Boston University OpenBU http://open.bu.edu

Theses & Dissertations Boston University Theses & Dissertations

2019 Pleistocene records of ice sheet processes and glacial history from Antarctica and Greenland

https://hdl.handle.net/2144/39515 Boston University BOSTON UNIVERSITY

GRADUATE SCHOOL OF ARTS AND SCIENCES

Dissertation

PLEISTOCENE RECORDS OF ICE SHEET PROCESSES AND

GLACIAL HISTORY FROM ANTARCTICA AND GREENLAND

by

ANDREW JAMES CHRIST

B.A., Hamilton College, 2011

Submitted in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

2019

© 2019 by ANDREW JAMES CHRIST All rights reserved except for chapter 2, which is © 2019 by the Geological Society of America

Approved by

First Reader ______

Paul R. Bierman, Ph.D. Professor of Geology University of Vermont

Second Reader ______

Christine Regalla, Ph.D. Assistant Professor of Earth & Environment

Third Reader ______

Sergio Fagherazzi, Ph.D. Professor of Earth & Environment

Fourth Reader ______

Duncan M. FitzGerald, Ph.D. Professor of Earth & Environment

DEDICATION

I dedicate this dissertation to my family.

iv ACKNOWLEDGMENTS

The pursuit – and completion – of this degree would have been impossible without the support of many along this journey. First, I thank the National Science Foundation and the Boston University College of Arts & Sciences for financially supporting my graduate education. I am grateful for the administrative staff in the Department of Earth &

Environment at BU and the Department of Geology at University of Vermont, who make things happen. I thank Joel Sparks for his laboratory assistance and patience. I thank Mark

Friedl for assisting my transition to UVM. Alissa Beideck always ensured that I was supported through challenges along the way.

I am grateful to my collaborators who spent time running analyses, editing drafts, teaching me in the field, and mentoring me: Marc Caffee, Lee Corbett, Eugene Domack,

Julie Fosdick, Brenda Hall, Alan Hidy, Paul Knutz, Jennifer Lamp, Amy Leventer, David

Marchant, Robert Michener, Joerg Schaefer, Kate Swanger, and Elizabeth Thomas. I thank my committee members, Duncan FitzGerald, Sergio Fagherazzi, and Jeremy Shakun for their service.

I owe my life to the staff at McMurdo Station. Helicopter pilots, Keith and Scott, and Search and Rescue teammates, Paul, Matt, and Abe, risked their lives to extract us from an extraordinarily dangerous situation on Mount Discovery. Bija Sasse organized a flawless operation while keeping me calm over the radio. If it were not for these people, I may never have returned from Antarctica, let alone finish my degree.

I’m thankful to BU Undergraduates Natalie Robinson, Daniel Rybarcyzk, and

Emelia Chamberlain for coming to Antarctica to help me with my research and who,

v despite almost getting blown away on the side of a volcano, were enthusiastic and courageous at every turn. Among the highlights of my graduate education was mentoring

Emelia on her honors thesis.

The graduate students at BU were a great group of people to learn with, commiserate with, and go Live Band Karaoke with, especially: Sam Levy, John Foster,

Luca Morreale, Julia Marrs, Drew Trlica, Sarah Garvey, Sarah Farron, Chloe Anderson,

Donovan Dennis, and Shivani Ehrenfeucht. My friends at UVM – Lee Corbett, Mae Kate

Campbell, and Chris Halsted – welcomed me into the lab to teach me chemistry. I am grateful for Jennifer Lamp, who helped analyze samples and guided me through challenging situations. I’m thankful for the gentlemen of 268 Broadway, Nick Green,

David Goldberg, and Aaron Bembenek, who put up with me (and various subletters) while

I was in Antarctica and made our apartment feel like home.

For my mentors: I thank Christine Regalla, who rushed to help me when I needed help the most, set me on a path towards success, and serves as an excellent role model. My advisor Paul Bierman at UVM has demonstrated a level of generosity that I didn’t know was possible. He inspires me to be the best scientist and person I can be inside and outside of the lab.

For my family: I’m so grateful for my Aunt Sarah and Uncle Jeff, who made home feel not so far away and encouraged me through the most difficult periods of my doctorate.

I am thankful to my Grandma Dorothy for always being moral compass, and to my

Grandma Betty for letting me flood her backyard with a hose and 50-pound bags of sand to ‘learn’ geomorphology hands-on as a kid. I am so grateful for my sister Allie, who I look

vi up to for how to persevere through challenges with a level head and treat others with kindness. My mother, Mary, who taught me to be passionate about my interests, fostered my creative thinking, and gave me a sense of humor. My father, Doug - who defeated cancer in less time than it did to write this dissertation - instilled in me a love of science and the outdoors, critical thinking skills, and strength. Finally, I thank Emily Schottenfels

- the single best thing to happen to me - who always loves me through the lowest lows and highest highs, and inspires me to be a better human.

vii PLEISTOCENE RECORDS OF ICE SHEET PROCESSES AND GLACIAL

HISTORY FROM ANTARCTICA AND GREENLAND

ANDREW JAMES CHRIST

Boston University, College of Arts and Sciences, 2019

Major Professor: Dr. Paul Bierman, Professor of Geology, University of Vermont

ABSTRACT

Anthropogenic climate change is transforming Earth’s Polar Regions. Oceanic and atmospheric warming threaten to melt portions of the Antarctic and Greenland ice sheets and raise sea level dramatically by the end of this century. As a guide for the future, we can examine the geologic record to understand the surface processes that govern ice sheet behavior and establish chronologies of past climate change. In this dissertation, I focus on the Pleistocene glacial history of two polar field sites: the presently ice-free region of

McMurdo Sound, Antarctica and formerly glaciated continental shelf of Melville Bugt,

Greenland. In my research, I use a combination of detailed geologic mapping, geochemical and isotopic analyses of glacial sediments, and several Quaternary geochronologic techniques.

I show that several marine-based ice sheets inundated McMurdo Sound and deposited a series of glacial sediments that record former ice sheet configuration and

Antarctic surface processes. Mapping results and a robust radiocarbon chronology indicate that grounded marine-based ice in the Ross Sea overflowed into McMurdo Sound and attained maximum thickness after the global Last Glacial Maximum; ice receded in

viii response to rising sea level and changing ocean circulation. Cosmogenic nuclide exposure ages of boulders from these glacial deposits do not record a simple history of ice extent, but instead the integrated effects of multiple surface processes operating below, along, and above the Antarctic ice sheets.

I build upon the insights gained from Antarctica to develop a novel approach to analyzing glacial marine diamict, a mixture of grain sizes and lithologies, deposited on the continental shelf by an expanded Greenland Ice Sheet. Using cosmogenic nuclides, low temperature thermochronometry, and lipid biomarkers, I show that the Greenland Ice Sheet during the Early Pleistocene was similar to today. Extremely low cosmogenic nuclide concentrations indicate the ice sheet already eroded pre-glacial soil cover, but thermochronometry data suggest deep fjords had not yet formed. Abundant lipid biomarkers recovered in this diamict indicate that ice-free areas supported vegetation, similar to the modern ice sheet. This dataset demonstrates how the novel application of proven techniques to glacial sediments can unlock new information about ice sheet history and process.

ix TABLE OF CONTENTS

DEDICATION ...... iv

ACKNOWLEDGMENTS ...... v

ABSTRACT ...... viii

TABLE OF CONTENTS ...... x

LIST OF TABLES ...... xiv

LIST OF FIGURES ...... xv

CHAPTER ONE: The need for polar paleoclimate records in a warming world ...... 1

1.1 Introduction ...... 1

1.2 Late Pleistocene glacial history of Ross Sea Embayment, Antarctica ...... 4

1.2.1. Chapter Two: The local Last Glacial Maximum in McMurdo Sound, Antarctica:

implications for ice-sheet behavior in the Ross Sea Embayment ...... 6

1.2.3. Chapter Three: Cosmogenic nuclide exposure age scatter in McMurdo Sound,

Antarctica record Pleistocene glacial history and polythermal ice sheet processes ... 7

1.3 Deep-time proxy records of the Greenland Ice Sheet ...... 9

1.3.1 Chapter Four: A deeply erosive Early Pleistocene Greenland ice sheet ...... 10

CHAPTER TWO: The local Last Glacial Maximum in McMurdo Sound, Antarctica: implications for ice-sheet behavior in the Ross Sea Embayment ...... 13

Abstract ...... 13

2.1 Introduction ...... 14

x 2.1.1 LGM and Deglacial History of the Ross Sea Embayment ...... 16

2.2 Study area...... 20

2.3 Methods...... 23

2.3.1 Glacial Geomorphologic Mapping ...... 23

2.3.2 Ice & Snow Stable Isotope Geochemistry ...... 24

2.3.3 Radiocarbon Dating ...... 26

2.4 Results ...... 27

2.4.1 Spatial Distribution of Ross Sea Drift in Southern McMurdo Sound ...... 27

2.4.2 Drift Limit Sedimentology & Stable Isotope Geochemistry ...... 34

2.4.1 Algae Radiocarbon Chronology ...... 41

2.5 Discussion ...... 45

2.5.1 Ice-marginal Paleoenvironments ...... 45

2.5.2 Timing of the local Last Glacial Maximum in McMurdo Sound ...... 50

2.5.3 Maximum Ice Sheet Configuration in McMurdo Sound ...... 52

2.5.4 Implications for Ice Sheet Interactions & Deglaciation in the Ross Sea

Embayment ...... 59

2.6 Conclusions ...... 62

CHAPTER THREE: Cosmogenic nuclide exposure age scatter in McMurdo Sound,

Antarctica records Pleistocene glacial history and polythermal ice sheet processes ...... 64

Abstract ...... 64

3.1 Introduction ...... 66

3.1.1 The problem of inherited cosmogenic nuclides ...... 66

xi 3.1.2 McMurdo Sound: a unique setting to understand nuclide inheritance ...... 69

3.2. Study area...... 71

3.3 Material and methods ...... 75

3.3.1 Glacial geologic mapping ...... 75

3.3.2 Sample collection ...... 77

3.3.3 Cosmogenic nuclide analyses ...... 77

3.4 Compiled exposure chronology ...... 78

3.4. Results ...... 79

3.4.1 Glacial geologic mapping ...... 79

3.4.2 Cosmogenic nuclide analyses ...... 80

3.4.3 Compiled cosmogenic nuclide ages ...... 81

3.5 Discussion ...... 87

3.5.1 Scattered exposure ages of Ross Sea drift ...... 87

3.5.2 Surface processes contributing to exposure age scatter ...... 88

3.5.3 New insights from inherited nuclides and exposure age scatter ...... 91

3.5.4 Upper Discovery deposit: a maximum glaciation in McMurdo Sound during

MIS 8? ...... 92

3.6 Conclusion ...... 95

CHAPTER FOUR: The northwest Greenland Ice Sheet during the Early Pleistocene was similar to today ...... 97

Abstract ...... 97

4.1 Introduction: a poorly known history of the Greenland Ice Sheet ...... 98

xii 4.2 Rationale and approach ...... 100

4.3 Results ...... 104

4.3.1 Extremely low cosmogenic 10Be & 26Al concentrations...... 104

4.3.2 Old (>160 Ma) detrital apatite (U-Th-Sm)/He ages ...... 110

4.3.3 Well-preserved lipid biomarkers ...... 112

4.4 Discussion ...... 115

4.4.1 Glacial erosion of pre-ice sheet regolith by the Early Pleistocene ...... 115

4.4.2 The absence of deeply carved fjords in Melville Bugt ...... 118

4.4.3 Evidence for Early Pleistocene terrestrial and marine vegetation ...... 119

4.5 Implications...... 121

4.6 Materials and Methods ...... 122

4.6.1 Core recovery ...... 122

4.6.2 Sample preparation ...... 123

4.6.3 Cosmogenic Nuclides ...... 124

4.6.4 Low-temperature thermochronology ...... 127

4.6.5 Lipid biomarkers ...... 128

CHAPTER FIVE: Conclusions...... 130

Final word ...... 132

BIBLIOGRAPHY ...... 134

CURRICULUM VITAE ...... 150

xiii LIST OF TABLES

Table 2.1 Ground ice stable geochemistry ...... 36

Table 2.2: Radiocarbon chronology ...... 42

3 Table 3.1: Hepyroxene exposure ages of Ferrar Dolerite ...... 83

10 Table 3.2: Bequartz exposure ages of granite ...... 84

Table 4.1: Meteoric 10Be extracted from sediment fractions ...... 106

Table 4.2: In situ cosmogenic 10Be in quartz ...... 107

Table 4.3: In situ cosmogenic 26Al in quartz ...... 108

Table 4.4: Apatite (U-Th-Sm)/He Ages ...... 112

Table 4.5: Lipid biomarkers ...... 114

Table 4.6: 10Be blank measurements ...... 126

Table 4.7: 26Al blank measurements ...... 126

xiv LIST OF FIGURES

Figure 1.1: Overview maps of study areas...... 3

Figure 2.1: Overview maps of the Ross Sea Embayment and McMurdo Sound, Antarctica.

...... 15

Figure 2.2: Aerial photographs of key field sites on the volcanic islands and peninsulas of

McMurdo Sound ...... 25

Figure 2.3: Integrated glacial geomorphologic and bedrock geology of McMurdo Sound,

Antarctica ...... 30

Figure 2.4: Glacial geologic maps of key field areas in McMurdo Sound ...... 33

Figure 2.5: Stable isotope measurements of ground ice in McMurdo Sound ...... 38

Figure 2.6: Characteristic soil pit stratigraphy in McMurdo Sound ...... 40

Figure 2.7: Compiled radiocarbon chronology of Ross Sea drift...... 44

Figure 2.8: Schematic of ice-marginal processes during maximum ice sheet extent...... 48

Figure 2.9: Local LGM grounded ice sheet reconstruction in McMurdo Sound...... 54

Figure 2.10: Paleoclimate from 31 kyr to present...... 61

Figure 3.1 Bedrock geology and stratigraphy of Southern , Antarctica ..... 70

Figure 3.2: Compiled cosmogenic nuclide exposure ages and radiocarbon sample locations

in McMurdo Sound...... 74

Figure 3.3 Compiled cosmogenic exposure ages and radiocarbon chronology of Ross Sea

drift...... 76

Figure 3.4: Glacial geology of Mount Discovery...... 82

xv Figure 3.5: Late Pleistocene glacial chronology of McMurdo Sound and paleoclimate since

500 ka...... 86

Figure 4.1: Study area maps of Greenland and Melville Bugt ...... 100

Figure 4.2: Micrographs of sorted diamict fractions...... 104

Figure 4.3: 1.9 Ma decay-corrected cosmogenic 10Be concentrations in diamict...... 105

Figure 4.4: In situ cosmogenic nuclide Monte Carlo simulations...... 109

Figure 4.5: Detrital apatite (U-Th-Sm)/He ages and analytical results...... 111

Figure 4.6: Lipid biomarker results...... 113

Figure 4.7: Synthesis of 10Be measurements from sedimentary archives in Greenland . 117

xvi 1

CHAPTER ONE: The need for polar paleoclimate records in a warming world

1.1 Introduction

Rapid climate change is transforming the cryosphere – including the Antarctic

(Mouginot et al., 2014; Rignot et al., 2011) and Greenland (Fichefet et al., 2003; Rignot et al., 2011) ice sheets – with global consequences. Atmospheric and oceanic warming is accelerating ice sheet thinning (Mouginot et al., 2014), grounding line retreat (Jenkins et al., 2010), and meltwater production (Boning et al., 2016), collectively threatening to disrupt global ocean circulation (Fichefet et al., 2003) and raise sea level (Bamber et al.,

2009; DeConto and Pollard, 2016). Marine-based portions of these ice sheets are dynamic and especially vulnerable to rapid thinning and raising sea level (Joughin et al., 2014). By the end of the 21st century, sea level could rise as much as 2.4 m due to the retreat of the

West Antarctic and Greenland ice sheets (DeConto and Pollard, 2016), posing dire consequences for coastal populations and human civilization at large.

We can examine the geologic record to investigate how ice sheets responded to past warming episodes as an analog for on-going and future climate warming. Extracting paleoclimate information from polar archives, however, remains challenging. The

Antarctic and Greenland ice sheets conceal terrestrial and marine landscapes below thick glacial ice, largely preventing direct geologic investigation of past climate variability.

Erosion from multiple glacial cycles has significantly modified pre-glacial and inter-glacial landscapes, complicating our reconstructions of polar climate during past warm periods.

2

Where climate archives are accessible and intact, robust chronologies of past climate events can be difficult to obtain and to interpret.

This dissertation investigates climate archives that record the past configuration and behavior of the Antarctic and Greenland ice sheets during the Pleistocene. I address several broad topics in polar paleoclimatology, surface processes, and the application of

Quaternary geochronological techniques in Antarctica and Greenland. The dissertation is divided into three principle projects from two polar field sites: McMurdo Sound, Antarctica and Melville Bugt, northwest Greenland (Figure 1.1). Chapters Two and Three comprise research conducted at Boston University. At the end of my 4th year of my graduate training,

David Marchant ceased to be my academic advisor due to a Title IX investigation and his subsequent termination. As a result, I became a Visiting Graduate Student at the University of Vermont (UVM), where I was advised by Professor Paul Bierman. At UVM I started a new project (Chapter Four) focused on glacial marine sediments from Greenland. In the remainder of this chapter, I briefly introduce the background, approach, and principle conclusions of each project.

3

Figure 1.1: Overview maps of study areas

A) Global overview of study areas in this dissertation, including B) Melville Bugt, northwest Greenland, and C) the Ross

Sea Embayment and McMurdo Sound, Antarctica. Base map data: ETOPO1, BedMachine v3, International Bathymetric

Chart of the South Ocean, and International Bathymetric Chart of the Arctic Ocean, and the Reference Elevation Mosaic of Antarctica. GrIS – Greenland Ice Sheet, WAIS – West Antarctic Ice Sheet, EAIS – East Antarctic Ice Sheet.

4

1.2 Late Pleistocene glacial history of Ross Sea Embayment, Antarctica

Approximately 25% of all Antarctic glacier ice drains into the Ross Sea

Embayment, making it the continent’s largest ice drainage basin (Halberstadt et al., 2016).

The Ross Ice Shelf and the marine-based West Antarctic Ice Sheet (WAIS) buttress the outflow of several large outlet glaciers that bisect the and drain interior ice from the East Antarctic Ice Sheet (EAIS). Collapse of the Ross Ice Shelf and retreat of the grounding line in the Ross Sea Embayment could induce accelerated flow and grounding line retreat of EAIS outlet glaciers, causing subsequent ice loss and significant contributions to sea-level rise from the EAIS (DeConto and Pollard, 2016).

The Pleistocene glacial history of the Ross Sea Embayment provides the opportunity to investigate the behavior and interaction between grounded marine-based ice in the Ross Sea and EAIS outlet glaciers. During glacial periods, EAIS outlet glaciers and the WAIS in the Ross Sea Embayment expand and advance northward towards the edge of the continental shelf as a marine-based ice sheet (Anderson et al., 2014; McKay et al.,

2012; Naish et al., 2009); this causes tributary EAIS outlet glaciers in the Transantarctic

Mountains to thicken as much as 1000 m in valley mouths (Denton et al., 1989; Hall et al.,

2013; Stuiver et al., 1981). The marine and terrestrial geomorphic features and sediments left behind following deglaciation preserve information about the configuration and dynamics of marine-based ice sheets and outlet glaciers during Pleistocene glacial- interglacial cycles (Halberstadt et al., 2016; Hall et al., 2013). This information is critical for understanding the Antarctic ice sheets’ response to atmospheric and oceanic warming and rising sea level (Clark et al., 2016).

5

Numerical chronologies of past changes in the Antarctic ice sheets are key to placing them in the context of global paleoclimate, but are complicated by processes unique to Antarctica. Deglacial chronologies from marine sediments are poorly constrained due to complications with marine radiocarbon, namely old, uncertain, and spatially and temporally variable 14C reservoir ages (Andrews et al., 1999; Domack et al., 1999). On land, organic matter available for radiocarbon dating is rare in glacial sediments or the age of the glacial deposit is too old (>50,000 years). As a result, terrestrial glacial chronologies often rely on in situ cosmogenic nuclide surface exposure ages of glacially transported boulders (Bromley et al., 2012; Bromley et al., 2010; Brook et al., 1995; Hall et al., 2017;

Joy et al., 2014; Spector et al., 2017; Storey et al., 2010; Todd et al., 2010). Exposure chronologies commonly yield scattered individual ages that vary widely across a contemporaneously deposited landform (i.e. moraine) and often exceed the expected depositional age. The scatter in these datasets suggests that these boulders carry an inventory of ‘inherited’ cosmogenic nuclides that accumulated during periods of prior exposure on the landscape, but were not subsequently removed by cold-based, non-erosive glaciation (Hein et al., 2014). The geologic processes contributing to nuclide inheritance is not well understood.

McMurdo Sound is an ideal field area to develop Antarctic glacier chronologies because it harbors well-preserved Last Glacial Maximum (LGM) sediments that can be dated using both radiocarbon ages of organic matter in ice-marginal sediments and cosmogenic nuclide exposure ages of glacial boulders. Cosmogenic nuclide exposure ages can be compared against this independent radiocarbon chronology to investigate the

6 distribution and sources of nuclide inheritance and scatter. With the patterns and possible processes responsible for nuclide inheritance, we can better evaluate exposure ages of older glacial deposits.

1.2.1. Chapter Two: The local Last Glacial Maximum in McMurdo Sound, Antarctica: implications for ice-sheet behavior in the Ross Sea Embayment

In Chapter 2, I reconstruct the former ice sheet extent and flow patterns in

McMurdo Sound during the LGM (Christ and Bierman, 2019). I integrate detailed geologic mapping and sedimentological descriptions of terrestrial landforms with the glacial marine geomorphology (Greenwood et al., 2018; Halberstadt et al., 2016; Lee et al., 2017) of the western Ross Sea to support this ice sheet reconstruction. I anchor the ice sheet history with a well-constrained radiocarbon chronology for landforms of previously unknown age, which can be correlated with other well-dated LGM terrestrial deposits in the region

(Denton et al., 1989; Hall and Denton, 2000; Hall et al., 2015; Jackson et al., 2018; Kellogg et al., 1990; Stuiver et al., 1981).

These data suggest that a marine-based ice sheet overflowed from the western Ross

Sea into McMurdo Sound and reached its local maximum thickness and extent between

19.6 to 12.3 calibrated thousands of years before present (cal. ka). The maximum elevation of LGM glacial deposits consistently declines in elevation from east to west across the volcanic islands of McMurdo Sound towards the coastal ice-free valleys of the Royal

Society Range; this suggests ice flow was from east to west with only minor contribution from local alpine glaciers. Calibrated radiocarbon ages of algae in LGM sediments from

7 the volcanic islands closely agree with similar regional datasets (Hall and Denton, 2000;

Hall et al., 2015; Jackson et al., 2018; Stuiver et al., 1981).

These data confirm that the local LGM in the western Ross Sea was delayed relative to the global LGM. While this dataset does not bear directly on ice thinning rates or deglaciation in McMurdo Sound, grounded ice at or near its maximum thickness persisted through and after Meltwater Pulse 1A (14.0 – 14.5 ka), a period of rapid sea level rise

(Clark et al., 2016; Lambeck et al., 2014). This suggests that this sector of the Antarctic

Ice Sheet was not a significant contributor to that sea level rise event.

1.2.3. Chapter Three: Cosmogenic nuclide exposure age scatter in McMurdo Sound,

Antarctica record Pleistocene glacial history and polythermal ice sheet processes

Building upon the mapping results and radiocarbon chronology from Chapter 2

(Christ and Bierman, 2019), I examine the issues of scatter and nuclide inheritance in exposure chronologies of glacial sediments in McMurdo Sound. I present a compilation of new1 and previously published (Anderson et al., 2017; Brook et al., 1995) cosmogenic nuclide surface exposure ages of glacially transported boulders in LGM sediments in

McMurdo Sound. I compare the scatter in the cosmogenic exposure ages against the independently constrained radiocarbon chronology of Ross Sea drift established in the previous chapter (Hall and Denton, 2000; Hall et al., 2015; Jackson et al., 2018; Stuiver et al., 1981). Using this information, I identify surface processes that enhance, prevent, or reduce the likelihood of inherited nuclides in glacially transported material in Antarctica. I

1 Samples were analyzed through collaboration with Joerg Schaefer, Gisela Winckler, and Jennifer Lamp in the Cosmogenic Nuclide Dating Group at Lamont-Doherty Earth Observatory, Columbia University.

8 then present a set of new exposure ages from the highest elevation glacial deposit on Mount

Discovery and discuss its possible age and context within global paleoclimate.

Exposure ages of glacial boulders on moraines and boulders from McMurdo Sound do not yield simple ages of emplacement on an ice margin; instead, they record the variety of surface processes operating on the landscape. Exposure ages of LGM glacial sediments are highly scattered and a majority of them exceed the time frame of maximum ice extent between 12.3 and 19.6 cal. ka established in Chapter 2. Exposure ages sort according lithology suggesting that bedrock source and glacial entrainment mechanisms and pathways explains the magnitude and patterns of nuclide inheritance. Clasts entrained via sub-glacial plucking below wet-based, erosive outlet glaciers may yield younger exposure ages that imply a lack of prior exposure or removal of inherited nuclides. Rocks that are exposed sub-aerially, either on cliff faces above glacier margins or on the volcanic landscape of McMurdo Sound, yield older exposure ages; this suggests that inherited nuclides are not removed during subsequent transport by cold-based, non-erosive glaciers, and possibly reworked across multiple glacial cycles.

While the exposure ages from LGM sediments do not provide precise ages of ice position, they tend to be within the last glacial period; therefore, it is possible to use exposure ages for assigning depositional ages at glacial/interglacial timescales. With the magnitude and sources of scatter in mind, I evaluate a new set of exposure ages of pre-

LGM glacial sediments on Mount Discovery that suggest maximum glaciation of

McMurdo Sound occurred during Marine Isotope Stage 8, a glacial period marked by warmer climate and higher sea level (Spratt & Lisiecki, 2016;.

9

1.3 Deep-time proxy records of the Greenland Ice Sheet

In Chapter Four, I transition from Antarctica to Greenland, where I use cosmogenic nuclides and other methods to investigate climate archives of polar ice sheet processes at longer timescales. The Greenland Ice Sheet (GrIS) may respond more quickly to on-going and future climate warming than the Antarctic ice sheets, contributing up to 0.34 m by the end of this coming century (Mengel et al., 2016). It is important to examine the behavior of the GrIS during earlier warmer-than-present climates, yet the history of the GrIS remains fragmentary (Bierman et al., 2016).

Much of the known deep time history of Greenland relies upon marine sediment cores collected in the north Atlantic Ocean (Bierman et al., 2016; Helland and Holmes,

1997; Reyes et al., 2014; St John and Krissek, 2002; Tripati et al., 2008). Most deep marine sediments are composed primarily of fine-grained material with only occasional clasts and sand transported by icebergs. The scarcity of sand-sized (and larger) grains and minerals in these types of sediment limits the application of chemical and isotopic analyses used in terrestrial environments such as cosmogenic radionuclides (Granger et al., 2013) and low- temperature thermochronology (Reiners and Ehlers, 2005).

Glacial marine diamict – a poorly sorted mixture of sediment of all grain sizes – deposited on the continental shelf remains a largely unexplored and promising archive of information. Historically, diamict has been under-utilized as a climate archive (relative to deep marine records) because it can be difficult to recover by drilling, to interpret stratigraphically, and to date. But, unlike ice-rafted debris contained in deep-marine

10 sediments, diamict often contains abundant sand-sized grains and a mixture of minerals allowing for the application of numerous and varied analytical techniques.

1.3.1 Chapter Four: A deeply erosive Early Pleistocene Greenland ice sheet

In this chapter, I apply a novel combination of proven analytical techniques – cosmogenic 10Be and 26Al (in situ and meteoric), apatite low temperature (U-Th)/He thermochronology, and leaf waxes – in a section of 1.8 – 2.0 Ma (Early Pleistocene) diamict in Core 344S-0110 recovered in Melville Bugt, west Greenland (made available to me by

Dr. Paul Knutz at the Geological Survey of Denmark and Greenland). This approach takes advantage of the diversity and abundance of grain sizes and minerals contained in diamict to apply analyses that are neither possible nor practical in fine-grained, ice-distal sediment cores. This unique combination of analyses enables more illustrative interpretations of the northwestern GrIS during the Early Pleistocene.

