EROSIONAL HISTORY OF EAST ANTARCTICA FROM DOUBLE

advertisement
EROSIONAL HISTORY OF EAST ANTARCTICA FROM DOUBLE
AND TRIPLE-DATING OF SINGLE GRAINS IN GLACIALLYDERIVED SEDIMENTS FROM PRYDZ BAY
by
Clare J. Tochilin
A Prepublication Manuscript Submitted to the Faculty of the
DEPARTMENT OF GEOSCIENCES
In Partial Fulfillment of the Requirements
for the Degree of
MASTER OF SCIENCE
In the Graduate College
THE UNIVERSITY OF ARIZONA
May 7th, 2012
STATEMENT BY THE AUTHOR
This manuscript, prepared for publication in Geochemistry Geophysics Geosystems, has
been submitted in partial fulfillment of requirements for the Master of Science degree at
The University of Arizona and is deposited in the Antevs Reading Room to be made
available to borrowers, as are copies of regular theses and dissertations.
Brief quotations from this manuscript are allowable without special permission, provided
that accurate acknowledgment of the source is made. Requests for permission for
extended quotation from or reproduction of this manuscript in whole or in part may be
granted by the Department of Geosciences when the proposed use of the material is in the
interests of scholarship. In all other instances, however, permission must be obtained
from the author.
(author 's signature)
(date)
APPROVAL BY RESEARCH COMMITTEE
As members of the Research Committee, we recommend that this prepublication
manuscript be accepted as fulfilling the research requirement for the degree of Master of
Science.
2.
Dr. Peter W. Reiners
Major Advisor
(signature)
(date)
(signature)
(date)
2otz_
Dr. George E. Gehrels
Dr. Stuart N . Thomson
Lf{., I
{J
1
(signature)
s/:r/1z..
.
1
(dat/i
Erosional history of East Antarctica from double and triple-dating of single
grains in glacially-derived sediments from Prydz Bay
Clare J. Tochilin1, Peter W. Reiners1, Stuart N. Thomson1, George E. Gehrels1, Sidney R.
Hemming1
1
Department of Geosciences, University of Arizona, Tucson, AZ
2
Lamont-Doherty Earth Observatory and Department of Earth and Environmental Sciences,
Columbia University, Palisades, NY
Abstract
Since its onset in the early Oligocene, the East Antarctic Ice Sheet has covered up to 98%
of the surface of East Antarctica. Because bedrock exposure is limited, little is known about its
tectonic and erosional history. To better constrain the subglacial erosional record, we analyzed
Oligocene through Quaternary sediments from Prydz Bay, as these sediments contain geo- and
thermochronologic information about the bedrock from which they are derived. A multi-dating
approach was applied in the analysis of these sediments, involving U-Pb, fission track, and (UTh)/He dating of single apatite and zircon grains to measure a range of crystallization and
cooling ages. Apatite and zircon U-Pb dates of grains spanning the entire sampled stratigraphic
section are characterized by a dominant ~500 Ma age peak, recording widespread metamorphism
and magmatism related to Pan-African orogenesis at this time. Zircon fission track dates record
two periods of cooling at ~250-300 Ma and ~120 Ma due to rifting in the Lambert Graben during
the Permian-Triassic and magmatic resetting during the Cretaceous. Mean apatite fission track
dates show a decrease from ~280 Ma in early Oligocene samples to ~180 Ma in late Miocene
samples, but show little change from ~180 Ma to ~150 Ma in Pliocene through Quaternary
samples. These ages demonstrate a rapid decrease in lag time from ~250-180 My throughout the
early Oligocene, indicating increased erosion rates during this time. The youngest measured
apatite He ages also decrease from ~100 Ma in the Oligocene sediments to ~20 Ma in the
Miocene sediments. Taken together, these results indicate a significant increase in erosion rates
!
1!
in the catchments draining to Prydz Bay between the early Oligocene and late Miocene, with
relatively steady rates of erosion since that time. This erosion was primarily achieved by glacial
incision into preexisting river valleys, based on glacial erosional modeling and observed valley
shapes. Glaciers must have incised to depths of ~2.8-3.0 km by the late Miocene. This is
consistent with models of the EAIS that show that it transitioned to less erosive cold-based
conditions following the middle Miocene climatic optimum.
1. Introduction
The tectonic and erosional history of East Antarctica is poorly understood due to
coverage of about 98% of the bedrock by the East Antarctic Ice Sheet (EAIS) (Barron et al.,
1991; Webb, 1990). The EAIS is the largest and longest-lived ice sheet on Earth and
emcompasses ~90% of the ice in Antarctica (Bamber et al., 2000). One nucleation point for the
onset of glaciation in East Antarctica in the early Oligocene is though to be the subglacial
Gamburtsev Mountains in the interior of the continent, which reach elevations of up to 4 km
above sea level, but are entirely covered by ice (Bo et al., 2009; Cox et al., 2007; Cox et al.,
2010; DeConto and Pollard, 2003a, 2003b; Siegert, 2008; Siegert et al., 2008; Taylor et al., 2004;
van de Flierdt et al., 2008). Rapid onset of the EAIS began at ~34 Ma as a result of global
cooling related to declining atmospheric CO2 levels and the development of the Antarctic
Circumpolar Current and consequent thermal isolation of Antarctica. Glaciation remained
dynamic and warm-based until ~14 Ma, after which time it underwent a shift to more stable,
cold-based, and potentially less erosive glaciation as a result of further cooling (Bo et al., 2009;
DeConto and Pollard, 2003a, 2003b; Jamieson and Sugden, 2008; Jamieson et al., 2010; Naish et
al., 2001; Sugden, 1996; Sugden et al., 1993; Young et al., 2011; Zachos et al., 1992).
!
2!
Approximately 10-20% of the EAIS is drained by the largest ice stream in Antarctica, the
Lambert Glacier, which flows through the north-south trending Lambert Graben and into Prydz
Bay on the northeastern margin of the continent (Fig. 1) (Bamber et al., 2000; Barrett et al.,
1996; Cox et al., 2010, 2007; DeConto and Pollard, 2003a; Hambrey and McKelvey, 2000;
Jamieson et al., 2005). The Lambert Graben was formed primarily by Permo-Triassic rifting
(Cox et al., 2010; Kurinin and Grikurov, 1982; Lisker, 2002; Lisker et al., 2007b) and extends
~500 km to the south of Prydz Bay into the interior of the continent (O’Brien et al., 2007).
Small-scale mafic magmatism associated with the break-up of Gondwana occurred in the
Lambert Graben/Prydz Bay area during the early to mid Cretaceous (Arne, 1994; Collerson and
Sheraton, 1986; Hambrey and McKelvey, 2000; Jamieson et al., 2005; Kurinin and Grikurov,
1982; Lisker et al., 2007a, 2007b, 2003).
In addition to hosting an ice stream currently draining a large portion of the EAIS, the
Lambert Graben is proposed to have been an important drainage route for fluvial systems prior to
the onset of glaciation in East Antarctica. Jamieson and Sugden (2008) modeled the preglacial
fluvial drainage patterns of this area (Fig. 1), showing that the Lambert Graben served as the
primary path for rivers carrying sediment out of the Gamburtsev Mountains. This drainage
system was also fed by sediments from the Prince Charles Mountains and Grove Mountains
along the flanks of the Lambert Graben near Prydz Bay (Cox et al., 2010; Jamieson et al., 2005;
O’Brien et al., 2007). The current ice streams draining this part of East Antarctica likely follow
these preexisting pre-glacial drainages within the Lambert Graben, Gamburtsev Mountains, and
surrounding areas, ultimately delivering sediment from the East Antarctic interior to Prydz Bay.
To address the erosional history beneath the EAIS, we applied a multi-dating approach to
sediments recovered from Prydz Bay by the Ocean Drilling Project (ODP). These sediments are
!
3!
ultimately derived from the inaccessible subglacial bedrock and can provide geo- and
thermochronologic records of their source areas as well as valuable information about the phases
of East Antarctic glaciation since its onset at 34 Ma (Barron et al., 2001; Whitehead et al., 2006).
The three dating methods applied to detrital minerals in this study are U-Pb (apatite and zircon),
Fission Track (apatite and zircon), and (U-Th)/He (apatite; due to the limited number of zircon
grains in our samples, zircon (U-Th)/He analysis was not conducted). This combination of
multiple high and low temperature chronometers is extremely powerful, as it allows for the
determination of both crystallization ages as well as a range of cooling ages on single grains,
providing a more complete provenance and cooling history than any one method alone. We use
this multi-dating method here to better constrain the subglacial exhumational history of East
Antarctica since the onset of the EAIS.
2. Geologic Setting
The East Antarctic Shield is composed primarily of an amalgamation of Archean-early
Cambrian crustal fragments, deformed and intruded during multiple episodes of tectonism
(Dalziel, 1992; Fitzsimons, 1997, 2000a, 2000b). The presence of Grenville-age and PanAfrican-age metamorphic belts and intrusive rocks indicates the continent’s integral involvement
in the assembly of supercontinents Rodinia and Gondwana, respectively (Boger et al., 2002;
Fitzsimons, 2003, 2000b; Veevers, 2003; Veevers et al., 2008). In some areas within the Lambert
Graben, Precambrian bedrock is overlain by mostly undeformed Devonian-Pliocene sediments
and sedimentary rocks (Fitzsimons, 2000a; McKelvey et al., 2001; O’Brien et al., 2006; Veevers
and Saeed, 2007; Whitehead et al., 2006).
!
4!
Several studies have measured U-Pb zircon dates of grains from the limited ice-free
bedrock exposures in East Antarctica, including samples from the Prydz Bay coast, the margins
of the Lambert Graben, the Prince Charles Mountains, and the Grove Mountains, in order to
characterize portions of the exposed material in this region by age and extrapolate these ages to
what may be beneath the ice (Fig. 2). These studies showed a wide range of crustal formation
ages in East Antarctica ranging from ~3.5 Ga-0.5 Ga, with dominant peaks at 0.4-0.6 Ga, 0.9-1.2
Ga, 1.6 Ga, 2.2 Ga, 2.6-2.7 Ga, 3.0-3.1 Ga, and 3.4-3.5 Ga. Peaks at 0.5-0.6 Ga and 0.9-1.2 Ga
from syntectonic intrusive rocks as well as metamorphic rims on some crystals correlate well
with large-scale Pan-African and Grenville orogenic events known to have affected the region
(Boger et al., 2001; Carson et al., 2000, 1996; Fitzsimons, 2000b; Fitzsimons et al., 1997;
Hemming et al., 2007; Liu et al., 2007; Mikhalsky et al., 2006; van de Flierdt et al., 2008;
Veevers and Saeed, 2007; Veevers et al., 2008; Williams et al., 2007).
Additionally, a number of studies have measured 40Ar/39Ar ages on hornblende and mica
from diamictite in Prydz Bay. In contrast with the variation observed in the U-Pb ages, 40Ar/39Ar
dates from this region of East Antarctica are almost exclusively ~500 Ma, providing further
evidence for a widespread tectonothermal event affecting the area during Pan-African time,
resulting in a strong metamorphic overprint. 40Ar/39Ar studies also show a minor peak of
Grenville age (1.1-1.3 Ma) (Hemming et al., 2007; Phillips et al., 2007; Roy et al., 2007).
Studies measuring apatite fission track and apatite and zircon (U-Th)/He dates of
Precambrian basement rocks, Paleozoic and Mesozoic sedimentary rocks, and sediment cores
from the Prydz Bay region have also been conducted. Cox et al. (2010, 2007) analyzed grains in
Eocene fluvial sediment from cores in Prydz Bay, reporting apatite fission track dates of 207-612
Ma and (U-Th)/He dates of 97-377 Ma and 197-397 Ma for apatite and zircon, respectively.
!
5!
They interpret these ages to reflect very slow erosion rates in the Gamburtsev Mountains and
very little tectonic activity in the interior of the continent since Pan-African time. Other studies,
however, have measured apatite fission track dates in Precambrian basement rocks and
Paleozoic-Mesozoic sedimentary rocks in the Prince Charles Mountains and the Lambert
Graben, showing age distributions primarily from ~100-300 Ma with an older late Paleozoicearly Mesozoic peak and a younger Cretaceous-Paleocene peak (Arne, 1994; Lisker et al., 2003,
2007a, 2007b). These data suggest two phases of cooling: an older peak associated with PermoTriassic rifting in the Lambert Graben area, and a younger peak that these authors propose to
correspond to Cretaceous rifting and exhumation during the breakup of Gondwana. Lisker et al.
(2003, 2007a, 2007b) interpret their data to be consistent with 2-4 km of erosion associated with
a Cretaceous rifting event, although independent geologic evidence for such an event is lacking.
In other areas of East Antarctica such as Dronning Maud Land (55°-15°E) and Wilkes
Land (155°-90°E) several of the same zircon U-Pb age peaks seen in the Prydz Bay/Lambert
Graben region are observed, especially the peaks of Grenville and Pan-African ages
(Fitzsimmons, 2000b; Roy et al., 2007). In Wilkes Land there is also a dominant zircon U-Pb age
peak from 1.6-1.8 Ga not seen in the Prydz Bay catchment (Goodge, 2007). Hornblende
40
Ar/39Ar studies show that all regions of East Antarctica are dominated by 500 Ma dates, with
the exception of Wilkes Land, which shows primarily Mesoproterozoic and Paleoproterozoic
dates. This implies that this area did not experience the Pan-African orogenic event as strongly as
the rest of East Antarctica (Roy et al., 2007).
3. Samples and Stratigraphy
!
6!
The samples analyzed in this study are early Oligocene-Quaternary sediments from ODP
core 739C (67°16.57’S, 75°04.91’) recovered from Prydz Bay in January, 1988 (Shipboard
Scientific Party, 1999). The presence of a thick section of diamictites down to the early
Oligocene recovered in this core and other cores in the vicinity provides evidence for a fully
developed EAIS by that time (Barron et al., 2001; Solheim et al., 1991). Volpi et al. (2009)
reanalyzed the stratigraphy of core 739C (Fig. 3), slightly redefining the stratigraphic and age
divisions previously determined by the Shipboard Scientific Party on the basis of new seismic
profiles, which they integrate with existing lithologic, biostratigraphic, and magnetostratigraphic
data. The stratigraphic ages in Figure 3 and any depositional ages of samples mentioned
throughout this study are those reported by Volpi et al. (2009). Absolute ages were determined
primarily through the dating of diatom fossils (Whitehead et al., 2006).
Core 739C consists of six main units of diamictites (Fig. 3). Unit A, the uppermost
segment of the core, consists of soft Quaternary diatomaceous sand, clay, and poorly
consolidated diamicton. Unit B, 24-80 meters below seafloor (mbsf), is a homogeneous, poorly
sorted diamictite that is late Pliocene in age. Unit C (80-110 mbsf) is a thick, stratified
diamictite, also late Pliocene in age. Unit D (110-140 mbsf) is a highly compacted diamictite of
early Pliocene age. Unit E (140-174 mbsf) is a very highly compacted diamictite divided into
two sections. E1 is diatomaceous and early Pliocene in age, while E2 is is late Miocene with few
to no diatoms. Core 739C is the only core drilled in this area to contain Miocene sediments.
Below the Miocene unit is an unconformity, underlain by a very thick section of early Oligocene
diamictites (174-486.8 mbsf) (Volpi et al., 2009).
We carried out geo/thermochronological analyes on twenty samples from this core,
including one Quaternary sample (001R55: 0-9.5 mbsf), three late Pliocene samples (004R21:
!
7!
24.1-28.7 mbsf; 005R115: 28.7-38.3 mbsf; 013R10: 105.9-115.5 mbsf), three early Pliocene
samples (014R29: 115.5-125.2 mbsf; 017R20: 135.0-140.0; 019R85: 144.7-149.7 mbsf), three
late Miocene samples (021R70: 154.3-159.3 mbsf; 022R74: 159.3-164.0 mbsf; 023R110: 164.0169.0 mbsf), and ten early Oligocene samples (025R10: 173.6-183.2 mbsf; 030R91: 221.8-231.4
mbsf; 031R85: 231.4-241.4 mbsf; 034R80: 260.4-270.0 mbsf; 038R60: 298.9-308.6 mbsf;
041R110: 327.8-337.5 mbsf; 047R55: 385.9-395.6; 052R20: 434.2-439.2; 056R35: 453.5-458.5;
062R0: 482.6-486.8 mbsf).
This dataset is complemented by that of Thomson et al. (2011), in which apatite grains
from Eocene fluvial sediments from ODP Leg 188, Hole 1166A and Oligocene-Quaternary
diamictites from Holocene Jumbo Piston Core 34 (JPC34) were analyzed using the same tripledating method (U-Pb, fission track, and (U-Th)/He dating) described in this study. However,
core 739C provides a more complete syn-glacial stratigraphic section of Oligocene through
Quaternary sediments, as cores 1166A and JPC34 do not include any Miocene sediments or as
extensive a section of early Oligocene sediments.
4. Methods
4.1. Fission Track Analysis
All samples were processed using standard mineral separation techniques for zircon and
apatite (Gehrels, 2000). The zircon yield for all samples was very low, and many of the zircon
grains were metamict, making them unsuitable for fission track and U-Pb analysis. Zircon and
apatite grains from each of the twenty samples were mounted in teflon and epoxy, respectively,
polished, and etched using the methods outlined in Hurford et al. (1991). The samples were
analyzed by the external detector method (Tagami and O’Sullivan, 2005) and irradiated at the
!
8!
Oregon State University Triga Reactor in Corvallis, OR, along with European Institute for
Reference Materials and Measurements uranium-dosed glasses IRMM 540R for apatite and
IRMM 541 for zircon to monitor neutron fluence. Spontaneous and induced tracks were counted
using an Olympus BX51 microscope at 1250x magnification at the University of Arizona Fission
Track Laboratory. Central ages were calculated using the zeta calibration approach of Hurford
and Green (1983). The zeta values used for zircon and apatite were 351.99 ± 8.94 and 113.21 ±
1.81, respectively. Analytical data for 1019 analyzed apatite grains and 218 analyzed zircon
grains can be found in Appendix A, Tables A1 and A2.
4.2. U-Pb Analysis
Fission track mounts for zircon and apatite were attached to glass slides adjacent to
standards mounted separately in epoxy or teflon for U-Pb analysis. Primary standards included a
Sri Lanka zircon (Gehrels et al., 2008), and Madagascar and McClure Mountain apatites
(Thomson et al., 2012). All grains dated by fission track analysis were located and analyzed at
the Arizona LaserChron Center by laser ablation-multicollector-inductively coupled plasma mass
spectrometry (LA-MC-ICPMS). Detailed procedures for zircon analysis are described by
Gehrels et al. (2008); procedures for apatite analysis are described by Thomson et al. (2012).
Three analyses were conducted in each apatite grain, depending on the size of the grain, to
account for compositional variation. Concordia ages were later calculated to determine one age
for each grain. In cases where very few zircons could be dated by fission track analysis, a small
number of additional suitable zircon grains were analyzed to obtain a greater number of U-Pb
analyses for those samples. The measured isotopic ratios for zircons are corrected for common
Pb using the measured 204Pb, assuming an initial Pb composition according to Stacey and
!
9!
Kramers (1975). Measured isotopic ratios for apatites are also corrected for common Pb
according to Stacey and Kramers (1975). However, due to the high common to radiogenic Pb
ratio in most apatite, a five-step iterative process is also applied to correct for this (Thomson et
al, 2012). Full analytical data for 982 apatite grains (multiple analyses per grain) and 262 zircon
grains yielding data of acceptable precision (apatite: <10% 206Pb/238U age and <10% 206Pb/207Pb
age; zircon: <15% 206Pb/238U age and <10% 206Pb/207Pb age) and discordance (<30% discordant)
are reported in Appendix A, Tables A3 and A4.
4.3. (U-Th)/He Analysis
Individual apatite grains from five representative samples from different stratigraphic age
divisions were analyzed by the (U-Th)/He method. Grains from one late Pliocene sample
(004R21), one early Pliocene sample (017R20), two late Miocene samples (022R74 and
023R110), and one early Oligocene sample (047R55) were analyzed, with a focus on the
Miocene samples due to their rarity in all other cores from this region. Double-dated grains were
handpicked from the fission track mounts and measured lengthwise and widthwise for alpha
ejection correction parameters. Each grain was then loaded into a Nb tube and analyzed using the
methodology described in Reiners et al. (2004) at the Arizona Radiogenic Helium Dating
Laboratory. Because the grains had previously been polished for fission track and U-Pb analysis,
a polished grain alpha ejection correction was applied to the raw age for each grain, assuming
that 30 µm of the grain had been removed (Farley, 2002; Reiners et al., 2007). However, because
all analyzed grains were anhedral and in many cases broken, it is unclear when the rims of the
grains were removed and whether the alpha ejection correction should actually be applied. For
this reason, a plot age was calculated for each grain by taking the average of the raw age and the
!
10!
fully corrected age. The uncertainty assumed for each plot age is the difference between the plot
age and the minimum (raw age minus analytical uncertainty) and maximum (fully corrected age
plus analytical uncertainty) ages for that grain. All ages plotted in the apatite (U-Th)/He
histograms are raw ages. Probability density function (PDF) curves (Ludwig, 2008) are shown
for raw ages as well as plot ages. Analytical data for 103 apatite grains can be found in Appendix
A, Table A5.
5. Results
5.1. Zircon U-Pb
All samples display a dominant age peak between ~450-550 Ma, and a smaller peak from
~800-1100 Ma (Fig. 4). Miocene through Quaternary samples contain a handful of grains
ranging from 1.5-3.2 Ga. The Oligocene samples contain far fewer zircons than the younger
samples overall, but have a much higher abundance of older Proterozoic and Archean grains.
5.2. Zircon Fission Track
Like the zircon U-Pb ages, zircon fission track ages are very consistent throughout the
stratigraphic section (Fig. 4). Each set of samples grouped by stratigraphic age contains two main
age populations – an older population at ~250-300 Ma, and a small younger population at ~120
Ma.
5.3. Apatite U-Pb
Apatite U-Pb ages are also consistent throughout the entire stratigraphic section (Fig. 5).
Like with the zircon U-Pb data, all samples display a dominant peak at ~450-550 Ma. Less than
!
11!
10% of grains in Miocene through Quaternary samples are older than 600 Ma. However, ~35%
of grains in the early Oligocene samples have ages ranging from 600-1300 Ma.
5.4. Apatite Fission Track and (U-Th)/He
Apatite fission track cooling ages display a clearly decreasing trend moving from the
early Oligocene through the Quaternary (Fig. 6). There are no distinct age populations within the
age spectra of any of the samples, but rather a central cooling age population whose probability
decreases fairly symetrically to both older and younger ages. Throughout the 5 myr period of the
early Oligocene represented in core 739C, there is a clear decrease in the mean cooling age of
each sample. The oldest Oligocene sample (062R0) has a mean age of ~280 Ma, and the
youngest Oligocene sample (025R10) has a mean age of ~210 Ma. Above the unconformity in
the 10 myr period from the late Miocene to the Quaternary, there is only a slight shift in mean
fission track age from ~180 Ma to ~150 Ma.
Apatite (U-Th)/He ages show a similar trend of decreasing low-temperature cooling ages
upsection (Fig. 7). Although the sample sizes are limited for all stratigraphic ages with the
exception of the late Miocene, early Oligocene minimum ages are clearly older than those in the
late Miocene. There are no grains with cooling ages younger than ~100 Ma in the early
Oligocene samples, and a large number of grains with cooling ages as young as ~20 Ma in the
Miocene-Quaternary samples.
6. Discussion
6.1. Zircon U-Pb
!
12!
The dominant peak at 500 Ma in samples of every stratigraphic age records the PanAfrican orogenic event that produced widespread magmatism and metamorphism throughout
East Antarctica, as noted in previous U-Pb geo/thermochronologic studies. Similarly, the peak
centered around 900-1000 Ma records an orogenic event of Grenville age, although this event
does not appear to have produced the same scale of magmatism as the Pan-Afican event, at least
as recorded in detrital zircon and apatite grains (Boger et al., 2001; Carson et al., 2000, 1996;
Fitzsimons, 2000b; Fitzsimons et al., 1997; Hemming et al., 2007; Liu et al., 2007; Mikhalsky et
al., 2006; van de Flierdt et al., 2008; Veevers and Saeed, 2007; Veevers et al., 2008; Williams et
al., 2007). All ages seen in the zircon age spectra are within the range of ages found in the
exposed bedrock outcrops in the vicinity of Prydz Bay and the Lambert Graben (Fig. 2) and most
likely also found in and around the Gamburtsev Mountains further into the interior of the shield.
The greater proportion of older grains in the Oligocene sediments as well as the sudden decrease
in zircons found in these sediments is indicative of a possible shift in source area between the
early Oligocene to the late Miocene. Additionally, the lack of zircons younger than ~400 Ma in
any of the samples suggests that the Gamburtsev Mountains, where a large portion of these
sediments are most likely sourced from based on known subglacial drainage patterns into Prydz
Bay, supports recent studies concluding that they are most likely an old feature (Cox et al., 2010;
Van de Flierdt et al., 2008) and not a young volcanic mountain range, as has been previously
hypothesized (Sleep, 2006).
6.2. Zircon Fission Track
All zircon fission track ages are from zircon grains with U-Pb ages that are Pan-African
and older, therefore representing later cooling and/or resetting events. The ~250-300 Ma
!
13!
population of fission track cooling ages most likely reflects cooling associated with PermoTriassic rifting in East Antarctica related to initial formation of the Lambert Graben (Arne,
1994). Normal faulting associated with this rifting would have exhumed the basement rock
underlying the Lambert Graben, with subsequent erosion of uplifted material producing the
Amery Group sediments of that age that now overlie the basement rock within the graben. If
exhumation by normal faulting is the mechanism responsible for the widespread ~250-300 Ma
zircon fission track ages, then this rifting would require tectonic exhumation to depths of ~8-12
km to reach basement temperatures of 200-300 °C, assuming a geothermal gradient of 25 °C/km.
If the geothermal gradient during rifting was elevated to ~30 °C/km in the area directly
surrounding the rift (Lisker et al., 2003), the basement rocks would only need to have been
exhumed through normal faulting to depths of ~6-10 km to cool through the range of fission
track closure temperatures. Some of this cooling is also likely related to erosion on uplifted rift
flanks.
A number of authors have proposed a second period of rifting in the Lambert Graben area
during the Cretaceous associated with the final breakup of Gondwana (Arne, 1994; Ferraccioli et
al., 2011; Lisker et al., 2003, 2007a, 2007b) that Lisker et al., (2003) use to explain their apatite
fission track data. However, the zircon fission track cooling ages of ~120 Ma form a single welldefined peak more consistent with local resetting related to alkaline mafic magmatism of this age
in the Beaver Lake area of the Northern Prince Charles Mountains (Coffin et al, 2002). This
magmatic activity is associated with widespread magmatism of this age related to initial surface
manifestation of more widespread Kerguelen hotspot magmatism (Coffin et al., 2002). A
regional cooling related to exhumation would be expected to yield partially reset zircon fission
track ages between 250-300 and 120 Ma, which we do not observe.
!
14!
6.3. Apatite U-Pb
The ~500 Ma apatite U-Pb age peak is dominant in all samples. However, the early
Oligocene samples have a greater proportion of Precambrian grains. In nine out of ten late
Miocene through Quaternary samples, ≥85% of analyzed grains are between 440-560 Ma. In
sample 023R110 (late Miocene), 75% of grains were within this range. In all ten early Oligocene
samples, however, only ≥65% of grains are within this age range. A Kolmogorov-Smirnov test
(Press et al., 1986) comparing the 440-560 Ma age population in each sample with the same
population in every other sample produced P-values greater than 0.05 in 178/190 possible sample
pairings. This indicates that there is a 95% level of confidence that this age population in each
sample is not statistically different from the same population in any other sample.
The peak at ~500 Ma likely includes both apatite grains that were reset during
widespread metamorphism associated with the Pan-African orogeny as well as grains that
crystallized during magmatic events at that time. However, whether the majority of grains of this
age are recording crystallization ages or ages of metamorphism cannot be elucidated from the
detrital grains alone. The presence of a small number of zircon and apatite U-Pb ages older than
~500 Ma indicates that there are older rocks in the source area, suggesting that some of the
apatite grains with ~500 Ma U-Pb ages represent Pan-African metamorphic resetting in line with
overprinting indicated by widespread monazite U-Th-Pb data of this age in the southern Prince
Charles Mountains (Phillips et al., 2009). However, the majority of the zircon U-Pb
crystallization ages in these samples are ~500 Ma, recording a magmatic event during which
many of the ~500 Ma apatite grains may have crystallized as well.
!
15!
The ~500 Ma age of metamorphism and crystallization is consistent with that seen in
40
Ar/39Ar hornblende ages from other sediments recovered from Prydz Bay (Hemming et al.,
2007; Phillips et al., 2007; Roy et al., 2007). The older grains seen mainly in the Oligocene
samples are most likely grains sourced from areas of East Antarctica that did not experience a
high enough degree of metamorphism at this time to be reheated to 450-550 °C, the range of
temperature at which the apatite U-Pb system is reset, depending on factors such as cooling rate
and grain radius (Cherniak et al., 1991; Thomson et al., 2012). Evidence that such areas of East
Antarctica exist can be seen in 40Ar/39Ar dates of hornblende from Wilkes Land, the only region
of East Antarctica dominated by 40Ar/39Ar dates older than Pan-African age (see Section 2) (Roy
et al., 2007).
6.4. Apatite Fission Track and (U-Th)/He
If erosion rates increase through time, the cooling ages measured in the detrital record of
this exhumation will become younger through time. We interpret the upsection-younging trends
in apatite fission track and (U-Th)/He cooling ages to reflect increasing erosion rates and
possibly localized glacial incision over time. The relationship between changes in cooling age
and depositional age can be addressed through the determination of lag time, or the difference
between the time of exhumation of a grain through its closure temperature for a given
thermochronometric system and its time of deposition (Brandon and Vance, 1992; Cerveny et al.,
1988). The rapid decrease in mean apatite fission track cooling ages during the early Oligocene
corresponds with a substantial decrease in lag time from ~240 My to ~180 My over a period of
only ~5 My (Figs. 7 and 8). This implies slow but increasing erosion rates in the source areas of
these sediments during this time. In contrast, fission track ages for the late Miocene-Quaternary
!
16!
samples show only a slight decrease from ~170 My to ~150 My over ~10 My, implying slow,
near steady-state erosion throughout this time. Likewise, apatite (U-Th)/He cooling ages are all
greater than 100 Ma in the early Oligocene sample. Beginning in the late Miocene, cooling ages
younger than 100 Ma become abundant and remain relatively constant throughout the Pliocene
and Quaternary, supporting the slow, steady erosion implied by fission track lag times.
Assuming that sediments were derived from the same general source areas during the
early Oligocene and the Miocene-Quaternary and that the depth-age structure in this region is
like that of a very slowly eroding region, the younger fission track and (U-Th)/He cooling ages
and near-constant lag times in the late Miocene-Quaternary samples indicate that the majority of
the erosion producing these younger cooling ages had already been achieved by the late
Miocene, with little additional exhumation taking place since that time. This, as well as the
change in erosion rates between the early Oligocene and late Miocene determined from fission
track cooling ages, is consistent with climatic evidence seen in the marine isotope record as well
as evidence from ice sheet erosional modeling (Bo et al., 2009) pointing to a transition from an
initial phase of unstable, warm-based glaciation from 34-14 Ma to a phase of predominantly
cold-based, less erosive glaciation from 14 Ma to the present (see Section 1). This suggests that
the majority of erosion associated with the EAIS in the Prydz Bay/Lambert Graben region most
likely occurred during the early Oligocene in the first 5 My of glaciation (based on the
depositional age of the early Oligocene sediments), with slower, limited glacial erosion since that
time.
Based on ice sheet erosional models and the characteristic U-shape of glacial valleys and
troughs observed in East Antarctica (Bo et al., 2009; Jamieson et al., 2010), the majority of the
erosion taking place during the early Oligocene in and around the Lambert Graben and
!
17!
Gamburtsev Mountains was likely in the form of glacial incision rather than widespread erosion
over the entire region. Glacial incision patterns generally follow the preglacial fluvial drainage
patterns along the margins of East Antarctica, leading to overdeepening of these preexisting river
valleys (Jamieson and Sugden, 2008; Jamieson et al., 2010).
6.5. Modeling Depths of Incision
The following discussion of modeled erosional depths assumes that erosion along the
eastern margin of Antarctica was accomplished primarily by glacial incision into preexisting
river valleys, as described previously in Section 6.4. To determine the depth of incision that must
have been achieved by the late Miocene based on our measured apatite fission track and (UTh)/He ages, we have modeled the predicted cooling ages of grains derived from various depths
beneath the surface of an area since the onset of glaciation at ~34 Ma using a time-temperature
history likely to be a good estimate for the Prydz Bay/Lambert Graben region. To determine the
predicted apatite fission track and (U-Th)/He cooling ages, we use a simple time-temperature
model produced in the HeFTy program of Ketcham (2005). The model of Ketchem et al. (2007a,
2007b) is used to predict fission track cooling ages, and the Durango model of Farley (2000) is
used to predict (U-Th)/He cooling ages. The parameters for this model include a period of rapid
cooling at 280 Ma followed by slow cooling until 34 Ma, with accelerated cooling to surface
temperatures between 34 Ma and the present day. During the period of slow cooling from 280-34
Ma, fission track and (U-Th)/He ages would have decreased steadily with depth, defining a
thermochronologic stratigraphy. Assuming a normal geothermal gradient of 25 °C/km, this
stratigraphy and the cooling ages predicted by the HeFTY model can be used to correlate our
measured cooling ages with depths of incision.
!
18!
The predicted cooling age curves generated by this model are shown in Figure 8, along
with histograms of measured fission track and uncorrected (U-Th)/He ages from two combined
late Miocene samples (022R74 and 023R110). By comparing the youngest measured cooling
ages in each histogram with the cooling ages predicted by the HeFTy model and their
corresponding incision depths, we can estimate the depth of incision that must have been
achieved by the late Miocene to produce the cooling ages measured in the Miocene samples. The
youngest measured fission track ages of ~80 Ma correspond to a predicted incision depth of ~2.8
km since the late Miocene. Measured (U-Th)/He ages indicate a slightly deeper predicted depth
of incision, with the youngest ages of ~25 Ma corresponding to ~3 km of incision by the late
Miocene.
The presence of young ages of apatite grains in late Miocene sediments that must have
come from ~2.8-3.0 km of depth supports other evidence indicating that the majority of glacial
incision occurred during the early Oligocene in the first 5 My after the onset of glaciation in East
Antarctica. During this time, the EAIS was primarily warm-based, making it erosive and capable
of incising valleys of these depths. After its transition to cold-based glaciation at ~14 Ma,
however, the EAIS would have become far less erosive. Therefore, the depths of incision
predicted by the HeFTY model must have been achieved before this transition to produce the
corresponding cooling ages in the late Miocene apatite grains.
7. Conclusions
Multi-dating of single apatite and zircon grains from diamictites in Prydz Bay allows for
the identification and interpretation of major tectonic and exhumational events throughout the
history of East Antarctica. Zircon U-Pb crystallization ages and apatite U-Pb crystallization and
!
19!
metamorphic ages record the widespread Pan-African orogeny at 0.5-0.6 Ga, which resulted in
deformation and intrusive activity throughout much of East Antarctica. Zircon fission track ages
record a rifting and exhumational event in the Permian-Triassic, as well as a magmatic heating
event during the Cretaceous (Coffin et al., 2002; Ferraccioli et al., 2011; Lisker et al., 2003).
From low temperature cooling ages measured on apatite grains, erosional history can be
constrained and estimated depths of glacial incision can be determined. Both the apatite fission
track and (U-Th)/He cooling ages show a rapidly decreasing trend throughout the early
Oligocene, and a more slowly decreasing trend from the late Miocene through the Quaternary.
Lag times determined from mean apatite fission track cooling ages decrease from ~250-180 Ma
during the early Oligocene, but remain relatively constant at ~160 Ma from the late Miocene to
Quaternary. This is indicative of increasing erosion rates throughout the early Oligocene,
followed by near steady-state erosion from the Miocene to Quaternary. Although a greater
number of apatite (U-Th)/He ages from Quaternary, Pliocene, and especially early Oligocene
samples would be necessary to make this dataset more robust, these trends are indicative of
accelerated glacial incision during the first 5 m.y. after the onset of the EAIS, followed by more
steady slower incision from the late Miocene to the present. HeFTy time-temperature modeling
compared with measured apatite fission track and (U-Th)/He cooling ages corresponds to ~2.83.0 km of incision that must have occurred by the late Miocene. Most of this incision likely
occurred during the early Oligocene before the transition from a warm-based, erosive EAIS to
the cold-based, stable ice sheet that exists today.
!
20!
Acknowledgements
This work was supported by NSF grants ANT-0838722 and EAR-1032156 (Arizona
LaserChron Center). Thanks to Mark Pecha, Nicky Giesler, Chelsi White, Intan Yokelson, and
Percival Gou in the Arizona LaserChron Center and Loren Hill and Uttam Chowdhuri in the
Arizona Radiogenic Helium Dating Laboratory for their assistance with sample preparation and
analysis.
References Cited
Arne, D.C., 1994. Phanerozoic exhumation history of northern Prince Charles Mountains (East
Antarctica). Antarctic Science 6, 69-84.
Bamber, J.L., Vaughan, D.G., Joughin, I., 2000. Widespread complex flow in the interior of the
Antarctic Ice Sheet. Science 287, 1248-1250.
Barrett, P.J., 1996. Antarctic paleoenvironment through Cenozoic times – A review. Terra
Antarctica 3, 103-119.
Barron, J., Larsen, B., Baldauf, J.G., 1991. Evidence for late Eocene to early Oligocene Antarctic
glaciation and observations on late Neogene glacial history of Antarctica: Results from
Leg 119, in: Barron, J., Larsen, B., Shipboard Scientific Party (Eds.), Proceedings of the
Ocean Drilling Project, Scientific results 119. Ocean Drilling Program, College Station,
Texas, pp. 869-891.
Bo, S., Siegert, M.J., Mudd, S.M., Sugden, D., Fujita, S., Xiangbin, C., Yunyun, J., Xueyuan, T.,
Yuansheng, L., 2009. The Gamburtsev Mountains and the origin and early evolution of
the Antarctic Ice Sheet. Nature 459, 690-693.
Boger, S.D., Carson, C.J., Fanning, C.M., Hergt, J.M., Wilson, C.J.L., Woodhead, J.D., 2002.
Pan-African intraplate deformation in the northern Prince Charles Mountains, East
Antarctica. Earth and Planetary Science Letters 195, 195-210.
Boger, S.D., Wilson, C.J.L., Fanning, C.M., 2001. Early Paleozoic tectonism within the East
Antarctic craton: The final suture between east and west Gondwana? Geology 29, 463
466.
Brandon, M.T., Vance, J.A., 1992. Tectonic evolution of the Cenozoic Olympic subduction
complex, Washington State, as deduced from fission track ages for detrital zircons.
American Journal of Science 292, 565-636.
!
21!
Carson, C.J., Fanning, C.M., Wilson, C.J.L., 1996. Timing of the Progress Granite, Larsemann
Hills: Further evidence for early Palaeozoic orogenesis within the East Antarctic Shield
and implications for Gondwana assembly. Australian Journal of Earth Sciences 43, 539
553.
Carson, C.J., Boger, S.D., Fanning, C.M., Wilson, C.J.L., Thost, D.E., 2000. SHRIMP U-Pb
geochronology from Mount Kirkby, northern Prince Charles Mountains, East Antarctica.
Antarctic Science 12, 429-442.
Cerveny, P.F., Naeser, N.D., Zeitler, P.K., Naeser, C.W., Johnson, N.M., 1988. History of uplift
and relief of the Himalaya during the past 18 million years: evidence from FT ages of
detrital zircons from sandstones of the Siwalik Group, in: Kleinspehn, K., Paola, C.
(Eds.), New Perspectives in Basin Analysis. Springer, New York, pp. 43-61.
Cherniak, D.J., Lanford, W.A., Ryerson, F.J., 1991. Lead diffusion in apatite and zircon using
ion implantation and Rutherford Backscattering techniques. Geochimica et
Cosmochimica Acta 55, 1663–1673.
Coffin, M.F., Pringle, M.S., Duncan, R.A., Gladczenko, T.P., Storey, M., Muller, R.D., Gahagan,
L.A., 2002. Journal of Petrology 43, 1121-1137.
Collerson, K.D., Sheraton, J.W., 1986. Bedrock geology and crustal evolution of the Vestfold
Hills, in: Pickard, J. (Ed.), Antarctic oasis, terrestrial environments and history of the
Vestfold Hills. Academic Press, Sydney, pp. 21-62.
Cox, S.E., Hemming, S.R., Reiners, P.W., Nicolescu, S., Thomson, S.N., Gehrels, G.E., van de
Flierdt, T., Goldstein, S.L., Brachfeld, S.A., 2007. Detrital apatite and zircon (U-Th)/He
evidence for early formation and slow erosion of the Gamburtsev Mountains, East
Antarctica. U.S. Geological Survey and The National Academies 10th International
Symposium on Antarctic Earth Sciences, extended abstract 193.
Cox, S.E., Thomson, S.N., Reiners, P.W., Hemming, S.R., van de Flierdt, T., 2010. Extremely
low long-term erosion rates around the Gamburtsev Mountains in interior East
Antarctica. Geophysical Research Letters 37, L22307, doi:10.1029/2010GL045106.
Dalziel, I.W.D., 1992. Antarctica: A tale of two supercontinents? Annual Reviews of Earth and
Planetary Science 20, 501-526.
DeConto, R.M., Pollard, D., 2003a. Rapid Cenozoic glaciation of Antarctica induced by
declining atmospheric CO2. Nature 421, 245-249.
DeConto, R.M., Pollard, D., 2003b. A coupled climate-ice sheet modeling approach to the Early
Cenozoic history of the Antarctic ice sheet. Palaeogeography, Palaeoclimatology,
Palaeoecology 198, 39-52.
!
22!
Farley, K.A., 2000. Helium diffusion from apatite: General behavior as illustrated by Durango
fluorapatite. Journal of Geophysical Research 105, 2903-2914.
Farley, K.A., 2002. (U-Th)/He dating: Techniques, calibrations, and applications: Reviews in
mineralogy and Geochemistry 47, 819-844.
Fitzsimons, I.C.W., 1997. The Brattstrand Paragneiss and the Seistrene Orthogneiss: A review of
Pan-African metamorphism and Grenville relics in southern Prydz Bay, in: Ricci, C.A.
(Ed.), The Antarctic region: Geological evolution and processes. Terra Antarctica
Publications, Siena, Italy, pp. 121-130.
Fitzsimons, I.C.W., 2000a. A review of tectonic events in the East Antarctic Shield and their
implications for Gondwana and earlier supercontinents. Journal of African Earth Sciences
31, 3-23.
Fitzsimons, I.C.W., 2000b. Grenville-age basement provinces in East Antarctica: Evidence for
three separate collisional orogens. Geology 28, 879-882.
Fitzsimons, I.C.W., 2003. Proterozoic basement provinces of southern and southwestern
Australia, and their correlation with Antarctica, in: Yoshida, M., Windley, B.F.,
Dasgupta, S. (Eds.), Proterozoic East Gondwana: Supercontinent Assembly and Breakup.
Geological Society, London, Special Publications 206, pp. 93-130
Fitzsimons, I.C.W., Kinny, P.D., Harley, S.L., 1997. Two stages of zircon and monazite growth
in anatectic leucogneiss: SHRIMP constraints on the duration and intensity of PanAfrican metamorphism in Prydz Bay, East Antarctica. Terra Nova 9, 47-51.
Gehrels, G.E., 2000. Introduction to detrital zircon studies of Paleozoic and Triassic strata in
western Nevada and northern California, in: Soreghan, M.J., Gehrels, G.E. (Eds.),
Paleozoic and Triassic Paleogeography and Tectonics of Western Nevada and Northern
California. Geological Society of America Special Paper 347, pp. 1-17.
Gehrels, G.E., Valencia, V.A., Ruiz, J., 2008. Enhanced precision, accuracy, efficiency, and
spatial resolution of U-Pb ages by laser ablation-multicollector-inductively coupled
plasma mass spectrometry. Geochemistry Geophysics Geosystems 9, Q03017, 13 p.
Goodge, J.W., 2007. Proxies of the East Antarctic Shield: Composition and Age of Ice-Covered
Basement from Sedimentary and Glacial Provenance, in: Cooper, A.K., Raymond, C.R.
(Eds.), Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th
ISAES, USGS Open-File Report 2007-1047, Extended Abstract 133, 5 p.
Hambrey, M.J., McKelvey, B., 2000. Major Neogene fluctuations of the East Antarctic ice sheet:
Stratigraphic evidence from the Lambert Glacier region. Geology 28, 887-890.
Hemming, S.R., Brachfeld, S.A., Gehrels, G.E., Goldstein, S.L., van de Flierdt, T., 2007.
Evidence from detrital hornblende 40Ar/39Ar and zircon U-Pb ages for the sources of
!
23!
glacial deposits in the Prydz Bay region. U.S. Geological Survey and The National
Academies 10th International Symposium on Antarctic Earth Sciences, extended abstract
089.
Hurford, A.J., Green, P.F., 1983. The zeta age calibration of fission-track dating. Isotope
Geoscience 1, 285–317.
Hurford, A.J., Hunziker, J.C., Stockhert, B., 1991. Constraints on the late thermotectonic
evolution of the western Alps: Evidence for episodic rapid uplift. Tectonics 10, 758-769.
Jamieson, S.S.R., Hulton, N.R.J., Sugden, D.E., Payne, A.J., Taylor, J., 2005. Cenozoic
landscape evolution of the Lambert basin, East Antarctica: the relative role of rivers and
ice sheets. Global and Planetary Change 45, 35-49.
Jamieson, S.S.R., Sugden, D.E., 2008. Landscape Evolution of Antarctica, in: Cooper, A.K.,
Barrett, P.J., Stagg, H., Storey, B., Stump, E., Wise, W., 10th ISAES editorial team (Eds.),
Antarctica: A Keystone in a Changing World. Proceedings of the 10th International
Symposium on Antarctic Earth Sciences. The National Academies Press, Washington,
DC, pp. 39-54.
Jamieson, S.S.R., Sugden, D.E., Hulton, N.R.J., 2010. The evolution of the subglacial landscape
of Antarctica. Earth and Planetary Science Letters 293, 1-27.
Ketcham, R.A., 2005. Forward and inverse modeling of low-temperature thermochronometry
data. Reviews in Mineralogy and Geochemistry 58, 275-314.
Ketcham, R.A., Carter, A.C., Donelick, R.A., Barbarand, J., Hurford, A.J., 2007a. Improved
measurement of fission-track annealing in apatite using c-axis projection. American
Mineralogist 92, 789-798.
Ketcham, R.A., Carter, A.C., Donelick, R.A., Barbarand, J., Hurford, A.J., 2007b. Improved
modeling of fission-track annealing in apatite. American Mineralogist 92, 799-810.
Kurinin, R.G., Grikurov, G.E., 1982. Crustal structure of part of East Antarctica from
geophysical data, in: Craddock, C. (Ed.), Antarctic Geoscience. University of Wisconsin
Press, Madison, pp. 895-901.
Lisker, F., 2002. Review of fission track studies in northern Victoria Land – passive margin
evolution versus uplift of the Transantarctic Mountains. Tectonophysics 349, 57-73.
Lisker, F., Brown, R., Fabel, D., 2003. Denudational and thermal history along a transect across
the Lambert Graben, northern Prince Charles Mountains, Antarctica derived from apatite
fission track thermochronology. Tectonics 22, doi:10.1029/2002TC001477.
Lisker, F., Gibson, H., Wilson, C.J., Laufer, A., 2007a. Denudation and uplift of the Mawson
Escarpment (eastern Lambert Graben, Antarctica) as indicated by apatite fission track
!
24!
data and geomorphological observation, in: Cooper, A.K., Raymond. C.R. (Eds.),
Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES,
USGS Open-File Report 2007-1047, Short Research Paper 105, 6 p.
Lisker, F., Wilson, C.J.L., Gibson, H.J., 2007b. Thermal history of the Vestfold Hills (East
Antarctica) between Lambert rifting and Gondwana break-up: Evidence from apatite
fission track data. Antarctic Science 19, 97-106.
Liu, X., Zhao, Y., Zhao, G., Jian, P., Xu, G., 2007. Petrology and geochronology of granulites
from the McKaskle Hills, Eastern Amery Ice Shelf, Antarctica, and implications for the
evolution of the Prydz Belt. Journal of Petrology 48, 1443-1470.
Ludwig, K.R., 2008. Isoplot 3.60. Berkeley, California, Berkeley Geochronology Center Special
Publication 4, 77 p.
McKelvey, B.C., Hambrey, M.J., Harwood, D.M., Mabin, M.C.G., Webb, P.-N., Whitehead,
J.M., 2001. The Pagodroma Group – a Cenozoic record of the East Antarctic ice sheet in
the northern Prince Charles Mountains. Antarctic Science 13, 455-468.
Mikhalsky, E.V., Beliatsky, B.V., Sheraton, J.W., Roland, N.W., 2006. Two distinct
Precambrian terranes in the Southern Prince Charles Mountains, East Antarctica:
SHRIMP dating and geochemical constraints. Gondwana Research 9, 291-309.
Naish, T.R., Woolfe, K.J., Barrett, P.J., Wilson, G.S., Atkins, C., Bohaty, S.M., Bucker, C.J.,
Claps, M., Davey, F.J., Dunbar, G.B., Dunn, A.G., Fielding, C.R., Florindo, F., Hannah,
M.J., Harwood, D.M., Henrys, S.A., Krissek, L.A., Lavelle, M., van der Meer, J.,
McIntosh, W.C., Niessen, F., Passchier, S., Powell, R.D., Roberts, A.P., Sagnotti, L.,
Scherer, R.P., Strong, C.P., Talarico, F., Verosub, K.L., Villa, G., Watkins, D.K., Webb,
P.-N., Wonik, T., 2001. Orbitally induced oscillations in the East Antarctic ice sheet at
the Oligocene/Miocene boundary. Nature 413, 719-723.
O’Brien, P.E., Goodwin, I., Forsberg, C.-F., Cooper, A.K., Whitehead, J., 2007. Late Neogene
ice drainage changes in Prydz Bay, East Antarctica and the interaction of Antarctic ice
sheet evolution and climate. Palaeogeography, Palaeoclimatology, Palaeoecology 245,
390-410.
Phillips, G., Kelsey, D.E., Corvino, A.F., Dutch, R.A., 2009. Continental reworking during
overprinting orogenic events, Southern Prince Charles Mountains, East Antarctica.
Journal of Petrology 50, 2017-2041.
Phillips, G., Wilson, C.J.L., Phillips, D., Szczepanski, S.K., 2007. Thermochronological
(40Ar/39Ar) evidence of Early Paleozoic basin inversion within the southern Prince
Charles Mountains, East Antarctica: implications for East Gondwana. Journal of the
Geological Society, London 164, 771-784.
Press, W.H., Flannery, B.P., Tenkolsky, S.A., Vetterling, W.T., 1986. Numerical recipes.
!
25!
Cambridge University Press, Cambridge, 818 p.
Reiners, P.W., 2007. Thermochronologic approaches to paleotopography. Reviews in
Mineralogy and Geochemistry 66, 243-267.
Reiners, P.W., Spell, T.L., Nicolescu, S., Zanetti, K.A., 2004. Zircon (U-Th)/He
thermochronometry: He diffusion and comparisons with 40Ar/39Ar dating. Geochimica et
Cosmochimica Acta 68, 1857-1887.
Reiners, P.W., Thomson, S.N., McPhillips, D., Donelick, R.A., Roering, J.J., 2007. Wildfire
thermochronology and the fate and transport of apatite in hillslope and fluvial
environments. Journal of Geophysical Research 112, F04001,
doi:10.1029/2007JF000759.
Roy, M., Van de Flierdt, T., Hemming, S.R., Goldstein, S.L., 2007. 40Ar/39Ar ages of hornblende
grains and bulk Sm/Nd isotopes of Circum-Antarctic glacio-marine sediments:
Implications for sediment provenance in the Southern Ocean. Chemical Geology 244,
507-519.
Siegert, M.J., 2008. Antarctic subglacial topography and ice-sheet evolution. Earth Surface
Processes and Landforms 33, 646-660.
Siegert, M.J., