., & Pelejero, C. (2007). Longterm sea surface temperature and climate change in the Australian-New Zealand region. Paleoceanography, 22, PA2215, doi:10.1029/2006PA001328. Barrows, T. T., Juggins, S., de Deckker, P., Thiede, J., & Martinez, J. I. (2000). Paleoceanography of the Southwest Pacific Ocean during the last glacial maximum. Paleoceanography, 15, 95–109. Bart, P. (2004). West-directed flow of the West Antarctic Ice Sheet across Eastern Ross Basin, Ross Sea during the Quaternary. Earth Planet. Sci. Lett., 228, 425–438. Becquey, S., & Gersonde, R. (2002). Past hydrographic and climatic changes in the subantarctic zone of the South Atlantic – The Pleistocene record from ODP site 1090. Palaeogeogr. Palaeoclimatol. Palaeoecol., 182, 221–239.
Late Pliocene–Pleistocene Antarctic Climate Variability
515
Berger, A., Gallee, H., Fichefet, T., et al. (1990). Testing the astronomical theory with a coupled climate ice-sheet model. Glob. Planet. Change, 89(1–2), 125–141. Berger, A., Li, X., & Loutre, M.-F. (1999). Modelling Northern Hemisphere ice volume over the last 3 Ma. Quaternary Sci. Rev., 18(1), 1–11. Bianchi, C., & Gersonde, R. (2004). Climate evolution at the last deglaciation: The role of the southern ocean. Earth Planet. Sci. Lett., 228, 407–424. Bintanja, R., Wal, R., & Oerlemans, J. (2005). A new method to estimate ice age temperatures. Climate Dyn., 24(2–3), 197–211. Birchfield, G. E., Weertman, J., & Lunde, A. T. (1981). A paleoclimate model of Northern Hemisphere ice sheets. Quaternary Res., 15(2), 126–142. Blunier, T., & Brook, E. J. (2001). Timing of millennial-scale climate change in Antarctica and Greenland during the last glacial period. Science, 291, 109–112. Blunier, T., Schwander, J., Stauffer, B., Stocker, T., Da¨llenbach, A., Indermu¨hle, J., Tschumi, J., Chappellaz, D., Raynaud, D., & Barnola, J.-M. (1998). Timing of the Antarctic cold reversal and the atmospheric CO2 increase with respect to the Younger Dryas event. Geophys. Res. Lett., 24, 2683–2686. Bond, G., Showers, W. J., Cheseby, M., Lotti, R., Almasi, P., deMenocal, P., Priore, P., Cullen, H., Hajdas, I., & Bonani, G. (1997). A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science, 278, 1257–1266. Broecker, W. (2000). Abrupt climate change: Causal constraints provided by the Paleoclimate record. Earth Sci. Rev., 51, 137–154. Broecker, W. S., Bond, G., Klas, M., Bonani, G., & Wolfi, W. (1990). A salt oscillator in the glacial North Atlantic? 1. The concept. Paleoceanography, 5, 469–477. Bromwich, D. H., Guo, Z., Bai, L., et al. (2004). Modelled Antarctic precipitation. Part I: Spatial and temporal variability. J. Climate, 17(3), 427–447. Brook, E., & Wolff, E. W. (2006). The future of ice core science. EOS Trans., 87, 39 pp. Brook, E. J., White, J. W. C., Schilla, A. S. M., Bender, M. L., Barnett, B., Severinghaus, J. P., Taylor, K. C., Alley, R. B., & Steig, E. J. (2005). Timing of millennial-scale climate change at Siple Dome, West Antarctica, during the last glacial period. Quaternary Sci. Rev., 24, 1333–1343. Calov, R., Ganopolski, A., Petoukhov, V., et al. (2002). Large-scale instabilities of the Laurentide Ice Sheet simulated in a fully coupled climate-system model. Geophys. Res. Lett., 29(24), doi:10.1029/2002gl016078. Calov, R., & Ganopolski, A. (2007). Simulation of glacial cycles with an Earth system model of intermediate complexity, Paper presented at European Geosciences Union General Assembly, Vienna, Austria, 15–20 August, 16077962/gra/EGU2007-A-02790. Carter, L., Manighetti, B., Ganssen, G., & Northcote, L. (2008). SW Pacific modulation of abrupt climate change during the Antarctic cold reversal – Younger Dryas. Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 284–298. Carter, L., McCave, I. N., & Williams, M. (this volume). Circulation and water masses of the Southern Ocean: A review. In: F. Florindo, & M. Siegert (Eds). Antarctic Climate Evolution. Elsevier, Amsterdam.
516
T. Naish et al.
Carter, L., Neil, H. L., & Northcote, L. (2002). Late Quaternary incursions of icebergs into the SW Pacific, off Eastern New Zealand. Mar. Geol., 191, 19–35. Carter, R. M., Carter, L., & Johnson, D. P. (1986). Submergent shorelines in the SW Pacific: Evidence for an episodic post-glacial transgression. Sedimentology, 33, 629–649. Chappell, J., Omura, A., East, T., McCulloch, M., Pandolfi, J., Ota, Y., & Pillans, B. (1996). Reconciliation of late Quaternary sea levels derived from coral terraces at Huon Peninsula with deep sea oxygen isotope records. Earth Planet. Sci. Lett., 141, 227–236. Chappell, J., & Shackleton, N. J. (1986). Oxygen isotopes and sea level. Nature (Lond.), 324, 137–140. Charbit, S., Ritz, C., & Ramstein, G. (2002). Simulations of Northern Hemisphere ice-sheet retreat: Sensitivity to physical mechanisms involved during the last deglaciation. Quaternary Sci. Rev., 21(1–3), 243–265. Charbit, S., Ritz, C., & Philippon, G. (2007). Numerical reconstructions of the Northern Hemisphere ice sheets through the last glacial-interglacial cycle. Climate Past, 3(1), 15–37. Charbit, S., Kageyama, M., Roche, D., Ritz, C., & Ramstein, G. (2005). Investigating the mechanisms leading to the deglaciation of past continental Northern Hemisphere ice sheets with the CLIMBER-GREMLINS coupled model. Glob. Planet. Change, 48, 253–273. Charles, C. D., & Fairbanks, R. G. (1992). Evidence from Southern Ocean sediments for the effect of North Atlantic deep water flux on climate. Nature, 355, 416–419. Clark, P. U., et al. (2007). Mechanisms for a B7-kyr climate and sea-level oscillation during marine isotope stage 3. American Geophysical Union Monograph (In review). Clark, P. U., & Pollard, D. (1998). Origin of the middle Pleistocene transition by ice sheet erosion of regolith. Paleoceanography, 13(1), 1–9. Conway, H., Hall, B., Denton, G., & Gades, A. (1999). Past and future groundingline retreat of the West Antarctic Ice Sheet. Science, 286(5438), 280–283. Cooke, D. W., & Hays, J. D. (1982). Estimates of Antarctic ocean seasonal sea-ice cover during glacial intervals. In: C. Craddock (Ed.). Antarctic Geoscience. University of Wisconsin Press, Madison, pp. 1017–1025. Cowan, E. A. (2002). Identification of the glacial signal from the Antarctic Peninsula since 3.0 Ma at site 1101 in continental rise sediment drift. In Ocean Drilling Program Leg 178, Scientific Results. doi:10.2973/odp.proc.sr.178.206.2001. CRST (Cape Roberts Science Team) (1999). Studies from the Cape Roberts Project, Ross Sea, Antarctica – Initial Report on CRP-2/2A. Terra Antartica, 6, 1–173 (with supplement). Crundwell, M., Scott, G., Naish, T. R., & Carter, L. (2008). Glacial–interglacial ocean climate variability spanning the mid-Pleistocene transition in the temperate Southwest Pacific, ODP site 1123. Palaeogeogr. Palaeoclimatol. Palaeoecol., 260, 202–229.
Late Pliocene–Pleistocene Antarctic Climate Variability
517
Dansgaard, W., Johnsen, S. J., Clausen, H. B., Dahl-Jensen, D., Gundestrup, N. S., Hammer, C. U., Hvidberg, C. S., Steffensen, J. P., Sveinbjornsdottir, A. E., Jouzel, J., & Bond, G. (1993). Evidence for general instability of past climate from a 250-kyr ice-core record. Nature (Lond.), 364, 218–220. Deblonde, G., & Peltier, W. R. (1991). A one-dimensional model of continental ice volume fluctuations through the Pleistocene-implications for the origin of the mid-Pleistocene climate transition. J. Climate, 4(3), 318–344. Deblonde, G., & Peltier, W. R. (1993). Late Pleistocene ice-age scenarios based on observational evidence. J. Climate, 6(4), 709–727. Delmotte, M., Raynaud, D., Morgan, V., & Jouzel, J. (1999). Climatic and glaciological information inferred from air-content measurements of a Law Dome (East Antarctica) ice core. J. Glaciol., 45, 255–263. deNoblet, N. I., Prentice, I. C., & Jousaumme, S. (1996). Possible role of atmosphere-biosphere interactions in triggering the last glaciation. Geophys. Res. Lett., 23(22), 3191–3194. Denton, G., & Hendy, C. (1994). Younger Dryas age advance of the Franz Joseph Glacier in the Southern Alps of New Zealand. Science, 264, 1434–1437. Denton, G., Hughes, T., & Karlen, W. (1986). Global ice sheet system interlocked by sea-level. Quaternary Res., 26, 3–36. Denton and Huybers, (2008). Interpolar climate symmetry at orbital timescales and the duration of summer hemisphere summer. Nature (in press). Ding, Z. L., Derbyshire, E., Yang, S. L., Yu, Z. W., Xiong, S. F., & Liu, T. S. (2002). Stacked 2.6-Ma grain size record from the Chinese loess based on five sections and correlation with the deep-sea d18O record. Paleoceanography, 17(3), 1–21. Elston, D. P., & Bressler, S. L. (1981). Magnetic stratigraphy of DVDP drill cores and late Cenozoic history of the Taylor Valley, Transantarctic Mountains, Antarctica, in Mc Guinnis L. D. (Ed.). Dry Valley Drilling Project. Antarctic Research Series 33. AGU, Washington, DC, pp. 413–426. EPICA Community Members. (2004). Eight glacial cycles from an Antarctic ice core. Nature, 429, 623–628. EPICA Community Members. (2006). One-to-one hemispheric coupling of millennial polar climate variability during the last glacial. Nature, 444, 195–198. Escutia, C., Cooper, A., Eittreim, S., Tanahashi, T., Ishihara, T., DeSantis, L., & O Brien, P. (2006). Cenozoic East Antarctic Ice Sheet history from Wilkes Land sediments, International Ocean Drilling Program, Vol. 482. Full Proposal. www.iodp.tamu.edu/scienceops/expeditions/wilkes_land.html Gallee, H., van Ypersele, J.-P., Fichefet, T., et al. (1992). Simulation of the last glacial cycle by a coupled, sectorially averaged climate-ice sheet model 2. Response to insolation and CO2 variations. J. Geophys. Res. Atmos., 97(D14), 15713–15740. Gates, W. L. (1976). Modelling ice-age climate. Science, 191(4232), 1138–1144. Gersonde, R., Abelmann, A., Brathauer, U., Becquey, S., Bianchi, C., Cortese, G., Grobe, H., Kuhn, G., Niebler, H-S., Segl, M., Sieger, R., Zielinski, U., & Futterer, D. K. (2003). Last glacial sea surface temperatures and sea-ice extent in
518
T. Naish et al.
the Southern Ocean (Atlantic-Indian Sector): A multiproxy approach. Paleoceanography, 18, 1061, doi:10.1029/2002/PA000809. Gersonde, R., Crosta, X., Abelmann, A., & Armand, L. (2005). Sea surface temperature and sea ice distribution of the Southern Ocean at the EPILOG last glacial maximum – A circum-Antarctic view based on siliceous microfossil records. Quaternary Sci. Rev., 24, 869–896. Gildor, H., & Tziperman, E. (2001). A sea ice climate switch mechanism for the 100-kyr glacial cycles. J. Geophys. Res. Oceans, 106(C5), 9117–9133. Gordon, A. L. (1986). Interocean exchange of thermocline water. J. Geophys. Res., 91, 5037–5056. Gordon, A. L., Molinelli, E., & Baker, T. (1978). Large scale relative dynamic topography of the Southern Ocean. J. Geophys. Res., 83, 3023–3032. Grobe, H., & Mackensen, A. (1992). Late Quaternary climatic cycles as recorded in sediments from the Antarctic continental margin. Antarct. Res. Ser., 56, 349–376. Grootes, P. M., Steig, E. J., Stuiver, M., Waddington, E. D., & Morse, D. L. (2001). The Taylor Dome Antarctic O-18 record and globally synchronous changes in climate. Quaternary Res., 56, 289–298. Haine, T. W. N., Watson, A. J., Liddicoat, M. I., & Dickson, R. R. (1998). The flow of Antarctic bottom water to the Southwest Indian Ocean estimated using CFCs. J. Geophys. Res., 103, 27,637–27,653. Hall, I. R., McCave, I. N., Shackleton, N. J., Weldon, G. P., & Harris, S. E. (2001). Glacial intensification of deep Pacific inflow and ventilation. Nature, 412, 809–812. Hambrey, M., Ehrmann, W. U., & Larsen, B. (1991). Cenozoic glacial record of the Prydz Bay Continental Shelf, East Antarctica. In: J. Barron, B. Larsen, et al. (Eds). Proceedings of the Ocean Drilling Program. Scientific Results, Vol. 119, pp. 77–128. Hammer, C., Mayewski, P. A., Peel, D., & Stuvier, M. (1997). Greenland summit ice cores. J. Geophys. Res., 102(C12), 26,315–26,886. Hargreaves, J. C., & Annan, J. D. (2002). Assimilation of paleo-data in a simple Earth system model. Climate Dyn., 19(5–6), 371–381. Hayes, D. E., Frakes, L. A., et al. (1975). Initial Reports of the Deep Sea Drilling Project. US Government Printing Office, Washington, DC, Vol. 28, 1012 pp. Hays, J. D., Imbrie, J., & Shackleton, N. J. (1976). Variations in the Earth’s orbit; pacemaker of the ice ages. Science, 194, 1121–1132. Hearty, P. J., Kindler, P., Cheng, H., & Edwards, R. L. (1999). A þ20 m middle Pleistocene sea-level highstand (Bermuda and the Bahamas) due to partial collapse of Antarctic Ice. Geology, 27, 375–378. Hewitt, C. D., Stauffer, R., Broccoli, A., & Mitchell, P. (2003). The effect of ocean dynamics in a coupled GCM simulation of the last glacial maximum. Climate Dyn., 20(2–3), 203–218. Hillenbrand, C-D., & Ehrmann, W. (2005). Late neogene to Quaternary environmental changes in the Antarctic Peninsula Region: Evidence from drift sediments. Glob. Planet. Change, 45, 165–191.
Late Pliocene–Pleistocene Antarctic Climate Variability
519
Hodell, D. A., Gersonde, R., & Blum, P. (2002). Leg 177 synthesis: Insights into Southern Ocean paleoceanography on tectonic to millennial timescales. In: R. Gersonde, D. Hodell, & P. Blum (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 177, pp. 1–54. Holland, D. M., Jacobs, S., & Jenkins, A. (2003). Modelling the ocean circulation beneath the Ross Ice Shelf. Antarct. Sci., 15(1), 13–23. Hostetler, S. W., Bartlein, P. J., Clark, P. U., et al. (2000). Simulated influences of Lake Agassiz on the climate of Central North America 11,000 years ago. Nature, 405(6784), 334–337. Howard, W. R., & Prell, W. L. (1992). Late Quaternary surface circulation of the Southern Indian Ocean and its relationship to orbital variations. Paleoceanography, 7, 79–117. Huber, C., Leuenberger, M., Spahni, R., Flu¨ckiger, J., Schwander, J., Stocker, T. F., Johnsen, S., Landals, A., & Jouzel, J. (2006). Isotope calibrated Greenland temperature record over marine isotope stage 3 and its relation to CH4. Earth Planet. Sci. Lett., 243, 504–519. Huybers, P. W. (2006). Early Pleistocene glacial cycles and the integrated summer insolation forcing. Science, 313(5786), 508–511. Huybers, P. W., & Wunsch, C. (2005). Obliquity pacing of the late Pleistocene glacial terminations. Nature, 434, 491–494. Huybrechts, P. (2002). Sea-level changes at the LGM from ice-dynamic reconstructions of the Greenland and Antarctic Ice Sheets during the glacial cycles. Quaternary Sci. Rev., 21(1–3), 203–231. Huybrechts, P., Letreguilly, A., & Reeh, N. (2004). Modelling Antarctic and Greenland volume changes during the 20th and 21st centuries forced by GCM time slice integrations. Glob. Planet. Change, 42(1–4), 83–105. Hyde, W. T., & Peltier, W. R. (1987). Sensitivity experiments with a model of the ice-age cycle-the response to Milankovitch forcing. J. Atmos. Sci., 44(10), 1351–1374. Jenkins, A., & Holland, D. M. (2002). A model study of ocean circulation beneath Filchner-Ronne Ice Shelf, Antarctica: Implications for bottom water formation. Geophys. Res. Lett., 29(8), doi:10.1029/2001gl014589. Johnson, J., & Fastook, J. L. (2002). Northern Hemispheric glaciaton and its sensitivity to basal melt water. Quaternary Int., 95–96, 65–74. Jouzel, J., Masson-Delmotte, V., Cattani, O., Dreyfus, G., Falourd, S., Hoffmann, G., Nouet, J., Barnola, J. M., Chappellaz, J., Fischer, H., Gallet, J. C., Johnsen, S., Leuenberger, M., Loulergue, L., Luethi, D., Oerter, H., Parrenin, F., Raisbeck, G., Raynaud, D., Schwander, J., Spahni, R., Souchez, R., Selmo, E., Schilt, A., Steffensen, J. P., Stenni, B., Stauffer, B., Stocker, T., Tison, J.-L., Werner, M., & Wolff, E. W. (2007). Orbital and millennial Antarctic climate variability over the last 800000 years. Science, 317, 793–796. Jouzel, J., Vimeux, F., Caillon, N., Delaygue, G., Hoffmann, G., Masson-Delmotte, V., & Parrenin, F. (2003). Magnitude of isotope/temperature scaling for
520
T. Naish et al.
interpretation of Central Antarctic Ice Cores. J. Geophys. Res. Atmos., 108, 4361, doi:10.1029/2002JD002677. Kanfoush, S. L., Hodell, D. A., Charles, C. D., Guilderson, T. P., Mortyn, P. G., & Ninnemann, U. S. (2000). Millennial scale instability of the Antarctic Ice Sheet during the last glaciation. Science, 288, 1815–1818. Kawamura, K., Parrenin, F., Lisiecki, L., Uemura, R., Vimeux, F., Severinghaus, J. P., Hutterli, M. A., Nakazawa, T., Aoki, S., Jouzel, J., Raymo, M. E., Matsumoto, K., Nakata, H., Motoyama, H., Fujita, S., Azuma, K., Fujii, Y., & Watanabe, O. (2007). Northern Hemisphere forcing of climatic cycles over the past 360,000 years implied by accurately dated Antarctic Ice Cores. Nature, 448, 912–916. Kennett, J. P., & Barker, P. F. (1990). Latest cretaceous to Cenozoic climate and oceanographic developments in the Weddell Sea, Antarctica: An ocean drilling perspective. In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 113, pp. 93–960. Kennett, J. P., Barker, P. F., et al. (2000). Proceedings of the Ocean Drilling Program, Initial Reports, Vol. 113 [Online]. Keys, J. R. (1990). Ice. In: G. P. Glasby (Ed.). Antarctic sector of the Pacific. Elsevier Oceanographic Series 51, Amsterdam, pp. 95–123. Kiefer, T., McCave, I. N., & Elderfield, H. (2006). Antarctic control on tropical Indian Ocean sea surface temperature and hydrography. Geophys. Res. Lett., 33, L24612, doi:10.1029/2006GL027097. Kim, S. J., Flato, G., & Boer, G. (2003). A coupled climate model simulation of the last glacial maximum, part 2: Approach to equilibrium. Climate Dyn., 20(6), 635–661. Krinner, G., Mangerud, J., Jakobsson, M., et al. (2004). Enhanced ice sheet growth in Eurasia owing to adjacent ice-dammed lakes. Nature, 427(6973), 429–432. Kroopnick, P. (1985). The distribution of 13C and SCO2 in the world oceans. Deep Sea Res., 32, 57–84. Kubatzki, C., Claussen, M., Calov, R., et al. (2006). Sensitivity of the last glacial inception to initial and surface conditions. Climate Dyn., 27(4), 333–344. Labeyrie, L. D., Pichon, J. J., Labracherie, M., Ippolito, P., Duprat, J., & Duplessy, J. C. (1986). Melting history of Antarctica during the past 60,000 years. Nature, 322, 701–706. Lamy, F., Kaiser, J., Ninnemann, U., Hebbeln, D., Arz, H. W., & Stoner, J. (2004). Antarctic timing of surface water off Chile and Patagonian ice sheet response. Science, 304, 1959–1962. Larter, R. D., & Barker, P. F. (1989). Seismic stratigraphy of the Antarctic Peninsula Pacific margin: A record of Pliocene-Pleistocene ice volume and paleoclimate. Geology, 17, 731–734. Larter, R. D., & Cunningham, A. P. (1993). The depositional pattern and distribution of glacial-interglacial sequences on the Antarctic Peninsula Pacific Margin. Mar. Geol., 109, 203–219. Leroy, S., & Dupont, L. (1994). Development of vegetation and continental aridity in northwestern Africa during the late Pliocene: The pollen record of ODP site 658. Palaeogeogr. Palaeoclimatol. Palaeoecol., 109, 295–316.
Late Pliocene–Pleistocene Antarctic Climate Variability
521
Lisiecki, L. E., & Raymo, M. E. (2005). A Pliocene-Pleistocene stack of 57 globally distributed Benthic Delta d18 records. Paleoceanography, 20, PA1003, doi:10.1029/2004PA001071. Lynch-Stieglitz, J., Fairbanks, R. G., & Charles, C. D. (1994). Glacial-interglacial history of Antarctic intermediate water: Relative strengths of Antarctic vs. Indian Ocean sources. Paleoceanography, 9, 7–29. MacAyeal, D. R. (1993). Binge/purge oscillations of the Laurentide ice-sheet as a cause of the North-Atlantics Heinrich events. Paleoceanography, 8(6), 775–784. Marshall, S. J., & Clark, P. U. (2002). Basal temperature evolution of North American Ice Sheets and implications for the 100-kyr cycle. Geophys. Res. Lett., 29(24), 4. Marshall, S. J., & Clarke, G. K. C. (1999a). Ice sheet inception: Subgrid hypsometric parameterization of mass balance in an ice sheet model. Climate Dyn., 15(7), 533–550. Marshall, S. J., & Clarke, G. K. C. (1999b). Modeling North American freshwater runoff through the last glacial cycle. Quaternary Res., 52(3), 300–315. Martinez, J. I. (1994). Late Pleistocene palaeoceanography of the Tasman Sea: Implications for the dynamics of the Warm Pool in the Western Pacific. Palaeogeogr. Palaeoclimatol. Palaeoecol., 112, 19–62. Mashiotta, T. A., Lea, D. W., & Spero, H. J. (1999). Glacial-interglacial changes in subantarctic sea surface temperature and d18O-water using foraminiferal Mg. Earth Planet. Sci. Lett., 170, 417–432. Matteucci, G. (1989). Orbital forcing in a stochastic resonance model of the LatePleistocene climatic variations. Climate Dyn., 3, 179–190. McCartney, M. S. (1977). Subantarctic mode water. In: M. V. Angel (Ed.). A Voyage of Discovery: George Deacon 70th Anniversary Volume, Supplement to Deep-Sea Research. Pergamon Press, Oxford, pp. 103–119. McGinnis, L. D. (Ed.) (1981). Dry Valley Drilling Project. Antarctic Research Series 81. AGU, Washington, DC, 465 pp. McKay, R., Browne, G., Carter, L., et al., The stratigraphic signature of Late Cenozoic Oscillations of the West Antarctic Ice Sheet in Ross Embayment. Bulletin of the Geological Society of America (accepted). McKelvey, B. C. (1981). The lithologic logs of DVDP Cores 10 and 11, Eastern Taylor Valley. In: L. D. Mc Guinnis (Ed.). Dry Valley Drilling Project. Antarctic Research Series 33. AGU, Washington, DC, pp. 63–94. McKelvey, B. C., Hambrey, M. J., Harwood, D. M., Mabin, M. C. G., Webb, P.-N., & Whitehead, J. M. (2001). The Pagadroma Group – A Cenozoic record of the East Antarctic Ice Sheet in the Northern Prince Charles Mountains. Antarct. Sci., 13, 455–468. Milankovitch, M. M. (1941). Canon of insolation and the ice age problem. Royal Serbian Academy Special Publication. Miller, M. F., & Mabin, M. C. G. (1998). Antarctic neogene landscapes – In the refrigerator or in the deep freeze? Geol. Soc. Am. GSA Today, 8(4), 1–8.
522
T. Naish et al.
Monnin, E., Indermuhle, A., Dallenbach, A., Fluckiger, J., Stauffer, B., Stocker, T. F., Raynaud, D., & Barnola, J. M. (2001). Atmospheric CO2 concentrations over the last glacial termination. Science, 291, 112–114. Moore, J. K., Abbott, M. R., & Richman, J. G. (1999). Location and dynamics of the Antarctic circumpolar front from satellite sea surface temperature data. J. Geophys. Res., 104, 3059–3073. Moreno, P. I., Lowell, T. V., Jacobson, G. L. Jr., & Denton, G. (1999). Abrupt vegetation and climate changes during the last glacial maximum and last termination in the Chilean Lake District: A case study from Canal De La Puntilla (411S). Geogr. Ann. Ser. A, 81, 285–311. Morgan, V. I., Wookey, C. W., Li, J., van Ommen, T. D., Skinner, W., & Fitzpatrick, M. F. (1997). Site information and initial results from deep ice drilling on Law Dome, Antarctica. J. Glaciol., 43, 3–10. Morigi, C., Capotondi, L., Giglio, F., Langone, L., Brilli, M., Turi, B., & Ravaioli, M. (2003). A possible record of the Younger Dryas event in deep-sea sediments of the Southern Ocean (Pacific Sector). Palaeogeogr. Palaeoclimatol. Palaeoecol., 198, 265–278. Mortyn, P. G., Charles, C. D., & Hodell, D. A. (2002). Southern ocean upper water column structure over the last 140 kyr with emphasis on the glacial terminations. Glob. Planet. Change, 34, 241–252. Moy, A. D., Howard, W. R., & Gagan, M. K. (2006). Late Quaternary palaeoceanography of the circumpolar deep water from the South Tasman Rise. J. Quaternary Sci., 21, 763–777. Mudelsee, M., & Schulz, M. (1997). The mid-Pleistocene climate transition: Onset of 100 ka cycle lags ice volume build-up by 280 ka. Earth Planet. Sci. Lett., 151(1–2), 117–123. Mulvaney, R., Arrowsmith, C., Barnola, J. M., McCormack, T., Loulergue, L., Raynaud, D., Lipenkov, V. Y., & Hindmarsh, R. C. A. (2007). A deglaciation climate and ice sheet history of the Weddell Sea region from the Berkner Island ice core (Abstract). Quaternary Int., 167–168(Suppl.), 294. Mulvaney, R., Ro¨thlisberger, R., Wolff, E. W., Sommer, S., Schwander, J., Hutterli, M. A., & Jouzel, J. (2000). The transition from the last glacial period in inland and near-coastal Antarctica. Geophys. Res. Lett., 27, 2673–2676. Naish, T., Powell, R., Levy, R., Krissek, L., Niessen, F., Pompilio, M., Scherer, R., Talarico, F., Wilson, G., Wilson, T., Browne, G., Carter, C., Cody, R., Cowan C., Crampton, J., Dunbar, G., Dunbar, N., Florindo, F., Gebhardt, C., Graham, I., Hannah, M., Hansaraj, D., Henrys, S, Helling, D., Kuhn, G., Kyle, P., La¨ufer, A., Maffioli, P., Magens, D., Mandernack, K., McIntosh, W., McKay, R., Millan, C., Morin, R., Ohneiser, C., Paulsen, T., Persico, D., Reed, J., Ross, J., Raine, I., Schmitt, D., Sagnotti, L., Sjunneskog, C., Strong, P., Taviani, M., Vogel, S., Wilch, T., Williams, T., & Winter, D. (Submitted). Late Cenozoic stability of the West Antarctic Ice Sheet. Nature. Naish, T. R., Abbott, S. T., Alloway, B. V., Beu, A. G., Carter, R. M., Edwards, A. R., Journeaux, T. D., Kamp, P. J. J., Pillans, B. J., Saul, G., & Woolfe, K. J. (1998).
Late Pliocene–Pleistocene Antarctic Climate Variability
523
Astronomical calibration of a Southern Hemisphere Plio-Pleistocene reference section, Wanganui Basin, New Zealand. Quaternary Sci. Rev., 17, 695–710. Naish, T. R., Levy, R. H., Powell, R. D., ANDRILL MIS Science, & Operations Teams (2006). ANDRILL McMurdo Ice Shelf Scientific Logistical Implementation Plan, ANDRILL Contribution No. 7, University of Nebraska, Lincoln, 117 pp. Naish, T. R., Powell, R. D., Barrett, P. J., Levy, R. H., Henrys, S., Wilson, G. S., Krissek, L. A., Niessen, F., Pompilio, M., Scherer, R., Talarico, F., & Pyne, A. (2008). Late Neogene climate history of the Ross Embayment from the AND-1B drill core: culmination of three decades of Antarctic margin drilling. In: Cooper et al. (Eds). Antarctica: A keystone in a changing world. Proceedings of the 10th International Symposium on Antarctic Earth Sciences, Santa Barbara. The National Academies Press, Washington, DC, pp. 71–82. Naish, T. R., Powell, R. D., & Levy, R. H. (Eds). (2007). ANDRILL-MIS science team, 2007. Initial science results from AND-B, ANDRILL McMurdo Ice Shelf Project, Antarctica. Terra Antartica, 14, 109–328. Neil, H., Carter, L., & Morris, M. (2004). Thermal isolation of Campbell Plateau, marginal to the Antarctic circumpolar current over the past 130 k.y. Paleoceanography, 19, PA4008, doi:10.1029/2003PA000975. Ninnemann, U. S., & Charles, C. D. (2002). Changes in the mode of Southern Ocean circulation over the last glacial cycled revealed by foraminiferal stable isotopic variability. Earth Planet. Sci. Lett., 201, 383–396. North Greenland Ice Core Project Members (2004). High-resolution record of Northern Hemisphere climate extending into the last interglacial period. Nature, 431, 147–151. O’Brien, P. E., Cooper, A. K., Richter, C., et al. (2001). Prydz Bay-Cooperation Sea, Antarctica. Proceedings of the Ocean Drilling Program. Initial Reports, Vol. 188, 66 pp. [Online]. O’Brien, P. E., Goodwin, I., Forsberg, C-F., Cooper, A. K., & Whitehead, J. (2007). Late neogene ice drainage changes in Prydz Bay, East Antarctica and the interaction of Antarctic Ice Sheet evolution and climate. Paleaoceanogr. Paleoclimatol. Paleoecol., 245, 390–410. Oerlemans, J. (1982a). A model of the Antarctic Ice-Sheet. Nature, 297(5867), 550–553. Oerlemans, J. (1982b). Glacial cycles and ice-sheet modeling. Climatic Change, 4(4), 353–374. Orsi, A. H., Whitworth, T. III, & Nowlin, W. D. Jr. (1995). On the meridional extent and fronts of the Antarctic circumpolar current. Deep Sea Res., 42, 641–673. Pahnke, K., & Sachs, J. P. (2006). Sea surface temperatures of southern midlatitudes 0–160 kyr BP. Paleoceanography, 21, PA2003, doi:10.1029/2005PA001191. Pahnke, K., & Zahn, R. (2005). Southern Hemisphere water mass conversion linked with North Atlantic climate variability. Science, 307, 1741–1746.
524
T. Naish et al.
Pahnke, K., Zahn, R., Elderfield, H., & Schulz, M. (2003). 340,000-year centennialscale marine record of Southern Hemisphere climatic oscillation. Science, 301, 948–952. Paillard, D. (1998). The timing of Pleistocene glaciations from a simple multiplestate climate model. Nature, 391(6665), 378–381. Parrenin, F., Barnola, J. M., Beer, J., Blunier, T., Castellano, E., Chappellaz, J., Dreyfus, G., Fischer, H., Fujita, S., Jouzel, J., Kawamura, K., Lemieux, B., Loulergue, L., Masson-Delmotte, V., Narcisi, B., Petit, J. R., Raisbeck, G., Raynaud, D., Ruth, U., Schwander, J., Severi, M., Spahni, R., Steffensen, J. P., Svensson, A., Udisti, R., Waelbroeck, C., & Wolff, E. W. (2007a). The EDC3 chronology for the EPICA Dome C ice core. Climate Past, 3, 485–497. Parrenin, F., Dreyfus, G., Durand, G., Fujita, S., Gagliardini, O., Gillet, F., Jouzel, J., Kawamura, K., Lhomme, N., Masson-Delmotte, V., Ritz, C., Schwander, J., Shoji, H., Uemura, R., Watanabe, O., & Yoshida, N. (2007b). 1-D-ice flow modelling at EPICA Dome C and Dome Fuji, East Antarctica. Climate Past, 3, 243–259. Pattyn, F., & Siegert, M. (2007). Mechanisms for subglacial lake drainage and outbursts Paper presented at European Geosciences Union General Assembly, Vienna, Austria, 15–20 April, 1607-7962/gra/EGU2007-A-01324. Payne, A. J., Vielli, A., Shephard, A., et al. (2004). Recent dramatic thinning of largest west Antarctic ice stream triggered by oceans. Geophys. Res. Lett., 31(23), doi:10.1029/2004gl021284. Peltier, W. R., & Marshall, S. (1995). Coupled energy-balance ice-sheetm model simulations of the glacial cycle – A possible connection between terminations and terrigenous dust. J. Geophys. Res. Atmos., 100(D7), 14269–14289. Petit, J. R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J. M., Basile, I., Bender, M., Chappellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V. M., Legrand, M., Lipenkov, V. Y., Lorius, C., Pepin, L., Ritz, C., Saltzman, E., & Stievenard, M. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok Ice Core, Antarctica. Nature, 399, 429–436. Pollard, D. (1978). Investigation of Astronomical theory of ice ages using a simple climate ice sheet model. Nature, 272(5650), 233–235. Pollard, D. (1982). A simple ice-sheet model yields realistic 100 kyr glacial cycles. Nature, 296(5855), 334–338. Pollard, D., & PMIP, Group Members (2000). Comparisons of ice-sheet surface mass budgets from paleoclimate modeling intercomparison project (PMIP) simulations. Glob. Planet. Change, 24(2), 79–106. Porter, S., & Beget, J. (1981). Provenance and depositional environments of late Cenozoic sediments in permafrost cores from lower Taylor Valley, Antarctica. In: L. D. Mc Guinnis (Ed.). Dry Valley Drilling Project. Antarctic Research Series 33. AGU, Washington, DC, pp. 351–364. Powell, R. D. (1981). Sedimentation conditions in Taylor Valley, Antarctica inferred from textural analysis of DVDP cores. In: L. D. Mc Guinnis (Ed.). Dry
Late Pliocene–Pleistocene Antarctic Climate Variability
525
Valley Drilling Project. Antarctic Research Series 33. AGU, Washington, DC, pp. 281–314. Pudesy, C. J. (1990). Grain size and diatom content of hemipelagic sediments at site 697, ODP Leg 113: Record of Pliocene-Pleistocene climate. In: P. F. Barker, J. P. Kennett, et al. (Eds). Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 113, pp. 111–120. Pudsey, C. J., Barker, P. F., & Hamilton, N. (1988). Weddell Sea abyssal sediments: A record of Antarctic bottom water flow. Mar. Geol., 81, 234–289. Pudsey, C. J., & Howe, J. A. (1998). Quaternary history of the Antarctic circumpolar current: Evidence from the Scotia Sea. Mar. Geol., 148, 83–112. Rahmstorf, S. (2002). Ocean circulation and climate during the past 120,000 years. Nature, 419(6903), 207–214. Ramstein, G., Fabre, A., Pinot, S., et al. (1997). Ice sheet mass balance at the last glacial maximum. Ann. Glaciol., 21, 343–347. Raymo, M. E. (1997). The timing of major terminations. Paleoceanography, 12, 577–585. Raymo, M. E., Lisiecki, L. E., & Nisancioglu, K. H. (2006). Plio-Pleistocene ice volume, Antarctic climate, and the global delta O-18 record. Science, 313(5786), 492–495. Rebesco, M., Camerlenghi, A., Geletti, R., & Canals, M. (2006). Margin architecture reveals the transition to the modern Antarctic Ice Sheet ca. 3 Ma. Geology, 34, 301–304. Rintoul, S. R., Hughes, C. W., & Olbers, D. (2001). The Antarctic circumpolar current system. In: G. Siedler, J. Church, & J. Gould (Eds). Ocean Circulation and Climate: Observing and Modelling the Global Ocean. Academic Press, London, pp. 271–302. Ritz, C., Rommelare, V., & Dumas, C. (2001). Modeling the evolution of Antarctic Ice Sheet over the last 420,000 years: Implications for altitude changes in the Vostok Region. J. Geophys. Res. Atmos., 106(D23), 31943–31964. Rosenthal, Y., Boyle, E. A., & Labeyrie, L. D. (1997). Late glacial maximum paleochemistry and deepwater circulation in the Southern Ocean: Evidence from foraminiferal cadmium. Paleoceanography, 12, 787–796. Ro¨thlisberger, R., Mulvaney, R., Wolff, E. W., Hutterli, M. A., Bigler, M., Sommer, S., & Jouzel, J. (2002). Dust and sea salt variability in Central East Antarctica (Dome C) over the last 45 kyrs and its implications for southern highlatitude climate. Geophys. Res. Lett., 29(20), 1963, doi:10.1029/2002GLO15186. Sachs, J. P., & Anderson, R. F. (2005). Increased productivity in the subantarctic ocean during Heinrich Events. Nature, 434, 1118–1121. Sachs, J. P., Anderson, R. F., & Lehman, S. J. (2001). Glacial surface temperatures of the Southeast Atlantic Ocean. Science, 293, 2077–2079. Saltzman, B., & Verbitsky, M. Y. (1993). Multiple instabilities and modes of glacial rhythmicity in the Pliopleistocene – A general-theory of late Cenozoic climaticchange. Climate Dyn., 9(1), 1–15.
526
T. Naish et al.
Sarmiento, J. L., Gruber, N., Brzezinski, M. A., & Dunne, J. P. (2004). High latitude controls of the global nutricline and low latitude biological productivity. Nature, 427, 56–60. Scherer, R. P., Aldahan, A., Tulaczyk, S., Possnart, G., Englehart, H., & Kamb, B. (1998). Pleistocene collapse of the West Antarctic Ice Sheet. Science, 281, 98–100. Scherer, R. P., Bohaty, S. M., Dunbar, R. B., Esper, O., Flores, J-A., Gersonde, R., Harwood, D. M., Roberts, A. P., & Taviani, M. (2008). Antarctic records of precession-paced, insolation-warming during Pleistocene marine isotope stage 31. Geophys. Res. Lett., 35, accepted proof. Schoof, C. (2007). Ice sheet grounding line dynamics: Steady states, stability and hysteresis. J. Geophys. Res. Earth Surface, 112, FO3S28, doi: 10.1029/ 2006JF000664. Schragg, D. P., Hampt, G., Murray, D. W., et al. (1996). Pore fluid constraints on the temperature and oxygen isotope composition of the glacial ocean. Science, 372, 1930–1932. Severinghaus, J. P., Sowers, T., Brook, E. J., Alley, R. B., & Bender, M. L. (1998). Timing of abrupt climate change at the end of the Younger Dryas Interval from thermally fractionated gases in polar ice. Nature (Lond.), 391, 141–146. Shackleton, N. J. (2000). The 100,000 year ice-age cycle identified and found to lag temperature, carbon dioxide, and orbital eccentricity. Science, 289, 1897–1902. Shackleton, N. J., & Opdyke, N. (1976). Oxygen isotopes and paleomagnetic stratigraphy of Equatorial Pacific Core V28-1238: Oxygen isotope temperatures and ice volumes on a 105 to 106 year scale. Quaternary Res., 3, 39–59. Shemesh, A., Hodell, D., Crosta, X., Kanfoush, S., Charles, C., & Guilderson, T. (2002). Sequence of events during the last deglaciation in Southern Ocean sediments and Antarctic Ice cores. Paleoceanography, 17, 1056, doi:10.1029/ 2000PA000599. Shulmeister, J., Goodwin, I., Renwick, J., Harle, K., Armand, L., McGlone, M. S., Cook, E., Dodson, J., Hesse, P. P., Mayewski, P., & Curran, M. (2004). The Southern Hemisphere westerlies in the Australasian sector over the last glacial cycle: A synthesis. Quaternary Int., 118–119, 23–53. Siegenthaler, U., Stocker, T. F., Monnin, E., Luthi, D., Schwander, J., Stauffer, B., Raynaud, D., Barnola, J. M., Fischer, H., Masson-Delmotte, V., & Jouzel, J. (2005). Stable carbon cycle-climate relationship during the late Pleistocene. Science, 310, 1313–1317. Siegert, M. J. (2005). Lakes beneath the ice sheet: The occurrence, analysis, and future exploration of Lake Vostok and other Antarctic Subglacial Lakes. Ann. Rev. Earth Planet. Sci., 33, 215–245. Sigman, D., & Boyle, E. (2000). Glacial/interglacial variations in atmospheric carbon dioxide. Nature, 407, 859–869. Sikes, E. L., Howard, W. R., Neil, H. L., & Volkman, J. K. (2002). Glacialinterglacial sea surface temperature changes across the subtropical front east of
Late Pliocene–Pleistocene Antarctic Climate Variability
527
New Zealand based on alkenone unsaturation ratios and foraminiferal assemblages. Paleoceanography, 17, 1–13. Spahni, R., Chappellaz, J., Stocker, T. F., Loulergue, L., Hausammann, G., Kawamura, K., Flu¨ckiger, J., Schwander, J., Raynaud, D., Masson-Delmotte, V., & Jouzel, J. (2005). Atmospheric methane and nitrous oxide of the Late Pleistocene from Antarctic ice cores. Science, 310, 1317–1321. Steig, E. J. (2001). No two latitudes alike. Science, 293, 2015–2016. Steig, E. J., Brook, E. J., White, J. W. C., Sucher, C. M., Bender, M. L., Lehman, S. J., Morse, D. L., Waddington, E. D., & Clow, G. D. (1998). Synchronous climate changes in Antarctica and the North Atlantic. Science, 282, 92–95. Stenni, B., Masson-Delmotte, V., Johnson, S., Jouzel, J., Longinelli, A., Monnin, E., Ro¨thlisberger, R., & Selmo, E. (2001). An oceanic cold reversal during the last deglaciation. Science, 293, 2074–2077. Stocker, T. F., & Johnsen, S. J. (2003). A minimum thermodynamic model for the bipolar seesaw. Paleoceanography, 18, 1087. Tarasov, L., & Peltier, W. R. (1997). Terminating the 100 kyr ice age cycle. J. Geophys. Res. Atmos., 102(D18), 21665–21693. Tarasov, L., & Peltier, W. R. (2004). A geophysically constrained large ensemble analysis of the deglacial history of the North American ice-sheet complex. Quaternary Sci. Rev., 23(3–4), 359–388. Taylor, K. C., White, J. W. C., Severinghaus, J. P., Brook, E. J., Mayewski, P. A., Alley, R. B., Steig, E. J., Spencer, M. K., Meyerson, E., Meese, D. A., Lamorey, G. W., Grachev, A., Gow, A. J., & Barnett, B. A. (2004). Abrupt climate change around 22 ka on the Siple Coast of Antarctica. Quaternary Sci. Rev., 23, 7–15. Tchernia, P., & Jeannin, P. F. (1983). Quelques Aspects de la circulation oce´anique Antarctique re´ve´le´s par l’observation de la de´rive d’icebergs (1972–1983). Centre National d’Etudes Spatiales, Expedition Polaires Francaises, Museum National d’Histoire Naturelle, Paris, 93 pp. Thiessen, K. M., Dunbar, R. B., Cooker, A. K., Mucciarone, D., & Hoffmann, D. (2003). The pleistocene evolution of the East Antarctic Ice Sheet in Prydz Bay region; stable isotopic evidence from ODP site 1167. Glob. Planet. Change, 39, 227–256. Thiede, J. (1979). Wind regimes over the late Quaternary Southwest Pacific Ocean. Geology, 7, 259–262. Thompson, S. L., & Pollard, D. (1997). Greenland and Antarctic mass balances for present and doubled atmospheric CO2 from the GENESIS version-2 global climate model. J. Climate, 10(5), 871–900. Toggweiler, J. R., Russell, J. L., & Carson, S. R. (2006). Midlatitude westerlies, atmospheric CO2, and climate change. Paleoceanography, 21, PA2005, doi:10.1029/2005PA001154. Turney, C. S. M., McGlone, M. S., & Wilmshurst, J. M. (2003). Asynchronous climate change between New Zealand and the North Atlantic during the last deglaciation. Geology, 31, 223–226.
528
T. Naish et al.
Tziperman, E., Raymo, M. E., Huybers, P., & Wunsch, C. (2006). Consequences of pacing the Pleistocene 100 kyr ice ages by nonlinear phase locking to Milankovitch forcing. Paleoceanography, 21, PA4206, doi:10.1029/ 2005PA001241. Uddstrom, M. J., & Oien, N. A. (1999). On the use of high resolution satellite data to describe the spatial and temporal variability of sea surface temperatures in the New Zealand Region. J. Geophys. Res., 104, 20,729–20,751. Valdes, P. J., Beerling, D. J., & Johnson, C. E. (2005). The ice age methane budget. Geophys. Res. Lett., 32, L02704, doi:10.1029/2004GL021004. Van Kreveld, S., Sarnthein, M., Erlenkeuser, H., Grootes, P., Jung, S., Nadeau, M., Pflaumann, U., & Voelker, A. (2000). Potential links between surging ice sheets, circulation changes, and the Daansgard-Oershger cycles in the Irminger Sea, 60-18-kyr. Paleoceanography, 15, 425–442. Venuti, A., Florindo, F., Michel, E., & Hall, I. R. (2007). Magnetic proxy for the deep (Pacific) western boundary current variability across the mid-Pleistocene climate transition. Earth Planet. Sci. Lett., 259, 107–118. Vettoretti, G., & Peltier, W. R. (2004). Sensitivity of glacial inception to orbital and Greenhouse gas climate forcing. Quaternary Sci. Rev., 23(3–4), 499–519. Villa, G., Lupi, C., Cobianchi, M., Florindo, F., & Pekar, S. F. (2007). A Pleistocene warming event at 1 Ma in Prydz Bay, East Antarctica: Evidence from ODP Site 1165. Paleogeog. Paleoclimatol. Paleoecol., 260, 230–255. Vogel, S. W., Tulaczyk, S., & Joughin, I. R. (2003). Distribution of basal melting and freezing beneath tributaries of ice stream C: Implication for the Holocene Decay of the West Antarctic Ice Sheet. Ann. Glaciol., 36, 273–282. Warner, R. C., & Budd, W. F. (1998). Modelling the long-term response of the Antarctic Ice Sheet to global warming. Ann. Glaciol., 27, 161–168. Watanabe, O., Jouzel, J., Johnsen, S., Parrenin, F., Shoji, H., & Yoshida, N. (2003). Homogeneous climate variability across East Antarctica over the past three glacial cycles. Nature, 422, 509–512. Weaver, A. J., Saenko, O. A., Clark, P. U., & Mitrovica, J. X. (2003). Meltwater pulse 1A from Antarctica as a trigger of the Bolling-Allerod warm interval. Science, 299, 1709–1713. Weaver, P. P. E., Carter, L., & Neil, H. L. (1998). Response of surface water masses and circulation to late Quaternary climate change, east of New Zealand. Paleoceanography, 13, 70–83. Webb, P. N., & Harwood, D. M. (1991). Late Cenozoic glacial history of the Ross Embayment, Antarctica. Quaternary Sci. Rev., 10(2–3), 215–223. Weertman, J. (1976). Milankovitch solar-radiation variations and ice age ice sheet sizes. Nature, 261(5555), 17–20. Wefer, G., Berger, W. H., Bijma, J., & Fischer, G. (1999). Clues to ocean history: A brief overview of proxies. In: G. Fischer, & G. Wefer (Eds). Use of Proxies in Paleoceanography. Springer-Verlag, Berlin Heidelberg, pp. 1–68. Whitehead, J. M., Harwood, D. M., McKelvey, B. C., Hambrey, M. J., & McMinn, A. (2004). Diatom biostratigraphy of the Cenozoic Fjordal Pagadroma
Late Pliocene–Pleistocene Antarctic Climate Variability
529
Group, northern Prince Charles Mountains, East Antarctica. Aust. J. Earth Sci., 51, 521–547. Whitehead, J. M., & McKelvey, B. C. (2001). The stratigraphy of the Pliocne – Lower Pleistocene Bardin Bluffs formation, Amery Oasis, northern Prince Charles Mountains, Antarctica. Antarct. Sci., 13, 79–86. Whitworth, T. III (1988). The Antarctic circumpolar current. Oceanus, 32, 53–58. Williams, P. W., King, D. N. T., Zhao, J.-X., & Collerson, K. D. (2004). Late Pleistocene to Holocene composite speleothem 18O and 13C chronologies from South Island, New Zealand – Did a global Younger Dryas really exist? Earth Planet. Sci. Lett., 230, 301–317. Wilson, G., Levy, R., Browne, G., Dunbar, N., Florindo, F., Henrys, S., Graham, I., McIntosh, W., McKay, R., Naish, T., Ohneiser, C., Powell, R., Ross, J., Sagnotti, L., Scherer, R., Sjunneskog, C., Strong, C. P., Taviani, M., & Winter, D., & ANDRILL-MIS Science Team. (2007). Preliminary integrated chronostratigraphy of the AND-1B core, ANDRILL McMurdo Ice Shelf Project, Antarctica. Terra Antartica, 14, 297–316. Wolff, E. W., Fischer, H., Fundel, F., Ruth, U., Twarloh, B., Littot, G. C., Mulvaney, R., Rothlisberger, R., de Angelis, M., Boutron, C. F., Hansson, M., Jonsell, U., Hutterli, M. A., Bigler, M., Lambert, F., Kaufmann, P., Stauffer, B., Stocker, T. F., Steffensen, J. P., Siggaard-Andersen, M. L., Udisti, R., Becagli, S., Castellano, E., Severi, M., Wagenbach, D., Barbante, C., Gabrielli, P., & Gaspari, V. (2006). Southern ocean sea-ice extent, productivity and iron flux over the past eight glacial cycles. Nature, 440, 491–496. Wunsch, C. (2003). The spectral description of climate change including the 100 ky energy. Climate Dyn., 20, 353–363. Zweck, C., & Huybrechts, P. (2005). Modeling of the Northern Hemisphere ice sheets during the last glacial cycle and glaciological sensitivity. J. Geophys. Res. Atmos., 110(D7), doi:10.1029/2004jd005489.
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00012-8
Chapter 12
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene A. P. Wright1,, D. A. White2, D. B. Gore2 and M. J. Siegert1 1
School of GeoSciences, Grant Institute, University of Edinburgh, The King’s Buildings, West Mains Road, Edinburgh, EH9 3JW 2 Department of Physical Geography, Macquarie University, Sydney, NSW 2109, Australia
ABSTRACT Recent technological advances in the study and dating of both land and marine glacial geologic features combined with improvements in both glaciological and post-glacial isostatic rebound modelling have led to significant improvements in our knowledge and understanding of the Antarctic Ice Sheets at the Last Glacial Maximum (LGM) and their subsequent changes throughout the Holocene. Here we review the geological evidence for the extent and timing of the maximum advance of the East and West Antarctic Ice Sheets and the ice cover of the Antarctic Peninsula during the most recent glacial cycle. We also discuss evidence for the rate and timing of Holocene Ice Sheet retreat. Geological data provide a very important ‘first-hand’ record of ice-sheet changes over a range of time periods. They are also useful for constraining and improving models which then have the potential to both fill in the gaps for which geological data are unavailable, and to make predictions of the future. Inspection of the geological record allows us to form qualitative scenarios concerning glacial history. Numerical modelling has been used on several occasions to test such hypotheses. We discuss such numerical studies, indicating both their importance and limitations in order to develop quantitative ideas about the late Quaternary history of the ice sheet. An important environmental aspect of Antarctica’s glacial history is its contribution to global sea level rise since the LGM. The past decade has seen the range of estimates (from reconstructions based solely on geological evidence, on glaciological Corresponding author. Tel.: +44(0)131 650 7339;
E-mail:
[email protected] (A.P. Wright).
532
A. P. Wright et al.
modelling constrained by geology and on modelling of the isostatic rebound) change from 0.5–38 m sea level equivalent to 5.9–19.2 m. Although the convergence of estimates is encouraging, there is a need for further glaciological modelling to take full account of the constraints placed on LGM ice expansion, particularly in East Antarctica, by new geological evidence.
12.1. Introduction Antarctica is comprised of two main grounded ice sheets; the West Antarctic and the East Antarctic Ice Sheets separated by the Transantarctic Mountains. Another significant region of grounded ice exists in the Antarctic Peninsula where a series of ice caps and glaciers are located (Fig. 12.1). The present volume of ice in Antarctica is around 25.4 106 km3 (equivalent to 57 m of global sea level) of which 91% is within the East Antarctic Ice Sheet (Lythe et al., 2001). The maximum ice sheet surface elevation of 4093 m occurs at Dome Argus which, along with several other subsidiary ice domes connected by ridges, forms a major ice divide through the centre of East Antarctica. Ice thickness in the central regions of the continent varies between 2800 and 4500 m mainly as a consequence of bedrock topography which is known to vary spatially by more than a vertical kilometre over just a few km. If the ice were to be removed from East Antarctica, the bedrock surface would be largely above sea level. However, if the same were to happen over West Antarctica, even accounting for isostatic rebound, the bedrock would remain below the modern sea level. Thus, the West Antarctic Ice Sheet is often referred to as a marine-based ice sheet. Ice drains from the interior domes via fast flowing rivers of ice known as ‘ice streams’, where the dominant method of flow is by sliding or basal sediment deformation. These contribute grounded ice to numerous floating ‘ice shelves’ that surround the grounded ice continent. The Ross and the Filchner-Ronne ice shelves are Antarctica’s largest, with areas of 472,000 and 430,000 km2, respectively. Icebergs, usually of tabular form, calve from the marine margins of ice shelves, where ice thickness is usually 250–300 m. This process represents an important mechanism by which ice is lost from the ice-sheet system, accounting for 75–85% of all ice lost from Antarctica (Jacobs et al., 1992). A significant amount is also lost due to bottom melting near the grounding lines of some outlet glaciers, particularly those where the grounding line is well below sea level (Rignot and Jacobs, 2002).
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
533
Figure 12.1: Location map and surface elevation of Antarctica with place names mentioned in the text. Contours are at 500 m intervals.
12.2. Response of the Ice Sheets to Glacial Climate and Late Quaternary Ice-Sheet Reconstructions The climate around Antarctica during ‘‘full glacial’’ periods such as the LGM is likely to have been highly conducive to the presence of ice-sheets. Therefore, the ‘minimum’ reconstruction for the LGM Antarctic Ice Sheets is similar to the present geometry. However, due to the sea-level reduction of around 120 m (e.g., Bard et al., 1990; Shackleton, 2000) and ocean temperature changes that occurred during the LGM, the ice-sheets in Antarctica are known to have grown out towards the continental shelf edge in several places. The ice shelves bordering the Antarctic Ice Sheet thickened, grounded and became part of the parent ice mass. A ‘maximum’
534
A. P. Wright et al.
reconstruction would therefore be an Antarctic Ice Sheet which expanded to reach the continental shelf break right around the continent. There is some ice core evidence from around Antarctica, including Taylor Dome and Law Dome, which suggests a much lower rate of accumulation at the LGM. The Taylor Dome core also shows that LGM storm paths came from a different direction to the modern day. This suggests a reorganisation of the climate system at least at this part of the ice sheet (Morse et al., 1998). While reductions in both sea-level and air temperature at the LGM would have been conducive to ice-sheet expansion, decreasing accumulation may well have had a restraining effect on icesheet growth. To resolve how the ice sheet responded to such complicated changes in forcing requires the application of numerical models, which we will return to later. Ice-sheet behaviour in the late Quaternary is coupled with the cryosphere, ocean and atmosphere. For example, if sea ice extent increased due to cooling of the air temperature over the Southern Ocean (Armand, 2000), then the moisture supply to the ice sheet may have decreased and so ice-sheet growth may have been impeded. This introduces the possibility that independent mountain glaciers in Antarctica may have responded differently to LGM conditions than areas connected to an ice-sheet. In other words, ice-sheet advance may have been related to sea level fall while, at the same time, glacier decay may be related to a significant reduction in rates of precipitation. Since the CLIMAP (1976) reconstruction of Ice Age Earth, there have been numerous reconstructions of LGM ice extent in Antarctica based on evidence from sediments and geomorphology, isostatic rebound and ice-flow modelling. These reconstructions provide estimates ranging from 0.5–2 to 38 m of sea level equivalent (e.g., Budd and Smith, 1982; Nakada and Lambeck, 1988; Colhoun et al., 1992). While the techniques used vary, many of the early estimates placed the grounding line at the edge of the continental shelf, and have relatively steep ice profiles. As such these are at the upper end of estimates of sea-level lowering. More recent work has advanced our understanding of the dynamics of ice flow (in particular the longitudinal profile of ice streams and their controls), produced sophisticated 3D thermomechanical ice-sheet models and provided reliable geomorphic evidence that indicates relatively limited ice expansion in some areas. These advances in understanding have rendered the upper estimates implausible, such that recent reconstructions indicate an increase of Antarctic ice volume at the LGM with respect to today of only 5.9–19.2 m of sea level equivalent (Bentley, 1999; Huybrechts, 2002).
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
535
12.3. Geological Information 12.3.1. Constraining Late Quaternary Ice-Sheet Extent, Volume and Timing The geometry of the Antarctic Ice Sheet throughout the last glacial cycle has been compiled using a number of different methods. In terrestrial environments, the aerial extent of former ice cover is generally recognised by mapping the lateral extent of ice marginal landforms such as moraines or ice marginal lake sediments, or by mapping subglacial landforms such as drumlins and striae. Ice volume can equally be reconstructed by mapping glacial debris striae, the locations of erratic boulders and the height/extent of mountain trimlines. The relative age of each mapped unit is usually constrained using a measure of exposure age, most often by the degree of weathering (such as soil formation or development of tafoni or iron staining on boulders), and more recently by measurement of isotopes produced by cosmogenic radiation. Stratigraphic correlations are generally difficult due to lack of natural exposures. Similar techniques are also used in marine environments, but rather than investigating the geomorphology of former glacial landscapes using field mapping and aerial photography, the ocean floor is imaged using techniques such as swath bathymetry and side-scan sonar. These techniques commonly identify ice marginal landforms such as moraines, or subglacial landforms such as mega scale glacial lineations, meltwater channels and drumlins. Stratigraphic techniques are more useful in the marine environment due to the ease of data acquisition via acoustic and seismic surveys, particularly when these sediment units are also sampled by gravity or piston cores. Establishing detailed late Quaternary glacial chronologies in Antarctica can be challenging (e.g., Anderson et al., 2002; Ingo´lfsson, 2004). The most direct dating is usually achieved in presently terrestrial environments, through cosmogenic exposure dating of moraines or radiocarbon dating of fossils contained within emergent marine sediments or ice-marginal lakes. However, factors such as production rate uncertainties (e.g., Gosse and Phillips, 2001), recycling of clasts with prior exposure and post-depositional reworking of glacial sediments (e.g., Brook et al., 1995) reduce the precision of cosmogenic chronologies to around 10% (Putkonen and Swanson, 2003). Where such materials are preserved, the algal mats commonly found in proglacial lake sediments can provide reasonably accurate radiocarbon ages, provided that the water column is well mixed and that at least part of the lake surface ice cover melts out each summer (Gore 1997b; Hendy and Hall, 2006). Limiting ages on glacial events in terrestrial environments can
536
A. P. Wright et al.
sometimes also be constrained by radiocarbon dating of the small amount of organic material sometimes found in ice-marginal or postglacial sediments (e.g., Burgess et al., 1994). Carbon sourced from marine environments, such as the bodies of seals and penguins or carbonate shells in former (and emergent) marine environments is subject to a marine reservoir effect that varies around the continent, but is on the order of 1.3 ka (e.g., Berkman and Forman, 1996; but also see Kiernan et al., 2003). On the continental shelf the majority of published data with regard to the timing of ice advance and retreat have been obtained by Atomic Mass Spectrometry (AMS) radiocarbon dating of the acid-insoluble (mostly diatom) fraction of organic material that is preserved immediately above or below a layer of glacially derived sediment. There are several potential problems with this technique, for example the influx of ‘‘dead’’ carbon from ice melt, fine-grained geogenic carbon such as graphite and recycling of older organic matter means that such material at the modern sediment/water interface provides radiocarbon ages of W2 ka, and sometimes W6 ka in areas of little biological productivity (e.g., Andrews et al., 1999). These values exceed the age derived from carbonate shells from living biota at such sites by over 1.3 ka. To some extent this problem has been rectified in more recent studies by AMS analysis of discrete carbonate shells (e.g., Leventer et al., 2006), but the lack of shells in many cores means that this approach is not always possible. Despite these difficulties, considerable efforts have been made in order to constrain the timing of ice advance and retreat across the continental shelf in many sectors of the continent. 12.3.2. Last Interglacial (120 ka BP) Data from a number of sources, including dated coral terraces and the oceanic d18O record, have fixed the timing of the Eemian interglacial at 128– 116 ka BP (Stirling et al., 1998; Shackleton et al., 2003). During this period there was less ice in the world than at present and global sea level was between 3 and 6 m higher than today (Stirling et al., 1998; Overpeck et al., 2006). Some believe that this was caused by a smaller West Antarctic Ice Sheet (e.g., Mercer, 1978; Scherer et al., 1998), while others suggest the decay of the Greenland Ice Sheet was responsible (e.g., Cuffey and Marshall, 2000; Tarasov and Peltier, 2003). Evidence to support the reduction in volume of at least part of the West Antarctic Ice Sheet at the last interglacial comes from observations of marine diatoms of Quaternary age and the measurement of high concentrations of 10Be in sediments recovered from beneath the
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
537
West Antarctic Ice Sheet (Scherer et al., 1998). 10Be and diatoms are thought to accumulate only under open marine conditions; when the sea is ice covered there is very little deposition of these materials. Moreover, if ice is grounded, any small amounts of 10Be and diatoms will be eroded. Sediments retrieved from beneath Ice Stream B, on the Siple Coast of the Ross Sea embayment, were probably emplaced during the last interglacial when marine conditions prevailed. The current dynamics of the ice-sheet may be affected by the location of these low shear strength, now subglacial, sediments (Anandakrishnan et al., 1998; Studinger et al., 2001). Similarly the Institute and Mo¨ller ice streams may have been free of grounded ice at the LGM, which according to Bingham and Siegert (2007) could account for B2 m of sea-level rise during the Eemian interglacial. 12.3.3. Last Glacial Maximum, the Holocene to the Present Day (B20–0 ka BP) Regardless of ice-sheet geometry at the last interglacial, there is persuasive evidence from the geological record to indicate that the Antarctic Ice Sheet was larger than present around the time of the global sea level lowstand at B20 ka BP, although the extent of this expansion is well constrained at only a few sites around the continental margin. Evidence of former ice surface elevations is particularly sparse, in part due to the lack of sites at which such information can be preserved, but also due to the difficulty in accessing remote inland mountain ranges where this evidence might occur. In the next sections we discuss glacial history by dividing Antarctica into seven sectors, based on lines of longitude (Fig. 12.2), which are approximately based on ice divides but centred on the geographical pole.
12.3.3.1. 0–601E: Queen Maud land/Enderby land Schirmacher Oasis (11.51E) is a 34 km2 area located behind Novolazarevskaya Ice Shelf. Infra-red stimulated luminescence dating of lake floor sediments yielded burial ages of B52 ka BP in Lake Glubokoye (Krause et al., 1997), which was in part corroborated by AMS 14C ages of 35 ka BP from 2 m above the glacial diamict at the base of the lake sediments. On the basis of these data, Schirmacher Oasis was ice free throughout the LGM. On the northern side of Fimbulheimen, the mountain range to the south of Shirmacher Oasis, deposition of mumijo (proventricular ejecta of the snow petrel, Pagodroma nivea) throughout the LGM indicates that ice thickening in
538
A. P. Wright et al.
Figure 12.2: Locations of the sectors discussed in this chapter; 1: 0–601E Queen Maud/Enderby Land. 2: 60–751E Mac.Robertson Land/Lambert Gl/Prydz Bay. 3: 75–1501E Princess Elizabeth/Queen Mary/Wilkes Land. 4: 1501E–1501W Transantarctic Mountains/Ross Sea Embayment. 5: 150–851W Marie Byrd Land/Ellsworth Land and the Amundsen and Bellingshausen Seas. 6: 85–601W Antarctic Peninsula/Bellingshausen Sea. 7: 60–01W Weddell Sea/Filchner Ice Shelf/Coats Land. this sector of the ice-sheet was also limited, with o80 m thickening occurring at the Insel Range (721S, 111E) during this time (Hiller et al., 1995). The well preserved 46–30 ka BP shorelines at Lutzow-Holm Bay (37.51E) indicate that this site probably remained ice-free through the LGM, constraining the expansion of the ice-sheet in this area (Igarashi et al., 1995; Igarashi et al., 1998). Holocene shorelines are also present at this site, reaching at least 15 m above present sea level and dating from 8 ka to the present day (Maemoku et al., 1997; Miura et al., 1998), suggesting a decrease in regional ice load in this area since the LGM (Nakada et al, 2000). Despite an abundance of small ice free areas that are suitable for Quaternary studies, including Riiser-Larsen (50.71E), Øygarden Group (57.51E) and Stillwell Hills (59.31E), there is a paucity of information regarding the geometry of this sector of the ice-sheet during or following the LGM.
12.3.3.2. 60–751E: Mac.Robertson Land/Lambert Glacier-Amery Ice Shelf/Prydz Bay 10
Be and 26Al cosmogenic isotope exposure ages on glacial erratics show that the ice-sheet thickness at Framnes Mountains (62.51E) was B350 m greater
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
539
than present during the LGM, had begun lowering at 13 ka BP, and had reached the modern ice margin by 6 ka BP (Mackintosh et al., 2007). This evidence agrees reasonably well with side-scan sonar/piston core data, which indicate that ice had retreated from the mid-outer continental shelf in Nielsen Basin B80 km to the east of Framnes Mountains by B11 cal ka BP and had reached the inner shelf by 6 cal ka BP (Harris and O’Brien, 1998). Former ice advances of the Lambert Glacier-Amery Ice Shelf system (701E) into Prydz Bay (Fig. 12.3) are relatively well constrained through geophysical investigations and marine sediment cores (Domack et al., 1998; Taylor and McMinn, 2002), and by drilling through the ice shelf (Hemer and Harris, 2003). The sedimentary record indicates that ice was grounded on shallow (o500 m water depth) banks in Prydz Bay and across most of the area currently occupied by the Amery Ice Shelf at the LGM. However, the base of an uninterrupted sequence of glacial marine and biogenic sediments containing a relative abundance of organic material has been AMS radiocarbon dated to at least 30 ka BP within the Prydz Channel, a 700 m deep trench that cuts across the continental shelf in Prydz Bay (Domack et al., 1998). This suggests that grounded ice may not have extended all the way to the continental shelf at the LGM. Ages from acid-insoluble organic matter offshore from Amery Ice Shelf suggest a retreat timing of 11.5 cal ka BP for the Lambert Glacier-Amery Ice Shelf System (Domack et al., 1998). Investigations into glacial and lake sediments deposited on mountains flanking the major outlet glaciers (Wagner et al., 2004; White and Hermichen, in press) also constrain the thickness of the ice sheet at the LGM. With the exception of the area around the southern tip of the modern Amery Ice Shelf, glacial sediments dating from the LGM are restricted to o200 m above the modern ice margin. Also, there are no emergent shorelines around epishelf Beaver Lake (Adamson et al., 1997), but there are subaerially deposited postglacial sediments on the lake bottom at 60 m below modern sea level (Wagner et al., 2007). These data support the results from the continental shelf, and indicate that both thickening and expansion of ice in this region at the LGM was limited. One reason for this may be the deep basin along which the outlet glaciers flow, which reduces the ability of the ice-sheet to advance to the continental shelf edge (Taylor et al., 2004).
12.3.3.3. 75–1501E: Princess Elizabeth Land/Queen Mary Land/Wilkes Land Larsemann Hills (76.21E) hosts lake sediments dated to 14C background (W42 ka BP), and lake sediment stratigraphy which has been used to infer continual exposure since the last interglacial (Hodgson et al., 2005).
540
A. P. Wright et al.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
541
An optically stimulated luminescence age of 21 ka BP was obtained from glaciofluvial sediments within 500 m of the ice margin (Hodgson et al., 2001), supporting interpretations of continual exposure through the LGM. Emergent shorelines are o3 m asl, consistent with relatively minor former ice loading. Vestfold Hills (781E) has emergent marine shorelines to o10 m asl (Zwartz et al., 1998), and cosmogenic 10Be analyses reveal that the ice margin had retreated to within 5 km of the present margin by 12.5–9 ka BP (Fabel et al., 1997). There has been a minor (o4 km) readvance of ice along the flank of Sørsdal Glacier during the late Holocene (Adamson and Pickard, 1983; Gore, 1997a). Gaussberg (89.21E) is a 370 m high, glacially striated volcano on the coast. The benched morphology and presence of palagonite encrusted pillow lavas indicates that this eruption occurred in a water filled subglacial vault, and that the ice-sheet has since retreated to its present position since the eruption at 56 ka BP (Tingey et al., 1983). Bunger Hills (1011E) has emergent marine shorelines to o11 m asl (Colhoun and Adamson, 1992; Colhoun et al., 1992). Optically stimulated luminescence ages from glacial lake shorelines and glaciofluvial sediments indicate that deglaciation commenced B40–30 ka BP (Gore et al., 2001), with the area largely deglaciated by B25 ka BP. Like Vestfold Hills, the oasis attained its present form around 11 ka BP. These data are corroborated by 14 C ages indicating exposure through the LGM (Hedges et al., 1996, cited in Krause et al., 1997). Windmill Islands (110.31E) has emergent marine shorelines to 35 m asl (Goodwin, 1993; Goodwin and Zweck, 2000), with deglaciation of the southern islands by 11 cal ka BP and the northern peninsulas by 8 cal ka BP (Kirkup et al., 2002). Offshore from the large Mertz and Ninnis glaciers (145–1501E) of George V Land swath bathymetry has been used to identify mega-scale glacial Figure 12.3: (A) LGM ice-sheet surface reconstruction in Mac.Robertson Land and eastern Princess Elizabeth Land. Grey lines indicate present icesheet contours (in m asl), brown polygons indicate ice-free areas, and grey areas indicating ice-sheet flow W300 ma 1 (Joughin, 2002). Solid green lines indicate LGM ice-sheet contours derived from field studies, with investigated sites indicated by red circles (Domack et al., 1998; Harris and O’Brien, 1998; Zwartz et al., 1998; Hodgson et al., 2001; Taylor and McMinn, 2002; Whitehead et al., 2003; Leventer et al., 2006; K. Lilly, Pers. Comm., 2007; White and Hermichen, in press). Dashed lines indicate areas where heights or grounding lines are poorly known. (B) Profile through the cross-section AB.
542
A. P. Wright et al.
lineations within the Mertz Trough which indicate LGM expansion of grounded ice to the outer continental shelf in this part of the ice-sheet (McMullen et al., 2006). Successive grounding zone wedge deposits have also been imaged in this region which record pauses in the retreat of the ice-sheet across the continental shelf, these however are yet to be conclusively dated (McMullen et al., 2006).
12.3.3.4. 1501E–1501W: Transantarctic Mountains/Ross Sea Embayment During the 1990s a comprehensive geophysical and sedimentological dataset was compiled across the Ross Sea using seismic profiles, swath bathymetry and side-scan sonar imagery of the sea-floor sediments in front of the present day Ross Ice Shelf (Anderson,1999; Domack et al., 1999; Shipp et al., 1999). Lineations, drumlins and large-scale grooves provide a firm basis for the identification of former fast-flowing, marine based ice streams situated in bathymetric troughs which extend right to the edge of the continental shelf in places but only to the mid-outer parts of the shelf in others (Fig. 12.4). Sediment cores taken from these mega-scale lineations have retrieved
Figure 12.4: LGM paleodrainage map for the Ross Sea based on geomorphic features on the shelf and including data from Shipp et al. (1999). Arrows are flow directions based on lineations and the dashed line is the maximum grounding line position. Taken from Mosola and Anderson (2006), and reproduced with permission from Elsevier.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
543
unconsolidated Plio-Pleistocene strata that are interpreted to have composed the readily deformable bed which allowed such extended ice streams to develop (Mosola and Anderson, 2006). Lateral moraines deposited by the many glaciers flowing through the Transantarctic Mountains and into the Ross Sea during the LGM have very similar profiles. These moraines are located high above the present day ice surface at the feet of these glaciers but merge toward the modern-day surface of the ice-sheet when traced upstream. LGM thickening indicated by these moraines therefore increases towards the marine margin, indicative of a thicker, grounded Ross Sea ice-sheet whereas the interior of the ice sheet in East Antarctica appears to have maintained a relatively constant thickness (Broecker and Denton, 1990). Deglaciation of the Ross Sea is thought to have begun first at the western margin and proceeded like a ‘swinging gate’ hinged at the eastern side of the basin, with the LGM Ross Sea ice-sheet largely composed of extensions of the present day Siple Coast ice streams (Conway et al., 1999). The contribution of ice from East Antarctic glaciers, flowing through the Transantarctic Mountains, to the Ross Sea ice-sheet remains contentious. The mineralogical composition of glacial diamicts from the floor of the eastern Ross Sea is similar to that of source areas on the Siple Coast, whereas diamicts from the western Ross Sea match with sources in the Transantarctic Mountains (Licht et al., 2005), indicating more complicated glacial flow patterns at the LGM than a simple expansion of the present day Siple Coast ice streams (Fig. 12.5). Recent model investigations, which have attempted to reconstruct the late Holocene change in the thickness of Siple Dome from the depth-age relationship of the Siple Dome ice core and plausible scenarios for the changes in accumulation rate since the LGM, have led to estimates of surface height increase of 200–400 m. For the ice-sheet to have extended B1000 km to the continental shelf edge, the required ice surface gradient would have been much shallower than that of present day ice streams, and thus only possible by invoking a very slippery bed (Waddington et al., 2005). This supports the interpretation that ice from the Transantarctic Mountains must have contributed significantly to the Ross Sea ice sheet. The timing of ice retreat in the eastern and central Ross Sea is poorly understood, but reworked foraminifera in glacial diamict on the outer shelf indicate that the maximum ice extent occurred after 16.5 cal ka BP. In contrast, there are numerous lines of both onshore and offshore evidence that constrain the timing of deglaciation in the western half of the Ross Sea. A recent review of the evidence from the marine record concluded that the maximum ice extent occurred after B17 cal ka BP (Licht, 2004),
544
A. P. Wright et al.
Figure 12.5: Flow lines for the Ross Ice Sheet during the last glaciation proposed by Licht et al. (2005) based on Ross Sea, East and West Antarctic till mineralogy and lithology. Dashed lines represent inferred flow due to lack of sample coverage. Reproduced with permission from Elsevier. with deglaciation beginning by 14 cal ka BP (Domack et al., 1999; Shipp et al., 1999). The grounding line retreated relatively slowly along the western margin of the Ross Sea, reaching the modern ice shelf edge by 7 cal ka BP (Domack et al., 1999). This marine-based deglacial chronology is supported by evidence from terrestrial studies, including the timing of penguin recolonisation along the coast (Baroni and Orombelli, 1994), the altitude of large proglacial lakes in the McMurdo Dry Valleys dammed by an expanded Ross Sea ice sheet (Hall et al., 2002), cosmogenic exposure ages from moraines deposited by ice in the McMurdo Dry Valleys (Brook et al., 1995) and moraines and proglacial lake sediments on the Hatherton Glacier (Bockheim et al., 1989).
12.3.3.5. 150–851W: Marie Byrd Land/Ellsworth Land and the Amundsen and Bellingshausen Seas Exposure age dated moraines and recessional deposits indicate that ice thickness in Marie Byrd Land was around 45 m greater than present when deglaciation began around 10 ka BP (Ackert et al., 1999). Further surface
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
545
exposure dating studies support the view that ice in this region was significantly thicker at the LGM and that surface lowering has continued steadily throughout the Holocene (Sugden et al., 2006). Deglaciation in Marie Byrd Land therefore happened much later than the retreat of the Ross Sea grounding line and may still be in progress (Stone et al., 2003). Large scale troughs have been identified offshore from all the major present day outlet glaciers of this sector of West Antarctica, extending across the shelf of the Amundsen Sea (Anderson and Shipp, 2001) and the Bellingshausen Sea (O’Cofaigh et al., 2005b). Swath bathymetry has been used to determine that the grounded ice-sheet reached the shelf edge in outer Pine Island Bay (Lowe and Anderson, 2002; Evans et al., 2006, Fig. 12.6) where elongate subglacial bedforms in the soft sediments deposited at the base of a palaeo-ice stream end in gullies incised seaward of the shelf break (Dowdeswell et al., 2006). Likewise the Belgica Trough, which extends the full width of the continental shelf in the Bellingshausen Sea, exhibits geomorphological evidence of an ice stream which is thought to have drained much of Ellsworth Land as well as some of the southern part of the the Antarctic Peninsula Ice Sheet at the LGM (O’Cofaigh et al., 2005b). AMS dating of foraminifera near the Getz Ice Shelf (Marie Byrd Land) indicates that an ice stream in this trough retreated across a mid-shelf position by B13.4 cal ka BP (Anderson et al., 2002). The Pine Island Bay ice stream retreated from the mid-outer shelf after 17.574 cal ka BP, and had reached a position near the present day grounding line by 10.270.5 cal ka BP (Lowe and Anderson, 2002). Retreat from the shelf edge in the Bellingshausen Sea probably occurred at around the same time with a mid-shelf position reached by B14 cal ka BP at the latest (Pope and Anderson, 1992; Pudsey et al., 1994; Heroy and Anderson, 2005). The ice edge had retreated to the inner shelf at Palmer Deep by 13 cal ka BP (Domack et al., 2001), and there is no evidence for the presence of grounded ice on the mid-outer continental shelf after this time.
12.3.3.6. 85–601W: Antarctic Peninsula The late Pleistocene extent of grounded ice in the Antarctic Peninsula has received a significant amount of attention during the past decade, partly due to ice-sheet models that predict a significant increase in ice volume during this period (e.g., Huybrechts, 1990). Geomorphic evidence including glacial erratics and striations on nunatak summits along the length of the peninsula indicate that the ice sheet in this region thickened significantly during the LGM. The centre of the ice-sheet thickened by up to 500 m in places,
546
A. P. Wright et al.
Figure 12.6: Geomorphological evidence for the extent and configuration of the West Antarctic Ice Sheet in Pine Island Bay during the last glaciation (Evans et al., 2006). Also shown are the locations of postglacial iceberg scours (Lowe and Anderson, 2002; Evans et al., 2006). Reproduced by permission of Elsevier. reaching a maximum height of 2350 m asl at Mt Jackson (721S), and there is evidence for at least two distinct ice domes along the spine of the southern part of the peninsula (Bentley et al., 2000, 2006; Ingo´lfsson, 2004). The striation orientations in the south-western part of the peninsula also suggest
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
547
that ice flow was deflected by an expanded West Antarctic Ice Sheet in the Weddell Sea, indicating that ice advance in these two regions was synchronous (Bentley et al., 2006). Geophysical investigations of the continental shelf around the Antarctic Peninsula using a combination of multibeam swath bathymetry, side-scan sonar and seismic surveys have revealed a comprehensive set of subglacial landforms. The sites of former ice streams that drained the expanded Antarctic Peneinsula Ice Sheet at the LGM have been located at Marguerite Bay (O’Cofaigh et al., 2002, 2005a; Dowdeswell et al., 2004), in the western Bransfield Basin to the north-west of the Peninsula (Canals et al., 2000, 2002; Evans et al., 2004) and the Robertson Trough on the eastern side of the Antarctic Peninsula (Evans et al., 2005) amongst others. In each case there is clear evidence that grounded ice streams extended all the way to the continental shelf break at the LGM. Post LGM retreat of the grounding line from the shelf edge position in the Robertson Trough region on the eastern side of the Antarctic Peninsula left no grounding zone wedges to indicate stillstands and is therefore interpreted to have been continuous and perhaps rapid (Evans et al., 2005). Slightly further north however, around the northern margin of the Larsen Ice Shelf A/southern part of the Prince Gustav Channel, ice retreat appears to have been much more gradual. The transition from subglacial to glacimarine deposition here has been shown by AMS radiocarbon dating to have occurred before 12 14C ka BP (Evans et al., 2005). A somewhat more complicated picture has emerged of the retreat history in the Marguerite Bay area but final retreat across the continental shelf is known to have been underway by 13 14C ka BP (Pope and Anderson, 1992; O’Cofaigh et al., 2005a). Terrestrial and marine records are in broad agreement that the various presently ice-free coastal and bay regions of the Antarctic Peninsula were mostly deglaciated between 8 and 6 ka BP (Ingolfsson, 2004). The base of the glaciomarine sedimentary strata has been AMS radiocarbon dated to B8 ka BP in both the Gerlache Strait (Harden et al., 1992) and in Lallemand Fjord (Shevenell, et al., 1996). The terrestrial record includes fossil molluscs from raised marine deposits on King George Island that indicate deglaciation by around 9–8 ka BP (Ma¨usbacher, 1991) and glaciomarine and sub-littoral deposits overlying till on James Ross Island which began to be deposited around 7.4 ka BP (Hjort et al., 1997). The timing of ice-sheet lowering inland is less well constrained, but has been attempted using cosmogenic exposure dating. At George VI Sound, at Moutonne´e Valley (on Alexander Island) and in the Batterby Mountains on the western margin of the peninsula, ice downwasting may have begun as early as 25 ka BP and been complete by 15–10 ka BP (Bentley et al., 2006).
548
A. P. Wright et al.
12.3.3.7. 60–01W: Weddell Sea Embayment/Filchner-Ronne Ice Shelf/Coats Land Despite the potential of the Weddell Sea sector to have contributed a similar magnitude of post-LGM sea-level rise to that of the Ross Sea embayment, neither the extent of the grounding line or the change in thickness of the inland ice is yet well constrained (see Fig. 12.7). Glacial trimlines more than 1000 m above the present ice surface in the Ellsworth Mountains (Denton et al., 1992) are probably much older than the LGM (Sugden et al., 2006;
Figure 12.7: Ice-sheet reconstruction based on field evidence from the Weddell Sea-Antarctic Peninsula region. Reproduced from Bentley et al. (2006) with permission from the GSA (see also Sugden et al., 2006).
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
549
Fogwill et al., 2007). Cosmogenic exposure-age dating of surface bedrock in the Shackleton Range has also been used to constrain the LGM thickening of the ice near the present day at grounding line of the Filchner Ice Shelf to a maximum of 750 m and a more probable limit of 340 m, consistent with limited LGM thickening (Fogwill et al., 2004). Marine geological evidence for the location of the LGM grounding line in the Weddell Sea is perhaps even more sparse than terrestrial data of ice-sheet thickness. Glacial diamict sampled from the bed of Crary Trough, which underlies the Filchner-Ronne Ice Shelf and extends out towards the edge of the continental shelf east of Berkner Island, has a composition indicative of a source in West Antarctica (Anderson et al., 1991). Radiocarbon dating, however, has so far been unable to confirm the age of these deposits. 12.3.4. Holocene Glacier and Climate Fluctuations Evidence for smaller ice volumes/ice extents during the mid-late Holocene relative to the present day is available at a number of sites. Sediments from the bed of epishelf Moutone´e Lake indicate the free movement of icebergs, and therefore the collapse of the George VI Ice Shelf between Alexander Island and the Antarctic Peninsula, between 9.6 and 8 cal. ka B.P (Bentley et al., 2005). Likewise the analysis of sediment cores recovered from the seabed of the Prince Gustav Channel on the eastern side of the peninsula, has determined that between 5 and 2 ka BP clasts were sourced from a range of distal locations, indicating open marine conditions and the collapse of the Prince Gustav Channel Ice Shelf (Pudsey and Evans, 2001). Neither is the mid-Holocene minimum confined to the Antarctic Peninsula. Goodwin (1996) provides evidence from a number of sources that Law Dome was at least 3–4 km smaller before B4 ka BP, while Baroni and Orombelli (1994) show that ice shelves in Terra Nova Bay were reduced by a similar amount between 7.5 and 5 ka BP. On the contrary, there is geological evidence to suggest that the Larsen Ice Shelf B had been in place for W11 ka before its collapse in 2002 (Domack et al., 2005). In East Antarctica, Verleyen et al. (2005) provide relative sea level (RSL) based evidence for increased ice load near the Larsemann Hills between B7 and 2.5 ka BP. Given the similarity of the RSL curve to that obtained from the Vestfold Hills (Zwartz et al., 1998) this may have been more than a local feature. This fluctuation may correlate with the glacial advances identified through striae patterns and weathering in the Vestfold Hills (Adamson and Pickard, 1986) and in the Rauer Group (White et al., submitted).
550
A. P. Wright et al.
12.3.5. Discussion: Pattern and Timing of Ice Retreat The geological record provides convincing evidence that the Antarctic Ice Sheet was more extensive at the LGM and has subsequently reduced in area and thinned near the margins. The magnitude of ice retreat varies around the continent, ranging from relatively little marginal change at some terrestrial sites such as the Larsemann Hills, to retreat of the grounding line by hundreds of km in the large embayments currently occupied by the Ross, Filchner-Ronne and Amery ice shelves. At most of the marine sites studied, the evidence implies that the LGM maximum ice extent was relatively shortlived, while some of the terrestrial sites (e.g., Bunger Hills, Lutzow-Holm Bay) suggest that the LGM was not necessarily the period of greatest ice extent during the last glacial cycle, and that ice-sheet margins were more advanced prior to 35 ka BP. Similarly, evidence for the timing of ice retreat following the LGM supports different retreat timings at different sites (Anderson, 1999). Some areas, such as the Antarctic Peninsula and the Bellingshausen Sea appear to have responded relatively rapidly to global climate and sea level changes following the LGM. At other locations, including the Ross Sea and the East Antarctic margin at Framnes Mountains, ice retreat appears to have begun relatively late and, in some places such as Marie Byrd Land, appears to have continued to the present day. While the evidence for this differential retreat timing cannot yet be considered definitive, the pattern may point to the differing sensitivities of each region to external forcing.
12.4. Numerical Modelling Reconstructions 12.4.1. Ice-Sheet Modelling Recent interest in the response of the Antarctic Ice Sheet to changing climates has been an important factor in driving the development of improved glaciological ice-sheet models. As well as three-dimensional ice flow by internal deformation, these models now incorporate a temperature calculation throughout the ice sheet. Thermo-mechanical coupling allows a more realistic representation of ice flow with the additional benefit that basal sliding can be parameterised over areas of the bed that reach the pressure melting point. Ice-sheet models require inputs of bedrock elevation, initial ice thickness, sea level and climate. They can be used to provide quantitative information about ice flow and the response of an ice sheet to changes in the
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
551
model parameter values or to a change in external forcing. Models are best used along with field data – both present day ice-sheet extent and past extents inferred from geological evidence. The ability of models to make predictions in areas of sparse data and for time periods which cannot be constrained by geological evidence means they are considered a valuable tool in reconstructing past ice-sheet configurations. The first comprehensive glaciological modelling of the Antarctic Ice Sheet over a full glacial-interglacial cycle was published by Huybrechts (1990). This work has now been updated with several improvements to both the model and the climatic input (Huybrechts, 2002). An independent study with an alternative model formulation by Ritz et al. (2001) provides a useful comparison for the case of reconstructing the extent and thickness of Antarctica at the LGM. This model has also recently been forced with output from a Global Climate Model providing a further comparison (Philippon et al., 2006). Both Huybrechts (2002) and Ritz et al. (2001) use time-dependent climatic input to force their models. In both studies a reference simulation is produced using paleoclimatic information derived from the Vostok ice core (Petit et al., 1999) to run the model through four glacial-interglacial cycles (B400 ka). Free parameters in each model are then tuned to match the results to observations of the present day ice-sheet. The ice-sheet produced at the end of the reference run is then used as the initial condition for a range of further simulations to assess the model sensitivity to changes in parameter values and forcings. 12.4.2. Reconstructions of the Present Day/Interglacial Ice Sheet Both model representations of the present day are similar, with both showing some differences from the observed geometry, mainly around the margins. Both models over-predict the extent of grounded ice in the Amundsen Sea and the Amery Ice Shelf regions. Ritz et al. (2001) also over-estimate the extent of grounded ice in the Weddell Sea, while both models underestimate the extent of grounded ice in the Ross Sea (Fig. 12.8A, B). Ritz et al. (2001) ascribe the discrepancies between their model and present day observations to the coarse (40 km) grid resolution of their model, as this hinders their ability to reproduce fast flowing ice streams and flow retarding bedrock features. The result is an ice-sheet which covers a similar area but has a 12% larger volume. Overall the present day ice sheet produced by Huybrechts (2002) is a better match to the observations and this is probably due to the higher resolution (20 km) grid cells of that model.
552
A. P. Wright et al.
Figure 12.8: Numerical model reconstructions of the present day ice sheet (A, B) and of the Last Glacial Maximum (C, D) by Huybrechts (2002) and Ritz et al. (2001), respectively. Reproduced with permission from Elsevier and the AGU. 12.4.3. Time Dependent Simulations Both Huybrechts and Ritz run their models over 400,000 years encompassing the last four glacial-interglacial cycles. The forcing in each case consists of time-series of air temperature change and eustatic sea-level variation. Both studies use temperature records derived from the dD (deuterium) signal of the Vostok ice core. In each case accumulation is based on the present day
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
553
pattern with the magnitude of variation calculated as a function of air temperature by the method described in Lorius et al. (1985). Huybrechts uses the SPECMAP reconstruction of sea-level change while Ritz et al. prefer that of Bassinot et al. (1994); the two records have some differences but sensitivity studies by Ritz et al. suggest that this does not affect their model results. Overall the two reconstructions agree on several important aspects of Quaternary Antarctic evolution: The present day ice sheet is slightly larger than the minimum reached during the last interglacial. Ritz suggests that this is because the present ice sheet is still in decline, while Huybrechts determines that the ice-sheet retreat from the Weddell Sea was more advanced during the last interglacial. General responses that dominate during a glacial period are grounding line advance and interior ice growth in West Antarctica with surface lowering and little margin advance in East Antarctica. Eustatic sea-level variation is the forcing factor driving change in the position of the ice sheet grounding line while changes in temperature (and consequently rate of accumulation) have the largest effect on surface altitude in the interior of East Antarctica. Grounding line advance occurs more readily in the Weddell Sea embayment than in the Ross Sea. Both models cite the differences in bed topography – the depth of the Ross Sea increases more rapidly with distance from the present coastline – as the cause of this asymmetry. The patterns of ice flow and positions of the ice domes in East Antarctica are not significantly changed from their present configuration at the LGM. Ice flow in West Antarctica would have been very different due to the advanced margins and thicker inland ice. There are several points at which results from the two models diverge. Huybrechts (2002) puts the timing of the LGM in Antarctica at 15 ka BP when the ice sheets would have contained 19.2 m sea level equivalent greater mass than at present, equivalent to an increase of 5 m sea level equivalent on the sea level contribution predicted with the same model in an earlier study (Huybrechts, 1991). Ritz et al. (2001) place the LGM at around 18 ka BP and find a difference in mass of just 5.9 m sea level equivalent from today. The Ritz model generally has a lower surface slope around the margins due to its inclusion of a grounded ‘dragging shelf ’ transition zone between inland and ice shelf flow intended to represent ice streaming. This factor is probably responsible for the majority of the differences between the two models, with the Ritz model placing thinner ice over much of West Antarctica at the
554
A. P. Wright et al.
LGM, especially in the Ross Sea and also notably in the present day area of Prydz Bay in East Antarctica (c.f. Fig. 12.8C and D). Interestingly, the reconstruction of Ritz et al. (2001) fails to predict any significant expansion of ice covering the Antarctic Peninsula except in the very southern part where it joins the West Antarctic Ice Sheet (Fig. 12.8D). This is in contrast to the results consistently obtained with Huybrechts model, which in agreement with geological evidence outlined earlier predicts an expansion of the Antarctic Peninsula Ice Sheet to the edge of the continental shelf under LGM conditions (Fig. 12.8C). 12.4.4. Ice-Sheet Sensitivity Both Huybrechts and Ritz analysed the sensitivity of their respective models to variations in climatic forcing and to uncertainty of the parameter values by repeating the time dependent simulations described above and comparing snapshots of the ice-sheet at fixed time-slices. These experiments dealt with aspects of the ice-sheet dynamics including bedrock adjustment, thermodynamic coupling, basal sliding and the response of ice shelves to ocean warming as well as to variations in air temperature and accumulation forcing. The elements of sensitivity common to both models include: The ice-sheet volume is relatively sensitive to changes in LGM snow accumulation. Ritz et al. found that reducing glacial accumulation values to 40% of present day (c.f. 50% for the standard simulation) decreased the area of the LGM ice-sheet by 18% and the volume by 43%. Grounding line advance during glaciations is sensitive to the way ice shelves respond to temperature change. Not allowing the ice temperature (which controls viscosity) to cool and hence stiffen the ice or preventing the basal melt rate from decreasing significantly reduces the advancement of the grounding line during the LGM especially in West Antarctica. Sea level lowering appears to be necessary for grounding line advance but so does a decrease in surface temperature. The rate of crustal response appears to have an intricate relationship with the amount of grounding line retreat occurring at the glacial-interglacial transition. A more rapid isostatic response leads to earlier re-grounding of a retreating ice shelf in either the Ross or Weddell Sea embayments thereby reducing the total retreat. Finally, a lower value for the geothermal heat flux reduces the temperature of basal ice and limits the area of the base reaching the pressure melting point. The colder ice flows more slowly and the result is to increase ice-sheet thicknesses at the LGM.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
555
12.4.5. Recent Modelling of Antarctica Recent studies using a variety of dating techniques have indicated that the former extent of the East Antarctic Ice Sheet at the LGM may have been less than previously thought (e.g., Gore et al., 2001; Hodgson et al., 2001; Mackintosh et al., 2007). Past modelling studies however, including both those by Ritz and Huybrechts described above, have placed the margin of the East Antarctic Ice Sheet at the edge of the continental shelf in agreement with the CLIMAP reconstruction (CLIMAP, 1976). With this in mind some new numerical modelling experiments have been undertaken in order to investigate LGM ice sheet configurations compatible with the newly available geological evidence from East Antarctica. The GLIMMER community ice-sheet model is an implementation of the shallow-ice approximation with full thermodynamic coupling and the inclusion of basal sliding (Payne, 1999). Using this model an ice sheet was created on a slightly modified and isostatically rebounded BEDMAP topography (Lythe et al., 2001) with 20 km grid resolution. Model parameter values were tuned to reproduce as closely as possible the present day extent of the Lambert Graben region of the East Antarctic Ice Sheet (Fig. 12.9A). The accumulation rate was calculated for each grid cell as a function of continentality (proximity to the coastline) following the method of Oerlemans (1982); surface air temperature was calculated using a fixed lapse rate and mass loss at the marine margins as a prescribed calving percentage. Table 12.1 shows the parameter values supplied to the model for the simulations of both the present day and LGM ice sheets. For the LGM these values are taken from the general consensus of the paleoclimate literature (e.g., Petit et al., 1999; Siegert, 2003; Van Ommen et al., 2004). The extent and thickness of the present day ice sheet produced by the model is a good match with observations in the Lambert Glacier region of East Antarctica (c.f. Fig. 12.9A with Fig. 12.3). The shape of the ice-sheet around the Lambert Graben and the surface height of the interior ice-sheet are reasonably well reproduced; the model used here is not capable of reproducing ice shelves and hence the Amery Ice Shelf does not appear in the results. A characteristic of this reconstruction of the LGM ice sheet (Fig. 12.9B) is the uneven nature of the glacial advance around the margins. The location of the grounding line at the LGM is predicted by the model to have advanced B80 km towards the middle of the continental shelf on the western flank of the Lambert Graben, by only B40 km on the eastern flank and for no significant advance on the present day margin along much of the rest of the coastline (Fig. 12.9B, C). This pattern is matched by negligible changes in
556
A. P. Wright et al.
Figure 12.9: Numerical model results for the Princess Elizabeth – Lambert Graben – Mac.Robertson Land area of the East Antarctic Ice Sheet. (A) Reference experiment created with parameter values tuned for the present day. (B) Ice sheet reconstruction with best-estimate parameter values for the LGM. (C) Predicted ice surface height change (B-A) at the LGM. The locations of the two longitudinal profiles in Fig. 12.10 are marked A–B and C–D.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
557
Table 12.1: Values of model parameters used for the simulation of both the present day and LGM East Antarctic Ice Sheet.
Sea level air temperature (1C) Air temperature lapse rate (1C m 1) Relative precipitation Eustatic sea level (m) Calving fraction at marine margins Basal traction factor Geothermal heat flux (W m 2)
Present day (0 ka BP)
LGM (B21 ka BP)
15 0.008 100% 0.0 90% 1.0 10 4 56 10 3
25 0.008 50% 120.0 90% 1.0 10 4 56 10 3
ice-sheet thickness along parts of the coast which are coincident with significant thickening in other areas (Fig. 12.9C). The patterns of advance and thickening seen here are strongly influenced by the bathymetry of the continental shelf. Where the shelf is shallow the 120 m sea-level drop at the LGM allows expansion of the ice sheet. Under this scenario the extent of the ice-sheet is controlled largely by eustatic sea-level and hence by the volume of northern hemisphere ice sheets (e.g., Denton et al., 1986). In contrast the interior of the ice-sheet experiences a general decrease in thickness of between 100 and 150 m under the LGM conditions prescribed here (Fig. 12.9C). This is a result of the decrease in accumulation which is not mitigated by the decrease in air temperature as surface melting is relatively unimportant here under either present day or interglacial conditions. The locations of the two profiles shown in Fig. 12.10 were chosen to allow direct comparisons to be made with the LGM reconstructions of the icesheet surface from cosmogenic dating of glacially abraded surfaces and transported clasts (Fig. 12.3 and Mackintosh et al., 2007). The AB-profile shown in Fig. 12.10A follows the route of the Fisher Glacier to its confluence with the Lambert Glacier and where the continuation of the latter discharges into the Amery Ice Shelf. Though the model is not capable of reproducing ice shelves, it can be seen (by comparing Figs. 12.3B and 12.10A) that grounded ice in the model extends B100 km past the position of the grounding line inferred from GPS measurements of tidal motion (Fricker et al., 2002). This over-advance under present day conditions occurs possibly because the depth of the topographic low, near the deepest point of which is located the present grounding zone, is underestimated in BEDMAP. The present day model terminus position of the glacier in the Lambert Graben is in fact very similar to the location of the LGM transition zone between steady icesheet flow and ice streaming inferred from geological evidence (Fig. 12.3B).
558
A. P. Wright et al.
Figure 12.10: Longitudinal profiles through the model reconstructed ice sheet for Fisher-Lambert Glacier and Framnes Mountains. These profiles demonstrate the main features of the LGM reconstruction of East Antarctica i.e. surface lowering in the interior with little or no advance of the margins. It is therefore perhaps unsurprising that no further terminus advance occurs in this area when the model is run under LGM climate/sea-level conditions (Fig. 12.10A). The LGM advance of the East Antarctic Ice Sheet margin along the CD profile (Fig. 12.9) has been shown by Mackintosh et al. (2007) to have been only around 5–10 km.This is beneath the grid resolution of the present model and hence the unchanged margin position along the CD-profile under LGM conditions (Fig. 12.10B) is in agreement with the field evidence. The model also predicts a slight LGM lowering of the ice-sheet surface along the entire length of this profile as a result of decreased accumulation, counter to the geological evidence which suggests a moderate increase in thickness near the margin (Mackintosh et al., 2007). The spatial extent of this thickening, however, is uncertain and may be too small to be resolved by the current model.
12.4.6. Isostatic Modelling of Post-Glacial Uplift and Contribution to Global Sea Level The Earth’s crust is displaced by the loading of ice sheets, which recovers when these loads are taken away. In Antarctica, the pattern of recovery
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
559
following removal of an ice load has been reconstructed by investigating and dating records of relative sea level, such as of emergent shorelines (e.g., Adamson and Pickard, 1986; Baroni and Orombelli, 1994), the transition from marine to freshwater sediment deposition in enclosed basins (Goodwin, 1993; Zwartz et al., 1998; Verleyen et al, 2005) and by determining the present day rate of crustal motion by measuring the change in height of bedrock surfaces over time using a Global Positioning System (GPS; e.g., Tregoning et al., 2000). These uplift data have been compared with the response predicted by numerical Earth models to constrain the spatial and temporal pattern of former ice loads. However, some of the parameters (such as mantle viscosity) used in the models are not independently known, and it is often possible to model the measured isostatic response at any one site with a range of former ice loads (e.g., Bassett et al., 2007). Thus, care must be taken when interpreting these results in regions like Antarctica where relative sea level reconstructions are few and far between. Despite these issues, the technique is useful in constraining former ice heights where evidence is ambiguous, or difficult to reconstruct by other means. This has resulted in numerous reconstructions of former ice heights across the continent (Nakada and Lambeck 1988; Tushingham and Peltier, 1991; Colhoun et al., 1992; Peltier, 1994), which provide a very large range of post-LGM sea level contributions from Antarctica, of between 2 and 37 m. The more recent models, which incorporate a greater degree of constraints derived from the geological studies detailed above, provide estimates of 17.3 m (ICE-5G model of Peltier, 2004) and 6.6–16.7 m (ANT5 and ANT6 of Nakada et al., 2000). These models have also been used to provide estimates of former ice loads at individual sites around the continent. Zwartz et al. (1998) used relative sea-level data from the Vestfold Hills to estimate that the ice sheet retreated 30–40 km, and thinned 600–700 m in this region since the LGM. Goodwin and Zweck (2000) used uplift information from raised beaches and the marine-lacustrine sediment transition in basins in the Windmill Islands and Budd Coast as input to an isostatic model of the Law Dome ice cap in East Antarctica. Their numerical Earth model predicts that the area near the present day margin of Law Dome formerly hosted an ice load of between 770 and 1000 m, and that ice expanded out 40 to 65 km across the continental shelf (Goodwin and Zweck, 2000). Goodwin and Zweck (2000) also used the similarity in the height of the marine limit along the coast between Wilkes and Oates land to postulate that the East Antarctic Ice Sheet in this area has retreated by a similar amount since the LGM.
560
A. P. Wright et al.
12.5. Summary There is persuasive marine geological evidence from the relatively well studied Ross Sea embayment, the Bellingshausen Sea, the Amundsen Sea and off the west coast of the Antarctic Peninsula for an LGM ice-sheet grounding line at, or very near to the edge of the continental shelf. A similar situation most likely existed on the eastern side of the peninsula and in the Weddell Sea embayment, though this region has been less well studied to date. LGM thickening of the West Antarctic Ice Sheet has been identified in several locations with the ice surface in the central regions likely to have been several hundred metres higher than today. Likewise ice in the Antarctic Peninsula can be shown to have been up to 500 m thicker at central sites, with the present day ice caps and glaciers coalescing to form the Antarctic Peninsula Ice Sheet at that time. The smaller number of studies undertaken on the East Antarctic Ice Sheet have found wide variations in the extent of LGM ice advance across the continental shelf. It appears that the grounding line in East Antarctica advanced to the continental shelf break in places, to a mid shelf position in others and for no significant advance to have occurred elsewhere. Furthermore there are several sites, including the Bunger Hills, where deglaciation appears to have begun at around 30 ka BP, well before the LGM. The timing and rate of deglaciation during the Holocene appears to have varied greatly across the continent. The Antarctic Peninsula Ice Sheet and the Bellingshausen Sea sector of the West Antarctic Ice Sheet may have been amongst the first regions to reach their present configurations while the Ross Sea sector along with many coastal areas in East Antarctica have responded less rapidly to climate and sea-level change since the LGM. At the end of the 20th century estimates of the reduction in the total volume of the Antarctic Ice Sheets since their maximum during the last glacial cycle ranged from 0.5 to 37 metres of equivalent sea level rise. Recent analysis of the geological evidence has made the upper estimates seem improbable with the consensus of opinion tending towards the upper end of the lower half of this range. Though there is as yet no single agreed value, recent estimates based on geological evidence (Bentley, 1999, 6.1–13.1 m; Denton and Hughes, 2002, 14 m) have converged with those based on numerical ice-sheet modelling (Ritz et al., 2001, 5.9 m; Huybrechts, 2002, 19.2 m; Philippon et al., 2006, 9.5–17.5 m) and those based on modelling of isostatic post-glacial uplift (Nakada et al., 2000, 6.6–16.7 m; Peltier, 2004, 17.3 m) to narrow the likely range to 5.9–19.2 m sea level equivalent.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
561
REFERENCES Ackert, R. P. Jr., Barclay, D. J., Borns, H. W. Jr., Calkin, P. E., Kurz, M. D., Fastook, J. L., & Steig, E. J. (1999). Measurements of Past Ice Sheet Elevations in Interior West Antarctica. Science, 286, 276–280. Adamson, D. A., Mabin, M. C. G., & Luly, J. G. (1997). Holocene Isostasy and Late Cenozoic Development of Landforms Including Beaver and Radok Lake Basins in the Amery Oasis, Prince Charles Mountains, Antarctica. Antarctic Science, 9(3), 299–306. Adamson, D. A., & Pickard, J. (1983). Late Quaternary Ice Movement Across the Vestfold Hills, East Antarctica. In: R. L. Oliver, P. R. James, & J. B. Jago (Eds). Antarctic Earth Science. Australian Academy of Science, Canberra, pp. 465–469. Adamson, D. A., & Pickard, J. (1986). Physiography and Geomorphology of the Vestfold Hills, Antarctica. In: J. Pickard (Ed.). Antarctic Oasis. Academic Press, Sydney, Australia, pp. 99–139. Anandakrishnan, S., Blankenship, D. D., Alley, R. B., & Stoffa, P. L. (1998). Influence of Subglacial Geology on the Position of a West Antarctic Ice Stream from Seismic Observations. Nature, 394(6688), 62–65. Anderson, J. B. (1999). Antarctic Marine Geology. Cambridge University Press, 289 pp. Anderson, J. B., Andrews, B. A., Bartek, L. R., & Truswell, E. M. (1991). Petrology and Palynology of Weddell Sea Glacial Sediments: Implications for Subglacial Geology. In: M. R. A. Thomson, J. A. Crame, & J. W. Thomson (Eds). Geological Evolution of Antarctica. Cambridge University Press, Cambridge, pp. 231–235. Anderson, J. B., & Shipp, S. S. (2001). Evolution of the West Antarctic Ice Sheet. In: R. B. Alley, & B. R. Binschadler (Eds). The West Antarctic Ice Sheet: Behaviour and Environments. American Geophysical Union, Washington, D.C., Vol. 77, pp. 45–57. Anderson, J. B., Shipp, S. S., Lowe, A. L., Wellner, J. S., & Mosola, A. B. (2002). The Antarctic Ice Sheet During the Last Glacial Maximum and Its Subsequent Retreat History: A Review. Quat. Sci. Rev., 21, 49–70. Andrews, J. T., Domack, E. W., Cunningham, W. L., Leventer, A., Licht, K. J., Jull, A. J. T., DeMaster, D. J., & Jennings, A. E. (1999). Problems and Possible Solutions Concerning Radiocarbon Dating of Surface Marine Sediments, Ross Sea, Antarctica. Quat. Res., 52(2), 206–216. Armand, L. K. (2000). An Ocean of Ice – Advances in the Estimation of Past Sea Ice in the Southern Ocean. GSA Today, 10, 1–7. Bard, E., Hamelin, B., & Fairbanks, R. G. (1990). U-Th Ages Obtained by Mass Spectrometry in Corals from Barbados: Sea Level During the Past 130,000 Years. Nature, 346, 456–458. Baroni, C., & Orombelli, G. (1994). Holocene Glacier Variations in the Terra Nova Bay Area (Victoria Land, Antarctica). Antarctic Sci., 6(4), 497–505.
562
A. P. Wright et al.
Bassett, S. E., Milne, G. A., Bentley, M. J., & Huybrechts, P. (2007). Modelling Antarctic Sea-Level Data to Explore the Possibility of a Dominant Antarctic Contribution to Meltwater Pulse IA. Quat. Sci. Rev., 26, 2113–2127. Bassinot, F. C., Labeyrie, L. D., Vincent, E., Quidelleur, X., Shackleton, N. J., & Lancelot, Y. (1994). The Astronomical Theory of Climate and the Age of the Brunhes-Matuyama Magnetic Reversal. Earth and Planetary Sci. Lett., 126, 91–108. Bentley, M. J. (1999). Volume of Antarctic Ice at the Last Glacial Maximum, and Its Impact on Global Sea Level Change. Quat. Sci. Rev., 18, 1569–1595. Bentley, M. J., Fogwill, C. J., Kubik, P. W., & Sugden, D. E. (2006). Geomorphological Evidence and Cosmogenic 10Be/26Al Exposure Ages for the Last Glacial Maximum and Deglaciation of the Antarctic Peninsula Ice Sheet. Geological Soc. America Bull., 118(9–10), 1149–1159. Bentley, M. J., Hodgson, D. A., Sugden, D. E., Roberts, S. J., Smith, J. A., Leng, M. J., & Bryant, C. (2005). Early Holocene Retreat of the George VI Ice Shelf, Antarctic Peninsula. Geology, 33(3), 173–176. Bentley, M. J., Kubik, P. W., Sugden and D. E. (2000). Geomorphological Evidence and Cosmogenic 10Be/26Al Exposure Ages for the Last Glacial Maximum Configuration of the Antarctic Peninsula Ice Sheet. In: C. Hjort (Ed.). Glaciation of the Weddell Sea Basin. ANTIME workshop, Abisko, Sweden. Berkman, P., & Forman, S. L. J. (1996). Pre-Bomb Radiocarbon and the Reservoir Correction for Calcareous Marine Species in the Southern Ocean. Geophys. Res. Lett., 23(4), 363–366. Bingham, R., & Siegert, M. J. (2007). Radar-Derived Bed Roughness Characterization of Institute and Mo¨ller Ice Streams, West Antarctica, and Comparison with Siple Coast Ice Streams. Geophys. Res. Lett., 34(L21504), doi:10.1029/ 2007GL031483. Bockheim, J. G., Wilson, S. C., Denton, G. H., Andersen, B. J., & Stuvier, M. (1989). Late Quaternary Ice-Surface Fluctuations of the Hatherton Glacier, Transantarctic Mountains. Quat. Res., 31, 229–254. Broecker, W. S., & Denton, G. H. (1990). The Role of Ocean-Atmosphere Reorganizations in Glacial Cycles. Quat. Sci. Rev., 9(4), 305–341. Brook, E. J., Kurz, M. D., Ackert, R. P., Raisbeck, G. M., & Yiou, F. (1995). Cosmogenic Nuclide Exposure Ages and Glacial History of Late Quaternary Ross Sea Drift in McMurdo Sound, Antarctica. Earth and Planetary Sci. Lett., 131(1–2), 41–56. Budd, W. F., & Smith, I. N. (1982). Large Scale Numerical Modelling of the Antarctic Ice Sheet. Ann. Glaciology, 3, 42–49. Burgess, J. S., Spate, A. P., & Shevlin, J. (1994). The Onset of Deglaciation in the Larsemann Hills, Eastern Antarctica. Antarctic Sci., 6, 491–495. Canals, M., Casamor, J. L., Urgeles, R., Calafat, A. M., Domack, E. W., Baraza, J., Farran, M., & De Batist, M. (2002). Seafloor Evidence of a Subglacial Sedimentary System Off the Northern Antarctic Peninsula. Geology, 30(7), 603–606.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
563
Canals, M., Urgeles, R., & Calafat, A. M. (2000). Deep Sea-Floor Evidence of Past Ice Streams Off the Antarctic Peninsula. Geology, 28(1), 31–34. CLIMAP (1976). The surface of the Ice Age Earth. Science, 191, 1131–1136. Colhoun, E. A. and Adamson, D. A. (1992). Raised Beaches of the Bunger Hills. Australian National Antarctic Research Expeditions Report, 136, 47 pp. Colhoun, E. A., Mabin, M. C. G., Adamson, D. A., & Kirk, R. M. (1992). Antarctic Ice Volume and Contribution to Sea Level Fall at 20 000 yr BP from Raised Beaches. Nature, 358, 316–319. Conway, H., Hall, B. L., Denton, G. H., Gades, A. M., & Waddington, E. D. (1999). Past and Future Grounding-Line Retreat of the West Antarctic Ice Sheet. Science, 286, 280–283. Cuffey, K. M., & Marshall, S. J. (2000). Substatial Contribution to Sea-Level Rise during the Last Interglacial from the Greenland Ice Sheet. Nature, 404, 591–594. Denton, G. H., Bockheim, J. G., Rutford, R. H., & Andersen, B. G. (1992). Glacial History of the Ellsworth Mountains, West Antarctica. In: G. F. Webers, C. Craddock, & J. F. Splettstoesser (Eds). Geology and Palaeontology of the Ellsworth Mountains, West Antarctica. Geological Society of America, Boulder, CO. Denton, G. H., & Hughes, T. P. (2002). Reconstructing the Antarctic Ice Sheet at the Last Glacial Maximum. Quat. Sci. Rev., 21, 193–202. Denton, G. H., Hughes, T. P., & Karlen, W. (1986). Global Ice-Sheet System Interlocked by Sea Level. Quat. Res., 26, 3–26. Domack, E. W., Duran, D., Leventer, A., Ishman, S., Doane, S., McCallum, S., Amblas, D., Ring, J., Gilbert, R., & Prentice, M. (2005). Stability of the Larsen-B Ice Shelf on the Antarctic Peninsula during the Holocene Epoch. Nature, 436(7051), 681–685. Domack, E. W., Jacobsen, E. A., Shipp, S. S., & Anderson, J. B. (1999). Late Pleistocene-Holocene Retreat of the West Antarctic Ice Sheet System in the Ross Sea: Part 2-Sedimentologic and Stratographic Signature. Geological Soc. America Bull., 111, 1517–1536. Domack, E. W., Leventer, A., Dunbar, R., Taylor, F., Brachfeld, S., Sjunneskog, C., & ODP Leg 178 Scientific Party. (2001). Chronology of the Palmer Deep Site, Antarctic Peninsula: A Holocene Palaeoenvironmental Reference for the CircumAntarctic. The Holocene, 11(1), 1–9. Domack, E. W., O’Brien, P., Harris, P. T., Taylor, F., Quilty, P. G., De Santis, L., & Raker, B. (1998). Late Quaternary Sediment Facies in Prydz Bay, East Antarctica and Their Relationship to Glacial Advance onto the Continental Shelf. Antarctic Sci., 10(3), 236–246. Dowdeswell, J. A., Evans, J., O’Cofaigh, C., & Anderson, J. B. (2006). Morphology and Sedimentary Processes on the Continental Slope Off Pine Island Bay, Amundsen Sea, West Antarctica. Geological Soc. America Bull., 118(5/6), 606–619. Dowdeswell, J. A., O’Cofaigh, C., & Pudsey, C. J. (2004). Thickness and Extent of the Subglacial till Layer Beneath An Antarctic Paleo-Ice Stream. Geology, 32, 13–16.
564
A. P. Wright et al.
Evans, J., Dowdeswell, J. A., & O’Cofaigh, C. (2004). Late Quaternary Submarine Bedforms and Ice-Sheet Flow in Gerlache Strait and on the Adjacent Continental Shelf, Antarctic Peninsula. J. Quat. Sci., 19(4), 397–407. Evans, J., Dowdeswell, J. A., O’Cofaigh, C., Benham, T. J., & Anderson, J. B. (2006). Extent and Dynamics of the West Antarctic Ice Sheet on the Outer Continental Shelf of Pine Island Bay During the Last Glaciation. Marine Geology, 230, 53–72. Evans, J., Pudsey, C. J., O’Cofaigh, C., Morris, P., & Domack, E. W. (2005). Late Quaternary Glacial History, Flow Dynamics and Sedimentation Along the Eastern Margin of the Antarctic Peninsula Ice Sheet. Quat. Sci. Rev., 24(5–6), 741–774. Fabel, D., Stone, J., Fifield, L. K., & Cresswell, R. G. (1997). Deglaciation of the Vestfold Hills, East Antarctica: Preliminary Evidence from Exposure Dating of Three Subglacial Erratics. In: C. A. Ricci (Ed.). The Antarctic Region: Geological Evolution and Processes. Terra Antarctica Publications, Siena, pp. 829–834. Fogwill, C. J., Bentley, M. J., Sugden, D. E., & Hubbard, A. G. (2007). Glacial history of the Ellsworth Mountains, Weddell Sea embayment, West Antarctica. Geophys. Res. Abstracts, 9(08271). Fogwill, C. J., Bentley, M. J., Sugden, D. E., Kerr, A. R., & Kubik, P. W. (2004). Cosmogenic Nuclides 10Be and 26Al Imply Limited Antarctic Ice Sheet Thickening and Low Erosion in the Shackleton Range for W1 m.y. Geology, 32(3), 265–268. Fricker, H. A., Allison, I., Craven, M., Hyland, G., Ruddell, A., Young, N., Coleman, R., King, M., Krebs, K., & Popov, S. (2002). Redefinition of the Amery Ice Shelf, East Antarctica, Grounding Zone. J. Geophys. Res., 107(B5), doi:10.1029/2001JB000383. Goodwin, I. D. (1993). Holocene Deglaciation, Sea-Level Change, and the Emergence of the Windmill Islands, Budd Coast, Antarctica. Quat. Res., 40(1), 55–69. Goodwin, I. D. (1996). A Mid to Late Holocene Readvance of the Law Dome Ice Margin, Budd Coast, East Antarctica. Antarctic Sci., 8(4), 395–406. Goodwin, I. D., & Zweck, C. (2000). Glacio-Isostasy and Glacial Ice Load at Law Dome, Wilkes Land, East Antarctica. Quat. Res., 53, 285–293. Gore, D. B. (1997a). Last Glaciation of Vestfold Hills: Extension of the East Antarctic Ice Sheet or Lateral Expansion of Sørsdal Glacier? Polar Record, 33(184), 5–12. Gore, D. B. (1997b). Blanketing Snow and Ice; Constraints on Radiocarbon Dating Deglaciation in East Antarctic Oases. Antarctic Sci., 9(3), 336–348. Gore, D. B., Rhodes, E. J., Augustinus, P. C., Leishman, M. R., Colhoun, E. A., & Rees-Jones, J. (2001). Bunger Hills, East Antarctica: Ice Free at the Last Glacial Maximum. Geology, 29(12), 1103–1106. Gosse, J. C., & Phillips, F. M. (2001). Terrestrial in situ Cosmogenic Nuclides: Theory and Application. Quat. Sci. Rev., 20, 1475–1560. Hall, B. L., Denton, G. H., Overturf, B., & Hendy, C. H. (2002). Glacial Lake Victoria, A High-Level Antarctic Lake Inferred from Lacustrine Deposits in Victoria Valley. J. Quat. Sci., 17(7), 697–706.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
565
Harden, S. L., DeMaster, D. J., & Nittrouer, C. A. (1992). Developing Sediment Geochronologies for High-Latitude Continental Shelf Deposits: A Radiochemical Approach. Marine Geology, 103, 69–97. Harris, P. T., & O’Brien, P. (1998). Bottom Currents, Sedimentation and Ice-Sheet Retreat Facies Successions on the Mac.Robertson Shelf, East Antarctica. Marine Geology, 151, 47–72. Hedges, R. E. M., Pettit, P. B., Bronk-Ramsey, C., & van Klinken, G. J. (1996). Radiocarbon Dates from the Oxford AMS System: Archeometry Datelist 22. Archeometry, 38(2), 391–415. Hemer, M., & Harris, P. (2003). Sediment Core from Beneath the Amery Ice Shelf, East Antarctica, Suggests Mid-Holocene Ice-Shelf Retreat. Geology, 31(2), 127–130. Hendy, C. H., & Hall, B. L. (2006). The Radiocarbon Reservoir Effect in Proglacial Lakes: Examples from Antarctica. Earth and Planetary Sci. Lett., 241(3–4), 413–421. Heroy, D. C., & Anderson, J. B. (2005). Ice-Sheet Extent of the Antarctic Peninsula Region During the Last Glacial Maximum (LGM)-Insights from Glacial Geomorphology. Geological Soc. America Bull., 117(11/12), 1497–1512. Hiller, A., Hermichen, W.-D., & Wand, U. (1995). Radiocarbon Dated Subfossil Stomach Oil Deposits from Petrel Nesting Sites: Novel Paleoenvironmental Records from Continental Antarctica. Radiocarbon, 37(2), 171–180. Hjort, C., Ingo´lfsson, O´., Mo¨ller, P., & Lirio, J. M. (1997). Holocene Glacial History and Sea-Level Changes on James Ross Island, Antarctic Peninsula. J. Quat. Sci., 12, 259–273. Hodgson, D. A., Noon, P. E., Vyverman, W., Bryant, C. L., Gore, D. B., Appleby, P., Gilmour, M., Verleyen, E., Sabbe, K., Jones, V. J., Ellis-Evans, J. C., & Wood, P. B. (2001). Were the Larsemann Hills Ice-Free Through the Last Glacial Maximum? Antarctic Sci., 13(4), 440–454. Hodgson, D. A., Verleyen, E., Sabbe, K., Squier, A. H., Keely, B. J., Leng, M. J., Saunders, K. M., & Vyverman, W. (2005). Late Quaternary Climate-Driven Environmental Change in the Larsemann Hills, East Antarctica, Multi-Proxy Evidence from A Lake Sediment Core. Quat. Res., 64(1), 83–99. Huybrechts, P. (1990). A 3-D Model for the Antarctic Ice Sheet: A Sensitivity Study on the Glacial-Interglacial Contrast. Climate Dynamics, 5, 79–92. Huybrechts, P. (2002). Sea Level Changes at the LGM from Ice-Dynamic Reconstructions of the Greenland and Antarctic Ice Sheets During the Glacial Cycles. Quat. Sci. Rev., 21, 203–231. Igarashi, A., Harada, N., & Moriwaki, K. (1995). Marine fossils of 30-40 ka in raised beach deposits, and late Pleistocene glacial history around Lu¨tzow-Holm Bay, East Antarctica. Proc. NIPR Symposium, Antarctic Geoscience, 8, 219–229. Igarashi, A., Miura, H. and Hart, C. (1998). Amino-Acid Racemization Dates of Fossil Molluscs from Raised Beach Deposits on East Ongul Island and Northern Part of Langhovde, Lu¨tzow-Holm Bay region, East Antarctica. The 18th Symposium on Antarctic Geosciences, Program and Abstracts. NIPR, pp. 59–61.
566
A. P. Wright et al.
Ingo´lfsson, O´. (2004). Quaternary Glacial and Climate History of Antarctica. In: J. Ehlers, & P. L. Gibbard (Eds). Quaternary Glaciations-Extent and Chronology Part III. Elsevier, pp. 3–43. Jacobs, S. S., Helmer, H. H., Doake, C. S. M., Jenkins, A., & Frohlich, R. M. (1992). Melting of ice shelves and the mass balance of Antarctica. J. Glaciology, 38(130), 375–387. Joughin, I. (2002). Ice-sheet velocity mapping: a combined interferometric and spekle tracking approach. Ann. Glaciology, 34, 195–201. Kiernan, K., McConnell, A., Colhoun, E., & Lawson, E. (2003). Radiocarbon Dating of Mumiyo from the Vestfold Hills, East Antarctica. Papers Proc. Royal Soc. Tasmania, 136, 141–144. Kirkup, H., Melles, M., & Gore, D. B. (2002). Late Quaternary Environment of the Southern Windmill Islands, East Antarctica. Antarctic Sci., 14, 385–394. Krause, W. E., Krbetschek, M. R., & Stolz, W. (1997). Dating of Quaternary Lake Sediments from the Schirmacher Oasis (East Antarctica) by Infra-Red Stimulated Luminescence (IRSL) Detected at the Wavelength of 560 NM. Quat. Sci. Rev., 16(3–5), 387–392. Leventer, A., Domack, E. W., Dunbar, R., Pike, J., Stickley, C., Maddison, E., Brachfeld, S., Manley, P., & McClennen, C. (2006). Marine Sediment Record from the East Antarctic Margin Reveals Dynamics of Ice Sheet Recession. GSA Today, 16(12), 4–10. Licht, K. J. (2004). The Ross Sea’s Contribution to Eustatic Sea Level During Meltwater Pulse 1 A. Sedimentary Geology, 165, 343–353. Licht, K. J., Lederer, J. R., & Swope, R. J. (2005). Provenance of LGM Glacial till (Sand Fraction) Across the Ross Embayment, Antarctica. Quat. Sci. Rev., 24, 1499–1520. Lorius, C., Jouzel, J., Ritz, C., Merlivat, L., Barkov, N. I., Korotkevich, T. S., & Kotlyakov, V. M. (1985). A 150,000 Year Climate Record from Antarctic Ice. Nature, 316, 591–596. Lowe, A. L., & Anderson, J. B. (2002). Reconstruction of the West Antarctic Ice Sheet in Pine Island Bay During the Last Glacial Maximum and Its Subsequent Retreat History. Quat. Sci. Rev., 21, 1879–1897. Lythe, M. B., & Vaughan, D. G., & BEDMAP Consortium. (2001). BEDMAP: A New Ice Thickness and Subglacial Topographic Model of Antarctica. J. Geophys. Res., 106, 11335–11351. Mackintosh, A., White, D., Fink, D., Gore, D. B., Pickard, J., & Fanning, P. C. (2007). Exposure Ages from Mountain Dipsticks in Mac, Robertson Land, East Antarctica, Indicate Little Change in Ice Sheet Thickness Since the Last Glacial Maximum. Geology, 35, 551–554. Maemoku, H., Miura, H., Saigusa, S., & Moriwaki, K. (1997). Stratigraphy of the Late Quaternary Raised Beach Deposits in the Northern Part of Langhovde, Lu¨tzow-Holm Bay, East Antarctica. Proc. NIPR Symposium. Antarctic Geoscience, 10, 178–186.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
567
Ma¨usbacher, R. (1991). Die Jungkvata¨re Relief- und Klimageschichte im Bereich der Fildeshalbinsel, Su¨d-Shetland-Inseln, Antarktis. Heidelberger Geographische Arbeiten, 89, 207. McMullen, K., Domack, E. W., Leventer, A., Olson, C., Dunbar, R., & Brachfeld, S. (2006). Glacial Morphology and Sediment Formation in the Mertz Trough, East Antarctica. Palaeogeography, Palaeoclimatology, Palaeoecology, 231, 169–180. Mercer, J. H. (1978). West Antarctic Ice Sheet and CO2 Greenhouse Effect: A Threat of Disaster. Nature, 271, 321–325. Miura, H., Maemoku, H., Seto, K., & Moriwaki, K. (1998). Late Quaternary East Antarctic Melting Event in the Soya Coast Region Based on Stratigraphy and Oxygen Isotopic Ratio of Fossil Molluscs. Polar Geosci., 11, 260–274. Morse, D. L., Waddington, E. D., & Steig, E. J. (1998). Ice Age Storm Trajectories Inferred from Radar Stratigraphy at Taylor Dome, Antarctica. Geophys. Res. Lett., 25(17), 3383–3386. Mosola, A. B., & Anderson, J. B. (2006). Expansion and Rapid Retreat of the West Antarctic Ice Sheet in Eastern Ross Sea: Possible Consequence of Over-Extended Ice Streams? Quat. Sci. Rev., 25, 2177–2196. Nakada, M., Kimura, R., Okuno, J., Moriwaki, K., Miura, H., & Maemoku, H. (2000). Late Pleistocene and Holocene Melting History of the Antarctic Ice Sheet Derived from Sea Level Variations. Marine Geology, 167, 85–103. Nakada, M., & Lambeck, K. (1988). The Melting History of the Late Pleistocene Antarctic Ice Sheet. Nature, 333, 36–40. O’Cofaigh, C., Dowdeswell, J. A., Allen, C. S., Hiemstra, J. F., Pudsey, C. J., Evans, J., & Evans, D. J. A. (2005a). Flow Dynamics and Till Genesis Associated with a Marine-Based Antarctic Palaeo-Ice Stream. Quat. Sci. Rev., 24, 709–740. O’Cofaigh, C., Larter, R. D., Dowdeswell, J. A., Hillenbrand, C. D., Pudsey, C. J., Evans, J., & Morris, P. (2005b). Flow of the West Antarctic Ice Sheet on the Continental Margin of the Bellingshausen Sea at the Last Glacial Maximum. J. Geophys. Res., 110(B11103), doi:10.1029/2005JB003619. O’Cofaigh, C., Pudsey, C. J., & Dowdeswell, J. A. (2002). Evolution of Subglacial Bedforms Along a Paleo-Ice Stream, Antarctic Peninsula Continental Shelf. Geophys. Res. Lett., 29(8), 1–4. Oerlemans, J. (1982). A Model of the Antarctic Ice Sheet. Nature, 297, 550–553. Overpeck, J. T., Otto-Bliesner, B. L., Miller, G. H., Muhs, D. R., Alley, R. B., & Kiehl, J. T. (2006). Paleoclimatic Evidence for Future Ice-Sheet Instability and Rapid Sea-Level Rise. Science, 311, 1747–1750. Payne, A. J. (1999). A Thermomechanical Model of Ice Flow in West Antarctica. Climate Dynamics, 15, 115–125. Peltier, W. R. (1994). Ice Age Paleotopography. Science, 265(5169), 195–201. Peltier, W. R. (2004). Global Glacial Isostasy and the Surface of the Ice-Age Earth: The ICE-56 (VM2) Model and Grace. Ann. Rev. Earth and Planetary Sci., 39, 114–149.
568
A. P. Wright et al.
Petit, J. R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J.-M., Basile, I., Bender, M., Chappellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V. M., Legrand, M., Lipenkov, V. Y., Lorius, C., Pepin, L., Ritz, C., Saltzman, E., & Stievenard, M. (1999). Climate and Atmospheric History of the Last 420,000 Years from the Vostok Ice Core, Antarctica. Nature, 399, 429–436. Philippon, G., Ramstein, G., Charbit, S., Kageyama, M., Ritz, C., & Dumas, C. (2006). Evolution of the Antarctic Ice Sheet Throughout the Last Deglaciation: A Study with a New Coupled Climate-North and South Hemisphere Ice Sheet Model. Earth and Planetary Sci. Lett., 248, 750–758. Pope, P. G., & Anderson, J. B. (1992). Late Quaternary Glacial History of the Northern Antarctic Peninsula’s Western Continental Shelf: Evidence from the Marine Record. Antarctic Res. Series, 57, 63–91. Pudsey, C. J., Barker, P. F., & Larter, R. D. (1994). Ice Sheets Retreat from the Antarctic Peninsula Shelf. Continental Shelf Res., 14, 1647–1675. Pudsey, C. J., & Evans, J. (2001). First Survey of Antarctic Sub-Ice Shelf Sediments Reveals Mid-Holocene Ice Shelf Retreat. Geology, 29(9), 787–790. Putkonen, J., & Swanson, T. (2003). Accuracy of Cosmogenic Ages for Moraines. Quat. Res., 59, 255–261. Rignot, E., & Jacobs, S. S. (2002). Rapid Bottom Melting Widespread near Antarctic Ice Sheet Grounding Lines. Science, 296(5575), 2020–2023. Ritz, C., Rommelaere, V., & Dumas, C. (2001). Modelling the Evolution of Antarctic Ice Sheet over the Last 420,000 Years: Implications for Altitude Changes in the Vostok Region. J. Geophys. Res., 106, 31943–31964. Scherer, R. P., Aldahan, A., Tulaczyk, S., Possnert, G., Engelhardt, H., & Kamb, B. (1998). Pleistocene Collapse of the West Antarctic Ice Sheet. Science, 281, 82–85. Shackleton, N. J. (2000). The 100,000-year Ice Age Cycle Identified and Found to Lag Temperature, Carbon Dioxide, And Orbital Eccentricity. Science, 289(5486), 1897–1902. Shackleton, N. J., Sanchez-Goni, M. F., Pailler, D., & Lancelot, Y. (2003). Marine Isotope Substage 5e and the Eemian Interglacial. Global and Planetary Change, 36(3), 151–155. Shevenell, A. E., Domack, E. W., & Kernan, G. M. (1996). Record of Holocene Palaeoclimate Change Along the Antarctic Peninsula: Evidence from Glacial Marine Sediments, Lallemand Fjord. Papers and Proc. Royal Soc. Tasmania, 130, 55–64. Shipp, S. S., Anderson, J. B., & Domack, E. W. (1999). Late Pleistocene-Holocene Retreat of the West Antarctic Ice Sheet system in the Ross Sea: Part 1-Geophysical Results. Geological Soc. America Bull., 111, 1486–1516. Siegert, M. J. (2003). Glacial-Interglacial Variations in Central East Antarctic Ice Accumulation Rates. Quat. Sci. Rev., 22, 741–750. Stirling, C. H., Esat, T. M., Lambeck, K., & McCulloch, M. T. (1998). Timing and Duration of the Last Interglacial: Evidence for a Restricted Interval of Widespread Coral Reef Growth. Earth and Planetary Sci. Lett., 160(3-4), 745–762.
Antarctica at the Last Glacial Maximum, Deglaciation and the Holocene
569
Stone, J. O., Balco, G. A., Sugden, D. E., Caffee, M. W., Sass, L. C. I., Cowdery, S. G., & Siddoway, C. (2003). Holocene Deglaciation of Marie Byrd Land, West Antarctica. Science, 299, 99–102. Studinger, M., Bell, R. E., Blankenship, D. D., Finn, C. A., Arko, R. A., Morse, D. L., & Joughin, I. (2001). Subglacial Sediments: A Regional Geological Template for Ice Flow in Antarctica. Geophys. Res. Lett., 28, 3493–3496. Sugden, D. E., Bentley, M. J., & O’Cofaigh, C. (2006). Geological and Geomorphological Insights into Antarctic Ice Sheet Evolution. Philosophical Trans. Royal Soc. A, 364, 1607–1625. Tarasov, L., & Peltier, W. R. (2003). Greenland Glacial Hstory, Borehole Constraints and Eemian Extent. J. Geophys. Res., 108(B3), 2143. Taylor, F., & McMinn, A. (2002). Late Quaternary Diatom Assemblages from Prydz Bay, Eastern Antarctica. Quat. Res., 57(1), 151–161. Taylor, J., Siegert, M. J., Payne, A. J., Hambrey, M. J., O’Brien, P. E., Cooper, A. K., & Leitchenkov, G. (2004). Topographic Controls on PostOligocene Changes in Ice-Sheet Dynamics, Prydz Bay Region, East Antarctica. Geology, 32(3), 197–200. Tingey, R. J., McDougall, I., & Gleadow, A. J. W. (1983). The Age and Mode of Formation of Gaussberg, Antarctica. J. Geological Soc. Australia, 30, 241–246. Tregoning, P., Welsh, A., McQueen, H., & Lambeck, K. (2000). The Search for Postglacial Rebound Near the Lambert Glacier, Antarctica. Earth, Planets Space, 52, 1037–1041. Tushingham, A. M., & Peltier, W. R. (1991). A New Global Model of late Pleistocene Deglaciation Based upon Geophysical Predictions of Post Glacial Relative Sea Level Change. J. Geophys. Res., 96, 4497–4523. Van Ommen, T. D., Morgan, V., & Curran, M. A. J. (2004). Deglacial and Holocene Changes in Accumulation at Law Dome, East Antarctica. Ann. Glaciol., 39, 359–365. Verleyen, E., Hodgson, D. A., Milne, G. A., Sabbe, K., & Vyverman, W. (2005). Relative Sea-Level History from the Lambert Glacier Region, East Antarctica, and Its Relation to Deglaciation and Holocene Glacier Readvance. Quat. Res., 63(1), 45–52. Waddington, E. D., Conway, H., Steig, E. J., Alley, R. B., Brook, E. J., Taylor, K. C., & White, J. W. C. (2005). Decoding the Dipstick: Thickness of Siple Dome, West Antarctica, at the Last Glacial Maximum. Geology, 33, 281–284. Wagner, B., Cremer, H., Hultzsch, N., Gore, D. B., & Melles, M. (2004). Late Pleistocene and Holocene History of Lake Terrasovoje, Amery Oasis, East Antarctica, and Its Climatic and Environmental Implications. J. Paleolimnology, 32(4), 321–339. Wagner, B., Hultzsch, N., Melles, M., & Gore, D. B. (2007). Indications of Holocene Sea-Level Rise in Beaver Lake, East Antarctica. Antarctic Science, 19, 125–128.
570
A. P. Wright et al.
Whitehead, J. M., Harwood, D. M., & McMinn, A. (2003). Ice-Distal Upper Miocene Marine Strata from Inland Antarctica. Sedimentology, 50(3), 531–552. Zwartz, D., Bird, M., Stone, J., & Lambeck, K. (1998). Holocene Sea-Level Change and Ice-Sheet History in the Vestfold Hills, East Antarctica. Earth Planetary Sci. Lett., 155(1–2), 131–145.
Subject Index AABW (see Antarctic bottom water (AABW)) AAIW (see Antarctic intermediate water (AAIW)) ACC (see Antarctic circumpolar current (ACC)) ACE programme (see Antarctic Climate Evolution (ACE) programme) Acoustic basement rocks, of Ross Sea region, 122 Acoustic stratification, wedge downlaps, 145 Acoustic stratigraphy, of Wilkes Land, 127–133 Age-depth model, for ODP drilling site, 141 Alkalic magmatic province, Ross Sea tectonic evolution, 283 AMS (see Atomic mass spectrometry (AMS)) Andean Orogen, Antarctic, 275 ANDRILL drilling, on Ross Ice shelf, 124 Angiosperms, fossil flora, 333–335 Antarctica climate change/warming, 103–107 future geoscience investigations in, 287–291 modelling, 555–558 Antarctic Andean Orogen, 275 Antarctic bottom water (AABW), 94–97
Antarctic circumpolar current (ACC), 97–99, 341, 573 Antarctic climate glacial-interglacial time scales, and, 481–485 millennial-scale changes, and, 485–486 Antarctic climate evolution, 509–513 ice sheet expansion, 509 late Pliocene cooling, 509 late Pliocene-early Pleistocene G-I climate cycles, 509–510 marine ice sheet margin development, 509 Marine Isotope Stage 31, 510–511 mid-Pleistocene climate transition, 511–512 millennial-scale climate variability, 512–513 Antarctic Climate Evolution (ACE) programme, 2–4 Antarctic coastal and continental shelf rock-drilling sites, 60–61 Antarctic coastal current, 100 Antarctic geological drilling programme (ANDRILL), 476–479 Antarctic glacial history, Cenozoic, 4–9 advancement in geological research during twenty-first century, 58–63 developments in late 1970s, 46–47
578
Subject Index
discoveries at offshore and on continent in 1980s, 47–53 drilling records during 1972–1975, 39–45 land-based drilling, 1973–1974, 43 sea ice drilling, 1975, 43–45 ship-based drilling, 1972–1973, 40–43 events in evolution of, 36–37 research advancement in mid-twentieth century, 38–39 research advancement in 1990s, 53–57 Antarctic glaciological data versus Ice-Sheet models, 245–249 Antarctic Ice Sheet DSDP–ODP drilling dates initiation of, 50–52 dynamic, late Oligocene, 49–50 future prospective of, 63–69 John Mercer’s hypothesis for behaviour of, 47 landscape modelling, 58–59 modelling of, 55–56 numerical modelling of, 235–253 EISMINT, 242–245 ice-sheet models, 238–241 ice-sheet models versus glaciological data, 245–249 ice-sheet processes, 236–238 ice-sheet reconstructions, 249–253 model input, 241–242 Pliocene record of, 7 pre-Quaternary history for, 39 proxies climate records, 50 thermal conditions, 247–248 velocity map of, 68 Antarctic Ice Sheet, 468, 533–535 over a full glacial-interglacial cycle, 551
Antarctic intermediate water (AAIW), 90–92, 491 Antarctic Isotopic Maxima (AIM), 486 Antarctic Offshore Stratigraphy (ANTOSTRAT) project, 52–53, 116 and ODP, 53–54 Antarctic Peninsula, 472–473, 532, 545–547, F12.1 eastern margin of bathymetric contours, 153 single-channel seismic line across, 154 fossil floras King George Island, 328–330 Late Palaeocene-Late Eocene, 326–327 Seymour Island, 330 geologic transition in, 170 geomorphic and seismic stratigraphic features of, 162–164 lithostratigraphic transitions in, 167–170 Pacific margin of depositional sequences, 159 mid-shelf basins, 161 ridge-crest segments, 158 sea-level and ice-volume changes, 170–172 sedimentary strata, 311–313 seismic stratigraphic history Cretaceous to early Oligocene, 165 Oligocene glaciomarine sections, 166 sediment drape, 167 South Orkney Islands region, 155–156 South Shetland Islands region Bransfield Strait, 156–157 scientific drilling in, 158
Subject Index
terrestrial environment, Middle Miocene to Pliocene, 417–421 Antarctic periphery climate, 486–487 Antarctic sea-ice drilling, 54–55 Antarctic slope front, 100 ANTOSTRAT (see Antarctic Offshore Stratigraphy (ANTOSTRAT) project) Apparent sea level, 379 Atlantic DWBC inflow, 101–102 Atmospheric carbon dioxide (CO2) at E/O boundary, computer modelling, 344–348 influence across Oligocene–Miocene boundary, 385 Atmospheric circulation, 467 Atmospheric variability, from ice cores, 479–488 Antarctic climate and glacial-interglacial time scales, 481–485 Antarctic periphery climate, 486–487 future challenges, for ice coring, 487–488 ice core record, 480–481 millennial-scale changes, 485–486 Atomic mass spectrometry (AMS), 536 Austro-Hungarian North Pole Expedition, 15–16 Basement rocks, acoustic of Ross Sea region, 122 Battye Glacier Formation, 412 Beacon Supergroup, 273–274 Beaver Lake, 539 BEDMAP topography, 555 Bellingshausen Sea Volcanic Group, 421 Biomagnetostratigraphic dating, 55
579
Borchgrevink orogeny, 272–273 Bransfield Strait, 156–157, 276 Bunger Hills, 541 Byrd, Rear-Admiral Richard E., 21–22 Byrd glacier, 248 Canonical Milankovitch theory, 467 Cape Roberts drilling project, 291 Cape Roberts Project (CRP), 54–55, 469 Carbon dioxide (CO2) (see Atmospheric carbon dioxide (CO2)) CDW (see Circumpolar deep water (CDW)) Cenozoic geological time periods, 2 magmatism in Ross Sea, during, 279–284 oxygen isotope records for, 343 tectonic evolution in Ross Sea, during, 279–284 Cenozoic climate history Antarctic Peninsula, 152–161 Prydz Bay, 135–144 Ross Sea, 118–126 Weddell Sea, 144–152 Wilkes Land, 126–135 Cenozoic investigations in the western Ross Sea (CIROS) Project, 47–48, 469 Circum-Antarctic geographical distribution maps dinocysts, in Late Paleocene–Miocene, 338 Circum East Antarctic Mobile Belt, 263 Circumpolar deep water (CDW), 92–94 types of, 94
580
Subject Index
CIROS (see Cenozoic investigations in the western Ross Sea (CIROS) Project) CIROS-1, 49–50 CIROS-1 drill core climate records, Oligocene–Miocene boundary, 381, 384–385 revised age models for, 383 western Ross Sea, 317–321 CLIMAP reconstruction, 555 Climate change/warming, of Antarctica and Southern Ocean (Subpolar region), 103–107 Climate modelling, 55–56 Climate records drilling, 316–321 Eocene/Oligocene, 310–349 from fossil plants, 325–331 from marine microfossils, 335–342 from palynomorphs, 331–335 from sedimentary record, 311–325 from terrestrial realm, 325–335 Middle Miocene to Pliocene marine record, 424–436 terrestrial environments, 404–424 Oligocene–Miocene boundary, 370–391 Coastal current, Antarctic, 100 Coats Land, 548–549 CO2 concentrations, 483 Collins Glacier deposit, fossil flora, 329–330 Comite´ Spe´cial de l’Anne´e Ge´ophysique Internationale (CSAGI), 23–24
Computer modelling CO2 and ice-sheet inception at E/O boundary, 344–348 Continental rise glacial stratigraphy, Wilkes Land, 132–133 Continental shelf glacial stratigraphy, Wilkes Land, 130–131 Continental slope deposits, 159 Continental slope glacial stratigraphy, Wilkes Land, 131 Contribution to global sea level, isostatic modelling, 558–559 Crary Mountains, 423 Crary Trough Mouth Fan channel-levee complexes, 151 seismic line NARE-8517 across, 148 Cratons, 261 Crozet basin, 102 CRP-2A drill core environmental proxy data, Oligocene–Miocene boundary, 381, 382 revised age models for, 383 CRP-3 drill core, 317–321 Cryosphere, 534 CSAGI (see Comite´ Spe´cial de l’Anne´e Ge´ophysique Internationale (CSAGI)) Cytadela leaf flora, 328–329 Dansgaard-Oeschger (D-O) interstadials, 467 dD (deuterium) signal, 552 Deep ice cores proxy climate records from, 53 Deep Sea Drilling Project (DSDP), 40–41 Wilkes Land margin, 133 Deep-sea isotopes Antarctic Ice Sheet proxy records, 50
Subject Index
Deep western boundary currents (DWBC), 101–103, 490 Depositional model for Late Oligocene/Early Miocene, 319–320 Depositional patterns, 159 of Antarctic Ice Sheets, 49 Dinocysts Late Eocene–Early Oligocene, 337–339 Middle Eocene, 339–340 Palaeocene–Middle Eocene, 336–337 D-O climate oscillations, 467 D-O interstadials (see Dansgaard-Oeschger (D-O) interstadials) Dragon Glacier Moraine, macroflora deposits, 328 Drake Passage, climate evolution, 346 Drake Plate, 275 Drill cores, in western Ross Sea, 317–321 Drilling climate records, 316–321 in Prydz Bay, 321–325 records of Ross Sea ice-sheet evolution, 122–126 on Wilkes Land margin, 133 Dronning Maud Land continental slope, 145 Dronning Maud Land margin, 152 seismic line AWI-90110 across, 147 Dry Valley Drilling Project (DVDP), 43, 469 DSDP (see Deep Sea Drilling Project (DSDP)) DSDP Leg 28, 40, 43 DSDP Leg 29, 40
581
DSDP–ODP drilling dates initiation of Antarctic Ice Sheet, 50–52 DVDP (see Dry Valley Drilling Project (DVDP)) DVDP-15, 45 DWBC (see Deep western boundary currents (DWBC)) EAIS (see East Antarctica ice sheet (EAIS)) Early glacial deposites during late Eocene, Prydz Bay, 138–139 ice-sheet evolution in Ross Sea region during late Oligocene to early Miocene, 123–124 Earth Models of Intermediate Complexity (EMIC), 505 Earth’s temperature, variation from cenozoic to present, 3 East Antarctica, 259–260 East Antarctic Craton, 261 East Antarctic Ice Sheet (EAIS), 403–404, 472, 573 marine record, Middle Miocene to Pliocene, 424–428 East terrestrial environments, Middle Miocene to Pliocene, 404–417 East wind drift, 100 EBM (see Energy-balance climate models (EBM)) Edward VII Peninsula (Westernmost Marie Byrd Land), 271 Eemian interglacial, 536–537 EISMINT (see European Ice Sheet Modelling Initiative (EISMINT)) ELA (see Equilibrium line altitude (ELA)) Ellsworth Land, 421–422 Ellsworth Orogeny, 274–275 Ellsworth–Pensacola Mountains, 274
582
Subject Index
EMIC (see Earth Models of Intermediate Complexity (EMIC)) Enderby land, 537–538, F12.2 Energy-balance climate models (EBM), 501 Environmental records from marine microfossils, 335–342 from palynomorphs, 331–335 Eocene environment in Ross Sea, 314–316 ocean isotope record, climate evolution, 342–344 Eocene erratic rocks found in coastal region of Ross Sea, 122–123 Eocene/Oligocene climate signals from sedimentary record, 311–325 EPICA (see European Programme for Ice Coring in Antarctica (EPICA)) Equilibrium line altitude (ELA), 250–251 Erebus volcanic province, Ross Sea tectonic evolution, 283–284 Erratic rocks, Eocene found in coastal region of Ross Sea, 122–123 Eucampia antarctica, 472 European Ice Sheet Modelling Initiative (EISMINT), 242–245 outcomes, 244–245 European Programme for Ice Coring in Antarctica (EPICA), 469 Eustatic sea-level curves, Oligocene–Miocene boundary, 378–379 Exhumation events, variation Transantarctic Mountains, 286
Ferrar Supergroup, 277 Filchner-Ronne Ice Shelf, 532, 548–549, 550, F12.1 Flow of ice, 236 in central regions, 245–247 processes controlling, 237 radar layering and internal, 247 Fluctuation, NH ice volume, 468 Fossil Hill Formation, 329 Fossil plants climate records, Eocene/Oligocene, 325–331 Fossil record, terrestrial of east Antarctic climate, 414–417 Framnes mountains, 538 Franz Joseph Land, 15 Gaussberg, 541 GCM (see Global Climate Model (GCM)) Geological information, 535–550 Holocene glacier, climate fluctuations, 549 Last Glacial Maximum, 537–549 last interglacial, 536–537 late Quaternary ice-sheet, 535–536 pattern and timing, of ice retreat, 550 Prydz Bay, 550 Geological map, 262 Geological time periods, Cenozoic, 2 Geomagnetic Polarity Time Scale calibration history, Oligocene– Miocene boundary, 373 Geotectonic settings before Gondwana amalgamation, 261–264 present-day, 259–260 George V Land, 541 Glacial climate, 533–534 Glacial history of Antarctic, 4–9
Subject Index
Glacial-interglacial cycles, 488–499 (see also Oceanic variability, from southern ocean sediment cores) iceberg discharge and migration, 498–499 intermediate waters, 491 ocean fronts, 496–497 surface ocean currents, 497 surface waters, 492–495 thermohaline circulation, 488–490 Glacial-interglacial (G-I) cycles, 468 Glacial-interglacial time scales, 481–485 Glacial sedimentation ancestral Lambert Glacier, advance and recession stages, 323–324 Glacial sequences on Continental shelf, 130–131 Glacial stratigraphy continental rise, 132–133 continental shelf, 130–131 continental slope, 131 Glacial variability, from continental margin geological record, 469–479 ANDRILL, 476–479 Antarctic Peninsula, 472–473 Prydz Bay, 473–475 Ross Sea, 475–476 Weddell Sea, 471–472 Glaciation record of Wilkes Land, 134–135 GLIMMER community ice-sheet model, 555 Global Circulation Model (GCM) simulation, 347, 348 south polar climate variations in response to a growing ice sheet, 388 Global Climate Model (GCM), 504, 551
583
Global ice volume evolution during Eocene to Oligocene, 342–344 Global positioning system (GPS), 559 GLOCHANT (see Group of Specialists on Global Change (GLOCHANT)) Glomar Challenger, 40, 42 Gondwana amalgamation, 270 geotectonic settings before, 261–264 break-up, 276–287 tectonic stages of, 276–279 dispersal of Southern Hemisphere Continents, 276–287 supercontinent, 265–276 Gondwana, evolution late precambrian–early palaeozoic (c. 600–450Ma), 266–272 from Rodinia, 267–272 upper palaeozoic–mesozoic (c. 450–180Ma), 272–276 GPS (see Global positioning system (GPS)) Grenvillian orogen, 263 Grounding-line fan model of glaciomarine sedimentation, Late Oligocene/Early Miocene, 319–320 Group of Specialists on Global Change (GLOCHANT), 54 Gyres, 99–100 ‘‘Heroic era’’, 35, 38 Holocene, 531 glacier, climate fluctuations, 549 Iceberg discharge and migration, 498–499 (see also Glacial-interglacial cycles) Iceberg-rafted debris (IRD), 473
584
Subject Index
Ice cores proxy climate records from deep, 53 Ice coring, future challenges, 487–488 Ice-sheet development mid-early Miocene to early Pliocene in Ross Sea region, 124–125 Ice-sheet evolution in Ross Sea early glacial, 123–124 during mid-early miocene to early pliocene, 124–125 polar ice sheet, 125–126 pre-ice-sheet, 122–123 Ice sheet expansion, 509 (see also Antarctic climate evolution) Ice-sheet fluctuations Late Miocene–Early Pliocene, 427–428 Ice-sheet hysteresis influence across Oligocene–Miocene boundary, 390–391 Ice-sheet inception at E/O boundary, computer modelling, 344–348 Ice sheet Model Intercomparison Programme (ISMIP), 245 Ice-sheet modelling, 55–56, 550–551 Ice-sheet models, numerical, 238–241 versus Antarctic glaciological data, 245–249 Ice-sheet processes, 236–238 Ice Sheets, Antarctic East (see East Antarctic Ice Sheet (EAIS)) modelling, 437–441 West (see West Antarctic Ice Sheet (WAIS)) Ice-sheet sensitivity, 554 Ice streams, 154
Ice volume, global evolution during Eocene to Oligocene, 342–344 Ice-volume variations, 170–172 ICSU (see International Council of Scientific Unions (ICSU)) IGY (see International Geophysical Year (IGY)) IMAGES programme (see International Marine Global Change Study (IMAGES) programme) IMO (see International Meteorological Organization (IMO)) Indian DWBC inflow, 102 Indo-Antarctica, 270 InSAR (see Interferometric Synthetic Aperture Radar (InSAR)) Integrated Ocean Drilling Programme (IODP), 65, 469 Interferometric Synthetic Aperture Radar (InSAR) for surface velocity of ice sheets, 248–249 Interglacial ice sheet, 551–552 International Council of Scientific Unions (ICSU), 23, 25 International Geophysical Year (IGY), 1957–1958, 22–26 establishment of SCAR, 25 estimation of ice volume in Antarctica, 25 International Marine Global Change Study (IMAGES) programme, 8 International Meteorological Conference, 1879, 17 International Meteorological Organization (IMO), 19 International Polar Commission, 17, 19–20
Subject Index
International Polar Conference, 17 International Polar Year (IPY) 1882–1883, 14–19 1932–1933, 19–22 1957–1958, 22–26 2007–2008, 27–30 themes, 29 IODP (see Integrated Ocean Drilling Programme (IODP)) Isostasy, 238 Isostatic modelling contribution to global sea level, 558–559 of post-glacial uplift, 558–559 Isotopic record Oligocene–Miocene boundary, 374–377 James Ross Island Volcanic Group, 417, 421 John Mercer’s hypothesis for behaviour of Antarctic Ice Sheet, 47 Jurassic Ferrar Large Igneous Province (FLIP), 277 Jutul Penck Graben, 279 K–Ar dating for Antarctic Cenozoic ice sheets, 39 King George Island fossil plants, 328–330 stratigraphy, 313–314 Krakowiak Glacial Member, 314 Lake glubokoye, 537 Lambert Glacier, ancestral glacial sedimentation, advance and recession stages, 323–324 Lambert glacier-Amery ice shelf, 7, 538–539
585
Lambert Glacier region Middle Miocene to Pliocene history, 410–414 Lambert Graben, 278–279 La Meseta Formation, 312 Land-based drilling to study Antarctic Cenozoic glaciation, 43 Landscape modelling of Antarctic Ice Sheet, 58–59 Larsemann Hills, 539, 549–550 Larsen Shelf sediments, 149 seismic line AWI-97051 across, 147 Last glacial maximum (LGM), 8–9, 473, 531, 537–549, 574 Antarctic Peninsula, 545–547 Coats Land, 548–549 Enderby Land, 537–538, F12.2 Filchner-Ronne Ice Shelf, 548–549 Lambert Glacier-Amery Ice Shelf, 538–539 Mac.Robertson Land, 538–539 Marie Byrd Land, 544–545 Princess Elizabeth Land, 539–542 Prydz Bay, 538–539, F12.3 Queen Mary Land, 539–542 Queen Maud Land, 537–538, F12.2 in Ross Sea, 125–126 Ross Sea Embayment, 542–544 Transantarctic Mountains, 542–544 Weddell Sea Embayment, 548–549 Wilkes Land, 539–542 Late Eocene–Early Oligocene dinocysts, 337–339 Late Eocene-early Oligocene cooling, 5 Late Miocene–Early Pliocene ice-sheet fluctuations, 427–428 marine records, 435–436
586
Subject Index
Late Oligocene/Early Miocene depositional model for, 319–320 Victoria Land coast in, 321 Late Palaeocene-Late Eocene fossil floras, Antarctic Peninsula region, 326–327 Late Pliocene cooling, 509 (see also Antarctic climate evolution) Late Pliocene-early Pleistocene G-I climate cycles, 509–510 (see also Antarctic climate evolution) Late Precambrian–Early Palaeozoic (c. 600–450 Ma) Gondwana, evolution, 266–272 palaeo-pacific margin, 270–272 Late Quaternary ice-sheet, 535–536 reconstructions, 533–534 Law dome, 105, 534 LCDW (see Lower circumpolar deep water (LCDW)) LGM (see Last glacial maximum (LGM)) Lithostratigraphic changes, in antarctic margin, 169 in drill cores from continental rise, 167–168 facies, 170 sea-level and ice-volume changes, 170–172 ‘‘Longyear 44’’ drilling rig, 43, 46 Lower circumpolar deep water (LCDW), 94 Lu¨tzow Holm–Prydz Bay Pan-African orogenic belt, 268–270 Mac.Robertson Land, 538–539 Magmatism in Ross Sea Sector, during the Cenozoic, 279–284
Magnesium/Calcium (Mg/Ca) data across Oligocene–Miocene boundary, 375–377, 380 Marie Byrd Land, 422–424, 544–545 Marine ice sheet margin development, 509 (see also Antarctic climate evolution) Marine Isotope Stage 31, 510–511 (see also Antarctic climate evolution) Marine microfossils climate records from, 335–342 Marine record, Middle Miocene to Pliocene Antartic Peninsula, 429–433 EAIS, 424–428 Southern Ocean, 433–436 WAIS, 429–433 Marine records, Late Miocene–Early Pliocene, 435–436 Mass balance, 237–238 Maud Rise, records from, 380 MCMC (see Monte Carlo Markov Chain (MCMC)) McMurdo Dry Valleys, 43 McMurdo Erratics, 315 McMurdo Ice Shelf (MIS), 471 McMurdo Sound climate data, Oligocene–Miocene boundary, 381–385 fossil flora, 330–331 map and cross-section of, 44 post-IGY surveys of glacial deposits in, 39 stratigraphy, 284 McMurdo Sound Sediment and Tectonic Studies Project (MSSTS), 46–47 MCS (see Multichannel seismic (MCS) reflection) MCS profiles, location of Wilkes Land margin, 128
Subject Index
Mertz glaciers, 541 Mesozoic–Cenozoic sedimentary basin, 152 Methane concentrations, 483 Meyer Desert Formation, fossil records, 416 Microfossils, marine, climate records from, 335–342 Middle-late Miocene cooling, 6 Middle Miocene Climate Transition (MMCT), 433–435 Middle Miocene cooling, marine records, 425–427 Middle Miocene to Pliocene terrestrial environments, 404–424 east, 404–417 west, 417–424 Mi-1 d18O event, origin, 386 Mid-Pleistocene Transition (MPT), 468, 505, 511–512 (see also Antarctic climate evolution; Pleistocene ice volume variation modelling) Mi-1 glaciations (see Miocene (Mi-1) glaciations) Millennial-scale climate variability, 485–486, 506, 512–513 (see also Antarctic climate evolution; Pleistocene ice volume variation modelling) Miocene climatic optimimum, 62–63 Miocene (Mi-1) glaciations, 5–6 Modelling Antartic climate and ice sheets, 437–441 CO2 and ice-sheet inception at E/O boundary, 344–348 Monte Carlo Markov Chain (MCMC), 502 Moraine complexes, 152
587
Mount Murphy, 422–423 Mount Wawel, macroflora deposits, 328 Moutone’e Lake, 549 Mozambique suture zone, 267–268 MPT (see Mid-Pleistocene Transition (MPT)) MSSTS (see McMurdo Sound Sediment and Tectonic Studies Project (MSSTS)) Multichannel seismic (MCS) reflection, 117 study of Ross Sea, 119–126 ice-sheet evolution, 122–126 NADW (see North Atlantic Deep Water (NADW)) Neogene age volcanic outcrops in West Antarctica, 418–420 Neogene (15-2 Ma) distribution terrestrial volcanic outcrops, West Antarctica, 413 Neogene stratigraphic section of Ross Sea, 126 Neogloboquadrina pachyderma, 474 New Zealand Antarctic Programme (see McMurdo Sound Sediment and Tectonic Studies Project (MSSTS)) NH (see Northern Hemisphere (NH)) Nimrod Glacier area, 270–271 Ninnis glaciers, 541 North Atlantic Deep Water (NADW), 488 Northern continental slope, 154 Northern Hemisphere (NH), 467 ice volume fluctuation, 468 summer insolation, 467 North Greenland Ice Core Project (North GRIP), 481
588
Subject Index
Novolazarevskaya ice shelf, 537 Nowaja Semlja, 15 Numerical ice-sheet models, 238–241 Numerical modelling of Antarctic Ice Sheet, 235–253 EISMINT, 242–245 ice-sheet models, 238–241 versus glaciological data, 245–249 ice-sheet processes, 236–238 ice-sheet reconstructions, 249–253 model input, 241–242 Numerical modelling reconstructions, 550–559 Antarctica modelling, 555–558 ice-sheet modelling, 550–551 ice-sheet sensitivity, 554 interglacial ice sheet, 551–552 time dependent simulations, 552–554 Ocean circulation, 467 influence across Oligocene–Miocene boundary, 385–389 Ocean Drilling Programme (ODP), 469 age-depth model for drilling site of, 141 and ANTOSTRAT, 53–54 deep-sea isotope data from sites in Pacific and Atlantic Oceans, 50 drilling sites of Prydz Bay, 136 Oceanic variability, from southern ocean sediment cores, 488–501 glacial-interglacial cycles, 488–499 millennial-scale cycles, 499–501 Ocean isotope record climate evolution during Eocene and Oligocene, 342–344
Ocean temperatures evolution during Eocene to Oligocene, 342–344 ODP (see Ocean Drilling Programme (ODP)) Oligocene glaciation, 158 ice sheets dynamic, 49–50 ocean isotope record, climate evolution, 342–344 Oligocene–Miocene boundary drivers of climate change across, 385–391 identification of, 371–374 proxy records, 374–379 records from Antartic margins, 379–385 Oligocene–Miocene boundary Mi-1 glaciation, 5–6 Orbital parameters influence across Oligocene–Miocene boundary, 389–390 Orogenic belts, 263 Oxygen isotope data across Oligocene–Miocene boundary, 374–376, 380 Oxygen isotope records for Cenozoic, 343 Pacific–Antarctic Ridge, initiation, 282 Pacific DWBC inflow, 102–103 Pagodroma Group, 7, 404–405 clay mineralogy, 412–414 glacial erosional features, 411 Middle Miocene to Pliocene history, 410–414 sedimentary characteristics, 411 stratigraphy, 408 Palaeoceanography, 339–342 Palaeocene–Middle Eocene dinocysts, 336–337
Subject Index
Palaeogene, Antarctic strata pollen assemblages, 332–333 Palaeogene fossil plants, 325–328 Palaeo-ice streams and S3/S2 boundary, 159 Palaeo-pacific margin Gondwana evolution, late precambrian-early palaeozoic, 270–272 Palaeo sea level records, Oligocene–Miocene boundary, 377–379 Palynomorphs climate records from, 331–335 from OPD drilling site in Prydz Bay, 139 Pattern and timing, of ice retreat, 550 Payer, Julius, 16 Plant macrofossils climate records from, 325–328 Plate-tectonically active part, 275–276 Pleistocene glacial cycles, 7–8 ice sheets, 251–253 Pleistocene ice volume variation modelling, 501–509 Antarctica, 506–509 millennial-scale variability, 506 MPT, 505 Northern Hemisphere, 501–505 Pliocene ice sheets, 249–251 record, of Antarctic Ice Sheet, 7 vegetation distributions, climate modelling, 439 Plio-pleistocene strata, 543 Point Hennequin Flora, 328 Polar ice sheet evolution in Prydz Bay, 142–143
589
evolution in Ross Sea region during early Pliocene through Quaternary, 125–126 Polar stereographic projection Oligocene–Miocene boundary locations, 372 Pollen assemblages in Antarctic Palaeogene strata, 332–333 Polonez Cove Formation, 314 Post-glacial uplift, isostatic modelling, 558–559 Pre-ice-sheet Cenozoic shelf edge, 148 evolution in Ross Sea region during pre-late-Oligocene, 122–123 palaeo-climate evolution of Prydz Bay during pre-late Eocene, 137–138 stratigraphy of Wilkes Land, 127–130 Prince Gustav Channel, 549 Princess Elizabeth Land, 539–542 Proxy records Oligocene–Miocene boundary, 374–379 Prydz Bay (PB), 473–475, 538–539, F12.3 cenozoic climate history, seismic interpretation and drilling, 135–144 ice-sheet development during Oligocene–Miocene, 139–142 drilling in, 321–325 geomorphic and seismic stratigraphic features of, 163 geomorphic features of, 163, 165 lithostratigraphic transitions in, 168–169 sea-level variations, 171
590
Subject Index
seismic lines and ODP drill sites of, 136 seismic stratigraphic features of, 165–167 Quaternary Antarctic evolution, 553–554 Quaternary climate dataset, 468 Queen Mary Land, 539–542 Queen Maud Land, 537–538 Radar layering Antarctic Ice Sheet, 247 RCM (see Regional Climate Model (RCM)) ‘‘Refrigerator trap’’, 342 Regional Climate Model (RCM), 504 Rennick Graben, 279 Rodinia Gondwana, evolution from, 267–272 Ross embayment McMurdo Sound coastal area of, 284 rifting in, 280–282 Ross gyre, 100 Ross Ice shelf ANDRILL drilling on, 124 Ross Ice Shelf Project (RISP), 46 Ross Orogeny, 271 Ross Sea, 475–476, 532, 542–544, F12.1 Cenozoic climate history, seismic interpretation and drilling data, 118–126 continental shelf edge, 166 Eocene environments in, 314–316 fossil flora from McMurdo Sound, 330–331 geomorphic and seismic stratigraphic features of, 165
lithostratigraphic transitions in, 168–170 magmatism, during Cenozoic, 279–284 sea-level variations in, 171 seismic data used in glacial sedimentary processes, 121–122 ship-based drilling at eastern, 40 stratigraphic correlation, 121 structural framework of, 119 tectonic evolution, during Cenozoic, 279–284 Ross Sea, South Western climate data, Oligocene–Miocene boundary, 381–385 Ross Sea, western drill cores in, 317–321 Ross Sea Rift, 279 RSL curve, 549 Salinity trends, of Southern Ocean, 105–106 SAM (see Southern Annular Mode (SAM)) SAMW (see Subantarctic mode water (SAMW)) Satellite interferometry to calculate ice velocity, 248–249 SCAR (see Scientific Committee on Antarctic Research (SCAR)) Schirmacher oasis, 537 Scientific Committee on Antarctic Research (SCAR) CIROS, 47–48 establishment during IGY, 25 planning for future research, 52–53 Sea ice drilling, 54–55 future prospective of, 64
Subject Index
to study Antarctic Cenozoic glaciation, 43–45 advancement during twenty-first century, 59–63 Sea-level variations, 170–172 Sea-riser technology, 471 Sedimentary record (Eocene/Oligocene) climate records from, 311–325 Seismic stratigraphy Antarctic Ice Sheet proxy records, 50 ice-sheet evolution in Ross Sea region, 122–126 sea level changes, 46–47 Seymour Island fossil flora, 330 Shelf aggradation, 145, 147 Shelf-ice drilling future prospective, 64 to study history of Antarctic Cenozoic Glaciation, 59–63 Ship-based drilling CIROS Project, 47–48 future prospective of, 64–66 to study Antarctic Cenozoic glaciation, 40–43 Siliceous biogenic facies, 149 Single-channel seismic (SCS) surveys, of Ross sea, 119 Siple coast, of Ross Sea, 537 Sirius Group, 7, 404–405 glacial erosional features, 407 Middle Miocene to Pliocene history, 406–410 sedimentary characteristics, 407 stratigraphy, 408 Skelton Glacier area, 270 Sørsdal Glacier, 541 Southern Annular Mode (SAM), 104 Southern Ocean circulation, 97–103 Antarctic circumpolar current, 97–99
591
subpolar circulation, 99–100 thermohaline circulation, 100–103 climate change/warming, 103–107 oceanographic elements of, 87 salinity trends of, 105–106 water mass formation and dispersal, 88–97 South Orkney Islands region, 155–156 South polar seasonal temperatures, sea ice and winds in response to a growing ice sheet, GCM simulation, 388 South Shetland Islands, 275 stratigraphy, 313–314 South Shetland Islands region Bransfield Strait, 156–157 scientific drilling in, 158 South Western Ross Sea climate data, Oligocene–Miocene Boundary, 381–385 SPECMAP reconstruction, 553 Sporomorphs, in Tasman Sea relative abundance, 335 S3/S2 boundary, 159 Stratigraphic correlation of Antarctic sedimentary strata with other Gondwana continents, 38–39 of Ross Sea area, 121 Stratigraphy continental shelf glacial, Wilkes Land, 130–131 Wilkes Land Acoustic, 127–133 Stratigraphy, seismic Antarctic Ice Sheet proxy records, 50 Strike–slip fault system in Victoria and Oates Lands, 279 Structural framework of Ross Sea, 119
592
Subject Index
Subantarctic mode water (SAMW), 90–92, 491 Subglacial lakes location of, 247–248 sediments deposites, 63 Subpolar circulation, 99–100 Summer insolation, NH, 467 Supercontinent, Gondwana, 265–276 Surface ocean currents, 497 (see also Glacial-interglacial cycles) Tasmanian Gateway climate evolution, 346 Tasman Sea sporomorphs, relative abundance, 335 Taylor Dome, 534 Taylor Dry Valley, 43 Tectonic evolution in Ross Sea Sector, during Cenozoic, 279–284 of Transantarctic Mountains, 284–287 Tectonic isolation influence across Oligocene–Miocene boundary, 385–389 Tectonic stages of Gondwana break-up, 276–279 Terrestrial environments, Middle Miocene to Pliocene east, 404–417 west, 417–424 Terrestrial realm climate records from, 325–335 Terrestrial volcanic outcrops in West Antarctica distribution of Neogene (15-2 Ma), 413 Tethyan closure, 389
Thermohaline circulation (THC), 100–103, 488–490 (see also Glacial-interglacial cycles) Time dependent simulations, 552–554 Transantarctic flora, 336 Transantarctic mountains, 259–260, 542–544 exhumation events, variation, 286 map of, 281 Middle Miocene to Pliocene history, 406–410 new records from, 63 reinterpretation of climate discoveries in, 56–57 tectonic evolution of, 284–287 wet-based glacial deposits, 48–49 UCDW (see Upper circumpolar deep water (UCDW)) Upper circumpolar deep water (UCDW), 94 Upper Palaeozoic–Mesozoic (c. 450–180 Ma) Gondwana, evolution, 272–276 Vegetation distributions Pliocene, climate modelling, 439 Velocity map of Antarctic Ice Sheet, 68 Vestfold Hills, 542, 549 Victoria Group, 274 Victoria Land coast in Late Oligocene/Early Miocene, 321 Volcanic outcrops of Neogene age in West Antarctica, 418–420 Vostok core, 53 Vostok ice core, 551–552
Subject Index
WAIS (see West Antarctic Ice Sheet (WAIS)) Water mass, formation and dispersal, 88–97 Antarctic bottom water, 94–97 circumpolar deep water, 92–94 Subantarctic mode water and Antarctic intermediate water, 90–92 surface water, 88–90 Weddell–Enderby Basin, 101 Weddell gyre, 99 Weddell Orogeny, 274–275 Weddell Sea, 471–472 Cenozoic climate history, seismic interpretation and drilling data, 144–152 continental margin sediments, 151 environmental change in Miocene sediments, 151 Oligocene sediments, 149 sedimentation rates, 150 glacigenic sediments, acoustic stratification of wedge of Crary Trough, 145, 148 Dronning Maud Land margin and Larsen Shelf, 147 ice-sheet history, 151–152 MCS lines, 145 pre-ice-sheet Cenozoic shelf edge, 148 pre-Oligocene section of inferred turbidites, 149 Weddell Sea Embayment, 548–549 West Antarctica, 259–260 West Antarctica, terrestrial volcanic outcrops distribution of Neogene (15-2 Ma), 413
593
West Antarctic Ice Sheet (WAIS), 118, 404, 536, 572 marine record, Middle Miocene to Pliocene, 429–433 West Antarctic Rift System (WARS), 259 map of, 281 Westernmost Marie Byrd Land (Edward VII Peninsula), 271 Western Ross Sea drill cores in, 317–321 West terrestrial environments, Middle Miocene to Pliocene, 417–424 West wind drift, 97 Weyprecht, Karl, 14–17 Whitmore Mountains, 274 Wilkes Land, 539–542 acoustic stratigraphy, 127–133 Cenozoic climate history, seismic interpretation and drilling data, 127–135 drilling on margins of, 133 geomorphic and seismic stratigraphic features of, 163, 165 inferred long-term record of glaciations, 134–135 seismic stratigraphic history, 166 Wilkes Land margin drilling on, 133 location of MCS profiles, 128 Wilkins ice shelf, 572, F13.2 Wind drift, 97 Windmill Islands, 541 WMO (see World Meteorological Organization (WMO)) World Meteorological Organization (WMO), 19, 27
Developments in Earth & Environmental Sciences, 8 F. Florindo and M. Siegert (Editors) r 2009 Elsevier B.V. All rights reserved DOI 10.1016/S1571-9197(08)00013-X
Chapter 13
Concluding Remarks: Recent Changes in Antarctica and Future Research Fabio Florindo1, and Martin Siegert2 1
Istituto Nazionale di Geofisica e Vulcanologia, via di Vigna Murata 605, 00143 Roma, Italy 2 School of GeoSciences, Grant Institute, University of Edinburgh, The King’s Buildings, West Mains Road, Edinburgh EH9 3JW, UK
ABSTRACT Antarctic Climate Evolution has been, and will be, hugely influential in the development of Earth’s environment. This book has detailed how Antarctica changed during several key stages in the Cenozoic. Here we take stock of past changes and consider how they may be helpful in evaluating future changes in Antarctica.
Warming of the climate system is unequivocal, as is now evident from observations of increases in global average air and ocean temperatures, widespread melting of snow and ice in Greenland and in the Antarctic Peninsula, and rising global sea levels. Eleven of the 12 years between 1995 and 2006 are among the 12 warmest years in the instrumental record since 1850 (IPCC, 2007). According to the latest IPCC worst-case scenario projections (i.e. continued greenhouse gas emissions at or above current rates), global annual mean temperatures by 2100 are likely to exceed those that have been experienced by the Earth in the last 40 myr, that is before the Antarctic Ice Sheet first developed (IPCC, 2007). If warming continues, the implication is that ice loss to the ocean may far outweigh gain from ice accumulation, with huge implications for global sea level (and associated Corresponding author. Tel.: þ39 0651860 383; Fax: þ39 0651860 397;
E-mail: fl
[email protected] (F. Florindo).
572
F. Florindo and M. Siegert
Figure 13.1: Ice-sheet surface elevation changes, recorded by satellite altimetric measurements, between 1992 and 2003. Note how ice loss is concentrated at the margins of ice streams in the Amundsen Bay region of East Antarctica, and the Totten and Cook glaciers in East Antarctica. These regions of ice loss are characterised by ice resting on a bed suppressed several hundred metres below sea-level, implying an ice-ocean connection to the current retreat (from Shepherd and Wingham (2007), with permission of the authors). major changes in coastlines and the inundation of low-lying areas), atmospheric composition and dynamics, and ocean circulation. Recent comprehensive, and near continuous, satellite altimetric observations indicate that the Antarctic Ice Sheet is losing mass, and that the rate of loss has increased steadily during the last few decades (Fig. 13.1, e.g. Davis et al., 2005; Shepherd and Wingham, 2007; Rignot et al., 2008). Most loss is from the West Antarctic Ice Sheet (WAIS), especially along the Bellingshausen and Amundsen seas (132760 Gt yr1 in 2006). During the past ca. 40 years many glaciers, and a series of ice-shelves, have retreated in the Antarctic Peninsula (e.g. Cook et al., 2005) and, since the new millennium, two large ice shelves, located either side of the Antarctic Peninsula (the Larsen B and Wilkins ice shelves, with extents of 3,250 km2 and over 570 km2, respectively; Fig. 13.2), have suddenly broken up in dramatic fashion and separated from the continent. The decay of these ice shelves will not have affected sea level because ice shelves are afloat and thus already displacing their weight of water. However, ice-shelf decay has the potential to greatly reduce buttressing forces on grounded ice that resist the flow from the continent to the ocean. In other words, ice-shelf decay may encourage further loss of grounded ice, which will affect global sea level. A number of factors contribute to ice-shelf collapse, including atmospheric and oceanic
Recent Changes in Antarctica and Future Research
573
Figure 13.2: High-resolution, enhanced-colour image of the Wilkins Ice Shelf in Antarctica, taken on 8 March 2008 by Taiwan’s Formosat-2 satellite (Credit: left, National Snow and Ice Data Center; right, National Snow and Ice Data Center/Courtesy Cheng-Chien Liu, National Cheng Kung University (NCKU), Taiwan and Taiwan’s National Space Organization (NSPO); processed at Earth Dynamic System Research Center at NCKU, Taiwan).
temperatures. If these continue to rise, we may see more ice shelves decaying across the Antarctic Peninsula and elsewhere in Antarctica (e.g. Morris and Vaughan, 2003; MacAyeal and others, 2006; Glasser and Scambos, 2008). It is worth noting that, over the past 50 years, the western part of the Antarctic Peninsula has experienced the greatest temperature increase on Earth, rising by nearly 31C (e.g. King et al., 2003; Turner et al., 2005). This is approximately 10 times the mean rate of global warming, as reported by the IPCC. The temperature increase is not confined to the Antarctic Peninsula; it is also occurring in and over the oceans surrounding Antarctica. For example, it is now well-established that the waters of the Antarctic Circumpolar Current (ACC) are warming more rapidly than the global ocean as a whole. A comparison of temperature measurements from the 1990s with data from earlier decades shows a large-scale warming of around 0.21C in the ACC waters at around 700–1,100 m depth (Gille, 2002). Across the East Antarctica Ice Sheet (EAIS) the mass balance is near a steady state with the flow of ice. Its interior is increasing in overall thickness, which is probably the result of increased ice accumulation due to warmer (and moister) atmospheric conditions over the continental interior (Church et al., 2001). This mass gain is balanced by ice thinning/loss across the potentially dynamic marine sectors in Wilkes Land (Davis et al., 2005). Further atmospheric and ocean warming may be incompatible with
574
F. Florindo and M. Siegert
sustaining mass balance in this region, implying it too is susceptible to further changes in a warming world. A greater understanding of past changes in this region of East Antarctica is clearly warranted. Comprehending glacial history since the Last Glacial Maximum (LGM), which culminated around 20,000 years ago, is also critical to knowing whether recent changes are a result of human activities and/or a result of icesheet relaxation since the LGM. The geological record tells us that the WAIS retreat occurred during the Holocene, far later than Northern Hemisphere ice sheets. It is certainly possible that changes observed in some locations today can be at least part explained by post-LGM ice retreat to a more stable interglacial condition. Geological evidence on the rate of ice loss since the LGM is therefore needed at several sites across the Antarctic continent. Understanding Antarctic climate evolution requires continent-wide as well as continent-to-deep-sea studies of past climate records, to decipher their separate but related histories. Some of these activities are long-term undertakings but commitment within the frame of IPY is essential to laying the groundwork for long-term success. Over the next few years ACE aims to integrate geoscience data provided by sediment cores (ANDRILL, andrill.org; SHALDRILL, www.shaldril.rice.edu; IODP, www.iodp.org) and ice cores (EPICA, www.esf.org; ITASE, www2.umaine.edu/itase/; TALDICE; WAIS Divide, www.waisdivide.unh.edu/) with the new generation of climate models that couple atmosphere, ocean, ice sheet and sediment modelling on key time periods from the distant past (i.e. tens of millions of years ago when global temperature was several degrees warmer than today) to the recent past (i.e. during the Holocene, prior to anthropogenic impacts as well as at the LGM). Acquisition of sedimentary records from the floors of subglacial lakes is also anticipated in the next few years. Such information may inform us about ice-sheet histories, and when West Antarctica was last deglaciated. Drilling, sampling and studying subglacial lakes remotely and without causing their contamination represent a considerable technological challenge. In this regard the exploration of sub-ice lakes represents a good analogue for the exploration of planets and satellites such as Europa. A number of subglacial lakes have been identified in Antarctica and, of these, one subglacial lake in West Antarctica, named Lake Ellsworth (from the American explorer Lincoln Ellsworth), is well suited to exploratory research (www.geos.ed.ac.uk/ellsworth). This book has demonstrated that the Antarctic Ice Sheet has undergone many changes, and has varied in size considerably over the Cenozoic. The nature of these changes has been shown to be associated with global climate conditions, forced externally or through interaction with ice and climate
Recent Changes in Antarctica and Future Research
575
conditions in Antarctica and the Southern Ocean. IPCC predictions of atmospheric greenhouse gases are pertinent to future ice volumes in Antarctica. If such gases continue to rise, in a few centuries time the value of atmospheric carbon dioxide may be greater than at any time in the Cenozoic. The obvious danger is that, in a warming world, the Antarctic Ice Sheet may respond to climate and ocean changes as it has done in the past. In other words, the palaeo-ice-sheet reconstructions highlighted in this book may be relevant to assessments of future changes in Antarctica. Over the next decade, ACE will be pursuing a broad range of objectives to better comprehend past Antarctic changes, through the organisation of workshops, where interdisciplinary research can be discussed, and through publications (e.g. Florindo et al., 2003, 2005, 2008; Barrett et al., 2006), allowing dissemination of results to a wide audience. It is only through such integration of geological data and numerical modelling that quantitative assessments of past changes, and possible future scenarios, can be achieved.
REFERENCES Barrett, P. J., Florindo, F., & Cooper, A. K. (Eds) (2006). Cenozoic paleoenvironments: Geologic record and models. Palaeogeogr. Palaeoclimatol. Palaeoecol., 231, 1–252. Church, J. A., Gregory, J. M., Huybrechts, P., Kuhn, M., Lambeck, K., Nhuan, M. T., Qin, D., Woodworth, P. L., Anisimov, O. A., Bryan, F. O., Cazenave, A., Dixon, K. W., Fitzharris, B. B., Flato, G. M., Ganopolski, A., Gornitz, V., Lowe, J. A., Noda, A., Oberhuber, J. M., O’Farrell, S. P., Ohmura, A., Opperheimer, M., Peltier, W. R., Raper, S. C. B., Ritz, C., Russell, G. L., Schlosser, E., Shum, C.K., Stocker, T. F., Stouffer, R. J., Van de Wal, R. S. W., Voss, R., Wiebe, E. C., Wild, M., Wingham, D. J., Zwally, H. J. (2001). Changes in Sea level. In: J. T. Houghton, et al (Eds). Climate Change 2001: The Scientific basis. Cambridge University Press, New York, pp. 639–694. Cook, A. J., Fox, A. J., Vaughan, D. G., & Ferrigno, J. G. (2005). Retreating glacier fronts on the Antarctic Peninsula over the past half-century. Science, 308, 541–544. Davis, C. H., Li, Y., McConnell, J. R., Frey, M. M., & Hanna, E. (2005). Snowfall driven growth in East Antarctic Ice Sheet Mitigates recent sea-level rise. Science, 308, 1898–1901. Florindo, F., Cooper, A. K., & O’Brien, P. E. (Eds) (2003). Antarctic Cenozoic palaeoenvironments: Geologic record and models. Palaeogeogr. Palaeoclimatol. Palaeoecol., 198(1–2), 1–278. Florindo, F., Harwood, D. M., & Wilson, G. S. (Eds) (2005). Long term changes in southern high-latitude ice sheets and climate, the Cenozoic history. Glob. Planet. Change, 45(1–3), 1–260.
576
F. Florindo and M. Siegert
Florindo F., Nelson A. E., & Haywood A. M. (Eds) (2008). Antarctic cryosphere and Southern Ocean climate evolution (Cenozoic–Holocene). Palaeogeogr. Palaeoclimatol. Palaeoecol., 260(1–2), 1–298. Gille, S. T. (2002). Warming of the Southern Ocean since the 1950s. Science, 295, 1275–1277. Glasser, N. F., & Scambos, T. A. (2008). A structural glaciological analysis of the 2002 Larsen B ice shelf collapse. J. Glaciol., 54(184), 3–16. IPCC (2007). Climate change 2007: Mitigation of climate change. In: B. Metz, O. R. Davidson, P. R. Bosch, R. Dave, & L. A. Meyer (Eds). Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, 851 pp. King, J. C., Turner, J., Marshall, G. J., Connolley, W. M., & Lachlan-Cope, T. A. (2003). Antarctic Peninsula climate variability and its causes as revealed by analysis of instrumental records. In: E. Domack, A. Leventer, A. Burnett, R. Bindschadler, P. Convey, & M. Kirby (Eds). Antarctic Peninsula Climate Variability: Historical and Paleoenvironmental Perspectives. Antarctic Research Series 79. American Geophysical Union, Washington, DC, pp. 17–30. MacAyeal, D., Okal, E., Aster, R., Bassis, J., Brunt, K., Cathles, L., Drucker, R., Fricker, H., Kim, Y.-J., Martin, S., Okal, M., Sergienko, O., Sponsler, M., & Thorn, J. (2006). Transoceanic wave propagation links iceberg calving margins of Antarctica with storms in tropics and Northern Hemisphere. Geophys. Res. Lett., 33(17), L17502, doi:10.1029/2006GL027235. Morris, E. M., & Vaughan, D. G. (2003). Spatial and temporal variation of surface temperature on the Antarctic Peninsula and the limit of viability of ice shelves. In: E. Domack, A. Leventer, A. Burnett, R. A. Bindschadler, P. Convey, & M. Kirby (Eds). Antarctic Peninsula Climate Variability-Historical and Paleoenvironmental Perspectives. Antarctic Research Series 79. American Geophysical Union, Washington, DC, pp. 61–68. Rignot, E., Bamber, J. L., van den Broeke, M. R., Davis, C., Li, Y., van de Berg, W. J., & van Meijgaard, E. (2008). Recent Antarctic ice mass loss from radar interferometry and regional climate modelling. Nat. Geosci., 1, 106–110, doi: 10.1038/ngeo102. Shepherd, A., & Wingham, D. J. (2007). Recent sea-level contributions of the Antarctic and Greenland ice sheets. Science, 315, 1529–1532. Turner, J., Colwell, S. R., Marshall, G. J., Lachlan-Cope, T. A., Carleton, A. M., Jones, P. D., Lagun, V., Reid, P. A., & Iagovkina, S. (2005). Antarctic climate change during the last 50 years. Int. J. Climatol., 25, 279–294.