Each method provides information about ice sheet process and history. The concentration of cosmogenic nuclides in rocks and sediments, both produced in the atmosphere (meteoric) and in mineral surfaces (in situ), reveals near-surface landscape histories of exposure, erosion, and burial (Lal, 1998). Under ice-free conditions, meteoric

10Be accumulates in soils and in situ 10Be is produced in rock surfaces. Sub-glacial erosion excavates into sub-surface material with low 10Be concentrations, while re-exposure of the land surface during periods with reduced ice cover increases the concentration of 10Be in surface materials (Bierman et al., 2016; Shakun et al., 2018). Low-temperature (U-Th)/He thermochronometry of detrital apatite mineral grains is used to resolve the thermal history of the upper 1-3 km of the Earth’s crust (Reiners, 2005). In glacial environments, (U-

11

Th)/He ages can resolve the depth of glacial denudation that can alter the geothermal gradient. Terrestrial flora produce long chain hydrocarbons, known as leaf waxes, that form a protective coating on plant leaves (Eglinton and Hamilton, 1967) that are abraded off of leaves and then transported, deposited, and preserved in sedimentary archives (Castaneda and Schouten, 2015; Eglinton and Eglinton, 2008). The abundance and type of different long-chain hydrocarbons reveals information about past terrestrial and marine vegetation in Greenland.

Primary results from this study suggest that by the Early Pleistocene, northwest

Greenland was similar today with an erosive ice sheet surrounded by vegetated ice-free areas. Extremely low cosmogenic isotope concentrations in this sediment, regardless of grain size or shape, indicate persistent ice cover and tens of meters of glacial erosion that had already removed the pre-glacial soil cover and regolith by 2.0 Ma. The absence of young (U-Th-Sm)/He ages indicate that glacial incision did not achieve sufficient depth to change the geothermal gradient of bedrock source areas, suggesting that glacial valleys and fjords greater than 1.3 km depth had not yet formed. Leaf waxes produced by land plants and marine algae indicate periods or areas of less than complete ice cover. These data suggest that by 1.8 to 2.0 million years ago, the Greenland Ice Sheet was a persistent, dynamic, erosive feature of Earth’s climate and landscape.

This novel approach and combination of analyses suggests that glacial marine diamict contains an as of yet un-tapped wealth of information about ice sheet history.

Multi-parameter analyses of diamict demonstrate its utility as an archive of past glacial activity in presently deglaciated margins, particularly where glacial stratigraphic units are

12 exposed at shallow depths below the seafloor. Diamict analysis has the potential to improve assessment of long-term changes in glacial erosion, ice cover, and vegetation throughout the polar regions. This unique approach could be applied to existing sediment core collections and future core drilling campaigns in glaciated margins.

13

CHAPTER TWO: The local Last Glacial Maximum in McMurdo Sound,

Antarctica: implications for ice-sheet behavior in the Ross Sea Embayment2

Abstract

During the Last Glacial Maximum (LGM), a grounded ice sheet filled the Ross Sea

Embayment in Antarctica and deposited glacial sediments on volcanic islands and peninsulas in McMurdo Sound and coastal regions of the Transantarctic Mountains. The flow geometry and retreat history of this ice are debated, with contrasting views yielding divergent implications for the interaction between and stability of the East and West

Antarctic ice sheets during late Quaternary time. Here, we present terrestrial geomorphologic evidence and reconstruct former ice-marginal environments, ice sheet elevations, and ice-flow directions in McMurdo Sound. Fossil algae in ice-marginal sediments provide a coherent radiocarbon chronology of maximum ice extent and deglaciation. We integrate these data with marine records to reconstruct grounded ice dynamics in McMurdo Sound and the western Ross Sea. The combined dataset suggests ice flow toward the Transantarctic Mountains in McMurdo Sound during peak glaciation, with thick, grounded ice at or near its maximum position between 19.6 and 12.3 ka.

Persistent grounded ice in McMurdo Sound and across the western Ross Sea after

Meltwater Pulse 1a (14.0-14.5 ka) suggests that this sector of Antarctica did not significantly contribute to this rapid sea-level rise event. Our data show no significant

2 This chapter is published: Christ, A.J., Bierman, P.R., 2019. The local Last Glacial Maximum in McMurdo Sound, Antarctica: implications for ice sheet behavior in the Ross Sea Embayment. Geological Society of America Bulletin. https://doi.org/10.1130/B35139.1

14 advance of locally derived ice from the Transantarctic Mountains into McMurdo Sound during the local LGM.

2.1 Introduction

Grounded marine-based portions of the Antarctic ice sheets are susceptible to rapid ice retreat and consequently contribute to sea-level rise (Bamber et al., 2009; DeConto and

Pollard, 2016). Recent glaciological observations of the West Antarctic Ice Sheet (WAIS) in the Amundsen Sea sector indicate ongoing ice sheet thinning, grounding line retreat, and accelerated ice flow (Helm et al., 2014; Joughin et al., 2014; Mouginot et al., 2014). These processes could lead to collapse of a large portion of the WAIS (Feldmann and Levermann,

2015) and sea-level rise >1 m by 2100 C.E. (DeConto and Pollard, 2016). Documenting the most recent ice-sheet expansion and subsequent retreat to modern ice configurations across the Antarctic continent can help predict future ice-sheet dynamics.

In the Ross Sea Embayment, the Ross Ice Shelf and the contiguous portion of the

WAIS buttress the outflow of several large outlet glaciers that bisect the Transantarctic

Mountains and drain interior ice from the East Antarctic Ice Sheet (EAIS) (Figure 2.1A).

Collapse of the Ross Ice Shelf and retreat of the grounding line in the Ross Sea Embayment may have occurred in the geologic past during “super inter-glacial” periods (i.e., Marine

Isotope Stages 11 and 31) (McKay et al., 2012; Naish et al., 2009; Scherer et al., 2008). A similar ice retreat scenario in the future could induce accelerated flow and grounding line retreat of EAIS outlet glaciers, causing subsequent ice loss and significant contributions to sea-level rise from the EAIS. Future collapse of the Ross Ice Shelf and WAIS has been

15

modelled under the worst-case climate scenario (RCP 8.5) for future CO2 emissions

(DeConto and Pollard, 2016). Paleoclimate modelling and proxy evidence suggest similar collapses for pre-Quaternary times, especially during the Pliocene, when global average temperatures were just ~3°C above present values (Haywood et al., 2013;

Naish et al., 2009; Pagani et al., 2011).

Figure 2.1: Overview maps of the Ross Sea Embayment and McMurdo Sound, Antarctica.

A) Ross Sea Embayment shown with reconstructed grounding line positions (Bentley et al., 2014) and the modern grounding line. B) The McMurdo Sound region showing the extent of Ross Sea drift in McMurdo Sound (yellow line) and locations discussed in text: CC – Cape Crozier, HH – Hjorth Hill, BP – Brown Peninsula, BI – Black Island, KG –

Koettlitz Glacier, MD – Mount Discovery, MBf – Minna Bluff, MM – Mount Morning, and WI – White Island. Hillshade base image is from IBSCO 2013, Landsat Image Mosaic of Antarctica, and (Howat et al., 2019).

16

The Last Glacial Maximum (LGM) and deglacial history of the Ross Sea

Embayment provide the opportunity to investigate the behavior and interaction between grounded marine-based ice in the Ross Sea and EAIS outlet glaciers. Here, we focus on the glacial sedimentology and geomorphology (e.g. moraines, limits of different glacial sediments) of the volcanic islands and peninsulas of McMurdo Sound (78°S) where southerly EAIS outlet glaciers, not local glaciers in the McMurdo Dry Valleys, fed a marine-based ice sheet in the western Ross Sea. We reconstruct former maximum ice extent and ice flow patterns during the LGM. A detailed sedimentological understanding of terrestrial glacial sediments integrated with glacial marine geomorphology (i.e. mega- scale glacial lineations, sub-glacial channels, grounding zone wedges) in the western Ross

Sea supports our ice sheet reconstruction. We anchor the ice sheet history with a well- constrained radiocarbon chronology for landforms of previously unknown age, which can be widely correlated with other well-dated LGM terrestrial deposits in the region. We discuss the implications of this regional ice sheet reconstruction in the wider context of

Antarctic contributions to deglacial sea-level rise and the fundamental behavior of the

Antarctic ice sheets.

2.1.1 LGM and Deglacial History of the Ross Sea Embayment

During the LGM, grounded ice in the Ross Sea Embayment expanded northward towards the edge of the continental shelf (Figure 2.1A) (Anderson et al., 2014; Bentley et al., 2014; Conway et al., 1999; Hall et al., 2013). Marine evidence for this expansion includes multi-beam swath bathymetry data that identify a suite of sub-glacial, ice- marginal, and pro-glacial geomorphic features (Greenwood et al., 2012; Halberstadt et al.,

17

2016; Shipp et al., 1999; Wellner et al., 2001), and shallow sediment cores that record the deglacial transition from sub-glacial to sub-ice shelf to open marine conditions (Domack et al., 1999; Licht et al., 1996; McKay et al., 2016; Prothro et al., 2018). Terrestrial data for past expansion includes erratic clasts perched at high elevations on nunataks (Ackert et al., 2007; Jones et al., 2015; Stone et al., 2003), and unweathered glacial sediments above present outlet glacier margins (Bromley et al., 2012; Bromley et al., 2010; Hall et al., 2017;

Joy et al., 2014; Spector et al., 2017; Todd et al., 2010) and along coastal ice-free areas

(Denton and Marchant, 2000; Hall et al., 2004; Hall and Denton, 2000; Hall et al., 2015;

Hall et al., 2000). Multiple hypotheses have emerged to explain the mechanisms and style of grounded ice expansion and retreat of the Antarctic ice sheets during and following the

LGM, particularly within the western Ross Sea and McMurdo Sound region (Ackert, 2008;

Anderson et al., 2017; Bart and Cone, 2012; Conway et al., 1999; Denton et al., 1989;

Denton and Marchant, 2000; Greenwood et al., 2018; Halberstadt et al., 2016; Hall and

Denton, 2000; Hall et al., 2015; Hall et al., 2013; Jackson et al., 2018; Lee et al., 2017;

McKay et al., 2016; Stuiver et al., 1981; Wilson, 2000).

Two hypotheses posit coeval expansion of the WAIS and the EAIS towards the edge of continental shelf, but diverge in their explanations for the mechanisms driving advance and retreat of East Antarctic ice in the Ross Sea. One hypothesis suggests that the expansion of WAIS buttressed the lower reaches of outlet glaciers in the Transantarctic

Mountains, allowing East Antarctic ice to thicken, ground, and expand into the Ross Sea

(Conway et al., 1999; Denton and Marchant, 2000; Hall et al., 2015; Stuiver et al., 1981).

Grounded ice in the Ross Sea Embayment isolated the EAIS plateau from ocean moisture

18 sources, reducing ice accumulation rates leading to negligible (potentially negative) changes in the EAIS plateau elevation (Denton and Marchant, 2000; Hall et al., 2015;

Stuiver et al., 1981). Under this scenario, East Antarctic ice expansion in the Ross Sea was dependent upon the expansion of grounded ice in the Ross Sea from West Antarctica.

A second, slightly different, but not mutually exclusive, hypothesis calls for the expansion of EAIS outlet glaciers along the of northern Victoria Land into the

Ross Sea during the LGM, an expansion that caused ice to persist at an extended position into the Early Holocene (Lee et al., 2017; McKay et al., 2016). This hypothesis hinges on the interpretation of west-east trending mega-scale glacial lineations in multi-beam bathymetry data that that indicate eastward ice flow from the Transantarctic Mountains, and undated back-stepping grounding zone wedges that represent the retreat of expanded outlet glaciers during the Early Holocene. Outlet glacier behavior during the during the

Early Holocene may be controlled by grounding line behavior, where grounding line retreat in the western Ross Sea may have triggered upstream thinning (Jones et al., 2015) or re- advance (Greenwood et al., 2018).

An alternative pair of hypotheses proposes that individual ice lobes comprised of local alpine glaciers, EAIS outlet glaciers, and marine-based ice in the Ross Sea varied in size and flow direction in McMurdo Sound during the last glacial period (Anderson et al., 2017;

Wilson, 2000). Wilson (2000) suggested that fossiliferous Eocene erratics in Ross Sea drift indicated expansion of local alpine glaciers during the LGM. Anderson et al., (2017) built upon this hypothesis and proposed that alpine glaciers and outlet glaciers expanded prior to the LGM, followed by ice flow reversal, and predominance of westward flow from the

19

Ross Sea during the LGM. Despite these differences, both hypotheses allow smaller, local ice masses to behave independently of wider grounded ice conditions in the Ross Sea, perhaps in response to changing local glacier mass balances driven by reduced accumulation and/or melting.

The style and timing of deglaciation in the Ross Sea Embayment following the

LGM also remains uncertain. Multiple hypotheses have emerged including the ‘swinging gate’ model (Conway et al., 1999), ‘saloon door’ model (Ackert, 2008), and asynchronous deglaciation – all controlled by complex factors such as local bathymetry, ice flow velocity and switching, individual glacier front dynamics, dynamic thinning, ocean melting, sea ice, among other effects (Bart and Cone, 2012; Bart et al., 2017; Greenwood et al., 2018;

Halberstadt et al., 2016). The deglacial chronology from marine sediments, however, remains poorly constrained. This is due to the scarcity of calcareous microfossils available for radiocarbon dating, uncertain and variable Antarctic marine reservoir ages, and excessively old radiocarbon ages of acid insoluble organic matter that suggest reworking

(Andrews et al., 1999; Domack et al., 1999; Hall et al., 2010b; Licht et al., 1996; McKay et al., 2016; Prothro et al., 2018). Ice thinning records from EAIS outlet glaciers along the

Transantarctic Mountains, however, suggest synchronous onset of rapid ice elevation lowering between 6 to 9 ka (Jones et al., 2015; Spector et al., 2017). In McMurdo Sound, ice initially thinned from an elevation of 512 m after 14 ka to 234 m at 10.3 ka, and reached the modern elevation of the McMurdo Ice Shelf (60 m) by 8.1 ka (Anderson et al., 2017).

Radiocarbon ages of marine fossils on the surface of the McMurdo Ice Shelf are

20 accordingly all less than 7 ka, marking the transition to floating ice shelf conditions.

Integrating terrestrial glacial chronologies with glacial marine data remains challenging.

At issue is the fundamental behavior and interaction of EAIS and WAIS during both glacial expansion and retreat. Did ice from East Antarctica expand independent of interactions with West Antarctic ice, implying increased precipitation as a driving force?

Alternatively, despite low ice accumulation rates, did East Antarctic ice reach its maximum extent during periods of lower sea-level when grounded ice in the Ross Embayment blocked the terminal margins of outlet glaciers? The interpretation of ice sheet dynamics in the western Ross Sea has significant implications for the origin and magnitude of

Antarctic contributions to past sea-level rise, including Meltwater Pulse 1a (MWP-1A; 14.0 to 14.5 ka) (Clark et al., 2009; Lambeck et al., 2014; Liu et al., 2016). Investigation of the

LGM to present glacial history of McMurdo Sound in the western Ross Sea offers the opportunity to resolve these questions.

2.2 Study area

McMurdo Sound is geographically bounded by the Transantarctic Mountains and

Late Cenozoic alkaline volcanic islands and peninsulas. McMurdo Sound is bordered by the and McMurdo Dry Valleys in the west, Ross Island in the north, the Ross Ice Shelf in the east, and Minna Bluff, Mount Discovery, and Mount Morning in the south (Figure 2.1B). The McMurdo Ice Shelf, an extension of the Ross Ice Shelf, flows into McMurdo Sound between Black Island, White Island, and Minna Bluff, and then northwestward between Black Island and Brown Peninsula towards the modern calving

21 front. Basal freezing-on of debris along with high surface ablation rates (Glasser et al.,

2006) and slow, constricted ice flow has accumulated glacio-tectonically deformed debris bands (Glasser et al., 2014) on the ice shelf surface east of Brown Peninsula. Koettlitz

Glacier, sourced exclusively from alpine glaciers in the southern Royal Society Range and on Mount Morning, and the smaller Discovery Glacier terminate as floating ice tongues locked in perennial sea ice between Brown Peninsula and the Royal Society Range.

The northern faces of Mount Morning, Mount Discovery, and Black Island and the entirety of Brown Peninsula are free of glacial ice. This is due to isolation from continental glacial ice sources and a regional ablation zone caused by southerly winds descending over

Mount Morning, Mount Discovery, and Minna Bluff (Denton and Marchant, 2000; Glasser et al., 2014). Below 770 m elevation, ice-free areas are mantled with glacial sediments composed primarily of rocks from the local McMurdo Volcanic Group and is locally rich in erratic lithologies – primarily Granite Harbor Intrusives, Ferrar Dolerite, Beacon

Sandstone, Koettlitz Group metamorphic rocks, and others – transported north by grounded ice sourced from Transantarctic Mountains south of McMurdo Sound (Talarico and

Sandroni, 2011). Earlier glacial mapping efforts in this area identified at least two different glacial sedimentary units: 1) a lower-elevation glacial sediment commonly referred to as

Ross Sea drift that is characterized by a higher-concentration of erratic lithologies containing un-weathered clasts that lack iron-staining or ventifaction, and 2) a higher elevation glacial sediment characterized by lower concentration of erratics and ventifacted and varnished boulders, presumably of older age (Denton and Marchant, 2000; Stuiver et al., 1981). Above these sediments, no erratic lithologies are present; only glacial sediments

22 deposited by local alpine glaciers exist. In this paper, we focus on the distribution and age of Ross Sea drift.

The age of Ross Sea drift on the volcanic islands and peninsulas of McMurdo

Sound is poorly constrained. Based on the weathering of boulders, the age of Ross Sea drift was initially interpreted as LGM, but an early attempt at cosmogenic surface exposure dating of erratic Ferrar Dolerite and local McMurdo volcanic boulders suggested a more complex exposure and perhaps depositional history due to large scatter in the dataset

(Brook et al., 1995). Recent cosmogenic 10Be surface exposure ages of erratic granite and sandstone boulders collected along an elevation transect on east Mount Discovery (~520 to 30 m) suggest ice thinning began after 14 ka and reached the modern elevation of the

McMurdo Ice Shelf by 8.3 ka (Anderson et al., 2017). This ice thinning chronology agrees with radiocarbon chronologies of fossil algae in Ross Sea drift in the mouths of Taylor

Valley (Hall and Denton, 2000; Hall et al., 2000) and the Royal Society Range (Hall et al.,

2015; Jackson et al., 2018) that indicate the maximum extent of grounded ice between 18.9 and 12.8 cal. ka. Nevertheless, the timing of the local LGM delineated by Ross Sea drift across the islands of McMurdo Sound remains unconstrained. Without sufficient age control, ice sheet reconstructions during the LGM and deglaciation in this region, along with their implications for ice sheet interactions in the wider Ross Sea Embayment, remain uncertain.

23

2.3 Methods

2.3.1 Glacial Geomorphologic Mapping

We conducted detailed field surveys to map and describe the distribution of glacial sediments and landforms (e.g. moraine ridges, sharp or diffuse limits of Ross Sea drift) across McMurdo Sound, specifically focusing efforts at two sites on Mount Discovery

(Figure 2.2A, B), two sites on Black Island (Figure 2.2C, D), and three sites on Brown

Peninsula (Figure 2.2E, F, G). GPS data points were collected using a handheld Garmin

GPS unit (precision <10 m) at soil sampling locations and using a Trimble Zephyr GNSS

Geodetic II differential GPS system (precision <0.1m) along continuous transects to map glacial landforms. Soil pits were excavated to the depth of underlying ice cement to describe the sedimentology and develop a stratigraphic framework of the landscape. On- the-ground field surveys were accompanied by several helicopter missions to take aerial photographs of field sites and other locations that were not sampled. To make a complete glacial geologic map, field observations were complemented with analysis of Landsat 8

OLI (30 m/pixel) and high-resolution (< 1.85 m/pixel) DigitalGlobe, Inc. imagery and several digital elevation models (DEMs) including: a stereo-photogrammetric DEM (2 m/pixel) created by the Polar Geospatial Center, a LiDAR DEM covering White Island by the Airborne Thematic Mapping campaign (Shenk et al., 2001), and a new LiDAR DEM

(1 m/pixel) covering the Denton Hills at the base of the Royal Society Range (Fountain et al., 2017). Previously published volcanic bedrock maps (Kyle, 1990) were digitized in

ArcGIS and integrated with geomorphologic mapping results.

24

2.3.2 Ice & Snow Stable Isotope Geochemistry

During soil pit excavation, buried massive ice (where encountered) and ice cement were sampled for stable isotopic analysis (δ18O & δD). To recover massive ice samples, we excavated ~ 5 cm into the ice with a pickaxe and recovered ice chips. Ice cement samples were recovered as chips of sediment with interstitial pore ice. Additionally, modern snow from storms that occurred during different field seasons was collected.

Samples were stored and shipped frozen from Antarctica to Boston University (BU), where samples were sub-sampled, melted at room temperature for 24 hours, and stored in glass scintillation vials in a refrigerator. Melted water from ice cement samples was decanted to avoid inclusion of sediment particles. Samples were then analyzed at the BU Stable Isotope

18 Laboratory. δ O was analyzed via CO2 equilibration. Each melt sample was analyzed for

δD via pyrolysis in a ChromeHD system (Morrison et al., 2001). Internal laboratory methods and quality control measures are summarized at https://www.bu.edu/sil/quality.htm. At least two measurements were made for each sample during each analysis and referenced to the Vienna Standard Mean Ocean Water (V-

SMOW) standard.

25

Figure 2.2: Aerial photographs of key field sites on the volcanic islands and peninsulas of McMurdo Sound

26

Aerial photographs of key field sites on the volcanic islands of McMurdo Sound with elevations of Ross Sea drift (yellow text) and view direction (bottom left of each panel). A) Sharp limit of erratic rich Ross Sea drift at 512 m on eastern

Mount Discovery. B) Moraine ridge along western Mount Discovery that ascends from 195 m (foreground) to 240 m

(background) to the southwest. C) Ice-cored moraine and erratic-rich Ross Sea drift at 490 m limit on northern Black

Island. D) Ross Sea drift moraine segments on west Black Island at 390 m. E) Small moraine ridge draped on a basanite cinder cone on south Brown Peninsula that descends from 375 to 350 m (foreground) and small moraine segments at 330 m (background). F) Moraine wrapping around Frame Ridge on northern Brown Peninsula that reaches a maximum elevation of 340 m. G) Kame terrace between the 340 m elevation moraine and cinder cone at Frame Ridge. H)

Continuous sharp crested moraine segment along western Brown Peninsula that ranges in elevation between 215 and 265 m. Photography credits: Andrew Christ (A-D, G), David Marchant (E, F, H).

Twenty-one samples of ground ice and three samples of modern snow were analyzed for δ18O and δD to investigate the origin of buried ice masses. From ground ice samples, we analyzed 11 samples of massive ice collected below massive diamict in five soil pits on Black Island and Mount Discovery and ten samples of ice cement in stratified sediments from three soil pits on Brown Peninsula (Table 2.1). Within one soil pit, DME-

13-40, six samples were collected from a depth profile 5 to 60 cm below the ice surface (or

35 to 90 cm below the till surface). Ice samples were collected from discrete depth intervals from 5-10 cm, 10-15 cm, 15-20 cm, 50-60cm, and 60-65 cm, and a bulk ice sample was collected from 20-60 cm.

2.3.3 Radiocarbon Dating

Bulk samples of fossil and modern algae fragments were collected using a clean metal spoon or trowel while wearing gloves to avoid contamination. Fossil algae samples were collected from within moraine ridges and ice-proximal sediments and where possible

27 at depths >10 cm below the surface to avoid potential contamination with modern, wind- blown algae. To determine if a local radiocarbon reservoir effect presently exists, modern algae were collected from active meltwater streams and the surface of frozen thermokarst ponds. Samples were stored and shipped frozen from Antarctica to BU. At BU, algae were dried in an oven at 50°C for 24 hours, cleaned to remove inorganic sediment, and submitted to the National Ocean Sciences Accelerator Mass Spectrometer Facility (NOSAMS) at the

Woods Hole Oceanographic Institution. At NOSAMS, samples were further cleaned, analyzed for δ13C, Δ14C, and radiocarbon dated; sample results were delivered in terms of

14C years. In order to account for temporally variable atmospheric 14/12C ratios, results were calibrated using the CALIB 7.1 program using the IntCal13.14c atmospheric radiocarbon calibration curve (IntCal13.14c) (Reimer et al., 2013). The same calibration dataset has been used in recent, similar terrestrial radiocarbon chronologies in Antarctica (Hall et al.,

2015; Jackson et al., 2018). In order to compare results from our field areas, radiocarbon ages of fossil algae from previous studies were compiled and re-calibrated using the

IntCal13.14c calibration curve.

2.4 Results

2.4.1 Spatial Distribution of Ross Sea Drift in Southern McMurdo Sound

In general, the maximum elevation of Ross Sea drift decreases from east to west across southern McMurdo Sound (Figure 2.3, 2.4). In the east, Ross Sea drift reaches a maximum elevation of at least 710 m at Cape Crozier on eastern Ross Island and at least

635 m on Minna Bluff (Denton and Marchant, 2000). While White Island is almost entirely

28 glaciated today, Ross Sea drift and erratic boulders are found up to 561 m, but may be present at higher elevations below the local ice cap (Denton and Marchant, 2000). In the following sections, I provide detailed descriptions of Ross Sea drift on the ice-free areas of volcanic islands and peninsulas of southern McMurdo Sound.

2.4.1.1 Mount Discovery

Ross Sea drift extends from the surface of the McMurdo Ice Shelf to a maximum elevation of 512 m at a sharp, erratic-rich (~50% surface lithologies) drift limit on a small parasitic basanite cinder cone on eastern Mount Discovery near Minna Saddle (Figure

2.2A, 2.4A), marking the highest elevation, well-defined limit of Ross Sea drift in southern

McMurdo Sound. Ross Sea drift at this location is occasionally ice-cored and pitted with thermokarst scarps and ponds that continues towards the surface of the McMurdo Ice Shelf.

Approximately 2 km north of the 512 m limit, the Ross Sea drift limit becomes more diffuse and descends in elevation to 300 m where it forms a small ~1 m tall moraine ridge composed primarily of erratic cobbles (~50% surface lithologies) and small boulders

(Figure 2.4A). Farther west, along the northern flank of Mount Discovery, the concentration of surface erratic lithologies in Ross Sea drift decreases. Here the Ross Sea drift is marked by hummocky, ice-cored thermokarst with seasonally frozen meltwater ponds. At lower elevations on Mount Discovery near Brown Saddle, older east-west trending debris bands and erratic-rich boulder lines are crosscut by the modern, presumably younger, tidal crack of the McMurdo Ice Shelf. The drift limit descends to 195 m in elevation where it intersects with a NE-SW trending, boulder-rich moraine with fewer erratic lithologies (~15% surface lithologies) (Figure 2.2B, 2.4A). This moraine ridge

29 wraps around the western flank of Mount Discovery, increasing in elevation from 195 m to 250 m to the southwest over a short distance of ~1 km where it encounters a small perennial snowfield. Farther west towards Discovery Glacier, the drift limit becomes discontinuous where it is disturbed by overflowing solifluction lobes.

2.4.1.2 Black Island

On Black Island Ross Sea drift extends from the coast to a maximum elevation of

490 m in the north where it terminates as an erratic-rich (~40% of surface lithologies), ice- cored moraine pitted with small, seasonally frozen thermokarst ponds (Figure 2.2C, 2.3,

2.4C). West of the ice-cored moraine, the drift limit becomes diffuse and no longer ice- cored in the lee of a perennial snow bank at 460 m elevation. Southwest of this snowbank, the limit of Ross Sea drift declines in elevation to 420 m to a sharp limit characterized by hummocky topography, locally rich in angular trachyte and phonolite cobbles and boulders. Farther southwest, the Ross Sea drift limit transitions into a moraine ridge at the foot of a large wind-swept snow field that declines to 400 m elevation with very few erratic boulders (Figure 2.2D, 2.4C). Westward there is a moraine ridge (390 m elevation) composed primarily of trachyte clasts (Figure 2.2D). South of this moraine, the Ross Sea drift limit becomes diffuse and along southwest Black Island modern alpine glaciers obscure any underlying glacial sediments. In general, the concentration of erratic lithologies at the surface of Ross Sea drift decreases from northeast to southwest on Black

Island.

30

Figure 2.3: Integrated glacial geomorphologic and bedrock geology of McMurdo Sound, Antarctica

31

This large overview map is a compilation of new and existing geologic and chronologic data. Glacial geomorphologic mapping and radiocarbon chronology draws on data from this study and previously published work (Denton et al., 1989;

Denton and Marchant, 2000; Hall et al., 2015; Hall et al., 2000; Jackson et al., 2018; Stuiver et al., 1981). Bedrock geologic mapping data was digitized and adapted from Kyle (1990) & Wright-Grassham (1987). Elevation data is from the Reference Elevation Mosaic of Antarctica (REMA) created by the Polar Geospatial Center from DigitalGlobe, Inc. imagery. Imagery data used to classify ice surfaces was from Landsat 8 OLI.

2.4.1.3 Brown Peninsula

Ross Sea drift envelops the lower elevations of Brown Peninsula, extending from the coastline and McMurdo Ice Shelf up to a maximum elevation of 375 m along a small

(0.25 km long) moraine draped upon a small basanite cinder cone on southeast Brown

Peninsula (Figure 2.2E, 2.3 2.4B). Here the moraine descends in elevation to 350 m and extends discontinuously around the southern flank of Brown Peninsula as small moraine fragments that reach elevations as low as 330 m. The surfaces of the moraine crests are characterized by desert pavement dominated by volcanic pebbles with rare erratic cobbles.

A small closed basin immediately northeast of Swyers Point, an eroded trachyte dome, contains erratic-rich, thermokarst sculpted Ross Sea drift, but lacks clearly defined drift limits (Figure 2.3).

On north Brown Peninsula, Ross Sea drift reaches a maximum elevation of 345 m where it wraps around the northeast face of Frame Ridge, a small basanite cinder cone

(Figure 2.2F, 2.2G, 2.4B). At this location, the moraine decreases in elevation to 250 m around the southern face of the cinder cone where Ross Sea drift terminates as an arcuate drift limit with its convex side oriented towards the center of northern Brown Peninsula

(Figure 2.2F, 2.2G, 2.4B). A kame terrace occupies a gently sloping plane composed of

32 bedded sands and gravel between the moraine and basanite cinder cone (Figure 2.2G). To the north, Ross Sea drift wraps continuously around Frame Ridge and continues as a prominent moraine ridge that descends in elevation to 215 m elevation along the northwest face.

Along the west coast of Brown Peninsula, Ross Sea drift extends from the coast up to a continuous moraine that traces southwest from the base of Frame Ridge for 6 km and varies in elevation between 215 and 265 m (Figure 2.2H). Similar to the moraine ridge on western Black Island and west Mount Discovery, very few erratic lithologies are present along this entire moraine segment; instead volcanic cobbles and gravel dominate surface lithologies. At the northern end, multiple moraine ridges exist. A higher elevation moraine ridge pitted by thermokarst scarps and meltwater ponds rises to 260 m elevation, nearly reaching the base of Frame Ridge. Below the thermokarst-pitted moraine ridge, another separate sharp crested moraine ridge extends continuously south for 6km. This sharp crested moraine reaches higher elevations where it is draped against bedrock slopes along

Rainbow Ridge and descends to lower elevations in small, intervening drainage catchments. Commonly snow-filled, ephemeral stream channels bisect the moraine ridge in each drainage catchment and drain to the sea ice at the coast of Brown Peninsula. The limit of Ross Sea drift reaches its highest elevation of 265 m at the southern end; beyond this location, the drift limit is diffuse.

33

Figure 2.4: Glacial geologic maps of key field areas in McMurdo Sound

34

A) north Mount Discovery, B) north Brown Peninsula, C) northwest Black Island, and D) north Mount Morning.

Geologic mapping symbology, color scheme, and nomenclature follows the format of Figure 2.3. Contour intervals are

100 m in panels A and D and 20 m in panels B and C. Base image is a digital elevation model (2 m/pixel) created by the

Polar Geospatial Center using DigitalGlobe Inc. imagery.

2.4.1.4 Mount Morning

On Mount Morning, volcanic bedrock geology maps identify (Kyle, 1990; Martin et al., 2010) – and analysis of satellite imagery and DEM data confirm – the presence of moraines and hummocky terrain similar to Ross Sea drift elsewhere in McMurdo Sound

(Figure 2.3; Figure 2.4D). Importantly, these moraines and drift limits mimic the shoreline geometries of Lake Morning and Lake Discovery; they do not parallel the geometry of the present margins of surrounding alpine glaciers. At Hurricane Ridge, Ross Sea drift extends from the shore of Lake Discovery (135 m elevation) to a continuous moraine ridge that ranges in elevation between 240 m and 320 m (+105 m and +185 m above Lake Discovery)

(Figure 2.3). Between Koettlitz and Vereyken glaciers, Ross Sea drift extends from Lake

Morning (370 m) to an apparent sharp limit that ascends from 470 m (+100 m above Lake

Morning) near Koettlitz Glacier to 570 m (+200 m above Lake Morning) on a cinder cone on Riviera Ridge and varies between 420 and 460 m (+50 and +90 m above Lake Morning) in Riu ō Te Ata Valley (Figure 2.4D).

2.4.2 Drift Limit Sedimentology & Stable Isotope Geochemistry

We identified two drift limit deposits – massive diamict and stratified ice-contact sediments – that both contained ground ice as well as in situ algae fragments that we radiocarbon dated. The stratigraphy, sedimentology, and ground ice isotope geochemistry

35 of these deposits lend insight into ice-marginal processes along a former grounded ice sheet in McMurdo Sound that place radiocarbon ages in geomorphic context and allow us to reconstruct former ice sheet geometry.

2.4.2.1 Massive Diamict

Along its maximum limits on east Mount Discovery and Black Island, Ross Sea drift is characterized by ice-cored thermokarst topography and increased concentrations of erratic lithologies relative to Ross Sea drift on northern Brown Peninsula and western

Mount Discovery. Excavations into Ross Sea drift along these drift limits reveals massive, unconsolidated diamict (Figure 2.6). Erratic cobbles within Ross Sea drift are often molded, striated, and/or gouged. Fragments of fossil coral, bivalves, and sponge spicules are common within the diamict matrix, similar to surface debris on the McMurdo Ice Shelf

(Kellogg et al., 1990; Stuiver et al., 1981). Along the 390 m elevation moraine on west

Black Island, fragments of fossil algae are found throughout the diamict, and are occasionally draped on buried cobbles. Below 40 to 50 cm of diamict, Ross Sea drift is commonly ice-cored. On east Mount Discovery between thermokarst depressions, lenses of massive ice containing clasts and debris bands occasionally underlie 30 to 40 cm of diamict. On north Black Island, Ross Sea drift is composed of erratic-rich diamict, but along thermokarst scarps, dry mirabillite powder fills the pore space between grains in diamict. Below 40 cm, the diamict contains a greater fraction of salts, and below 50 cm, there is massive ice with isolated bands of sandy gravel (Figure 2.6).

36

Table 2.1 Ground ice stable geochemistry

37

Ground ice from soil pits on Mount Discovery and Black Island yielded δ18O ratios that range between -14.86‰ and -28.94‰ and δD ratios that range between -122.83‰ and

-238.23‰ (Figure 2.5 A,B; Table 2.1). In general, stable isotope ratios plot below the global meteoric water line (GMWL), yet yield a slope of 8.49 (R2 = 0.9877). D-excess values ranged from 1.38‰ to -19.49‰. When the samples from soil pit DME-13-40 are not considered, the range of stable isotope ratios are more depleted (δ18O: -26.30‰ to -

28.94‰, δD: -188.69‰ to -238.23‰) and D-excess values are similarly negative (D- excess: -5.90‰ to -19.49‰).

Along the depth profile of ice samples from pit DME-13-40 on east Mount

Discovery, isotope ratios are increasingly depleted with depth, with δ18O and δD decreasing from -15.35‰ and -122.83‰ at 5-10 cm to -20.29‰ and -160.93‰ below 60 cm depth, respectively (Figure 2.5 C-E). δ18O and δD vary strongly and inversely with depth (R2 = 0.85 and R2 = 0.81, respectively). D-excess values, however, are essentially invariable with depth (R2 = 0.47, slope = 0.06), increasing from -3.97‰ (5-10 cm) to

1.38‰ (below 60 cm). In general, samples from DME-13-40 were more isotopically enriched and had higher D-excess values than ice from soil pits DME-13-41 and ACE-13-

04, despite being collected from the same general location on east Mount Discovery. A modern snow sample from a nearby snowbank yielded similar δD (-129.86 ‰) and δ18O (-

17.67‰) ratios, but a slightly more positive D-excess value (11.47‰).

38

Figure 2.5: Stable isotope measurements of ground ice in McMurdo Sound

Ground ice stable isotope geochemistry results from McMurdo Sound and regional comparisons. Bivariate plots of A.)

δ18O and δD and B.) δD and deuterium-excess of massive ice (dark gray), ice cement in stratified sediments (light gray),

39 and modern snow (white) from McMurdo Sound. Depth profiles of C) δ18O, D) δD, and E) deuterium-excess of massive ice recovered from soil pit DME-13-40. Vertical error bars denote the depth range from which ice was collected; circle is the median depth of the excavated interval. Comparisons of F.) δ18O and δD and G.) δD and deuterium-excess to other regional records. F) Normalized histogram of δ18O ratios of regional glacial ice, ground ice, ice shelf ice, and modern snow. Histogram blocks are transparent, not stacked, to show overlapping frequency of δ18O ratios. Superscripts: θthis study, ΛHall et al. (2010), †Pollard et al. (2002), ‡Kellogg et al. (1990), *Levy et al. (2013), #Steig et al. (2000).

2.4.2.2 Stratified Deposits

On Brown Peninsula and west Mount Discovery, excavations into moraines and drift limits along Ross Sea drift commonly revealed stratified sediments bearing in situ algae fragments. Along moraines, the sedimentology can generally be described by: 1) basal ice-cemented pebbly gravel overlain by 2) cm-scale bedded sands with algae that were occasionally capped by 3) blocky silts and fine sand that are occasionally deformed, and 4) unconformable slumped, unconsolidated diamict (Figure 2.6). Bedding commonly dips into the moraine, and occasionally, the entire sequence of sediments was deformed, with folds well preserved by ice-cement. Importantly, fossil algae were only recovered from within ice-cemented pebbly gravel and bedded sands, below blocky sandy silt or the capping slumped diamict. Sediments capping a prominent moraine segment on west Mount

Discovery lacked bedded sands. Instead, fossil algae draped directly onto underlying pebbly gravel and is capped by massive sandy silt. Along the kame terrace bordering the moraine at Frame Ridge, soil excavations revealed a slightly different stratigraphy characterized by occasionally deformed dark gray cm-scale bedded sand containing algae fragments overlain by tan pebbly gravel.

40

Stable isotope ratios of ice cement from Brown Peninsula vary widely and generally plot below the GMWL with a slope of 6.75 (R2 = 0.88). δ18O ratios ranged between -

16.41‰ and -32.74‰ and δD ratios ranged between -126.28‰ and -273.65‰ (Figure 2.5

A, B; Table 2.1). D-excess values ranged widely from as high as 8.31‰ (ACI-14-28A) to as low as -41.92‰ (ACE-14-27A), but were mostly negative. Compared to the massive ice samples, stable isotope ratios from ice cement range more widely and have more negative d-excess values. Two samples of modern snow from Brown Peninsula yielded ratios of

δ18O (-24.96‰, -27.41‰) and δD (-203.04‰, -222.17‰) that fell within the range of ratios for ice cement samples but had higher D-excess values (-3.33‰, -2.86‰).

Figure 2.6: Characteristic soil pit stratigraphy in McMurdo Sound

Soil pit stratigraphic columns characteristic of ice marginal environments in McMurdo Sound for massive diamict (DME-

13-40 and ACE-15-21) and stratified sediments, which include bedded deltaic ice-contact sediments (ACE-14-29, ACE-

14-05, ACE-14-06) and kame terrace deposits (ACE-14-25). Colors scheme corresponds to sediment color.

41

2.4.1 Algae Radiocarbon Chronology

Four samples of modern algae collected from ephemeral streambeds and frozen thermokarst ponds yielded “modern” ages indicating that today no local reservoir effect exists in McMurdo Sound (Table 2.2, 2.3). Additionally, we do not recognize the signature of marine-sourced carbon in these samples for the following reasons. Radiocarbon ages from our study area fall within the range of other terrestrial algae radiocarbon dates from

Ross Sea drift elsewhere in McMurdo Sound. Additionally, Ross Ice Shelf proximal sediment from the modern sediment-water interface produces highly depleted 14C (-

1000‰) and δ13C (<-20‰) ratios (Ohkouchi & Eglinton, 2006); the respective isotopic ratios of 14C (-672.46‰ to -860.47‰) and δ13C (-4‰ to +6‰) of algae in Ross Sea drift is more enriched, despite a much older, true depositional age. If algae were incorporating marine carbon, the responding radiocarbon ages and 14C and δ13C ratios should be older and more depleted, respectively. As such, fossil algae radiocarbon ages were calibrated using the CALIB 7.1 program (Reimer et al., 2013) using the IntCal13.14c calibration curve with no additional reservoir effect (ΔR value), as conducted in similar recent radiocarbon chronologies of Ross Sea drift from the Royal Society Range (Hall et al.,

2015). All calibrated ages are reported with 1 sigma error.

42

Table 2.2: Radiocarbon chronology

43

Median calibrated ages of fossil algae recovered from Ross Sea drift ranged from

10.0 to 19.0 calibrated thousands of years before present (cal. ka) (Table 2.3; Figure 2.7).

On west Mount Discovery, sample ACA-15-02B, collected from a moraine at 207 m elevation, yielded a median age of 15.6  0.1 cal. ka. On Black Island, radiocarbon ages of seven samples recovered from within the unsorted till matrix the moraine at 390 m elevation ranged between 13.5  0.2 and 16.4  0.2 cal. ka. On Brown Peninsula, median calibrated radiocarbon ages cluster according to landform (i.e. western moraine ridge vs. kame terrace below Frame Ridge) and are generally older with increasing elevation. Fossil algae recovered from the kame terrace adjacent to the moraine ridge on Frame Ridge yielded some of the oldest radiocarbon ages in our dataset, 17.9  0.1 cal. ka (ACA-14-25; elevation: 259 m) and 18.8  0.1 cal. ka (ACA-14-08; elevation: 275 m). Algae samples

(n=10) recovered from stratified sediments along the western moraine of Brown Peninsula yield radiocarbon ages that range from 10.0  0.1 to 19.0  0.1 cal. ka.

44

Figure 2.7: Compiled radiocarbon chronology of Ross Sea drift.

Normal kernel density estimate plots of A) all (new and published) calibrated radiocarbon ages of sub-fossil algae collected from the maximum limit of Ross Sea drift in McMurdo Sound including the B) volcanic islands (this study),

C) Taylor Valley (Hall and Denton, 2000), and D) the Royal Society Range (Hall et al., 2015; Jackson et al., 2018).

Individual thin light gray Gaussian curves represent the median calibrated radiocarbon age with 2-sigma error. Thick black line is the cumulative probability for all radiocarbon ages within each class. Note: radiocarbon ages from Taylor

Valley exclude those collected from lacustrine, lake-ice conveyor, or stream sediments, which do not necessarily record the presence of grounded ice at its maximum position in McMurdo Sound.

45

2.5 Discussion

Using this compiled dataset of glacial geologic mapping, ground ice stable isotope geochemistry, and calibrated radiocarbon ages, we interpret the former grounded ice sheet at its maximum thickness and position in McMurdo Sound. We discuss ice marginal paleoenvironments and their significance for interpreting radiocarbon ages of fossil algae and by extension the timing of maximum ice positions that can be compared to other regional records. Then, we describe our interpretations for the former ice flow geometry in southern and northern McMurdo Sound and wider implications for marine ice sheet dynamics in the western Ross Sea during the local LGM and deglaciation.

2.5.1 Ice-marginal Paleoenvironments

2.5.1.1 Buried Glacial Ice & Ablation Till

The characteristics of Ross Sea drift along the eastern and northern flanks of Mount

Discovery and northwest Black Island suggest a similar ice source and depositional setting.

The concentration of erratic lithologies in massive diamict along eastern Mount Discovery and northwest Black Island is higher than elsewhere on the volcanic islands of McMurdo

Sound. Additionally, erratic cobbles and boulders in Ross Sea drift along eastern Mount

Discovery and northwest Black Island are commonly striated, gouged, or glacially molded.

Together, we infer that the erratics were sub-glacially sourced and transported by more southern outlet glaciers, specifically the Skelton, Mullock, and Carlyon glaciers, to

McMurdo Sound following the provenance study of erratic lithologies in the nearby AND-

46

1B drill core and coastal moraines on eastern Mount Discovery (Talarico et al., 2012;

Talarico et al., 2013).

The prevalence and geochemical similarity of massive ice below diamict on Mount

Discovery, Black Island, and elsewhere in McMurdo Sound, further suggests a similar ice source. δ18O ratios of this massive ice are comparable to some buried ice masses in

Garwood Valley (Levy et al., 2013; Pollard et al., 1999), Taylor Valley (Hall et al., 2000), and in the McMurdo Ice Shelf (-25‰ to -30‰) that are interpreted as remnant ice blocks from a grounded ice sheet (Kellogg & Kellogg, 1990) (Figure 2.5). These δ18O ratios, however, are not as depleted as LGM-age (10-20 kyr) ice from Taylor Dome (average -

40.33 ± 2.68‰) (Steig et al., 2000). The less depleted δ18O ratios from massive ice and old portions of the McMurdo Ice Shelf could indicate isotopic alteration from meteoric deposition and/or repeated sublimation or melt cycles (Kellogg et al., 1990). This alteration is particularly apparent for the depth profile from pit DME-13-40 where isotope ratios near the ice-diamict interface (5-10 cm: -15.35‰ [δ18O], -122.83‰ [δD]) are similar to modern snow from nearby (-17.67‰ [δ18O], -129.86‰ [δD]), but become depleted with depth (60-

65 cm: -20.29‰ [δ18O] and -160.93‰ [δD]; Table 2.1, Figure 2.5). We interpret these massive ice bodies on Mount Discovery and Black Island as remnant glacial ice preserved in Ross Sea drift derived from non-local sources modified by post-depositional precipitation, melting, and sublimation.

We attribute the similar characteristics of the drift limits and the preservation of glacial ice within Ross Sea drift on eastern Mount Discovery and northwest Black Island to glaciological conditions similar to the modern McMurdo Ice Shelf. Today, southerly

47 adiabatic winds cause enhanced ablation of the McMurdo Ice Shelf leading to upward advection of basal ice and entrained debris (Glasser et al., 2006; Kellogg et al., 1990).

Continued surface ablation accumulates debris at the ice shelf surface to form an extensive blanket of supraglacial till (Glasser et al., 2014). Given the prominence of the volcanic islands of McMurdo Sound, this adiabatic wind pattern, and therefore a similarly localized ablation zone, likely persisted during the last glacial period even as a grounded ice sheet occupied McMurdo Sound (Golledge et al., 2013). We suggest that enhanced ablation led to debris accumulation along the ice sheet margin that mantled underlying ice, preserving it since the LGM through deglaciation and into the Holocene. The prevalence and continued decay of sub-surface massive ice developed the extensive thermokarst topography that characterizes the surface of Ross Sea drift across much of McMurdo Sound today.

At the 390 m elevation moraine on west Black Island, Ross Sea drift contains fossil algae which we interpret as thriving in supraglacial melt ponds and was later buried in ablation till. The accumulated debris on the ice sheet surface lowered local ice sheet surface albedo, enhanced melt, and formed small supraglacial ponds that supported algal mats, similar to modern seasonal meltwater ponds on debris-laden portions of the McMurdo Ice

Shelf. The absence of stratified sediment superposed on or within the moraines, yet presence of algae within the unsorted till matrix, suggests that algae was buried in ablation till as the ice down-wasted. In one pit (ACE-15-22), radiocarbon ages of algae samples ranged from 16.4 cal. ka (fragments in till matrix 10-15 cm depth) to 14.1 cal. ka (algae draped on an adjacent cobble at 15 cm), suggesting burial and reworking of algae as

48 ablation till accumulated and slumped while underlying ice melted. Under this model, radiocarbon ages recovered from massive till along this moraine provide an age for the presence of grounded ice at or near its maximum position on Black Island where small supraglacial ponds supported algae communities, but not necessarily the depositional age of the moraine itself (Figure 2.8).

Figure 2.8: Schematic of ice-marginal processes during maximum ice sheet extent.

A) Debris on the ice surface enhanced ablation that formed B) supraglacial ponds harboring algal communities. C)

Meltwater drains through supraglacial streams to D) ice-contact deltas with algae. E) Ice contact ponds form between the ice sheet and along F) volcanic bedrock topography mantled with older, weathered glacial sediment. Radiocarbon ages of algae in Ross Sea drift record the presence of a grounded ice sheet at or near its maximum geometry.

49

2.5.1.2 Ice-contact Meltwater Streams

We interpret that stratified sediments in Ross Sea drift along west Mount Discovery and northern Brown Peninsula were deposited by meltwater streams and supported algae communities along the ice sheet margin. Meltwater streams deposited these stratified sediments as they consist of bedded medium-grained sands that overlie glacial diamict; therefore, the radiocarbon ages of these algae record the presence of active meltwater systems at the ice margin similar to Ross Sea drift in the Royal Society Range and Taylor

Valley (Figure 2.8) (Denton and Marchant, 2000; Hall and Denton, 2000; Hall et al., 2015;

Jackson et al., 2018; Stuiver et al., 1981). Erratic clasts are rare in Ross Sea drift on Brown

Peninsula and western Mount Discovery; as a result, primarily dark volcanic debris would have accumulated on the ice surface. We interpret that this lowered the surface albedo and formed surface meltwater streams that drained toward the ice margin and deposited small ice contact deltas that likely supported algal communities. Meltwater streams then ponded in basins bounded by volcanic bedrock slopes and the ice margin that deposited fine- grained lacustrine sediments. Subsequent fluctuations of grounded ice overrode and deformed underlying bedded sands and lacustrine silts. Ice cement 20 to 40 cm below the surface preserves delicate folds in these soft sediments. As ice retreated and down-wasted, the stratified sediments were capped by a structureless ablation till that presently characterizes the surface of Ross Sea drift. Algal fragments were exclusively recovered from gravel and bedded sands within the moraine or kame terrace sediments; therefore, radiocarbon ages record when grounded ice was positioned at the moraine ridge and

50 supplied ample meltwater to support algal communities in ice-marginal stream and ice- contact deltas.

2.5.2 Timing of the local Last Glacial Maximum in McMurdo Sound

Given this sedimentological interpretation, the range of calibrated radiocarbon ages of fossil algae in Ross Sea drift from the volcanic islands (n = 20), the Royal Society Range

(n = 149) (Hall et al., 2015; Jackson et al., 2018), and Taylor Valley (n = 67) (Hall and

Denton, 2000), (total n = 236) brackets the window of maximum ice sheet extent in

McMurdo Sound between 12.3 and 19.6 cal. ka. Normal kernel density estimates (Figure

2.7) provide a nonparametric representation of the probability density function of individual calibrated radiocarbon ages according to field site and the entire region (Figure

2.7). The overlapping frequency of calibrated radiocarbon ages provides an estimate of the age of Ross Sea drift while also considering the variability of individual ages with respect to analytical uncertainty as well as within and across different field sites in McMurdo

Sound. We attribute the variability of individual calibrated radiocarbon ages to small changes in the position of the ice margin or the activation of individual ice-contact streams, deltas, and other transient ice-marginal features that supported algal communities, rather than a major reorganization of the local ice sheet geometry. We interpret features delineating former maximum ice sheet extent and surface elevation as deposited across the local LGM in McMurdo Sound; under this assumption, we reconstruct the former geometry and ice flow directions of the grounded ice sheet.

We note a single radiocarbon age from west Brown Peninsula that yielded an anomalously young age (ACA-14-04: 10.0  0.1 cal. ka). While this algae fragment was

51 recovered from stratified sediments similar to other soil pits, it is possible that incomplete pretreatment of the sample prior to radiocarbon analysis generated an anomalously young age. The anomalous radiocarbon age could also reflect an algae fragment that subsisted in a small pond that filled the small basin bounded by the moraine ridge and Rainbow Ridge during deglaciation, similar to lacustrine deposits in Garwood Valley (Levy et al., 2017); however, another algae sample from bedded sands embedded in the same moraine ridge generated the oldest age in our dataset (DME-14-01: 19.0  0.1 cal. ka). Additionally, if we compare this age to ice thinning records from east Mount Discovery where the ice elevation thinned to 234 m by 10.3 ka (Anderson et al., 2017), the anomalously young algae radiocarbon age was collected from a moraine ridge at 210 m elevation. Rapid thinning in eastern McMurdo Sound would be presumably accompanied by comparable thinning between Brown Peninsula and the Royal Society Range at that time. Given these reasons, we are hesitant to interpret the young age as meaningful.

This compiled radiocarbon chronology of sub-fossil algae in Ross Sea drift is compatible with other chronological constraints of glacial sediments in McMurdo Sound deposited during the local Last Glacial Maximum and ice thinning and grounding line retreat. Cosmogenic nuclide exposure ages of local and erratic lithologies across McMurdo

Sound contain age scatter, most likely due to nuclides inherited from prior and complex exposure histories. Some individual exposure ages exceed the range of algae radiocarbon ages in ice-marginal sediments such as those from south Brown Peninsula (Anderson et al., 2017: 20.1 ka) and the Royal Society Range (Brook et al., 1995) where several 3He exposure ages of volcanic clasts in Ross Sea drift are more than 100 ka ‘too old’. At the

52 maximum limit of Ross Sea drift at the 510 m on east Mount Discovery, however, 10Be surface exposure ages of granite boulders record initial ice thinning after 14.0 ka – which coincides with the younger end of radiocarbon ages. With one lone exception, the range of radiocarbon ages from fossil algae that record the maximum ice configuration all pre-date the timing of ice thinning records in McMurdo Sound. The subsequent ice thinning record from east Mount Discovery – rapid thinning from 512 m to 234 m from 14.1 to 10.3 ka, then to 60 m by 8.3 ka at the modern elevation of the McMurdo Ice Shelf (Anderson et al.,

2017). Exposure age samples of erratic boulders (<6.1 ka) and radiocarbon ages of calcareous marine fossils (<7.6 ka) on the McMurdo Ice Shelf (Kellogg et al., 1990; Hall et al., 2010), which record the transition to a floating ice shelf in McMurdo Sound, post- date the radiocarbon chronology of fossil algae from the limit of Ross Sea drift. The timing of maximum ice configuration in McMurdo Sound similarly agrees the pulse of rapid thinning at Mackay Glacier that initiated after 8 ka (Jones et al., 2015) and the transition to open marine conditions north of Ross Island by 8.6 ka (Mackay et al., 2016).

2.5.3 Maximum Ice Sheet Configuration in McMurdo Sound

2.5.3.1 Terrestrial Evidence for Westward Flow in Southern McMurdo Sound

The westward decline in elevation of age-equivalent Ross Sea drift limits and moraines suggest ice flowed north and then westward around the volcanic islands towards the Transantarctic Mountains during the local LGM (Figure 2.9). Former ice surface elevation declines from 510 m to 195 m on Mount Discovery and 490 m to 390 m on Black

Island. Moraines on the eastern side of Brown Peninsula – the 375 to 350 m moraine ridge draped on a cinder cone on south Brown Peninsula, and 340 m moraine on Frame Ridge –

53 are higher than the continuous moraine ridge along west Brown Peninsula (220 to 260 m).

Ross Sea drift terminates at the foot of the Royal Society Range at headland moraines and drift limits that decrease in elevation between 350 m near Blue Glacier to as low as 250 m in Miers Valley (Hall et al., 2015; Jackson et al., 2018). This regional pattern suggests that the ice surface sloped, and therefore ice flowed, to the west across southern McMurdo

Sound when ice reached its maximum extent.

The orientation and slope of specific moraines further suggest westward flow across southern McMurdo Sound. The moraine draped on Frame Ridge on north Brown Peninsula decreases in elevation from 340 m to 250 m around the southern face of the cinder cone where Ross Sea drift terminates as an arcuate limit with its convex side oriented west towards the center of northern Brown Peninsula (Figure 2F, G). On South Brown

Peninsula, a moraine decreases in elevation from 375 m to 350 m and small moraine segments to the west reach only 330 m (Figure 2E). Within the ice-free valleys of the Royal

Society Range, Ross Sea drift similarly terminates as arcuate moraines and limits with convex margins oriented westward and up-valley (Denton and Marchant, 2000; Hall et al.,

2015; Jackson et al., 2018). These regional geomorphic patterns of suggest that Ross Sea drift was deposited by a westward flowing ice sheet in southern McMurdo Sound.

54

Figure 2.9: Local LGM grounded ice sheet reconstruction in McMurdo Sound.

This map based upon data from this study and previous publications (Denton et al., 1989; Denton and Marchant, 2000;

Greenwood et al., 2012; Hall and Denton, 2000; Hall et al., 2015; Hall et al., 2000; Jackson et al., 2018; Lee et al., 2017;

Stuiver et al., 1981). Former ice sheet elevations correspond to the maximum limits of Ross Sea drift on ice-free areas across McMurdo Sound. Ice flow directions were determined based on the implied ice surface slope, the distribution of erratic lithologies, and flow diagnostic features in the western Ross Sea. BP – Brown Peninsula, BI – Black Island, CB

– Cape Bird, CC – Cape Crozier, HH – Hjorth Hill, MD – Mount Discovery, , MBf – Minna Bluff, and MM – Mount

Morning, and WI – White Island. Base map elevation data is from IBSCO 2013 and the Reference Elevation Mosaic of

Antarctica (REMA) created by the Polar Geospatial Center from DigitalGlobe, Inc. imagery.

55

Our interpretation of westward flow contradicts the ice sheet reconstructions that propose local grounded ice filling McMurdo Sound, with individual ice lobes centered over

Koettlitz Glacier, the Ross Sea near Ross Island, and Minna Bluff that varied in size and flow direction during the last glacial period (Anderson et al., 2017; Wilson, 2000). Each reconstruction hinges upon the distribution of specific lithologies in Ross Sea drift – fossiliferous Eocene erratics (Wilson, 2001) and local trachyte cobbles (Anderson et al.,

2017) – to suggest that Koettlitz Glacier expanded and flowed north during the last glacial period. In both hypotheses, neither the Eocene erratics nor trachyte cobbles have direct age constraints to determine if they were deposited by the ice sheet at its maximum configuration or during deglaciation. Additionally, trachyte crops out extensively across

McMurdo Sound (mapped generally as “Phonolite” in Figure 2.3 (Kyle, 1990)); thus, the presence of trachyte cobbles in Ross Sea drift cannot be definitively attributed to a single lithological source to infer ice flow direction of Koettlitz Glacier.

Our mapping is inconsistent with the interpretation of debris ridges on the

McMurdo Ice Shelf as relict medial moraines formed by the convergence of an expanded, northward flowing Koettlitz Glacier with ice lobes from Minna Bluff and the Ross Sea.

(Anderson et al., 2017; Glasser et al., 2014; Wilson, 2000). The orientation of moraines draped onto cinder cones on Frame Ridge and south Brown Peninsula indicate westward flow of a larger ice mass into McMurdo Sound. Furthermore, age-equivalent moraines on the west coast of Brown Peninsula (10.0  0.1 - 19.0  0.1 cal. ka; 210 - 250 m) have a lower elevation than those in the east (17.9  0.1 and 18.8  0.1 cal. ka; 259 - 340 m), again indicating westward ice flow. Koettlitz Glacier could not expand and flow north over

56

Brown Saddle and continue around the eastern side of Brown Peninsula at a higher elevation than in the west. Additionally, east-west trending erratic-rich debris stripes on

Brown Saddle and southern Brown Peninsula that trace onto the surface of the McMurdo

Ice Shelf are crosscut by the modern tidal crack and folded debris ridges (Stuiver et al.,

1981); this suggests that the formation and folding of these debris ridges post-dates maximum grounded ice extent when the east-west debris stripes formed.

The distribution and orientation of Ross Sea drift limits on Mount Morning suggest only minor contribution of local ice to the grounded ice sheet during the LGM (Figure 2.3,

2.4). Although presently lacking age control, the moraines and limits of Ross Sea drift on

Mount Morning do not mimic the present geometries of Koettlitz Glacier and surrounding alpine glaciers. This suggests that the present flow regime of local glaciers is different from the LGM ice flow pattern. On west Mount Discovery, moraines and limits of Ross Sea drift increase slightly in elevation from 195 m to 245 m towards Mount Morning; this could reflect the upstream damming effect of Koettlitz Glacier as it converged with a wider grounded ice mass. Importantly, Koettlitz Glacier is not an outlet glacier fed by the EAIS, but rather a marine-terminating glacier fed exclusively by small alpine glaciers in the southern Royal Society Range and on Mount Morning. At least some land-terminating alpine glaciers in the MDV region, including the nearby Salmon Glacier in the Royal

Society Range (Jackson et al., 2018), retreated during the LGM, possibly in response to reduced ice accumulation rates when a grounded ice sheet occupies McMurdo Sound siphoning off oceanic moisture sources (Hall et al., 2000; Marchant et al., 1994; Swanger et al., 2017). Although Koettlitz Glacier is a marine-terminating ice mass with different

57 dynamics than land-terminating alpine glaciers, we find it unlikely that Koettlitz Glacier could expand or thicken independently while its tributary local alpine glaciers likely retreated during the LGM.

2.5.3.2 Northern McMurdo Sound & Integration with Glacial Marine Records

In northern McMurdo Sound, our ice sheet reconstruction agrees with earlier efforts

(Denton and Marchant, 2000; Hall and Denton, 2000; Hall et al., 2015; Hall et al., 2000;

Jackson et al., 2018; Stuiver et al., 1981) that suggest a larger ice mass overflowed southward into northern McMurdo Sound during the local LGM (Figure 2.9). The elevation of Ross Sea drift in northern McMurdo Sound is higher than further south, reaching at least

540 m on Cape Bird and 420 m at the moraine ridge on Hjorth Hill at the mouth of Taylor

Valley (Denton and Marchant, 2000; Hall et al., 2000). This compares to <300 m in the

Royal Society Range and <265 western Brown Peninsula. These drift limit elevations suggests southwestward ice flow when the ice sheet was in its maximum configuration

(Denton and Marchant, 2000). This reconstruction additionally satisfies the distribution of kenyite erratics in Ross Sea drift in Taylor Valley and the Royal Society Range, which require a component of southwestward flow through McMurdo Sound to transport kenyite clasts from several known outcrops on west Ross Island (Denton and Marchant, 2000; Hall et al., 2000; Stuiver et al., 1981). Kenyite may also be sourced from unknown submarine and/or presently glaciated bedrock that could be reworked into Ross Sea drift from earlier glaciations, as noted on Black Island (Vella, 1969) and in the AND-1B core (Lee et al.,

2017). However, the distribution of kenyite clasts in Ross Sea drift in the mouths of the

Dry Valleys requires some component of flow towards the Transantarctic Mountains in

58

McMurdo Sound (Stuiver et al., 1981; Denton & Marchant, 2000; Hall & Denton, 2000;

Hall et al., 2015).

We argue that our maximum ice sheet configuration in McMurdo Sound is compatible with eastward flowing, expanded EAIS outlet glaciers in the western Ross Sea during the LGM (Lee et. al, 2017). Improved and continued multi-beam bathymetric surveys of the western Ross Sea revealed sub-glacial landforms such as mega-scale glacial lineations north of Beaufort Island that indicate eastward flow of EAIS ice from the

Transantarctic Mountains. Between Beaufort Island and Cape Bird on Ross Island, east- west trending lineations remain ambiguous with respect to their ice flow direction. We suggest that some of the ice draining the EAIS from the Scott Coast from via Mackay

Glacier overflowed south into McMurdo Sound during the local LGM, while most flowed east and then northward into the western Ross Sea. Channel features in McMurdo Sound are of unknown age, but channels incised into bedrock may be relict features carved by

Miocene sub-glacial outburst floods (Lewis et al., 2006; Sugden and Denton, 2004), and braided channel systems may be post-glacial marine features formed by turbidity currents

(Greenwood et al., 2012) or sub-glacial meltwater draining to the northeast (Lee et al.,

2017; Simkins et al., 2017; Greenwood et al., 2018).

During deglaciation, ice flow patterns in McMurdo Sound likely diverged from the maximum configuration we have proposed. Ribbed and DeGeer moraines on the submarine slopes of Ross Island and offshore of New Harbor formed as ice flow direction reorganized as the grounding line encroached into southern McMurdo Sound during deglaciation

(Greenwood et al., 2012). Further north, back-stepping grounding zone wedges seaward of

59 present termini of Mackay and Mawson Glaciers (Lee et al., 2017) remain undated but are likely recessional features. Recent, more detailed geomorphic and stratigraphic analysis of multi-beam bathymetric data suggest a more complex deglaciation pattern, perhaps local re-advance of EAIS outlet glaciers following reorganization of the ice sheet in the western

Ross Sea during deglaciation in the Early Holocene to deposit these features (Greenwood et al., 2018; Halberstadt et al., 2016). As such, radiocarbon ages of in situ algae fragments recovered from maximum limits of Ross Sea drift provide age control on the maximum ice sheet configuration in McMurdo Sound, and not necessarily the rate and pattern of deglaciation of EAIS outlet glaciers during the Early Holocene.

2.5.4 Implications for Ice Sheet Interactions & Deglaciation in the Ross Sea Embayment

At a wider regional perspective, our data provide support for previous interpretations of ice sheet interactions in the Ross Sea Embayment. Maximum ice extent in McMurdo Sound was delayed relative to the global LGM (Hall et al., 2015), yet coincides with atmospheric warming and enhanced ice accumulation rates at WAIS Divide

(Buizert et al., 2015; Cuffey et al., 2016) while local air temperatures and ice accumulation rates at Taylor Dome remained near their minima (Figure 2.10). We suggest that mass gains over WAIS Divide drove expansion of WAIS that buttressed ice sourced from EAIS outlet glaciers in the Ross Sea Embayment, causing outlet glaciers to thicken upstream (Denton et al., 1989; Stuiver et al., 1981). This buttressing effect may also have allowed for enhanced flow and sediment discharge from large outlet glaciers (such as Byrd Glacier) where these EAIS ice masses became marine-based in the Ross Sea Embayment (Golledge et al., 2013). In the McMurdo Sound region, we interpret that contemporaneously low local

60 ice accumulation rates and temperatures caused local ice masses to retract (Denton and

Marchant, 2000; Hall et al., 2000; Jackson et al., 2018), allowing buttressed ice in the Ross

Sea to overflow into McMurdo Sound towards the Transantarctic Mountains. We suggest that as sea-level and ocean heat flux increased during deglaciation (Denton et al., 2010;

Elderfield et al., 2012; Lambeck et al., 2014), the initial precipitation-driven gains over

WAIS (Buizert et al., 2015) gave way to marine ice-sheet retreat (Anderson et al., 2014;

Hall et al., 2015).

61

Figure 2.10: Paleoclimate from 31 kyr to present

A) cumulative probability density functions of all calibrated radiocarbon ages of sub-fossil algae from the maximum limits of Ross Sea drift in McMurdo Sound that record maximum ice extent (Hall and Denton, 2000; Hall et al., 2015;

Jackson et al., 2018), B) generalized glaciologic setting in McMurdo Sound (Anderson et al., 2017; Denton and Marchant,

2000; Hall et al., 2015; Hall et al., 2010a; Jackson et al., 2018; Kellogg et al., 1990; McKay et al., 2016; Stuiver et al.,

1981), ice core accumulation rates (gray lines) and paleotemperatures (black line) from the C) Taylor Dome (TD2015 timescale [Baggenstos et al., 2018; Steig et al., 2000]) and D) WAIS Divide (Buizert et al., 2015; Cuffey et al., 2016), and E) eustatic sea-level (black line) and change in sea-level (gray line) (Lambeck et al., 2014). Light gray box shows timing of maximum ice extent in McMurdo Sound and dark gray strip shows the MWP-1A sea-level rise event (Lambeck et al., 2014).

62

Marine records suggest that deglaciation of the Ross Sea Embayment was complex and possibly asynchronous, with the central and eastern portions retreating earlier than the western Ross Sea (Bart et al., 2017; Greenwood et al., 2018; Halberstadt et al., 2016).

While our dataset does not directly address the deglacial dynamics of the western Ross

Sea, the compiled radiocarbon chronology provides a maximum age (12.3 cal. ka) of ice thinning in McMurdo Sound related to grounding line retreat in the western Ross Sea.

Recent evidence for Early Holocene re-advance of EAIS outlet glaciers in the western Ross

Sea (Greenwood et al., 2018) suggests re-organization of ice flow north of McMurdo

Sound during deglaciation; therefore, the radiocarbon chronology of Ross Sea drift does not directly relate to outlet glacier thinning histories along the Scott Coast as implied by

Lee et al. (2017). Additionally, the persistence of grounded ice at or near its maximum elevation and position until after Meltwater Pulse-1A (14.5 – 14.0 ka) (Lambeck et al.,

2014), suggests that this sector of the Antarctic Ice Sheet in the western Ross Sea was not a significant contributor to that sea level rise event.

2.6 Conclusions

We reconstructed the geometry and timing of a grounded ice sheet across McMurdo

Sound during the local Last Glacial Maximum to address competing hypotheses about ice sheet behavior and interactions in the Ross Sea Embayment. Our data show that grounded ice sourced from southern EAIS outlet glaciers flowed north and westward into southern

McMurdo Sound depositing erratic-rich Ross Sea drift. The elevation of Ross Sea drift limits and orientation of moraines suggest the local ice sheet surface sloped – and therefore

63 ice flowed – westward around volcanic islands and peninsulas during the local LGM.

Depressed ice accumulation rates caused local alpine glaciers, including Koettlitz Glacier, to retreat during the LGM, resulting in only minor contributions of local ice masses to the grounded ice sheet. The radiocarbon chronology of Ross Sea drift in McMurdo Sound, both new data from this study and data from prior publications, brackets maximum ice extent between 12.3 and 19.6 cal. ka., and, by extension, provides a maximum age of ice thinning in McMurdo Sound of 12.3 cal. ka. The persistence of grounded ice in McMurdo

Sound during and after Meltwater Pulse 1A suggests that this sector of the Antarctic ice sheet in the western Ross Sea was not a significant contributor to that rapid sea level rise event.

64

CHAPTER THREE: Cosmogenic nuclide exposure age scatter in McMurdo Sound,

Antarctica records Pleistocene glacial history and polythermal ice sheet processes

Abstract

The preservation of cosmogenic nuclides that accumulated during periods of prior exposure, but were not subsequently removed by erosion or radioactive decay, complicates interpretation of exposure, erosion, and burial ages used for a variety of geomorphological applications. In glacial settings, cold-based, non-erosive glacier ice may fail to remove inventories of inherited nuclides in glacially transported material. As a result, individual exposure ages can vary widely across a single landform (e.g. moraine) and exceed the expected or true depositional age. The surface processes that contribute to this inheritance signal remain poorly understood, thus limiting interpretations of cosmogenic nuclide studies in glacial environments.

Here, I present a compilation of new and previously published exposure ages of clasts using multiple lithologies in Last Glacial Maximum (LGM) and older Late Pleistocene glacial sediments in McMurdo Sound, Antarctica. Unlike most Antarctic exposure chronologies, we are able to compare exposure ages of LGM sediments directly against a robust independent radiocarbon chronology of fossil algae from the same sedimentary unit that brackets the age of the local LGM between 12.3 and 19.6 ka. Cosmogenic exposure ages vary by lithology, suggesting that bedrock source and surface processes prior to, during, and after glacial entrainment explain scatter. 10Be exposure ages of quartz in granite, sourced from the base of the stratigraphic section in the Transantarctic Mountains,

65 are scattered but younger, suggesting that clasts entrained by sub-glacial plucking can generate reasonable apparent exposure ages. 3He exposure ages of pyroxene in Ferrar

Dolerite, which outcrops above outlet glaciers in the Transantarctic Mountains, are older, which suggests that clasts initially exposed on cliff faces and glacially entrained by rock fall carry greater nuclide inheritance. 3He exposure ages of olivine in basalt from local volcanic bedrock in McMurdo Sound contain many excessively old ages, but also have a bimodal distribution with peak probabilities that slightly pre-date and post-date the local

LGM; this suggests that glacial clasts from local bedrock record local landscape exposure.

With the magnitude and geological processes contributing to age scatter in mind, we examine exposure ages of older Late Pleistocene glacial deposits and suggest that most extensive Late Pleistocene ice sheet inundated McMurdo Sound during Marine Isotope

Stage 8. These results underscore how surface processes operating in the Transantarctic

Mountains are preserved and manifested in the cosmogenic nuclide inventories held in

Antarctic glacial sediments.

66

3.1 Introduction

3.1.1 The problem of inherited cosmogenic nuclides

Cosmogenic nuclides revolutionized geomorphologists’ ability to quantitatively constrain and describe Earth surface processes. The basic premise of this method – that in situ cosmogenic stable nuclides (such as 3He, 21Ne) and radionuclides (including 10Be, 14C,

26Al, 36Cl) accumulate in near-surface materials through exposure to cosmic radiation – can be applied across a wide range of environments and timescales to quantify erosion rates

(Portenga and Bierman, 2011), date landforms and surfaces (Wells et al., 1995), reconstruct changes in climate over millions of years (Bierman et al., 2016; Shakun et al., 2018), among many other applications.

Despite the wealth of information unlocked by cosmogenic nuclides, a consistent feature of this method continues to confound geomorphologists: inherited nuclides.

Inherited nuclides are those that accumulated in surface material during periods of prior exposure, but were not subsequently removed by erosion or through the decay of radioactive nuclides, and can produce results (i.e., surface exposure ages, erosion rates, burial ages) that are difficult to interpret. Nuclide inheritance affects nearly all geomorphologic and climatic settings – it complicates the interpretation of alluvial fan ages in deserts (Owen et al., 2011), river terrace exposure ages to determine canyon incision rates (Cook et al., 2009), shore platform erosion rates in coastal settings (Trenhaile, 2018) and glacial chronologies using exposure ages of boulders on moraines or glacially sculpted bedrock (Briner et al., 2014; Corbett et al., 2016; Young et al., 2016).

67

Inherited nuclides are prevalent in glacial sediments – from alpine glacier moraines on the Tibetan plateau (Heyman et al., 2011) to glacial erratics transported by polar ice sheets (Briner et al., 2014; Corbett et al., 2016). Surface exposure ages of glacial deposits are assumed to record the duration of exposure following emplacement and glacial retreat

(Dunai, 2010), allowing a reconstruction of former ice extent through time. In practice, however, exposure chronologies commonly yield highly scattered exposure ages that vary widely across a single, contemporaneously deposited landform (i.e., moraine) and often greatly exceed the expected or true depositional age (Balco, 2011; Hein et al., 2014;

Heyman et al., 2011). This scatter is likely due to inherited nuclides that accumulated during periods of prior exposure but were not subsequently removed by non-erosive, cold- based glacial ice (Corbett et al., 2016; Hein et al., 2014).

Cold-based glaciation occurs where ice thickness and basal ice temperatures are insufficient to produce meltwater and basal sliding, both of which facilitate subglacial erosion, and therefore removal of inherited nuclides in formerly exposed surfaces (Bennett and Glasser, 1996). The basal thermal regime below ice can be highly heterogeneous, resulting in selective erosion of the landscape (Briner et al., 2006; Sugden et al., 2005). As a result, glacial sediments can be entrained, transported, and deposited by both warm-based and cold-based glacial ice, and ultimately produce scattered cosmogenic surface exposure ages (Heyman et al., 2011). Deciphering past glacial conditions, entrainment processes, and transport trajectories for clasts in glacial sediment is often difficult as clast source areas and transport histories are unknown.

68

The problem of inherited nuclides and exposure age scatter is exacerbated in

Antarctica, where large polythermal ice sheets (Sugden and Denton, 2004; Sugden et al.,

2006), hyper-arid polar desert climate conditions (Lewis et al., 2008), and extremely low sub-aerial erosion rates (<1 m/Ma) (Marchant et al., 2013) have persisted since the Miocene

(Lewis and Ashworth, 2016; Shakun et al., 2018). Terrestrial glacial sediments in the ice- free areas of Antarctica (Baroni and Hall, 2004; Bromley et al., 2012; Bromley et al., 2010;

Christ and Bierman, 2019; Denton and Marchant, 2000; Hall et al., 2015; Hall et al., 2000;

Spector et al., 2017; Todd et al., 2010), preserve direct geological constraints on former ice sheet elevation, and therefore, ice sheet volume, from Holocene to possibly Miocene time

(Balco et al., 2013; Kong et al., 2010). Cosmogenic nuclides are often the only method for constraining a numerical chronology of Antarctic glacial deposits; radiocarbon dating is often unavailable due to the absence of organic remains and/or the sediments are >50,000 years old. Stable, hyper-arid polar climate conditions and low erosion rates has resulted in long-lived exposure of rock surfaces that are later covered and/or entrained by cold-based glacier ice that fails to erode and remove inherited nuclides (Hein et al., 2014; Stone et al.,

2003). At the same time, large outlet glaciers and ice streams have deeply eroded and modified parts of the landscape (Sugden et al., 2005) and transported eroded sediments into the ocean and terrestrial margins (Anderson, 1999; Anderson et al., 2014). As a result, many cosmogenic nuclide exposure chronologies from Antarctic terrestrial glacial deposits contain scatter with exposure ages that can span several hundred thousand years for individual landforms.

69

3.1.2 McMurdo Sound: a unique setting to understand nuclide inheritance

Here, I focus on cosmogenic nuclide analyses from McMurdo Sound, Antarctica

(78°S, 165°E), as it is uniquely suited to study scatter in exposure age datasets from glacial sediments (Figure 3.1A). McMurdo Sound hosts one of the largest ice-free areas in

Antarctica and contains well-preserved Late Pleistocene glacial sediments on volcanic islands and coastal valleys of the Transantarctic Mountains (Christ and Bierman, 2019;

Denton and Marchant, 2000; Hall et al., 2015; Jackson et al., 2018; Stuiver et al., 1981)

(Figure 3.1, Figure 3.2). The topographic prominence of the Royal Society Range prevents the East Antarctic Ice Sheet (EAIS) from directly discharging into McMurdo Sound. As a result, during glacial periods when southerly EAIS outlet glaciers expanded into the Ross

Sea, grounded ice circumvented volcanic features and overflowed into McMurdo Sound from the east (Christ and Bierman, 2019; Denton and Marchant, 2000; Hall et al., 2015)

(Figure 3.1).

These incursions of grounded ice into McMurdo Sound transported and deposited lithologies from the entire stratigraphic column of the Transantarctic Mountains and other rocks from the East Antarctic interior (Figure 3.1B). The diversity of lithologies in glacial sediments in McMurdo Sound provided targets for multiple cosmogenic nuclides

10 3 3 ( Bequartz, Hepyroxene, Heolivine) to calculate exposure ages; yet, results still yield a complex exposure, and possibly depositional history (Anderson et al., 2017; Brook et al., 1995).

70

Figure 3.1 Bedrock geology and stratigraphy of Southern Victoria Land, Antarctica

A. Regional overview map showing the generalized geologic units of Southern Victoria Land, Antarctica. Blue arrows show the general flow directions of outlet glaciers when expanded into the Ross Sea. Base map uses data from the 2013

International Bathymetric Chart of Southern Ocean and the Reference Elevation Mosaic of Antarctica generated by the

Polar Geospatial Center (Howat et al., 2018); bedrock geology adapted from (Grindley and Laird, 1969; Warren, 1969).

B. Generalized stratigraphy of Southern Victoria Land, Antarctica adapted from (Fitzgerald, 2002)

71

There are several advantages to examining exposure ages of glacial deposits in

McMurdo Sound. First, exposure ages of clasts in the LGM glacial deposit, known locally as Ross Sea drift, can be compared against an independently constrained radiocarbon chronology of fossil algae embedded in moraines and ice-marginal sediments from the same deposit (Christ and Bierman, 2019; Hall and Denton, 2000; Hall et al., 2015; Jackson et al., 2018) (Figure 3.2). This radiocarbon chronology, which consists of 236 individual calibrated radiocarbon ages, constrains when grounded ice was present at or near its local

LGM extent between 12.3 and 19.6 calibrated thousands of years before present (cal. ka)

(Christ and Bierman, 2019; Hall and Denton, 2000; Hall et al., 2015; Jackson et al., 2018).

Second, the different lithologies in glacial sediments from McMurdo Sound can be tied to their bedrock sources in stratigraphic sections of the Transantarctic Mountains. By examining a compilation of new and previously published exposure ages of different lithologies in Ross Sea drift, we are able quantify the magnitude of exposure age scatter and investigate potential surface processes that contribute to, prevent, or reduce nuclide inheritance across the Transantarctic Mountains and western Ross Sea. With the magnitude and pattern of exposure age scatter known, we then examine older Late Pleistocene glacial deposits that delineate maximum ice extent in McMurdo Sound, but are as of yet unmapped and undated.

3.2. Study area

McMurdo Sound (78°S, 165°E) hosts an ice-free area geographically bounded by the Royal Society Range and Late Cenozoic volcanic islands and peninsulas (Figure 32).

In the Royal Society Range, local alpine glaciers do not extend all the way to coast, leaving

72 the valley mouths ice-free. The northern faces of several volcanic features – including

Mount Morning, Mount Discovery, and Black Island – and the entirety of Brown Peninsula are free of glacial ice. These ice-free conditions are caused by the topographic prominence of the Royal Society Range that prevents the EAIS from flowing directly into McMurdo

Sound. In addition, southerly adiabatic winds descending over Mount Morning, Mount

Discovery, and Minna Bluff create a regional ablation zone that prevents the accumulation of glacial ice (Denton and Marchant, 2000; Glasser et al., 2014).

Below 775 m elevation, these ice-free areas are covered by glacial sediments comprised of local McMurdo Volcanic rocks and erratic lithologies sourced from the

Transantarctic Mountains and interior East Antarctica. These erratic rocks include Granite

Harbor Intrusives, Ferrar Dolerite, Koettlitz Group metamorphic rocks, Beacon Sandstone, and others (Figure 3.1) (Christ and Bierman, 2019; Denton and Marchant, 2000; Talairico et al., 2012). Earlier mapping efforts identified at least two different glacial sedimentary units that contain erratic lithologies in McMurdo Sound. At lower elevations (< 510 m),

Ross Sea drift is characterized by a high concentration of erratic lithologies and un- weathered clasts that lack iron-staining or ventifaction (Christ and Bierman, 2019; Hall et al., 2015; Hall et al., 2000). In general, the maximum elevation of Ross Sea drift decreases from east to west across McMurdo Sound, from 510 m on eastern Mount Discovery and

490 m on northern Black Island to less than 250 m along the coast of the Royal Society

Range (Christ and Bierman, 2019; Denton and Marchant, 2000; Hall et al., 2015). This trend in elevation, combined with the radiocarbon chronology of fossil algae in ice- marginal deposits, suggests that ice flowed from east to west across McMurdo Sound

73 during the local LGM between 12.3 and 19.6 cal. ka (Christ and Bierman, 2019; Denton

10 and Marchant, 2000; Hall et al., 2015) (Figure 3.2). An elevation transect of Bequartz exposure ages of granite boulders along eastern Mount Discovery indicates that ice thinned from its maximum extent no earlier than 14.1 ka, and reached present conditions 7.6 ka

(Anderson et al., 2017).

At higher elevations in McMurdo Sound, other glacial sediments remain largely unmapped and undated. These glacial deposits are characterized by a lower concentration of erratic lithologies (relative to Ross Sea drift) and are presumably older, as indicated by enhanced surface weathering characteristics, including chemically-altered surfaces, and ventifacted and cavernously weathered boulders (Denton and Marchant, 2000). Earlier reconnaissance mapping on Mount Discovery indicated these deposits extend up to 770 m elevation (Denton and Marchant, 2000; Stuiver et al., 1981). Other similar, higher elevation glacial deposits exist above Ross Sea drift in the Royal Society Range and, while they have not been mapped in detail, have some exposure age constraints (Brook et al., 1995).

74

Figure 3.2: Compiled cosmogenic nuclide exposure ages and radiocarbon sample locations in McMurdo Sound.

75

Colored boxes show the number of samples from each location color-coded according to the symbology listed in the legend. The extent of Ross Sea drift from (Christ and Bierman, 2019) is shown. All new samples and previously published data (^Brook et al., 1995, *Anderson et al., 2017) were recalculated using the same scaling scheme (Stone, 2000) and shown here as the exposure age ± 1σ internal uncertainty from previous studies. For detailed view and exposure ages on

Mount Discovery, see Figure 3.4. The location of the AND-1B drill core (McKay et al., 2016; Naish et al., 2009) is shown for reference. Base map: Landsat Image Mosaic of Antarctica and 500 m contours derived from DEM(s) created by the Polar Geospatial Center from DigitalGlobe, Inc. imagery (Howat et al., 2019).

3.3 Material and methods

3.3.1 Glacial geologic mapping

I integrate the detailed mapping and descriptions of Ross Sea drift (Christ and

Bierman, 2019; Denton and Marchant, 2000; Hall et al., 2015) with higher elevation glacial deposits on Mount Discovery (hereafter referred to as the Upper Discovery deposit).

Mapping of Ross Sea drift (Christ and Bierman, 2019) focused on several field sites on

Mount Discovery, Black Island, and Brown Peninsula. The Upper Discovery deposit was mapped during helicopter missions to several sites on Mount Discovery at elevations between 450 and 775 m. GPS data points were collected using a handheld Garmin GPS unit (precision <10 m) on moraines and glacial limits. We complemented on-the-ground field surveys with analysis of aerial photographs, high-resolution (< 1.85 m/pixel)

DigitalGlobe, Inc. imagery, and a stereo-photogrammetric digital elevation model (2 m/pixel) created by the Polar Geospatial Center.

76

Figure 3.3 Compiled cosmogenic exposure ages and radiocarbon chronology of Ross Sea drift.

A) The timing of Marine Isotope Stages since 130 ka (Lisiecki and Raymo, 2005). Probability density functions of: A) calibrated radiocarbon ages of fossil algae from ice marginal sediments in Ross Sea drift that bracket the timing of the local LGM (Christ and Bierman, 2019; Hall and Denton, 2000; Hall et al., 2015; Jackson et al., 2018); B) all exposure

3 ages of boulders in Ross Sea drift (this study; (Anderson et al., 2017; Brook et al., 1995); C) Heolv in basalt (Brook et

3 10 al., 1995); D) Hepyx in Ferrar Dolerite (this study); and E) Beqtz in granite (this study; Anderson et al., 2017; Brook et al., 1995). Gray box outlines the timing of the local LGM established by the radiocarbon chronology. Peak probability modes are labeled in each panel of exposure ages. Color corresponds to bedrock geology symbology in Figure 3.1.

77

3.3.2 Sample collection

During geomorphologic mapping campaigns on the volcanic islands of McMurdo

Sound, erratic clasts of Granite Harbor Intrusives (granite) and Ferrar Dolerite were

10 10 3 3 collected for cosmogenic Be in quartz ( Beqtz) and He in pyroxene ( Hepyx) exposure dating, respectively, from moraine crests and the limits of Ross Sea drift (Christ and

Bierman, 2019) and the Upper Discovery deposit. Clasts that were clearly perched, either on top of other clasts or on the surface of glacial sediment, were collected over those embedded into sediment to avoid ‘young’ exposure ages due to exhumation. Samples were collected from local topographic highs (i.e. moraine ridge crests) to avoid local shielding, clast overturning, rotation, exhumation, or downslope movement. The GPS coordinates, topographic shielding, clast orientation, clast dimensions, weathering characteristics, and geomorphic setting were recorded for each sample. Samples from large boulders were collected by plucking the upper 10 cm of rock using a hammer-drill and wedges and shims.

Whole-rock samples of smaller cobble-sized clasts were taken.

3.3.3 Cosmogenic nuclide analyses

The upper ~5 cm of each sample was cut using a rock saw at Boston University.

Samples were shipped to the Cosmogenic Nuclide Dating Group at the Lamont-Doherty

Earth Observatory for further sample processing and in situ nuclide extraction.

3.3.3.1 Cosmogenic 3He in pyroxene

Samples were crushed to 63-125 μm to separate pyroxene mineral grains, rinsed thoroughly with deionized (DI) water to remove fines, then leached in a 10% phosphoric acid solution overnight to remove oxidation and other contaminants from mineral surfaces.

78

The samples were then density-separated in a centrifuge using sodium polytungstate heavy liquid at a density of 3.0 g∙cm-3; the sinking fraction was rinsed with DI water and dried overnight. Magnetic minerals were removed using a Frantz magnetic separator set between

0.4 and 0.5 amps. The non-magnetic fraction was then leached in a 2% HNO3/2% HF solution on a shaker table overnight, then rinsed, dried, and packed in aluminum foil for analysis. Helium isotopes were measured with the LDEO MAP 215–50 noble gas mass

3 3 spectrometer using an air standard. He in pyroxene ( Hepyx) exposure ages were calculated the version 3 of the online exposure age calculator hosted by the University of Washington

(https://hess.ess.washington.edu/) assuming density of 2.9 gcm-3, corrected for shielding and thickness, and used the St schaling scheme (Stone, 2000).

3.3.3.2 Cosmogenic 10Be in quartz

Samples were crushed to 500-125 μm, then froth flotation was used to separate the quartz grains. Quartz separates were purified by leaching in a HNO3/HF acid solution, and

10 then Beqtz was extracted from purified quartz, both according to standard protocols in the

Cosmogenic Nuclide Dating Group. 10Be/9Be ratios were measured at Lawrence Livermore

10 National Laboratory. Beqtz exposure ages were calculated using the version 3 of the online exposure age calculator hosted by the University of Washington

(https://hess.ess.washington.edu/) assuming a rock density of 2.7 g cm-3 corrected for shielding and thickness, and the St scaling scheme (Stone, 2000).

3.4 Compiled exposure chronology

New and previously published cosmogenic nuclide concentrations from previous studies in McMurdo Sound were compiled from the ICE-D database (antarctica.ice-d.org).

79

Updated exposure ages for all samples were calculated version 3 of the online exposure age calculator hosted by the University of Washington (https://hess.ess.washington.edu/) using the St scaling scheme (Stone, 2000). Only samples collected from maximum extent of Ross Sea drift were selected from previous studies; we do not include exposure ages explicitly used for deglacial thinning chronologies. All exposure ages of older glacial deposits, however, were included as geomorphologic mapping of these deposits is less

10 3 3 3 complete. Only Beqtz in granite, He in olivine ( Heolv) in basalt, and Hepyx in dolerite are considered here; too few exposure ages in other erratic lithologies (such as gneiss and

26 sandstone) and Alqtz exist to be compared against the larger dataset. Additionally, we

3 exclude previously published Hequartz exposure ages due to diffusion uncertainties inherent to that specific nuclide system (Brook et al., 1995).

3.4. Results

3.4.1 Glacial geologic mapping

The Upper Discovery deposit reaches a maximum elevation of 776 m and decreases in elevation from east to west (Figure 3.4). The limit of the upper glacial deposit is nearly continuous and well-defined across the northern face of Mount Discovery, and is draped onto parasitic cinder cones, volcanic vents, and intervening valleys. Several large windswept snow fields and glaciers presently cover and cross over the glacial limit in a few locations. The elevation of the glacial limit descends to ~450 m elevation on western

Mount Discovery above Koettlitz Glacier where it becomes poorly defined due to disturbance from solifluction. Between the limits of the Upper Discovery deposit and Ross

Sea drift on Mount Discovery, no clear moraines or intermediate glacial limits were found.

80

Above the Upper Discovery deposit, no erratic lithologies are present – only local volcanic rocks exist.

The Upper Discovery deposit is characterized by low-relief moraines (< 1 m) where clasts are mostly embedded in the soil matrix. Large boulders (>1 m) are rare, and most clasts are cobble-sized. Observed erratic lithologies included Ferrar Dolerite and Granite

Harbor Intrusives. Clasts are characterized by varnished surfaces, ventifaction, and cavernous weathering. Patterned ground is well developed on the surface, and, unlike Ross

Sea drift, thermokarst topography and features are absent along much of the upper elevation limit. Solifluction lobes are present along the northwestern face of upper Mount

Discovery.

3.4.2 Cosmogenic nuclide analyses

3 3.4.2.1 New Hepyx in Ferrar Dolerite exposure ages

3 Hepyx exposure ages of Ferrar Dolerite generally sort according to the stratigraphy of mapped glacial deposits, but display wide scatter within these stratigraphic groupings

3 (Table 3.1). Four of the five Hepyx exposure ages of Ferrar Dolerite in Ross Sea drift on eastern Mount Discovery and Brown Peninsula cluster between 27.51.0 ka to 53.31.08 ka, while the remaining exposure age from a sharp-crested moraine at on Brown Peninsula

3 was much older, 382.810.9 ka. Hepyx exposure ages (n=4) from the Upper Discovery deposit were older; three ages are between 245.68.1 ka and 262.513.9 ka; the other exposure age was older (366.2 11.6 ka).

81

10 3.4.2.2 New Beqtz in Granite exposure ages

10 Beqtz exposure ages of granite sort according to stratigraphic age and display less scatter

3 10 than Hepyx exposure ages (Table 3.2). Beqtz exposure ages of clasts from Ross Sea drift on Mount Discovery and Brown Peninsula had two younger and similar ages (13.40.3 ka,

13.60.3 ka); and two older ages (21.60.6 ka, 37.60.7 ka). A single exposure age of a cavernously weathered boulder in the Upper Discovery deposit yielded an age of 153.32.0 ka.

3.4.3 Compiled cosmogenic nuclide ages

3.4.3.1 Ross Sea drift

Exposure ages of Ross Sea drift (n = 38) vary widely, ranging from 6.81.4 ka to

382.810.9 ka and skew towards older ages (skewness: 2.88). A probability density function of all exposure ages returns primary, secondary, and tertiary modes at 13.5 ka,

10.4 ka, and 21.4 ka, respectively (Figure 3.3). Although the primary mode falls within the radiocarbon-constrained timeframe of the local LGM, only 7 of the 38 exposure ages from

Ross Sea drift fall within this window of maximum ice extent between 12.3 and 19.6 ka.

Another 7 exposure ages are younger than the local LGM (<12.3 ka); 24 exposures ages are older than the local LGM (>19.6 ka). Most of these excessively old ages are within the last glacial period (MIS 2 - 4), but several exposure ages date to the last interglacial (MIS

5) and two are much older (211.61.4 ka [MIS 7], and 262.513.9 ka [MIS 11]).

82

Figure 3.4: Glacial geology of Mount Discovery.

A-D) Annotated field photos from the upper limit of the Upper Discovery deposit. Inset detail pictures of exposure age samples included in panels B and C. E) Integrated glacial geomorphologic and bedrock geologic map of Mount Discovery

83 with plotted exposure ages from Ross Sea drift and the Upper Discovery deposit (Anderson et al., 2017). Photo locations shown on map by corresponding panel letter. Bedrock geology adapted from (Kyle, 1990; Wright-Grassham, 1987). Base map: 100 m contours and hillshade derived from a digital elevation model created by the Polar Geospatial Center from

DigitalGlobe, Inc. imagery (Howat et al., 2019).

3 Table 3.1: Hepyroxene exposure ages of Ferrar Dolerite

84

10 Table 3.2: Bequartz exposure ages of granite

85

Exposure ages of Ross Sea drift sort and vary according to lithology and nuclide

3 (Figure 3.3). Heolv in basalt exposure ages (Brook et al., 1995, n = 23) vary widely (range:

6.7 – 211 ka, skewness: 2.34), but display two modes, 9.6 ka and 25.0 ka, before and after

3 the local LGM, respectively. Only three Heolv basalt exposure ages are within the local

3 LGM timeframe. Hepyx in dolerite exposure ages (n = 5) are consistently older (range: 27.5 to 382.8 ka; skewness: 1.33); none of these ages fall within the timing of the local LGM.

10 Beqtz in granite exposure ages are scattered (range: 10.5 to 50.1 ka, skewness: 1.69) and tend to be younger. The primary (13.4 ka) and secondary modes (15.1 ka) coincide with the latter half of the local LGM, while an additional secondary mode slightly pre-dates maximum ice extent (21.1 ka).

4.3.2 Upper-elevation glacial sediments

Exposure ages (n=13) from the higher elevation glacial deposits (including the

Upper Discovery deposit) are highly scattered, but nearly all pre-date the onset of the last glacial period (MIS 4) (Figure 2.5A, B). One age fell within the last glacial period during

MIS 3 (53.31.8 ka). Several ages clusters coincide with previous glacial periods MIS 6 (n

=4, 133.0 – 178.8 ka) and MIS 8 (n = 4, 219.3 – 262.9 ka), with several ages dating to MIS

10 (366.211.6 ka) and MIS 12 (455.327.6 ka). Three ages date to earlier interglacial periods, including two ages during MIS 5 (84.62.2 ka, 87.55.3 ka) and one age during

MIS 7 (219.35.9 ka).

86

Figure 3.5: Late Pleistocene glacial chronology of McMurdo Sound and paleoclimate since 500 ka.

A) Marine Isotope Stages (Lisiecki and Raymo, 2005); blue boxes shown in each panel mark Late Pleistocene glacial periods. B) Exposure ages (±1σ) of Ross Sea drift (black), pre-LGM glacial deposits (blue), and the Upper Discovery deposit (orange) versus elevation. Gray box with dotted line delineates proposed age of Upper Discovery deposit. C)

Paleoenvironmental reconstruction of McMurdo Sound from the AND-1B drill core (McKay et al., 2012) where I: sub- glacial, P: ice-proximal, D: ice-distal, M: open marine. Gray-patterned boxes mark intervals of sub-glacial till. Curvy lines delineate glacial erosional unconformities. D) Paleotemperature data from the Vostok (red) and EPICA Dome C

(black) ice cores (Augustin et al., 2004; Petit et al., 1999). E) Global sea level reconstruction based on the benthic marine stack (Spratt and Lisiecki, 2016).

87

3.5 Discussion

3.5.1 Scattered exposure ages of Ross Sea drift

The range of exposure ages, albeit scattered, that date to within the last glacial period (Anderson et al., 2017; Brook et al., 1995), and the narrow range of calibrated radiocarbon ages (12.3 – 19.6 cal. ka) (Hall and Denton, 2000; Hall et al., 2015; Jackson et al., 2018) confirm that Ross Sea drift is indeed the youngest glacial deposit recording ice sheet expansion into McMurdo Sound. The compiled cosmogenic nuclide exposure chronology derived from multiple lithologies is highly scattered, and a majority of individual exposure ages exceed the timeframe of local LGM ice extent established by the radiocarbon chronology. While many ages are ‘too old’ and suggest that many clasts carry an inventory of inherited nuclides, most of the exposure ages (with 3 exceptions) date to the early Holocene (MIS 1) or last glacial period (MIS 2 - 4). Upper elevation, weathered deposits yield much older Pleistocene ages (MIS 5 – 12) that are incompatible with an

LGM depositional age.

This exposure age dataset underscores the difficulty of using surface exposure dating in Antarctica to constrain precise ages of individual landforms or glacial deposits from the LGM. The most extreme example of scatter comes from two exposure ages from the same moraine ridge on northern Brown Peninsula. This sharp-crested moraine ridge is independently dated to 17.9 cal. ka (Christ and Bierman, 2019); yet, nearly adjacent clasts yield exposure ages that are both ‘too old’ and widely different. A perched granite cobble produced an exposure age of 21.60.6 ka (ACX-14-08) and a nearby dolerite cobble has an exposure age of 382.810.9 ka (ACX-14-05), the oldest of any sample from Ross Sea

88 drift. Elsewhere in McMurdo Sound, there is similar exposure age scatter across a single landform. On eastern Mount Discovery (512 m), the glacial limit is marked by a stark lithological contrast between erratic-rich Ross Sea drift (~75% of surface lithologies) below volcanic-rich (only 20% erratic lithologies) glacial deposits above (Anderson et al.,

2017; Christ and Bierman, 2019; Denton and Marchant, 2000). Here, granite exposure ages coincide with the end of the local LGM in McMurdo Sound (13.4 to 15.1 ka, n = 3; this study; (Anderson et al., 2017)), but exposure ages of dolerite from the same location are consistently older (27.5 to 45.5 ka, n = 3). Clearly, exposure ages from individual landforms from the same glacial deposit in McMurdo Sound contain clasts with significantly different exposure histories.

3.5.2 Surface processes contributing to exposure age scatter

What is causing the wide scatter in exposure ages of Ross Sea drift? We suggest that scatter is related to bedrock source and that specific surface processes prevent, enhance, or reduce nuclide inheritance prior to and during glacial entrainment and

10 deposition. Exposure age scatter sorts according to lithology and nuclide: granite Beqtz

3 exposure ages are younger, but scattered, dolerite Hepyx ages are consistently too old, and

3 basalt Heolv ages slightly pre-date and post-date the local LGM. These patterns can be explained by examining how clasts are glacially entrained and incorporated into Ross Sea drift.

3.5.2.1 Sub-glacial plucking reduces nuclide inheritance

Granite clasts in Ross Sea drift, sourced from Granite Harbor Intrusives, are likely entrained by sub-glacial plucking, and therefore produce younger, less-scattered exposure

89 ages. Granite Harbor Intrusives comprise the base of the stratigraphic section for much of the Transantarctic Mountains in Southern Victoria Land (Figure 3.1). The large EAIS outlet glaciers that bisect the Transantarctic Mountains south of McMurdo Sound – Byrd

Glacier, Mullock Glacier, and Skelton Glacier – incised deep troughs that cut into Granite

Harbor Intrusives (Figure 3.1). Ice sheet models simulate that during glacial periods these outlet glaciers not only thicken and expand into the Ross Sea, but are also wet-based and erode large volumes of sediment (Golledge et al., 2013; Golledge et al., 2014). Rock surfaces at the bottom of outlet glacier troughs are up to 1000 m below sea level and could not experience exposure to cosmic rays. In addition, the EAIS and outlet glaciers in the Transantarctic Mountains have been stable and persistent throughout the Pleistocene, preventing prior exposure of the bedrock troughs (Kaplan et al., 2017). Taken together, this means that Granite Harbor Intrusives plucked from the base of outlet glacier troughs should not contain cosmogenic nuclides inherited from periods of exposure prior to LGM exposure in Ross Sea drift.

3.5.2.2 Cliff exposure, rockfall, and supraglacial transport increases nuclide inheritance

Old exposure ages of several granite and all dolerite clasts in Ross Sea drift may reflect prior cliff exposure and entrainment by rockfall along outlet glacier margins.

Granite Harbor Intrusive rocks are exposed in cliff faces above the glacier surface closer to the coast and along outlet glacier margins in the Transantarctic Mountains (Figure 3.1).

During previous interglacial periods or glacial low-stands when outlet glaciers were thinner

(Jones et al., 2017), similar exposure on granite cliff faces would have accumulated cosmogenic nuclides prior to later glacial entrainment. The effect of prior cliff exposure

90 could be more pronounced for Ferrar Dolerite because these rocks form many of the peaks and cliff faces above glacier surfaces in the Transantarctic Mountains at elevations >1000 m (Figure 3.1). Prolonged cliff exposure combined with low erosion rates of ~<1m Ma-1

(Marchant et al., 2013) at high elevations – where the cosmogenic nuclide production rate is higher (Stone, 2000) – will increase cosmogenic nuclide inventories in dolerite more

3 than granite clasts. Additionally, Hepyx is a stable cosmogenic nuclide that does not decay during burial (Dunai, 2010); this further increases the possibility of nuclide inheritance in dolerite clasts.

Despite increased outlet glacier surface elevations during glacial periods, cliff faces may not be subjected to wet-based glacial erosion due to insufficient ice thickness.

Selective linear erosion below polythermal ice sheets can lead to wet-based, erosive glacial conditions in outlet glacier troughs, but cold-based, non-erosive glacial conditions along valley walls (Fogwill et al., 2004; Stone et al., 2003; Sugden et al., 2005). Clasts sourced from rockfall or reworked from older ice marginal sediments in these regions of cold-based ice will not be sufficiently eroded to remove inherited nuclides. South of the McMurdo Dry

Valleys, Ferrar Dolerite may be particularly susceptible to cold-based, non-erosive glacial transport because many high-elevation cliff faces are exposed over thin alpine and valley glaciers that are tributaries to larger outlet glaciers. Dolerite clasts sourced from these areas likely carry inherited nuclides that are not removed by glacial erosion.

When rockfalls deliver granite or dolerite from cliff faces onto glacier surfaces below, inherited nuclides still remain in some clasts and supraglacial transport may cause additional nuclide inheritance. While rockfall onto glaciers may shatter large volumes of

91 rock and expose fresh surfaces (Mackay et al., 2014) some clasts may retain the original surface exposed on the cliff face. Clasts that land directly onto the ice surface can be exposed during supraglacial transport, where additional nuclides accumulate (Swanger et al., 2011). Rockfall that does not land directly on the glacier surface, but accumulates along outlet glacier margins, could be subjected to additional exposure prior to a glacial high- stand that entrains these clasts.

3.5.2.3 Ice sheet reworking of local volcanic rocks causes and prevents prior exposure

3 Heolv exposure ages of basalt clasts, sourced from the local McMurdo Volcanic

Suite, record local landscape exposure prior to and following ice sheet advance (Brook et

3 al., 1995). The bimodal distribution of Heolv in basalt exposure ages predates maximum ice sheet extent in McMurdo Sound (25.0 Ka) and post-dates ice thinning (9.6 ka). Clasts with inherited nuclides and excessively old exposure ages may have been initially sub- aerially exposed during a glacial low-stand prior to the LGM or the last interglacial (MIS

5). As ice advanced into McMurdo Sound during the local LGM, exposed volcanic clasts may have been reworked into Ross Sea drift. Clasts that lack previous exposure could be entrained from submarine sources that lack any previous exposure, and transported along the base of the ice sheet in McMurdo Sound. As ice thinned and down-wasted, basal clasts would be exposed last, possibly producing exposure ages younger than the onset of ice thinning (12.3 ka).

3.5.3 New insights from inherited nuclides and exposure age scatter

In former and active ice sheet environments, cosmogenic nuclides in glacial sediments often do not yield simple, easy-to-interpret exposure ages; instead, cosmogenic

92 nuclides record the integrated effect of many surface processes operating on a landscape through time. Rather than focusing exclusively on the age of ice extent or retreat, cosmogenic nuclide studies of glacial landscapes can be used to disentangle how glacial sediment is entrained from bedrock source areas, transported by glacier ice, and deposited on the landscape. By taking advantage of the relatively simple geology of the

Transantarctic Mountains, an independent (albeit rare) chronology, and analysis of multiple nuclides and lithologies in Ross Sea drift, we can better quantify the magnitude of and processes responsible for exposure age scatter. While scatter in exposure ages from

Ross Sea drift prevents a precise age of maximum ice extent, it is possible to assign ages at glacial-interglacial timescales. With this information, we now consider the age of older

Late Pleistocene glacial deposits in McMurdo Sound

3.5.4 Upper Discovery deposit: a maximum glaciation in McMurdo Sound during MIS

8?

The Upper Discovery deposit marks the maximum limit of Late Pleistocene glaciation of McMurdo Sound, possibly between ~243 and 270 ka at the end of MIS 8 (243

– 300 ka). The decreasing elevation of the Upper Discovery limit from 776 m in the east to ~450 m in the west around Mount Discovery suggests a larger ice sheet in the Ross Sea overflowed into McMurdo Sound during a glacial period prior to the LGM. As there are no erratic lithologies present above this limit, the Upper Discovery deposit delineates the largest and thickest ice sheet in the western Ross Sea to inundate McMurdo Sound since

Mount Discovery formed 5.5 to 4.5 Ma (Kyle, 1990).

93

We assign an age of late MIS 8 (245 to 270 ka) to the Upper Discovery deposit with

3 consideration of exposure age scatter and nuclide inheritance. Three Hepyx exposure ages in Ferrar Dolerite from the Upper Discovery deposit cluster near the end of MIS 8 (ACX-

13-48: 262.513.9 ka, ACX-13-52: 245.6 13.9 ka, and ACX-13-68: 248.68.5 ka). While dolerite exposure ages in Ross Sea drift produce ages that are ‘too old’, they still tend to be within the last glacial period. At the LGM limit on Mount Discovery, Ross Sea drift

3 10 Hepyx ages exceed the Beqtz ages by 14.1 to 32.1 ka. If a similar magnitude of nuclide inheritance is assumed for the Upper Discovery deposit, these older exposure ages still fall

3 within MIS 8. The older Hepyx dolerite exposure age from this deposit (ACX-13-61:

366.211.6 ka) is compatible with an age of MIS 8; as observed with in Ross Sea drift,

3 10 outlier Hepyx exposure ages can be preserved in the same deposit. The Beqtz exposure age

(ACX-13-52: 153.32.0 ka, MIS 6), however, is too young. We note that this particular sample was collected above a cavernous weathering pit on a granite boulder, which clearly marks that the boulder was eroding, and thus possibly losing the original inventory of 10Be that accumulated since deposition, which would yield a younger exposure age.

A maximum ice sheet expansion during the end of MIS 8 agrees with other Late

Pleistocene glacial records in McMurdo Sound. Many of the exposure ages from pre-LGM glacial sediments at lower elevation are either younger or older than those from the Upper

Discovery deposit; this suggests that other Late Pleistocene ice sheets in McMurdo Sound were thicker than during the LGM, but thinner than during MIS 8. The Late Pleistocene stratigraphy from the AND-1B core in McMurdo Sound records ~ 5 glaciations during the last 500 ka (McKay et al., 2012). While poor core recovery in the upper part of this core

94 limits the interpretation of ice sheet behavior during the last glaciation and interglacial, deformed till recovered from the third-to-last glacial period could reflect the greater ice thickness in McMurdo Sound during MIS 8 (McKay et al., 2012).

Outside of McMurdo Sound, however, there is little evidence for increased outlet glacier elevations in the Transantarctic Mountains during MIS 8, which was globally different than other Late Pleistocene glaciations. Only one moraine deposited by upper

Taylor Glacier in Kennar Valley has an interpreted age of early MIS 8 (290 ka) (Swanger et al., 2011). Elsewhere in the McMurdo Dry Valleys or the Transantarctic Mountains, exposure ages record several fluctuations in the elevation of the ice sheet margin (Jones et al., 2017), but general stability of EAIS plateau elevation throughout the Middle and Late

Pleistocene (Kaplan et al., 2017). Exposure ages of pre-LGM glacial sediments above outlet glacier margins record glacial high-stands during MIS 6, the Early Pleistocene, and earlier, but there is no apparent evidence for elevated outlet glaciers during MIS 8 (Staiger et al., 2006, Joy et al., 2014; Storey et al., 2010, Bromley et al., 2010).

The expansion of the maximum ice sheet in McMurdo Sound during MIS 8 is confounding, as this glacial period was globally different than other Late Pleistocene glaciations. Between 243 and 270 ka, Antarctic ice cores record warmer average temperatures (Petit et al., 1999; Jouzel et al., 2007) and δ18O values from the global benthic marine stack are less depleted (Lisiecki and Raymo, 2005) than during MIS 2, suggesting warmer global temperature and reduced global ice volume. Average eustatic sea level was in fact higher between 243-270 ka (-8412 m [relative to today]) than during the LGM (-

11415m), even as a marine-based ice sheet occupied McMurdo Sound that was larger

95 than its LGM configuration (Spratt and Lisiecki, 2016). Lower sea level plays an important control on marine ice sheet stability as it reduces the buoyant force acting on grounded ice masses. This is especially true in the western Ross Sea because tectonic extension and glacial erosion along the front of the Transantarctic Mountains resulted in bathymetry more than 700 m below sea level (Tinto et al., 2019). A thicker marine-based ice sheet in

McMurdo Sound despite lower sea level and warmer temperatures during MIS 8, may reflect the spatially variable ice sheet responses to different orbital parameters during the

Late Pleistocene.

3.6 Conclusion

Exposure ages of glacial deposits in McMurdo Sound record multiple surface processes operating in the Transantarctic Mountains prior to, during, and following glacial entrainment throughout the Late Pleistocene. Ross Sea drift contains clasts that yield exposure ages during the early Holocene and last glacial period (MIS 1-4), but generally exceed the timing of local LGM ice extent (12.3 – 19.6 ka) established by an independent radiocarbon chronology. We attribute the scatter and prior exposure in this dataset to surface processes related to the stratigraphic position of different bedrock sources for glacially eroded and transported material. Granite clasts, sourced via sub-glacial plucking and/or rock fall, produce scattered exposure ages that are less likely to record prior exposure. Ferrar Dolerite clasts, sourced from exposed cliff faces and high elevation peaks in the Transantarctic Mountains above outlet glaciers, are more likely to contain inherited nuclides, due to entrainment along cold-based glacier margins. Local volcanic rocks in

96

McMurdo Sound produce exposure ages that record landscape exposure prior to ice sheet advance and following ice sheet thinning, and reflect reworking of previously exposed clasts along with clasts derived from sub-marine sources that lack any previous exposure.

3 Hepyx exposure ages of higher elevation, weathered glacial deposits on Mount

Discovery suggest that the largest Pleistocene ice sheet advanced into McMurdo Sound during the latter half of MIS 8. Although prior exposure and boulder erosion limits a precise age of such older deposits, it may be only direct geologic evidence of Antarctic ice sheet volume during MIS 8, a glacial period marked by generally warmer Antarctic temperature and higher global sea level.

97

CHAPTER FOUR: The northwest Greenland Ice Sheet during the Early Pleistocene

was similar to today

Abstract

The multi-million-year history of the Greenland Ice Sheet (GrIS) remains poorly known. Offshore, deep-ocean sedimentary archives preserve a continuous, long-term record of GrIS history, but far-field marine sediments often lack coarse grains, precluding analyses critical for understanding weathering history and sediment sources. In contrast, ice-proximal glacial marine diamict, a poorly sorted mixture of grain sizes, deposited directly from ice onto continental shelves, provides direct but discontinuous records of ice sheet and paleoclimate history that can be deciphered using a range of chemical and isotopic analyses.

Our multi-proxy analysis of an Early Pleistocene diamict cored in Melville Bugt, northwestern Greenland, indicates the existence of a persistent, erosive GrIS prior to 1.8

Ma. Extremely low cosmogenic nuclide concentrations (meteoric 10Be and in situ 10Be and

26Al) indicate the diamict is dominated by sediment with minimal near-surface exposure, similar to terrestrial sediment discharged by the GrIS today. (U-Th-Sm)/He ages of detrital apatite grains all pre-date glaciation by >150 million years, suggesting this sector of the ice sheet had not yet incised fjords required to excavate grains from sufficient depth to have young (U-Th-Sm)/He ages when the diamict was deposited. The diamict contains well- preserved leaf waxes produced by terrestrial plants, demonstrating that the ice sheet margin left exposed land surfaces either as nunataks and/or at least for short periods during

98 interglacial periods. These data indicate that a persistent, dynamic ice sheet existed in northwestern Greenland prior to 1.8 Ma and that diamict from glaciated margins provides a useful archive of ice sheet and climate history.3

4.1 Introduction: a poorly known history of the Greenland Ice Sheet

Understanding the deep-time history of the Greenland Ice Sheet (GrIS) during warmer-than-present climates is important given on-going warming (Clark et al., 2016). In the coming century, the GrIS could contribute up to 0.34 m of sea-level rise (Mengel et al.,

2016). During the Pliocene (5.33 to 2.58 Ma), CO2 concentrations were similar to today

(Pagani et al., 2010); yet, global temperature was 3C higher (Haywood et al., 2013) and sea level was 10-30 m higher (Rovere et al., 2014), with the GrIS contributing up to 6.4 m

(Koenig et al., 2014). The behavior and state of the GrIS during this time and others (e.g.,

Pleistocene “super” interglacial periods) remain poorly understood (Bierman et al., 2016).

Due to present ice cover and on-going glacial erosion, terrestrial records of Greenland climate and ice sheet history prior to the Holocene are rare (Funder et al., 2001; McFarlin et al., 2018).

Much of the known deep-time climate history of Greenland relies on seismic surveys of the continental shelf (Knutz et al., 2019) and marine sediment cores collected around the island and in the North Atlantic Ocean (Bierman et al., 2016; Helland and

3This project was conducted through collaboration with Paul Bierman and Lee Corbett (University of Vermont; cosmogenic nuclide extraction), Paul Knutz (Geological Survey of Denmark and Greenland, core access), Julie Fosdick (University of Connecticut; thermochronometry), Elizabeth Thomas (University at Buffalo; leaf waxes), Alan Hidy (Lawrence Livermore National Laboratory; 10Be analysis), and Marc Caffee (Purdue University, 26Al analysis).

99

Holmes, 1997; Reyes et al., 2014; Tripati et al., 2008) (Figure 4.1) where sediment accumulation rates are slow and core chronology can be constrained. Many deep-marine sediment cores in the North Atlantic are distal from Greenland; thus, they record not only local changes in the GrIS, but changes in global climate and glaciers on adjacent landmasses (e.g., Baffin Island and eastern North America). Most deep marine sediment contains primarily fine-grained material with only occasional clasts and sand transported by icebergs. The paucity of coarse material in deep-marine sediment limits the application of techniques used in terrestrial and near-shore environments such as cosmogenic radionuclides (Dunai, 2010) and low-temperature thermochronometers (Farley, 2002).

Glacial marine deposits from Greenland present a promising – but largely unexplored – deep time archive. On the continental shelf of Greenland, repeated ice sheet expansions deposited diamict (Knutz et al., 2019), a poorly sorted mixture of sediment.

Relative to deep marine records, diamict is under-utilized as an archive of past ice sheet behavior and landscape change because diamict can be difficult to recover by drilling

(Barker et al., 1999), to interpret stratigraphically, and to date. However, unlike ice-rafted debris contained in deep-marine sediments, diamict is deposited proximal to ice sheets, reducing uncertainty about sediment provenance. In addition, diamict often contains abundant sand-sized grains and a mixture of minerals, allowing for the application of numerous and varied analytical techniques.

100

Figure 4.1: Study area maps of Greenland and Melville Bugt

A) Greenland and the locations of this study (yellow star) and other 10Be studies from marine cores (white circles)

(Bierman et al., 2016; Shakun et al., 2017), bedrock (Schaefer et al., 2016) and basal ice (Bierman et al., 2014) from the

10 10 GISP2 ice core (green circles), and modern terrestrial sediment (orange: Bem, blue: Bei) (Bierman et al., 2014; Corbett et al., 2017a; Graly et al., 2018; Nelson et al., 2014); colors correspond to Figure 4.7. Elevation data: International

Bathymetric Chart of the Arctic Ocean; ice velocity data: ESA Sentinel-1. Coastline and grounding line shown by gray and dark gray lines, respectively. B) Bedrock topography of the northwest Greenland and Melville Bugt region with core

344S-U0110 (yellow star). Elevation data: BedMachine3.

4.2 Rationale and approach

We apply a novel combination of proven analytical techniques – meteoric 10Be and in situ produced 10Be and 26Al, (U-Th-Sm)/He thermochronology, and long-chain lipid biomarkers – to an 80 cm section of diamict in Core 344S-U0110 recovered by the JOIDES

101

Resolution in Melville Bay on the west-central Greenland shelf (Figure 4.1). Site U0110 was drilled through 124 m of glaciogenic deposits that primarily consist of compacted, clast-rich, muddy diamict with 30% recovery (Acton and Scientists, 2012). Seismic reflection data tied to exploration well chronologies and paleomagnetic age constraints date this diamict section to 1.8 – 2.0 Ma (Olduvai) (Knutz et al., 2019). The glaciogenic deposits drilled at site U0110 form part of a sedimentary unit deposited during the early phase of the Melville Bugt trough-mouth fan, a large paleo-ice sheet drainage system on the NW

Greenland margin (Figure 4.1B) (Knutz et al., 2019). Visual inspection of core material identified bivalve fragments in the diamict matrix. In the 850-2000 m fraction, ~20% of quartz grains were rounded and occasionally iron-stained compared to the remaining angular quartz grains. The abundance of sand-sized material and diversity of minerals

(including quartz and apatite) allow us to make analyses that are neither possible nor practical in fine-grained, ice-distal sediment.

The concentration of cosmogenic 10Be and 26Al in rocks and sediments reveals near-surface landscape histories of exposure, erosion, and burial (Dunai, 2010). Meteoric

10 10 Be ( Bem) is produced in the atmosphere and delivered to the Earth’s surface via wet or dry deposition where it accumulates in soil, regolith, and pelagic sediments (Graly et al.,

10 2010). Thus, Bem in glacial marine diamict reflects both the near-surface history of glacially eroded and transported ancient regolith from pre-glacial and/or interglacial landscapes (Graly et al., 2018) as well as ice cover in glacial marine settings (Yokoyama et al., 2016).

102

10 10 26 26 On ice-free landscapes, in situ Be ( Bei) and Al ( Ali) are produced primarily in the upper 2-3 meters of near-surface rock and sediment by neutron spallation with minor production by muons extending tens of meters deep (Dunai, 2010). Wet-based glaciers

10 26 excavate into sub-surface material with low Bei and Ali concentrations, while re- exposure of the land surface during periods with reduced ice cover increases the

10 26 concentration of Bei and Ali in surface materials (Bierman et al., 2016; Shakun et al.,

10 26 2018). Bei and Ali are produced in Greenland at a ratio of 7.3 ± 0.3 by neutron spallation in surface materials (Corbett et al., 2017b) and higher ratios at depth by fast muons (Balco,

26 10 2017). The half-life of Ali is shorter than that of Bei; thus, measured deviations from the production ratio reflect complex histories of exposure, erosion, and/or burial in glacially sourced surface materials. Because 10Be and 26Al are radionuclides, their concentrations measured in marine sediment can be decay-corrected to concentrations at the time of deposition and thus used to infer sediment histories.

In glacial settings, low-temperature thermochronometers, such as (U-Th-Sm)/He, quantify the timing and magnitude of erosion depth within sediment source areas, and thus, place constraints on ice-sheet incision that excavates glacial valleys and fjords (Christeleit et al., 2017; Ehlers et al., 2015). The apatite (U-Th-Sm)/He thermochronometer relies on production of 4He from radioactive decay of U and Th (Farley, 2002) and thermal diffusion of 4He at low closure temperatures (Farley, 2000). He diffusion closure generally occurs at

~70-90 °C, corresponding to shallow depths (<3 km) in Earth’s crust (Farley, 2002;

Flowers et al., 2009), but closure may be lower (~30-40 °C) depending on mineral chemistry and cooling rate (Reiners and Brandon, 2006). Compared to bedrock

103 thermochronology studies in once-glaciated settings, detrital apatite studies of glacial sediments are less common but can detect signals of glaciation (Tochilin et al., 2012). In rifted continental margins with elevated coastal regions sculpted by glaciation, such as

Greenland, (U-Th-Sm)/He thermochronometry can be used to infer sediment source area denudation by glacial erosion rather than tectonic activity (Thomson et al., 2010).

Terrestrial flora produce long-chain hydrocarbons, known as leaf waxes, that form a protective coating on plant leaves (Eglinton and Eglinton, 2008). These compounds are abraded off leaves and then transported, deposited, and preserved with minimal alteration in sedimentary archives for millions of years (Castaneda and Schouten, 2011). Long-chain n-alkanoic acids are indicative of contemporaneous terrestrial plant productivity, as the n- alkanoic acids are more labile, and therefore less likely than n-alkanes to be preserved in and eroded from surrounding sedimentary rocks (Huang et al., 1999). In addition, alkenones and short chain-length n-alkanoic acids can be indicative of marine phytoplankton (Moros et al., 2016). These lipid biomarkers have been applied widely to climate archives including loess, permafrost, lake deposits, and ocean sediments (Eglinton and Eglinton, 2008). On formerly glaciated continental shelves, where stratified pelagic sediment is rare, analyzing biomarkers in glacial marine diamict is one of the few opportunities to assess primary production both on and surrounding Greenland through time.

To satisfy grain size, target mineral, and sample masses for each of these methods, we sorted 1.2 kg of core material into different grain sizes (<125, 125-250, 250-850, 850-

2000, and >2000 μm), shapes (rounded vs. angular quartz), and mineral fractions (quartz,

104 apatite, mixed, mafic minerals) (Figure 4.2). Multi-proxy analysis across these sedimentological and lithological categories enables interpretation of the northwest GrIS at ~1.8 Ma. Refer to section 4.6 for a full description of methods and materials.

Figure 4.2: Micrographs of sorted diamict fractions.

Samples are organized by grain size (columns), lithologic fraction and shape (rows). All photos taken prior to any acid treatment.

4.3 Results

4.3.1 Extremely low cosmogenic 10Be & 26Al concentrations

Cosmogenic 10Be and 26Al concentrations, decay-corrected for 1.9 Ma of post- depositional ocean-floor burial, are very low across all sediment sizes, shapes, and mineral

10 5 7 -1 fractions, with Bem (10 – 10 atomsg ) at least two orders of magnitude greater than

10 3 -1 10 10 Bei (10 atomsg ) (Figure 4.2; Tables 4.1, 4.2). Neither Bem nor Bei concentrations

105 are related to grain size, except the fine fraction (GEUS01: <125 um), which had the

10 7 -1 highest decay-corrected Bem concentration (4.1±0.01 x10 atomsg ) (Table 4.1, Figure

10 4.2). Decay-corrected Bem adhered to mafic mineral fractions (weighted mean ± standard error [SE]: 1.7±0.6 x106 atomsg-1, n = 3) was an order of magnitude greater than for quartz fractions (3.8±0.1 x105 atomsg-1, n = 5). Weighted-mean decay-corrected concentrations

10 3 -1 26 4 -1 of Bei (5.3±0.8 x10 atomsg ) and Ali (4.8±0.9 x10 atomsg ) were also extremely low (n= 6, 1 SE), displayed no discernible grain size trend, and yield a weighted-mean

26 10 Ali/ Bei ratio of 9.1±1.5 (Figure 4.3, 4.4, Table 4.3, S4, n= 6, 1 SE). A Monte Carlo simulation (25,000 realizations) returned similar 1.9 Ma decay-corrected concentrations of

10 26 26 10 Bei Ali, and Ali/ Bei ratio (Figure 4.4).

Figure 4.3: 1.9 Ma decay-corrected cosmogenic 10Be concentrations in diamict.

Results are shown according to nuclide system, grain size, and mineral fraction. All samples labeled with numerical suffix (i.e., GEUS 01). Error bars are less than the symbol size.

106

Table 4.1: Meteoric 10Be extracted from sediment fractions

107

Table 4.2: In situ cosmogenic 10Be in quartz

108

Table 4.3: In situ cosmogenic 26Al in quartz

109

Figure 4.4: In situ cosmogenic nuclide Monte Carlo simulations.

10 26 26 10 Simulated values of A) Bei, B) Ali, and C) Ali / Bei. In each panel, the red circle and error bar shows the weighted

10 26 26 10 average (by quartz mass) and standard error (n=6) of the measured Bei and Ali concentrations, and Ali / Bei ratio, and with 1.8, 1.9, and 2.0 Ma decay-corrections applied. The blue bar with a black line shows the simulated mean and standard deviation of 25,000 simulations, represented by the probability density function below.

110

4.3.2 Old (>160 Ma) detrital apatite (U-Th-Sm)/He ages

Detrital apatite yielded single-grain apatite (U-Th-Sm)/He ages from 159 to 686

Ma (Figure 4.5, Table 4.4) with no Cenozoic ages. The ten apatite (U-Th-Sm)/He ages fall into four groups (Figure 4.5A): Jurassic (159-181 Ma, n =3), Late Triassic (221-228 Ma, n

=2), Late Carboniferous-Permian (267-305 Ma, n = 4), and Neoproterozoic (686 Ma, n

=1), with no apparent correlations between age and grain size (~70-82 µm) (Figure 4.5B) or Th/U (0.03-1.6) (Figure 4.5D). Effective uranium concentrations (eU), a proxy for natural radiation damage to crystal lattices, range from ~2 to 78 ppm (Figure 4.5C). A negative correlation between age and eU suggests some influence of radiation damage on

He diffusion, and thus closure temperature and apatite (U-Th-Sm)/He dates (Flowers et al.,

2009). For apatite with this range of eU (using a radiation diffusion model after (Flowers et al., 2009) and a 1°C/Ma monotonic cooling rate), we estimate closure temperatures of

~45-70 °C. Corresponding closure depths of ~1.3-3.0 km (assuming a 20-28°C/km crustal geotherm (Martos et al., 2018) and 10°C average surface temperature) provide an upper limit on erosion depth within the sediment source areas.

111

Figure 4.5: Detrital apatite (U-Th-Sm)/He ages and analytical results.

Normal kernel density estimate of (U-Th-Sm)/He ages of detrital apatites. Thin lines: individual Gaussian age distributions color-coded by age grouping; thick line: population (n=10) normal kernel density estimate. Text above refers to the thermal-tectonic events in West Greenland (Alsulami et al., 2015; Gregersen et al., 2019; Henriksen et al.,

2009). B) effective radius of apatite grains used in He volume calculations; C) effective uranium (eU) concentration, a useful proxy for the magnitude of radiation damage in apatite crystals; and D) measured Th/U ratio.

112

Table 4.4: Apatite (U-Th-Sm)/He Ages

4.3.3 Well-preserved lipid biomarkers

The diamict contains well-preserved lipid biomarkers. Alkenones are present in the

-1 sediment, but at low concentration (< 50 ngg dry sediment for C37:2 and C37:3 combined).

-1 Short-chain n-alkanoic acids are the most abundant (C16: 990±152 ngg , C18: 1449±197 ngg-1; Figure 4.6A; Table 4.5), while mid- and long-chain n-alkanoic acids are present at

-1 lower concentrations: C28 is most abundant, 482±96 ngg (Table 4.5). All n-alkanoic acids have a strong even-over-odd preference (4.3±0.6), indicating minimal degradation, as plant biosynthesis produces even-chain n-alkanoic acids (Eglinton and Eglinton, 2008).

Concentrations of n-alkanes are lower than n-alkanoic acids (C29 is most abundant [309±33 ngg-1]; Figure 4B; Table 4.5), with a lower odd-over-even preference (2.4±0.1), indicating potential contamination by organic matter from sedimentary rocks.

113

Figure 4.6: Lipid biomarker results.

A) Methylated acid fraction, even-chain length fatty acid methyl ester peaks labeled with carbon chain number. B) Apolar fraction, odd-chain length n-alkane peaks labeled with carbon chain number.

114

Table 4.5: Lipid biomarkers

115

4.4 Discussion

4.4.1 Glacial erosion of pre-ice sheet regolith by the Early Pleistocene

10 10 26 Extremely low Bem, Bei, and Ali concentrations in diamict – regardless of grain size, shape, or mineralogy – suggest that the northwestern GrIS had already eroded tens of meters of previously exposed sediment source areas by the Early Pleistocene. Such low decay-corrected cosmogenic nuclide concentrations require minimal contribution from sediment exposed at or near Earth’s surface; thus, sub-glacial material sourced from >10 m dominates sediment in this diamict (Nelson et al., 2014). Even grains with apparent surface weathering characteristics, such as rounded, iron-stained sand (GEUS06), yield

10 3 -1 extremely low Bei concentrations (1.72.0 x10 atomsg ); these grains were most likely glacially scoured from Cretaceous sedimentary rocks (Gregersen et al., 2019).

26 10 The 1.9 Ma decay-corrected Ali/ Bei ratio of this diamict (9.1±1.6; weighted mean ± standard error) is substantially higher than the contemporary Greenland surface production ratio of 7.3±0.3 (Corbett et al., 2017b). The decay-corrected Monte Carlo

26 10 simulated Ali/ Bei ratio (9.2±1.3; weighted mean ± 1 standard deviation) indicates with

26 10 high probability that the Ali/ Bei ratio of this diamict at deposition was greater than the

26 10 observed Greenland surface production ratio. Ali/ Bei ratios >7 are indicative of muon- induced production at depths of tens of meters for short (~100 kyr) periods (Balco, 2017).

Thus, this Early Pleistocene diamict is comprised primarily of deeply eroded material. The low-concentration, high-ratio material in the 344S core requires stripping by the early GrIS

10 26 of tens of meters of rock and regolith (which would have higher Bei and Ali concentrations and a lower ratio) followed by hundreds of thousands of years of exposure

116 without ice cover before rapid excavation via glacial erosion and transport to the core site

(Balco, 2017). Such erosion by early ice followed by exposure is consistent with warmth indicated by the nearby Kap København Formation which, with an age of ~2 Ma, predates

26 the diamict we analyzed (Funder et al., 2001). The presence of Ali in core 344S sediment supports the paleomagnetic age model (Knutz et al., 2019). If the diamict were much older

26 5 than 2.0 Ma, Ali would not be preserved due to its 7 x10 year half-life.

The Early Pleistocene GrIS transported and deposited glacial sediment isotopically similar to that discharged at the ice sheet margin today (Harden et al., 1992) (Figure 4.7).

10 5 7 -1 10 Bem concentrations in Early Pleistocene diamict (10 -10 atomsg ) are similar to Bem measurements in modern englacial sediment (106 – 108 atomsg-1) and sub-glacial and glacio-fluvial sediment (105-107 atomsg-1) from the edge of the GrIS (Graly et al., 2018).

10 3 -1 Bei concentrations (10 atomsg ) more closely resemble those in modern terrestrial sediment from glacierized regions (103-104 atomsg-1) than from ice-free areas (104-105 atomsg-1) (Goehring et al., 2010; Nelson et al., 2014). Englacial and outwash clasts

10 3 - emerging from the ice sheet today have an average Bei concentration of 8.2 x 10 atomsg

1 (Corbett et al., 2017a). Thus, we infer Early Pleistocene GrIS erosional processes and rates were similar to today.

117

Figure 4.7: Synthesis of 10Be measurements from sedimentary archives in Greenland

Left panel shows the 1.9 Ma decay-corrected 10Be concentrations in diamict (this study). Right panels show previous measurements from various archives; data points color coded by archive. Each box plot shows the median (red line), 25th and 75th percentiles (black boxes), and whiskers extend to values not considered outliers. Light gray box outlines the

10 range of Be measurements from this study for comparison. Superscript numbers correspond to references.

10 10 The range of Bem and Bei in diamict and in material sourced from the base of the ice sheet (Bierman et al., 2014; Bierman et al., 2016; Graly et al., 2018; Schaefer et al.,

2016) highlights how spatially heterogeneous thermal regimes below the GrIS have controlled erosion since the Early Pleistocene, a pattern also observed for other continental

10 7 8 -1 ice sheets (Gjermundsen et al., 2015). Bem in silt in basal ice (10 -10 atomsg ) (Bierman

10 4 -1 et al., 2016) and Bei in sub-glacial bedrock (10 atomsg ) (Schaefer et al., 2016) at the base of the GISP2 ice core are at least an order of magnitude greater than decay-corrected

118

10 5 6 -1 10 3 -1 Bem (10 -10 atomsg ) and Bei (10 atomsg ) concentrations in the diamict we analyzed (Figure 4.7). These data suggest that the GrIS, where cold-based and frozen to its bed, can preserve pre-glacial soils and bedrock surfaces or those exposed during ice-free interglacial periods; yet, where the base of the ice sheet reaches the pressure-melting point,

10 10 the GrIS excavates into the underlying rock where little Bem and Bei has accumulated.

Diamict on the continental shelf appears to have a different exposure and erosional history than sediment transported by icebergs and recovered in deep marine cores (Figure

10 4.7) (Bierman et al., 2016; Shakun et al., 2018). Decay-corrected Bei concentrations in diamict (103 atomsg-1, this study) are an order of magnitude less than in ice-rafted debris

(103 -104 atomsg-1) in sections of similar age in deep marine cores surrounding Greenland

10 (Bierman et al., 2016). Elevated decay-corrected Bei concentrations in ice-rafted debris could be the result of incorporating sub-aerially exposed sediment sources, such as Tertiary regolith pre-dating glaciation or rockfall from exposed highlands. The decay-corrected

10 3 -1 Bei concentrations in 344S diamict (10 atomsg ) are similar to those from the

ANDRILL marine core (103 – 104 atomsg-1) where a stable, yet erosive East Antarctic Ice

Sheet with only limited periods of reduced ice cover allowed for only minor accumulation

10 26 of Bei and Ali since the Middle Miocene, 14 Ma (Shakun et al., 2018).

4.4.2 The absence of deeply carved fjords in Melville Bugt

Consistently old (>160 Ma) apatite (U-Th-Sm)/He ages likely reflect the long-term thermos-tectonic history of Greenland and provide no evidence for fjord-scale glacial erosion (>1.3 km) (Figure 4.5). Apatites with a Late Carboniferous-Permian (U-Th-

119

Sm)/He age may correspond to sediment sources from the Ellesmerian Orogenic belt in northernmost Greenland (Henriksen et al., 2009). The timing of deformation resulting from this event is not well known, but post-dated the late Devonian (370 Ma). Apatites of

Jurassic age pre-date Cretaceous rifting and Paleogene basalt volcanism (Alsulami et al.,

2015) in West Greenland; there are no known sources or thermogenic events that can explain these grain ages (Gregersen et al., 2019).

The lack of apatite (U-Th-Sm)/He ages <150 Ma suggests that the diamict was derived from sediment sources that have undergone little Pliocene or Pleistocene denudation. Such sources may include the higher topographic levels of present day fjords and intervening ridges, and/or the low-relief plateau areas within the Greenland continental margin (Figure 4.1B). Sediment derived from deep erosion of glacial valleys and fjords would yield (U-Th-Sm)/He ages younger than known thermal events, similar to the incision of the Andes by the Patagonian Ice Sheet (Christeleit et al., 2017; Fosdick et al., 2013).

4.4.3 Evidence for Early Pleistocene terrestrial and marine vegetation

Despite cosmogenic evidence for widespread glacial erosion of at least tens of meters, biomarkers recovered from the diamict indicate sediment input from regions with terrestrial vegetation (Figure 4.6); this requires ice-free conditions around the margins of the ice sheet, either spatially and/or temporally, prior to 1.8-2.0 Ma. Modern terrestrial plant samples and modern and Holocene lake sediments on Greenland do not contain high concentrations of short chain (

120

Thomas et al., 2018). The diamict is dominated by longer chain length n-alkanoic acids

(C26 and C28) than in Greenland Holocene lake sediments, which have additional input from aquatic macrophytes and are dominated by C24 (6). We infer that the C20 to C32 n-alkanoic acids are derived from terrestrial plants growing in exposed areas around the margins of

Greenland, much like during the Holocene. The strong even-over-odd preference of n- alkanoic acids indicates that these compounds have been well preserved and were likely not sub-aerially exposed for long periods prior to deposition in this diamict. Given the concentration and preservation of these biomarkers, it is unlikely that they are derived from windblown transport. The presence of well-preserved long-chain n-alkanoic acids in this diamict indicates that vegetation covered ice-free areas of northwestern Greenland before transport by later ice sheet advance(s) and deposited in this diamict.

10 Short-chain lipids, shell fragments, and elevated Bem in the <125 µm fraction

(GEUS01) suggest grounded ice reworked marine sediments from earlier periods of reduced ice cover on the continental shelf (Figure 4.6). The presence of short-chain n- alkanoic acids, produced by marine algae throughout the Arctic Ocean today (Sachs et al.,

2018), and small concentrations of alkenones – biomarkers of marine coccolithophores

(Moros et al., 2016) – indicate marine sediment sources (Eglinton and Eglinton, 2008;

10 Sachs et al., 2018). Higher decay-corrected Bem concentrations in GEUS01 <125 µm

(4.10.01 x107 atomsg-1), relative to all other measurements on diamict fractions >125 µm

(average: 8.97.1 x105 atomsg-1) suggest the fine fraction is a mixture of pelagic and sub- glacially derived sediment. In modern glacial marine environments such as the Ross Sea,

Antarctica, the transition from sub-ice shelf or sub-glacial conditions to open marine

121

10 conditions during the Holocene increased Bem concentrations in marine sediment (up to

109 atomsg-1) (Graly et al., 2010). Yet, decay-corrected concentrations in 344S diamict are still two orders of magnitude lower than the Holocene measurements in the Ross Sea.

10 Elevated Bem in the fine sediment fraction of Early Pleistocene 344S diamict likely

10 reflects a mixture of reworked pelagic sediments ( Bem-rich) and sub-glacially derived

10 ( Bem-poor) fine sediment.

4.5 Implications

Our data suggest that during the Early Pleistocene, the landscape of northwestern

Greenland was similar to today, with substantial ice cover, glacial erosion below warm- based ice, and coastal ice-free areas that harbored vegetation. These data suggest and lend support for the presence of a largely persistent, yet dynamic, ice sheet in NW Greenland by the Early Pleistocene (Knutz et al., 2019). Decay-corrected, 10Be concentrations in diamict most closely resemble 10Be measurements of sediment discharged from the GrIS

26 10 today (Figure 4.7). The decay-corrected Ali/ Bei ratio is higher than the observed

Greenland surface production ratio (Corbett et al., 2017b). Together, these observations suggest that sub-glacial material eroded from tens of meters depth and discharged to the ice margin effectively dilutes 10Be in sediment sourced from ice-free areas (Nelson et al.,

2014). By 1.8 Ma, sub-glacial erosion in northwestern Greenland had already removed

-1 10 26 sufficient surface materials that only a few thousand atomsg of Bei and Ali were present in quartz, most likely nuclides produced over time by muon reactions at tens of meters depth. Ice-free conditions were so limited – either spatially or temporally – that 10Be

122 did not accumulate significantly in rocks or sediment before glacial erosion. In contrast, consistently old detrital apatite (U-Th-Sm)/He ages indicate that glacial denudation of sediment source areas, while sufficient to remove 10Be, had not yet carved deep fjords before 1.8 Ma in northwest Greenland. The abundance of well-preserved terrestrial and

10 marine lipid biomarkers, bivalve fragments, and elevated Bem in <125 μm sediment suggest that the ice sheet transported and reworked sediments originally deposited under open-marine and ice-free conditions.

Diamict offers a promising and largely unexplored archive of climate and ice sheet history. Because diamict is a mixture of sediment from an eroding landscape, it contains information about multiple surface processes integrated over time and space. Lipid biomarkers provide insight into terrestrial and marine life whereas cosmogenic nuclides and (U-Th-Sm)/He ages of detrital apatite in diamict provide information about erosion at different depth scales. Multi-parameter analyses of diamict demonstrate its utility as an archive of past glacial activity in presently deglaciated margins, particularly where glacial stratigraphic units are exposed at shallow depths below the seafloor. Diamict analysis has the potential to improve assessment of long-term changes in glacial erosion, ice cover, and vegetation throughout the polar regions.

4.6 Materials and Methods

4.6.1 Core recovery

Site U0110 in NE Baffin Bay (74.75537°N, 62.55702°W, 318.8 m water depth) was drilled as part of an industry sponsored coring expedition (344S; (Acton and Scientists,

123

2012)). The site penetrated a Quaternary section of 124.53 m, mainly consisting of over- consolidated muddy diamict, with a core recovery of ~27%. Scarcity of biogenic components precludes a fossil-based chronology but paleomagnetic measurements

(inclination data) in combination with seismic stratigraphy and tie to nearby boreholes suggests an Early Pleistocene age, probably Olduvai (1.8-2.0 Ma), for most of the diamict interval (Knutz et al., 2019).

4.6.2 Sample preparation

Sediment was prepared at the University of Vermont (UVM) in fall 2018. Material was disaggregated and wet sieved using de-ionized water. We dried all size fractions >125

μm for 24 hours at 65C. We allowed the <125 μm fraction to settle and decanted the supernatant. We transferred the <125 μm fraction into a 1 L HDPE bottle, decanted again, froze at -10C, and then freeze-dried. We sub-sampled the bulk <125 μm fraction and the

125-250 μm mafic fraction for lipid biomarker and low temperature thermochronology analyses, respectively. We isolated mineralogic and morphologic fractions from the 125-

250 μm, 250-850 μm, and 850-2000 μm size fractions. The 125-250 um and 250-850 um fractions were density separated using lithium polytungstate to isolate mafic (>2.85g/cm3) and felsic (<2.85 g/cm3) fractions. Grains from the 850-2000 μm fraction were hand-picked to isolate angular and rounded quartz, and mafic minerals.

124

4.6.3 Cosmogenic Nuclides

4.6.3.1 Meteoric 10Be

All samples were twice leached in 6N HCl to strip grain coatings. Leachates were decanted into 1L HDPE bottles and isotopically diluted using ~400 g 9Be. Each leachate

10 was dried; remaining sediment was rinsed and processed for Bei extraction. GEUS01 was re-spiked with an additional ~300 g of Be carrier because recovery of original carrier was low (38 μg). We added Na2SO4 and 8N HF and heated to dryness twice before HClO4 fuming. Samples were water extracted and precipitated with NaOH. Supernatants from the

NaOH precipitation were neutralized to pH 7-8 with concentrated HNO3. Samples were re- precipitated using NH3OH at pH 8-9. Be was purified from samples using column chemistry (Corbett et al., 2016) (Table 4.1).

4.6.3.2 In situ 10Be and 26Al

Quartz-bearing samples were prepared for in situ 10Be and 26Al extraction. The 850-

2000 μm and >2000 μm fractions were crushed and ground to <1000 μm prior to chemical treatment. Samples were then spiked with ~250 µg 9Be and 27Al (only for samples with insufficient total Al) and processed for in situ 10Be and 26Al extraction following established protocols (Corbett et al., 2016; Kohl and Nishiizumi, 1992) with special procedures for low concentrations (Shakun et al., 2018). GEUS04 was split into two sub- samples for replicate measurement. 27Al concentrations were measured with ICP-OES.

Four procedural blanks were prepared with the samples (one for meteoric, three for in situ) during nuclide extraction to estimate backgrounds from both laboratory sample processing and acceleratory mass spectroscopy (AMS) analysis. The blank uncertainties

125

10 9 were propagated in quadrature. Be/ Be ratios in samples and blanks were measured at the

Center for Accelerator Mass Spectrometry, Lawrence Livermore National Laboratory and normalized to 07KNSTD3110, which has an assumed 10Be/9Be ratio of 2.85 × 10−12

(Nishiizumi et al., 2007). Blank corrections were between 0.6% and 9.6% of the adjusted

10Be values (Table 4.6). 27Al/26Al ratios in samples and blanks were measured at the Purdue

Rare Isotope Measurement Laboratory and normalized to KNSTD, which has an assumed

10Be/9Be ratio of 1.818 × 10−12 (Nishiizumi, 2004). Blank corrections were 1.3% to 14.0%

(Table 4.7).

126

Table 4.6: 10Be blank measurements

Table 4.7: 26Al blank measurements

127

4.6.3.3 Monte Carlo Simulation

10 26 We made a Monte Carlo simulation of Bei and Ali concentrations in Matlab

10 26 (Figure 4.3). In each simulation, a Bei and Ali concentration was randomly selected from the mean and 1 sigma analytical uncertainty of each measured sample. The weighted

10 26 26 average (by quartz mass) was then calculated for Bei and Ali in each simulation. A Ali

10 / Bei was calculated from that simulated weighted average. The 25,000 simulations of

10 26 10 Bei, 26Ali, and Ali / Bei were then decay-corrected for 1.8 Ma, 1.9 Ma, and 2.0 Ma of burial time.

4.6.4 Low-temperature thermochronology

Samples were prepared at the University of Connecticut. Apatite grains were extracted from the 125-250 µm fraction using standard heavy mineral separation procedures. Individual crystals were handpicked in ethyl alcohol, screened, and measured at 120x magnification under polarized light for crystal size, shape, and inclusions.

Individual crystals were placed into 1 mm diameter Nb foil packets and placed in a laser planchette under an ultra-high vacuum (~5 x10-9 torr). Nb packets were heated with a 45W diode laser to 1000°C for 5 minutes to extract the radiogenic 4He. The degassed 4He was spiked with ~7 ncc of pure 3He, gettered and cryogenically purified, and analyzed on a

Balzers PrismaPlus QME 220 quadrupole mass spectrometer. This gas extraction procedure was repeated at least once to ensure complete mineral degassing. Degassed aliquots were shipped to the University of Colorado for chemical dissolution and U-Th-Sm

235 230 145 measurement. Nb packets were spiked with a U- Th- Nd tracer in HNO3 and heated

128 at 80°C for 2 hours. Once dissolved, minerals were diluted with doubly-deionized water.

Sample solutions were analyzed for U, Th, and Sm content using a Thermo Element 2 magnetic sector mass spectrometer. U, Th, and Sm contents were combined with He measurements and grain data to calculate He dates using published methods (Ketcham et al., 2011). Laboratory standards (Durango Apatite) and procedural blanks were analyzed with each planchette.

4.6.5 Lipid biomarkers

Organic geochemical analyses were conducted at the University at Buffalo. Lipids were extracted from multiple aliquots of the <125 μm fraction, and the n-alkanoic acids were isolated and derivitized to Fatty Acid Methyl Esters (FAMEs) following established procedures (58). Apolar and ketone fractions were isolated from the neutral fraction by sequential elution over activated Al2O3 using Methylene Chloride (DCM):hexane (9:1, v:v)

(apolar), DCM:hexane (1:1, v:v) (ketone), and DCM:methanol (1:1, v:v) (polar) as the eluents. FAMEs, n-alkanes, and ketones were quantified on a Thermo Scientific Trace1310

Gas Chromatograph equipped with two flame ionization detectors run simultaneously using AI 1310 autosamplers. HP-1MS columns (Agilent) were used for all analyses. All samples were injected on split/splitless injectors at 250°C, run in splitless mode until 0.75 minutes with a split flow of 14.0 mL∙min-1. Column flow was constant at 3.60 mL∙min-1 using hydrogen carrier gas. Oven temperature was initially 60°C for 1 minute, ramped to

220°C at a rate of 20°C∙min-1, followed by 6°C∙min-1 to 315°C with a final hold at 315°C for 10 minutes. Peaks were identified by retention time referenced to external standards while masses were calculated using external standard calibration curves determined for

129 octacosanoic acid and heptacosane. All aliquots of the diamict were treated as separate samples (true replicates carried through the entire sample preparation process), and mean and standard deviation concentration values for the sample were calculated from the multiple replicate analyses.

130

CHAPTER FIVE: Conclusions

Pleistocene ice sheets left behind landscapes, landforms, and sediments that hold clues to the climate of the past. In this dissertation, I sought to unlock information from these climate archives by using an array of methods – ranging from traditional geologic mapping of Antarctic landscapes to novel applications of cosmogenic nuclides to glacial sediments deposited on the Greenland continental shelf two million years ago.

McMurdo Sound, Antarctica contains glacial sediments that record the history of ice sheets in the Ross Sea. During Late Pleistocene glacial periods, the West Antarctic Ice

Sheet (WAIS) and East Antarctic Ice Sheet (EAIS) outlet glaciers expanded as marine- based ice sheets into the Ross Sea. During multiple glaciations, these marine ice sheets overflowed into McMurdo Sound and deposited glacial sediments that contain far-travelled erratic rocks from the Transantarctic Mountains and outline former ice sheet elevation and flow directions.

Despite a colder and drier climate at the peak of the last glacial period than today,

McMurdo Sound still harbored life – algae – that thrived in small streams and ponds along the margin of a thick ice sheet. Buried below only tens of centimeters of sand and till, fragments of fossil algae yield radiocarbon ages that bracket the timing of maximum ice sheet extent between 19.6 and 12.3 thousand years ago. The marine ice sheet in the western

Ross Sea achieved and maintained its maximum extent and thickness, even as earth’s climate warmed, the northern hemisphere ice sheets retreated, and global sea level rose following the global Last Glacial Maximum. The ice sheet in the western Ross Sea persisted through the Meltwater Pulse 1A event, suggesting that this sector of Antarctica

131 did not contribute significantly to this period of rapid sea level rise. Instead, geologic data from McMurdo Sound and elsewhere in the Ross Sea and Transantarctic Mountains suggest that the interplay between rising sea level, changing ocean heat circulation, and local ice dynamics led to the demise of this ice sheet during the Early Holocene.

Cosmogenic nuclides in glacially transported boulders in McMurdo Sound record surface processes operating across the Antarctic landscape during the Late Pleistocene. In

McMurdo Sound, boulder exposure ages do not yield simple ages of emplacement along an ice margin. Instead, exposure ages reflect the different pathways of glacial entrainment below, along, and above a polythermal ice sheet according to bedrock source. Rocks sourced by sub-glacial plucking below wet-based, erosive EAIS outlet glaciers tend to lack inherited nuclides. Rocks formerly exposed on cliff faces in the Transantarctic Mountains that later fell onto glacier surfaces below contain a larger inventory of inherited cosmogenic nuclides. Subsequent entrainment and transport by cold-based, non-erosive glacial ice failed to remove these cosmogenic nuclides. Rocks from the local volcanic bedrock in

McMurdo Sound – both those previously exposed on the ice-free, interglacial landscape and others from the ocean floor – were incorporated into advancing marine ice sheets that failed to erode and remove inherited nuclides. While the Antarctic ice sheets expanded and decayed in-sync with orbital eccentricity cycles of the past 800,000 years, the preservation of inherited nuclides in Antarctic glacial sediments suggests that surface processes have remained stable and predictable throughout the Late Pleistocene.

We can use cosmogenic nuclides in concert with other well-established methods to under-utilized climate archives in Greenland, where the deep-time climate history is poorly

132 known. Analysis of glacial marine diamict deposited on the continental shelf revealed that by the Early Pleistocene, the northwest Greenland Ice Sheet was similar to today.

Extremely low cosmogenic nuclide concentrations indicate the existence of a persistent, erosive ice sheet that had already removed ancient regolith and pre-glacial soil cover, yet the absence of young (U-Th-Sm)/He ages of detrital apatite, however, indicate that the ice sheet had yet to incise deep (>1.3 km) glacial valleys and fjords. Well-preserved biomarkers in diamict suggest that, similar to today, ice-free regions flanked the ice sheet and supported terrestrial vegetation. By the early Pleistocene, the Greenland Ice Sheet was already a dynamic, yet persistent feature of Earth’s climate system.

Final word

The waxing and waning of polar ice sheets defined earth’s climate for the past several million years. Our species only evolved within the last few of these glacial- interglacial cycles, adapting to slow immense changes in climate for hundreds of thousands of years. Yet, within only the past couple of centuries, humans – through industrialization and emission of greenhouse gases to the atmosphere – are rapidly driving earth’s climate in a new direction.

As we look towards a warmer future, there is no greater existential threat to civilization than the melting of polar ice sheets and rising sea level. Just as our human ancestors moved with ice sheet-driven changes in sea level during the Late Pleistocene, future generations may face a similar climate-driven migration away from disappearing coastlines and rising seas. The geologic record cannot offer an exact roadmap for the fate

133 of polar ice sheets – and humanity – in a warming world, but it does detail how ice sheets could behave and change within Earth’s dynamic climate system.

134

BIBLIOGRAPHY

Ackert, R. P., Mukhopadhyay, S., Parizek, B. R., and Borns, H. W., 2007, Ice elevation near the West Antarctic Ice Sheet divide during the Last Glaciation: Geophysical Research Letters, v. 34, no. 21.

Ackert, R. P., Swinging gate or Saloon doors: do we need a new model of Ross Sea deglaciation, in Proceedings Fifteenth West Antarctic Ice Sheet Meeting, Sterling, VA, 2008.

Acton, G., and Scientists, E. S., 2012, Expedition 344S: International Ocean Discovery Program.

Alsulami, S., Paton, D. A., and Cornwell, D. G., 2015, Tectonic variation and structural evolution of the West Greenland continental margin: AAPG Bulletin, v. 99, no. 9, p. 1689-1711.

Anderson, J. B., 1999, Antarctic marine geology, New York, New York : Cambridge University Press.

Anderson, J. B., Conway, H., Bart, P. J., Witus, A. E., Greenwood, S. L., McKay, R. M., Hall, B. L., Ackert, R. P., Licht, K., Jakobsson, M., and Stone, J. O., 2014, Ross Sea paleo-ice sheet drainage and deglacial history during and since the LGM: Quaternary Science Reviews, v. 100, p. 31-54.

Anderson, J. T. H., Wilson, G. S., Fink, D., Lilly, K., and Levy, R. H., 2017, Reconciling marine and terrestrial evidence for post LGM ice sheet retreat in southern McMurdo Sound: Quaternary Science Reviews, v. 157, p. 1-13.

Andrews, J. T., Domack, E. W., Cunningham, W. L., Leventer, A., Licht, K. J., Jull, A. J. T., DeMaster, D. J., and Jennings, A. E., 1999, Problems and possible solutions concerning radiocarbon dating of surface marine sediments, Ross Sea, Antarctica: Quaternary Research, v. 52, no. 2, p. 206-216.

Baggenstos, D., Severinghaus, J. P., Mulvaney, R., McConnell, J. R., Sigl, M., and Maselli, O., 2018, A Horizontal Ice Core From Taylor Glacier, its implications for Antarctic climate history, and an improved Taylor Dome ice core time scale: Paleoceanography and Paleoclimatology, v. 33, p. 778-794.

Balco, G., 2017, Production rate calculations for cosmic-ray-muon-produced 10Be and 26Al benchmarked against geological calibration data: Quaternary Geochronology, v. 39, p. 150 - 173.

135

Bamber, J. L., Riva, R. E. M., Vermeersen, B. L. A., and LeBrocq, A. M., 2009, Reassessment of the Potential Sea-Level Rise from a Collapse of the West Antarctic Ice Sheet: Science, v. 324, no. 5929, p. 901-903.

Barker, P. F., Camerlenghi, A., Acton, G. D., and Party, S. S., 1999, Leg 178 Summary.

Bart, P. J., and Cone, A. N., 2012, Early stall of West Antarctic Ice Sheet advance on the eastern Ross Sea middle shelf followed by retreat at 27,500 C-14 yr BP: Palaeogeography Palaeoclimatology Palaeoecology, v. 335, p. 52-60.

Bart, P. J., Krogmeier, B. J., Bart, M. P., and Tulaczyk, S., 2017, The paradox of a long grounding during West Antarctic Ice Sheet retreat in Ross Sea: Scientific Reports, v. 7.

Bentley, M. J., et al., 2014, A community-based geological reconstruction of Antarctic Ice Sheet deglaciation since the Last Glacial Maximum: Quaternary Science Reviews, v. 100, p. 1-9.

Berke, M. A., Cartagena Sierra, A., Bush, R., Cheah, D., and O'Connor, K., 2019, Controls on leaf wax fractionation and δ2H values in tundra vascular plants from western Greenland: Geochimica et Cosmochimica Acta, v. 244, p. 565-583.

Bierman, P. R., Corbett, L. B., Graly, J. A., Neumann, T. A., Lini, A., Crosby, B. T., and Rood, D. H., 2014, Preservation of a Preglacial Landscape Under the Center of the Greenland Ice Sheet: Science, v. 344, no. 6182, p. 402-405.

Bierman, P. R., Shakun, J. D., Corbett, L. B., Zimmerman, S. R., and Rood, D. H., 2016, A persistent and dynamic East Greenland Ice Sheet over the past 7.5 million years: Nature, v. 540, no. 7632, p. 256-260.

Boning, C.W., Behrens, E., Biastoch, A., Getzlaff, K., Bamber, J.L., 2016. Emerging impact of Greenland meltwater on deepwater formation in the North Atlantic Ocean. Nature Geoscience v. 9, no. 7, p. 523-527.

Bromley, G. R. M., Hall, B. L., Stone, J. O., and Conway, H., 2012, Late Pleistocene evolution of Scott Glacier, southern Transantarctic Mountains: implications for the Antarctic contribution to deglacial sea level: Quaternary Science Reviews, v. 50, p. 1-13.

Bromley, G. R. M., Hall, B. L., Stone, J. O., Conway, H., and Todd, C. E., 2010, Late Cenozoic deposits at Reedy Glacier, Transantarctic Mountains: implications for former thickness of the West Antarctic Ice Sheet: Quaternary Science Reviews, v. 29, no. 3-4, p. 384-398.

136

Bromley, G.R.M., Hall, B.L., Stone, J.O., Conway, H., 2012. Late Pleistocene evolution of Scott Glacier, southern Transantarctic Mountains: implications for the Antarctic contribution to deglacial sea level. Quaternary Science Reviews 50, 1-13.

Brook, E. J., Kurz, M. D., Ackert, R. P., Raisbeck, G., and Yiou, F., 1995, Cosmogenic nuclide exposure ages and glacial history of Late Quaternary Ross Sea Drift in McMurdo Sound, Antarctica: Earth and Planetary Science Letters, v. 131, no. 1-2, p. 41-56.

Buizert, C., et al., 2015, Precise interpolar phasing of abrupt climate change during the last ice age: Nature, v. 520, no. 7549, p. 661-169.

Castaneda, I. S., and Schouten, S., 2011, A review of molecular organic proxies for examining modern and ancient lacustrine environments: Quaternary Science Reviews, v. 30, p. 2851-2891.

Castaneda, I. S., and Schouten, S., 2015, A review of molecular organic proxies for examining modern and ancient lacustrine environments (vol 30, pg 2851, 2011): Quaternary Science Reviews, v. 125, p. 174-176.

Christ, A.J., Bierman, P.R., 2019. The local Last Glacial Maximum in McMurdo Sound, Antarctica: implications for ice sheet behavior in the Ross Sea Embayment. Geological Society of America Bulletin. https://doi.org/10.1130/B35139.1

Christeleit, E. C., Brandon, M. T., and Shuster, D. L., 2017, Miocene development of alpine glacial relief in the Patagonian Andes, as revealed by low-temperature thermochronometry: Earth and Planetary Science Letters, v. 460, p. 152-163.

Clark, P. U., Dyke, A. S., Shakun, J. D., Carlson, A. E., Clark, J., Wohlfarth, B., Mitrovica, J. X., Hostetler, S. W., and McCabe, A. M., 2009, The Last Glacial Maximum: Science, v. 325, no. 5941, p. 710-714.

Clark, P. U., Shakun, J. D., Marcott, S. A., Mix, A. C., Eby, M., Kulp, S., Levermann, A., Milne, G. A., Pfister, P. L., Santer, B. D., Schrag, D. P., Solomon, S., Stocker, T. F., Strauss, B. H., Weaver, A. J., Winkelmann, R., Archer, D., Bard, E., Goldner, A., Lambeck, K., Pierrehumbert, R. T., and Plattner, G. K., 2016, Consequences of twenty-first-century policy for multi-millennial climate and sea-level change: Nature Climate Change, v. 6, no. 4, p. 360-369.

Conway, H., Hall, B. L., Denton, G. H., Gades, A. M., and Waddington, E. D., 1999, Past and future grounding-line retreat of the West Antarctic Ice Sheet: Science, v. 286, no. 5438, p. 280-283.

137

Corbett, L. B., Bierman, P. R., and Rood, D. H., 2016, An approach for optimizing in situ cosmogenic Be-10 sample preparation: Quaternary Geochronology, v. 33, p. 24- 34.

Corbett, L. B., Bierman, P. R., Neumann, T. A., Graly, J. G., Shakun, J. D., Marc, W., Caffee, M. W., and Dunai, T., Analysis of three cosmogenic isotopes in subglacial cobbles helps unravel Greenland’s exposure and erosion history., in Proceedings Geological Society of America Abstracts with Programs, Seattle, WA, 2017a, Volume 49.

Corbett, L. B., Bierman, P. R., Rood, D. H., Caffee, M. W., Lifton, N. A., and Woodruff, T. E., 2017b, Cosmogenic 26Al/10Be surface production ratio in Greenland: Geophysical Research Letters, v. 44, no. 3, p. 1350-1359.

Cuffey, K. M., Clow, G. D., Steig, E. J., Buizert, C., Fudge, T. J., Koutnik, M., Waddington, E. D., Alley, R. B., and Severinghaus, J. P., 2016, Deglacial temperature history of West Antarctica: Proceedings of the National Academy of Sciences of the United States of America, v. 113, no. 50, p. 14249-14254.

DeConto, R. M., and Pollard, D., 2016, Contribution of Antarctica to past and future sea- level rise: Nature, v. 531, no. 7596, p. 591-597.

Denton, G. H., and Marchant, D. R., 2000, The geologic basis for a reconstruction of a grounded ice sheet in McMurdo Sound, Antarctica, at the last glacial maximum: Geografiska Annaler Series a-Physical Geography, v. 82A, no. 2-3, p. 167-211.

Denton, G. H., Anderson, R. F., Toggweiler, J. R., Edwards, R. L., Schaefer, J. M., and Putnam, A. E., 2010, The Last Glacial Termination: Science, v. 328, no. 5986, p. 1652-1656.

Denton, G. H., Bockheim, J. G., Wilson, S. C., and Stuiver, M., 1989, Late Wisconsin and early Holocene glacial history, inner Ross Embayment, Antarctica: Quaternary Research, v. 31, no. 2, p. 151-182.

Domack, E. W., Jacobson, E. A., Shipp, S., and 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: Geological Society of America Bulletin, v. 111, no. 10, p. 1517-1536.

Domack, E. W., Taviani, M., and Rodriguez, A., 1999, Recent sediment remolding on a deep shelf, Ross Sea: implications for radiocarbon dating of Antarctic marine sediments: Quaternary Science Reviews, v. 18, no. 13, p. 1445-1451.

138

Dunai, T. J., 2010, Cosmogenic Nuclides: Principles, Concepts and Applications in the Earth Surface Sciences, Cambridge, Cambridge University Press.

Eglinton, G., and Hamilton, R. J., 1967, Leaf epicuticular waxes: Science, v. 156, no. 3780, p. 1322-1335.

Eglinton, T. I., and Eglinton, G., 2008, Molecular proxies for paleoclimatology: Earth and Planetary Science Letters, v. 275, no. 1-2, p. 1-16.

Ehlers, T. A., Szameitat, A., Enkelmann, E., Yanites, B. J., and Woodsworth, G. J., 2015, Identifying spatial variations in glacial catchment erosion with detrital thermochronology: Journal of Geophysical Research-Earth Surface, v. 120, no. 6, p. 1023-1039.

Elderfield, H., Ferretti, P., Greaves, M., Crowhurst, S., McCave, I. N., Hodell, D., and Piotrowski, A. M., 2012, Evolution of Ocean Temperature and Ice Volume Through the Mid-Pleistocene Climate Transition: Science, v. 337, no. 6095, p. 704- 709.

Farley, K. A., 2000, Helium diffusion from apatite: General behavior as illustrated by Durango fluorapatite: Journal of Geophysical Research-Solid Earth, v. 105, no. B2, p. 2903-2914.

Farley, K.A., 2002, (U-Th)/He dating: Techniques, calibrations, and applications: Noble Gases in Geochemistry and Cosmochemistry, v. 47, p. 819-844.

Feldmann, J., and Levermann, A., 2015, Collapse of the West Antarctic Ice Sheet after local destabilization of the Amundsen Basin: Proceedings of the National Academy of Sciences of the United States of America, v. 112, no. 46, p. 14191-14196.

Fichefet, T., Poncin, C., Goosse, H., Huybrechts, P., Janssens, I., and Le Treut, H., 2003, Implications of changes in freshwater flux from the Greenland ice sheet for the climate of the 21st century: Geophysical Research Letters, v. 30, no. 17.

Flowers, R. M., Ketcham, R. A., Shuster, D. L., and Farley, K. A., 2009, Apatite (U– Th)/He thermochronometry using a radiation damage accumulation and annealing model: Geochimica et Cosmochimica Acta, v. 73, no. 8, p. 2347-2365.

Fosdick, J. C., Grove, M., Hourigan, J. K., and Calderon, M., 2013, Retroarc deformation and exhumation near the end of the Andes, southern Patagonia: Earth and Planetary Science Letters, v. 361, p. 504-517.

Fountain, A. G., Fernandez-Diaz, J. C., Obryk, M., Levy, J., Gooseff, M., Van Horn, D. J., Morin, P., and Shrestha, R., 2017, High-resolution elevation mapping of the

139

McMurdo Dry Valleys, Antarctica, and surrounding regions: Earth System Science Data, v. 9, no. 2, p. 435-443.

Funder, S., Bennike, O., Bocher, J., Israelson, C., Petersen, K. S., and Simonarson, L. A., 2001, Late Pliocene Greenland - The Kap Kobenhavn Formation in North Greenland: Bulletin of the Geological Society of Denmark, v. 48, p. 117-134.

Gjermundsen, E. F., Briner, J. P., Akçar, N., Foros, J., Kubik, P. W., Salvigsen, O., and Hormes, A., 2015, Minimal erosion of Arctic alpine topography during late Quaternary glaciation: Nature Geoscience, v. 8, p. 789 – 792.

Glasser, N. F., Goodsell, B., Copland, L., and Lawson, W., 2006, Debris characteristics and ice-shelf dynamics in the ablation region of the McMurdo Ice Shelf, Antarctica: Journal of Glaciology, v. 52, no. 177, p. 223-234.

Glasser, N. F., Holt, T., Fleming, E., and Stevenson, C., 2014, Ice shelf history determined from deformation styles in surface debris: Antarctic Science, v. 26, no. 6, p. 661- 673.

Goehring, B. M., Kelly, M. A., Schaefer, J. M., Finkel, R. C., and Lowell, T. V., 2010, Dating of raised marine and lacustrine deposits in east Greenland using beryllium- 10 depth profiles and implications for estimates of subglacial erosion: Journal of Quaternary Science, v. 25, no. 6, p. 865-874.

Golledge, N. R., Levy, R. H., McKay, R. M., Fogwill, C. J., White, D. A., Graham, A. G. C., Smith, J. A., Hillenbrand, C. D., Licht, K. J., Denton, G. H., Ackert, R. P., Maas, S. M., and Hall, B. L., 2013, Glaciology and geological signature of the Last Glacial Maximum Antarctic ice sheet: Quaternary Science Reviews, v. 78, p. 225-247.

Graly, J. A., Bierman, P. R., Reusser, L. J., and Pavich, M. J., 2010, Meteoric 10Be in soil profiles - A global meta-analysis: Geochimica Et Cosmochimica Acta, v. 74, no. 23, p. 6814-6829.

Graly, J. A., Corbett, L. B., Bierman, P. R., Lini, A., and Neumann, T. A., 2018, Meteoric 10Be as a tracer of subglacial processes and interglacial surface exposure in Greenland: Quaternary Science Reviews, v. 191, p. 118-131.

Granger, D. E., Lifton, N. A., and Willenbring, J. K., 2013, A cosmic trip: 25 years of cosmogenic nuclides in geology: Geological Society of America Bulletin, v. 125, no. 9-10, p. 1379-1402.

Greenwood, S. L., Gyllencreutz, R., Jakobsson, M., and Anderson, J. B., 2012, Ice-flow switching and East/West Antarctic Ice Sheet roles in glaciation of the western Ross Sea: Geological Society of America Bulletin, v. 124, no. 11-12, p. 1736-1749.

140

Greenwood, S. L., Simkins, L. M., Halberstadt, A. R. W., Prothro, L. O., and Anderson, J. B., 2018, Holocene reconfiguration and readvance of the East Antarctic Ice Sheet: Nature communications, v. 9, no. 1, p. 3176-3176.

Gregersen, U., Knutz, P. C., Nohr-Hansen, H., Sheldon, E., and Hopper, J. R., 2019, Tectonostratigraphy and evolution of the West Greenland continental margin: Bulletin of the Geological Society of Denmark, v. 67, p. 1-21.

Halberstadt, A. R. W., Simkins, L. M., Greenwood, S. L., and Anderson, J. B., 2016, Past ice-sheet behaviour: retreat scenarios and changing controls in the Ross Sea, Antarctica: Cryosphere, v. 10, no. 3, p. 1003-1020.

Hall, B. L., and Denton, G. H., 2000, Radiocarbon chronology of Ross Sea drift, eastern Taylor Valley, Antarctica: Evidence for a grounded ice sheet in the Ross Sea at the last glacial maximum: Geografiska Annaler Series a-Physical Geography, v. 82A, no. 2-3, p. 305-336.

Hall, B. L., Baroni, C., and Denton, G. H., 2004, Holocene relative sea-level history of the Southern Victoria Land Coast, Antarctica: Global and Planetary Change, v. 42, no. 1-4, p. 241-263.

Hall, B. L., Bromley, G., Stone, J., and Conway, H., 2017, Holocene ice recession at Polygon Spur, Reedy Glacier, Antarctica: The Holocene, v. 27, no. 1, p. 122-129.

Hall, B. L., Denton, G. H., and Hendy, C. H., 2000, Evidence from Taylor Valley for a grounded ice sheet in the Ross Sea, Antarctica: Geografiska Annaler Series a- Physical Geography, v. 82A, no. 2-3, p. 275-303.

Hall, B. L., Denton, G. H., Heath, S. L., Jackson, M. S., and Koffman, T. N. B., 2015, Accumulation and marine forcing of ice dynamics in the western Ross Sea during the last deglaciation: Nature Geoscience, v. 8, no. 8, p. 625-628.

Hall, B. L., Denton, G. H., Stone, J. O., and Conway, H., 2013, History of the grounded ice sheet in the Ross Sea sector of Antarctica during the Last Glacial Maximum and the last termination: Antarctic Palaeoenvironments and Earth-Surface Processes, v. 381, p. 167-181.

Hall, B. L., Henderson, G. M., Baroni, C., and Kellogg, T. B., 2010, Constant Holocene Southern- Ocean 14C reservoir ages and ice-shelf flow rates: Earth and Planetary Science Letters, v. 296, no. 1, p. 115-123.

Harden, S. L., Demaster, D. J., and Nittrouer, C. A., 1992, Developing sediment geochronologies for high latitude continental-shelf deposits- a radiochemical approach: Marine Geology, v. 103, no. 1-3, p. 69-97.

141

Haywood, A. M., Hill, D. J., Dolan, A. M., Otto-Bliesner, B. L., Bragg, F., Chan, W. L., Chandler, M. A., Contoux, C., Dowsett, H. J., Jost, A., Kamae, Y., Lohmann, G., Lunt, D. J., Abe-Ouchi, A., Pickering, S. J., Ramstein, G., Rosenbloom, N. A., Salzmann, U., Sohl, L., Stepanek, C., Ueda, H., Yan, Q., and Zhang, Z., 2013, Large-scale features of Pliocene climate: results from the Pliocene Model Intercomparison Project: Climate of the Past, v. 9, no. 1, p. 191-209.

Hein, A. S., Fogwill, C. J., Sugden, D. E., and Xu, S., 2014, Geological scatter of cosmogenic-nuclide exposure ages in the Shackleton Range, Antarctica: Implications for glacial history: Quaternary Geochronology, v. 19, p. 52-66.

Hein, A.S., Fogwill, C.J., Sugden, D.E., Xu, S., 2014. Geological scatter of cosmogenic- nuclide exposure ages in the Shackleton Range, Antarctica: Implications for glacial history. Quaternary Geochronology 19, 52-66.

Helland, P. E., and Holmes, M. A., 1997, Surface textural analysis of quartz sand grains from ODP Site 918 off the southeast coast of Greenland suggests glaciation of southern Greenland at 11 Ma: Palaeogeography Palaeoclimatology Palaeoecology, v. 135, no. 1-4, p. 109-121.

Helm, V., Humbert, A., and Miller, H., 2014, Elevation and elevation change of Greenland and Antarctica derived from CryoSat-2: Cryosphere, v. 8, no. 4, p. 1539-1559.

Henriksen, N., Higgins, A. K., Kalsbeek, F., and Pulvertaft, C. R., 2009. Greenland from Archaean to Quaternary Descriptive text to the 1995 Geological map of Greenland, 1:2,500,000.: Geological Survey of Denmark and Greenland Bulletin, v. 18, p. 126.

Howat, I. M., Porter, C., Smith, B. E., Noh, M. J., and Morin, P., 2019, The Reference Elevation Model of Antarctica: The Cryosphere, v. 13, p. 665-674.

Huang, Y., Li, B., Bryant, C., Bol, R., and Eglinton, G., 1999, Radiocarbon Dating of Aliphatic Hydrocarbons: A New Approach for Dating Passive-Fraction Carbon in Soil Horizons: Soil Science Society of America Journal, v. 63, no. 5, p. 1181-1187.

Jackson, M. S., Hall, B. L., and Denton, G. H., 2018, Asynchronous behavior of the Antarctic Ice Sheet and local glaciers during and since Termination 1, Salmon Valley, Antarctica: Earth and Planetary Science Letters, v. 482, p. 396-406.

Jenkins, A., Dutrieux, P., Jacobs, S. S., Mcphail, S. D., Perrett, J. R., Webb, A. T., and White, D., 2010, Observations beneath Pine Island Glacier in West Antarctica and implications for its retreat: Nature Geoscience, v. 3, no. 7, p. 468-472.

Jones, R. S., Mackintosh, A. N., Norton, K. P., Golledge, N. R., Fogwill, C. J., Kubik, P. W., Christl, M., and Greenwood, S. L., 2015, Rapid Holocene thinning of an East

142

Antarctic outlet glacier driven by marine ice sheet instability: Nature Communications, v. 6.

Joughin, I., Smith, B. E., and Medley, B., 2014, Marine Ice Sheet Collapse Potentially Under Way for the Thwaites Glacier Basin, West Antarctica: Science, v. 344, no. 6185, p. 735-738.

Joy, K., Fink, D., Storey, B., and Atkins, C., 2014, A 2 million year glacial chronology of the Hatherton Glacier, Antarctica and implications for the size of the East Antarctic Ice Sheet at the Last Glacial Maximum: Quaternary Science Reviews, v. 83, p. 46- 57.

Kellogg, T. B., Kellogg, D. E., and Stuiver, M., 1990, Late Quaternary history of the southwestern Ross Sea: evidence from debris bands on the McMurdo Ice Shelf, Antarctica: Contributions to Antarctic Research, v. 50, p. 25-56.

Ketcham, R. A., Gautheron, C., and Tassan-Got, L., 2011, Accounting for long alpha- particle stopping distances in (U–Th–Sm)/He geochronology: Refinement of the baseline case: Geochimica et Cosmochimica Acta, v. 75, no. 24, p. 7779-7791.

Knutz, P. C., Newton, A. M. W., Hopper, J. R., Huuse, M., Gregersen, U., Sheldon, E., and Dybkjær, K., 2019, Eleven phases of Greenland Ice Sheet shelf-edge advance over the past 2.7 million years: Nature Geoscience, v. 12, p. 361-368.

Koenig, S. J., DeConto, R. M., and Pollard, D., 2014, Impact of reduced Arctic sea ice on Greenland ice sheet variability in a warmer than present climate: Geophysical Research Letters, v. 41, no. 11, p. 3933-3942.

Kohl, C. P., and Nishiizumi, K., 1992, Chemical isolation of quartz for measurement of in- situ-produced cosmogenic nuclides: Geochimica Et Cosmochimica Acta, v. 56, no. 9, p. 3583-3587.

Kyle, P. R., 1990, McMurdo volcanic group, western Ross Embayment, Volcanoes of the Antarctic Plate and Southern Oceans, American Geophysical Union.

Lal, D., 1998, Cosmic ray produced isotopes in terrestrial systems: Proceedings of the Indian Academy of Sciences Earth & Planetary Sciences, v. 107, no. 4, p. 241-249.

Lambeck, K., Rouby, H., Purcell, A., Sun, Y. Y., and Sambridge, M., 2014, Sea level and global ice volumes from the Last Glacial Maximum to the Holocene: Proceedings of the National Academy of Sciences of the United States of America, v. 111, no. 43, p. 15296-15303.

143

Lee, J. I., McKay, R. M., Golledge, N. R., Yoon, H. I., Yoo, K. C., Kim, H. J., and Hong, J. K., 2017, Widespread persistence of expanded East Antarctic glaciers in the southwest Ross Sea during the last deglaciation: Geology, v. 45, no. 5, p. 403-406.

Levy, J. S., Fountain, A. G., O'Connor, J. E., Welch, K. A., and Lyons, W. B., 2013, Garwood Valley, Antarctica: A new record of Last Glacial Maximum to Holocene glaciofluvial processes in the McMurdo Dry Valleys: Geological Society of America Bulletin, v. 125, no. 9-10, p. 1484-1502.

Levy, J.S., Rittenour, T.M., Fountain, A.G., and O’Connor, J.E., 2017, Luminescence dating of paleolake deltas and glacial deposits in Garwood Valley, Antarctica: Implications for climate, Ross ice sheet dynamics, and paleolake duration. Geological Society of America Bulletin, v. 129, no. 9-10, p. 1071-1084.

Lewis, A. R., Marchant, D. R., Kowalewski, D. E., Baldwin, S. L., and 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, v. 34, no. 7, p. 513-516.

Licht, K. J., Jennings, A. E., Andrews, J. T., and Williams, K. M., 1996, Chronology of late Wisconsin ice retreat from the western Ross Sea, Antarctica: Geology, v. 24, no. 3, p. 223-226.

Liu, J., Milne, G. A., Kopp, R. E., Clark, P. U., and Shennan, I., 2016, Sea-level constraints on the amplitude and source distribution of Meltwater Pulse 1A: Nature Geoscience, v. 9, no. 2, p. 130-134.

Marchant, D. R., Denton, G. H., Bockheim, J. G., Wilson, S. C., and Kerr, A. R., 1994, Quaternary changes in level of the upper Taylor Glacier, Antarctica - implications for paleoclimate and East Antarctic Ice Sheet dynamics: Boreas, v. 23, no. 1, p. 29- 43.

Martin, A. P., Cooper, A. F., and Dunlap, W. J., 2010, Geochronology of Mount Morning, Antarctica: two-phase evolution of a long-lived trachyte-basanite-phonolite eruptive center: Bulletin of Volcanology, v. 72, no. 3, p. 357-371.

Martos, Y. M., Jordan, T. A., Catalan, M., Jordan, T. M., Bamber, J. L., and Vaughan, D. G., 2018, Geothermal Heat Flux Reveals the Iceland Hotspot Track Underneath Greenland: Geophysical Research Letters, v. 45, no. 16, p. 8214-8222.

McFarlin, J. M., Axford, Y., Osburn, M. R., Kelly, M. A., Osterberg, E. C., and Farnsworth, L. B., 2018, Pronounced summer warming in northwest Greenland during the Holocene and Last Interglacial: Proceedings of the National Academy of Sciences of the United States of America, v. 115, no. 25, p. 6357-6362.

144

McKay, R., Golledge, N. R., Maas, S., Naish, T., Levy, R., Dunbar, G., and Kuhn, G., 2016, Antarctic marine ice-sheet retreat in the Ross Sea during the early Holocene: Geology, v. 44, no. 1, p. 7-10.

McKay, R., Naish, T., Powell, R., Barrett, P., Scherer, R., Talarico, F., Kyle, P., Monien, D., Kuhn, G., Jackolski, C., and Williams, T., 2012, Pleistocene variability of Antarctic Ice Sheet extent in the Ross Embayment: Quaternary Science Reviews, v. 34, p. 93-112.

Mengel, M., Levermann, A., Frieler, K., Robinson, A., Marzeion, B., and Winkelmann, R., 2016, Future sea level rise constrained by observations and long-term commitment: Proceedings of the National Academy of Sciences of the United States of America, v. 113, no. 10, p. 2597-2602.

Miller, K. G., Wright, J. D., Browning, J. V., Kulpecz, A., Kominz, M., Naish, T. R., Cramer, B. S., Rosenthal, Y., Peltier, W. R., and Sosdian, S., 2012, High tide of the warm Pliocene: Implications of global sea level for Antarctic deglaciation: Geology, v. 40, no. 5, p. 407-410.

Moros, M., Lloyd, J. M., Perner, K., Krawczyk, D., Blanz, T., de Vernal, A., Ouellet- Bernier, M. M., Kuijpers, A., Jennings, A. E., Witkowski, A., Schneider, R., and Jansen, E., 2016, Surface and sub-surface multi-proxy reconstruction of middle to late Holocene palaeoceanographic changes in Disko Bugt, West Greenland: Quaternary Science Reviews, v. 132, p. 146-160.

Morrison, J., Brockwell, T., Merren, T., Fourel, F., and Phillips, A. M., 2001, On-line high- precision stable hydrogen isotopic analyses on nanoliter water samples: Analytical Chemistry, v. 73, no. 15, p. 3570-3575.

Mouginot, J., Rignot, E., and Scheuchl, B., 2014, Sustained increase in ice discharge fromthe Amundsen Sea Embayment, West Antarctica, from 1973 to 2013: Geophysical Research Letters, v. 41, no. 5, p. 1576-1584.

Naish, T., Powell, R., Levy, R., Wilson, G., Scherer, R., Talarico, F., Krissek, L., Niessen, F., Pompilio, M., Wilson, T., Carter, L., DeConto, R., Huybers, P., McKay, R., Pollard, D., Ross, J., Winter, D., Barrett, P., Browne, G., Cody, R., Cowan, E., Crampton, J., Dunbar, G., Dunbar, N., Florindo, F., Gebhardt, C., Graham, I., Hannah, M., Hansaraj, D., Harwood, D., Helling, D., Henrys, S., Hinnov, L., Kuhn, G., Kyle, P., Laeufer, A., Maffioli, P., Magens, D., Mandernack, K., McIntosh, W., Millan, C., Morin, R., Ohneiser, C., Paulsen, T., Persico, D., Raine, I., Reed, J., Riesselman, C., Sagnotti, L., Schmitt, D., Sjunneskog, C., Strong, P., Taviani, M.,

145

Vogel, S., Wilch, T., Williams, T., 2009, Obliquity-paced Pliocene West Antarctic ice sheet oscillations. Nature, v. 458, p. 322-328.

Nelson, A. H., Bierman, P. R., Shakun, J. D., and Rood, D. H., 2014, Using in situ cosmogenic 10Be to identify the source of sediment leaving Greenland: Earth Surface Processes and Landforms, v. 39, no. 8, p. 1087-1100.

Nishiizumi, K., 2004, Preparation of 26Al AMS standards: Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, v. 223-224, p. 388 - 392.

Nishiizumi, K., Imamura, M., Caffee, M. W., Southon, J. R., Finkel, R. C., and McAninch, J., 2007, Absolute calibration of 10Be AMS standards: Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, v. 258, no. 2, p. 403-413.

Ohkouchi, N., and T. I. Eglinton, 2006, Radiocarbon constraint on relict organic carbon contributions to Ross Sea sediments, Geochemistry Geophysics Geosystems, v. 7, no. 4, Q04012.

Pagani, M., Huber, M., Liu, Z. H., Bohaty, S. M., Henderiks, J., Sijp, W., Krishnan, S., and DeConto, R. M., 2011, The Role of Carbon Dioxide During the Onset of Antarctic Glaciation: Science, v. 334, no. 6060, p. 1261-1264.

Pagani, M., Liu, Z. H., LaRiviere, J., and Ravelo, A. C., 2010, High Earth-system climate sensitivity determined from Pliocene carbon dioxide concentrations: Nature Geoscience, v. 3, no. 1, p. 27-30.

Pollard, W., Doran, P., and Wharton, R., The nature and significance of massive ground ice in Ross Sea Drift, Garwood Valley, McMurdo Sound, in Proceedings Antarctica at the close of a millenium: proceedings of the 8th International Symposium on Antarctic Earth Sciences, Wellington, New Zealand, 1999, Royal Society of New Zealand, p. 397-404.

Portenga, E.W., Bierman, P.R, 2011, Understanding earth’s eroding surface with 10Be: GSA Today, v. 21, no. 8, p. 4- 10.

Prothro, L. O., Simkins, L. M., Majewski, W., and Anderson, J. B., 2018, Glacial retreat patterns and processes determined from integrated sedimentology and geomorphology records: Marine Geology, v. 395, p. 104-119.

Reimer, P. J., et al., 2013, INTCAL13 and MARINE13 radiocarbon age calibration curves 0-50,000 years CAL BP: Radiocarbon, v. 55, no. 4, p. 1869-1887.

146

Reiners , P. W., and Ehlers, T. A., 2005, Low-Temperature Thermochronology: Techniques, Interpretations, and Applications, Chantilly, VA, The Mineralogical Society of America, Reviews in Mineralogy & Geochemistry, 622 p.:

Reiners , P.W., Ehlers, T.A., 2005. Low-Temperature Thermochronology: Techniques, Interpretations, and Applications. The Mineralogical Society of America, Chantilly, VA.

Reiners, P. W., and Brandon, M. T., 2006, Using thermochronology to understand orogenic erosion: Annual Review of Earth and Planetary Sciences, v. 34, p. 419-466.

Reyes, A. V., Carlson, A. E., Beard, B. L., Hatfield, R. G., Stoner, J. S., Winsor, K., Welke, B., and Ullman, D. J., 2014, South Greenland ice-sheet collapse during Marine Isotope Stage 11: Nature, v. 510, no. 7506, p. 525-528.

Reyes, A. V., Carlson, A. E., Beard, B. L., Hatfield, R. G., Stoner, J. S., Winsor, K., Welke, B., and Ullman, D. J., 2014, South Greenland ice-sheet collapse during Marine Isotope Stage 11: Nature, v. 510, no. 7506, p. 525-528.

Rignot, E., Velicogna, I., van den Broeke, M. R., Monaghan, A., and Lenaerts, J., 2011, Acceleration of the contribution of the Greenland and Antarctic ice sheets to sea level rise: Geophysical Research Letters, v. 38.

Rovere, A., Raymo, M. E., Mitrovica, J. X., Hearty, P. J., O'Leary, M. J., and Inglis, J. D., 2014, The Mid-Pliocene sea-level conundrum: Glacial isostasy, eustasy and dynamic topography: Earth and Planetary Science Letters, v. 387, p. 27-33.

Sachs, J. P., Stein, R., Maloney, A. E., Wolhowe, M., Fahl, K., and Nam, S. I., 2018, An Arctic Ocean paleosalinity proxy from delta H-2 of palmitic acid provides evidence for deglacial Mackenzie River flood events: Quaternary Science Reviews, v. 198, p. 76-90.

Schaefer, J. M., Finkel, R. C., Balco, G., Alley, R. B., Caffee, M. W., Briner, J. P., Young, N. E., Gow, A. J., and Schwartz, R., 2016, Greenland was nearly ice-free for extended periods during the Pleistocene: Nature, v. 540, no. 7632, p. 252-255.

Scherer, R. P., Bohaty, S. M., Dunbar, R. B., Esper, O., Flores, J. A., Gersonde, R., Harwood, D. M., Roberts, A. P., and Taviani, M., 2008, Antarctic records of precession-paced insolation-driven warming during early Pleistocene Marine Isotope Stage 31: Geophysical Research Letters, v. 35, no. 3, L03505

Shakun, J. D., Corbett, L. B., Bierman, P. R., Underwood, K., Rizzo, D. M., Zimmerman, S. R., Caffee, M. W., Naish, T., Golledge, N. R., and Hay, C. C., 2018, Minimal

147

East Antarctic Ice Sheet retreat onto land during the past eight million years: Nature, v. 558, no. 7709, p. 284-287.

Shakun, J., Corbett, L., Bierman, P., and Zimmerman, S., Pliocene Greenland ice sheet growth recorded by in situ 10Be decrease in multiple marine sediment, in Proceedings GSA Abstracts with Programs, Seattle, WA, 2017, Volume 49.

Shipp, S., Anderson, J., and Domack, E., 1999, Late Pleistocene-Holocene retreat of the West Antarctic Ice-Sheet system in the Ross Sea: Part 1 - Geophysical results: Geological Society of America Bulletin, v. 111, no. 10, p. 1486-1516.

Simkins, L.M., Anderson, J.B., Greenwood, S.L., Gonnerman, H.M., Prothro, L.O., Halberstadt, A.R.W., Stearns, L.A., Pollard, D., Deconto, R.M., 2017, Anatomy of a meltwater drainage system benearth the ancestral East Antarctic Ice Sheet, Nature Geoscience, v. 10, p 691 – 697.

Spector, P., Stone, J., Cowdery, S. G., Hall, B., Conway, H., and Bromley, G., 2017, Rapid early-Holocene deglaciation in the Ross Sea, Antarctica: Geophysical Research Letters, v. 44, no. 15, p. 7817-7825.

St John, K. E. K., and Krissek, L. A., 2002, The late Miocene to Pleistocene ice-rafting history of southeast Greenland: Boreas, v. 31, no. 1, p. 28-35.

Steig, E. J., Morse, D. L., Waddington, E. D., Stuiver, M., Grootes, P. M., Mayewski, P. A., Twickler, M. S., and Whitlow, S. I., 2000, Wisconsinan and Holocene climate history from an ice core at Taylor Dome, western Ross Embayment, Antarctica: Geografiska Annaler Series a-Physical Geography, v. 82A, no. 2-3, p. 213-235.

Stone, J. O., Balco, G. A., Sugden, D. E., Caffee, M. W., Sass, L. C., Cowdery, S. G., and Siddoway, C., 2003, Holocene deglaciation of Marie Byrd Land, West Antarctica: Science, v. 299, no. 5603, p. 99-102.

Storey, B. C., Fink, D., Hood, D., Joy, K., Shulmeister, J., Riger-Kusk, M., and Stevens, M. I., 2010, Cosmogenic nuclide exposure age constraints on the glacial history of the Lake Wellman area, Darwin Mountains, Antarctica: Antarctic Science, v. 22, no. 6, p. 603-618.

Stuiver, M., Denton, G. H., Hughes, T. J., and Fastook, J. L., 1981, History of the marine ice sheet in West Antarctica during the last glaciation: A working hypothesis, in Denton, G. H., and Hughes, T. J., eds., The Last Great Ice Sheets: New York, John Wiley and Sons.

148

Sugden, D., and Denton, G., 2004, Cenozoic landscape evolution of the Convoy Range to Mackay Glacier area, Transantarctic Mountains: Onshore to offshore synthesis: Geological Society of America Bulletin, v. 116, no. 7-8, p. 840-857.

Swanger, K. M., Lamp, J. L., Winckler, G., Schaefer, J. M., and Marchant, D. R., 2017, Glacier advance during Marine Isotope Stage 11 in the McMurdo Dry Valleys of Antarctica: Scientific Reports, v. 7., 41433

Talarico, F. M., and Sandroni, S., 2011, Early Miocene basement clasts in ANDRILL AND-2A core and their implications for paleoenvironmental changes in the McMurdo Sound region (western Ross Sea, Antarctica): Global and Planetary Change, v. 78, no. 1-2, p. 23-35.

Talarico, F. M., McKay, R. M., Powell, R. D., Sandroni, S., and Naish, T., 2012, Late Cenozoic oscillations of Antarctic ice sheets revealed by provenance of basement clasts and grain detrital modes in ANDRILL core AND-1B: Global and Planetary Change, v. 96-97, p. 23-40.

Talarico, F. M., Pace, D., and Levy, R. H., 2013, Provenance of basement erratics in Quaternary coastal moraines, southern McMurdo Sound, and implications for the source of Eocene sedimentary rocks: Antarctic Science, v. 25, no. 5, p. 681-695.

Thomas, E. K., Briner, J. P., Ryan-Henry, J. J., and Huang, Y., 2016, A major increase in winter snowfall during the middle Holocene on western Greenland caused by reduced sea ice in Baffin Bay and the Labrador Sea: Geophysical Research Letters, v. 43, no. 10, p. 5302-5308.

Thomas, E. K., Castaneda, I. S., McKay, N. P., Briner, J. P., Salacup, J. M., Nguyen, K. Q., and Schweinsberg, A. D., 2018, A Wetter Arctic Coincident With Hemispheric Warming 8,000 Years Ago: Geophysical Research Letters, v. 45, no. 19, p. 10637- 10647.

Thomson, S. N., Brandon, M. T., Tomkin, J. H., Reiners, P. W., Vasquez, C., and Wilson, N. J., 2010, Glaciation as a destructive and constructive control on mountain building: Nature, v. 467, no. 7313, p. 313-317.

Tochilin, C. J., Reiners, P. W., Thomson, S. N., Gehrels, G. E., Hemming, S. R., and Pierce, E. L., 2012, Erosional history of the Prydz Bay sector of East Antarctica from detrital apatite and zircon geo- and thermochronology multidating: Geochemistry Geophysics Geosystems, v. 13, no. 11, p. 1-21.

Todd, C., Stone, J., Conway, H., Hall, B., and Bromley, G., 2010, Late Quaternary evolution of Reedy Glacier, Antarctica: Quaternary Science Reviews, v. 29, no. 11- 12, p. 1328-1341.

149

Tripati, A. K., Eagle, R. A., Morton, A., Dowdeswell, J. A., Atkinson, K. L., Bahe, Y., Dawber, C. F., Khadun, E., Shaw, R. M. H., Shorttle, O., and Thanabalasundaram, L., 2008, Evidence for glaciation in the Northern Hemisphere back to 44 Ma from ice-rafted debris in the Greenland Sea: Earth and Planetary Science Letters, v. 265, no. 1-2, p. 112-122.

Vella, P., 1969, Surficial geological sequence, Black Island and Brown Peninsula, McMurdo Sound, Antarctica, New Zealand Journal of Geology and Geophysics, v. 12, no. 4, p. 761-770.

Wellner, J. S., Lowe, A. L., Shipp, S. S., and Anderson, J. B., 2001, Distribution of glacial geomorphic features on the Antarctic continental shelf and correlation with substrate: implications for ice behavior: Journal of Glaciology, v. 47, no. 158, p. 397-411.

Wilson, G. S., 2000, Glacial geology and origin of fossiliferous-erratic bearing moraines, southern McMurdo Sound, Antarctica - an alternative ice sheet hypothesis, in Stilwel, J. D., and Feldman, R. M., eds., Paleobiology and paleoenvironments of Eocene rocks, McMurdo Sound, East Antarctica, Volume 76: Washington DC, American Geophysical Union, p. 19-37.

Wright-Grassham, A., 1987, Volcanic geology, mineralogy, and petrogenesis of the Discovery volcanic subprovince, South Victoria Land [PhD]: New Mexico Institute of Mining and Technology, 512 p.

Yokoyama, Y., Anderson, J. B., Yamane, M., Simkins, L. M., Miyairi, Y., Yamazaki, T., Koizumi, M., Suga, H., Kusahara, K., Prothro, L., Hasumi, H., Southon, J. R., and Ohkouchi, N., 2016, Widespread collapse of the Ross Ice Shelf during the late Holocene: Proceedings of the National Academy of Sciences of the United States of America, v. 113, no. 9, p. 2354-2359.

150

CURRICULUM VITAE

151 152

153 154

155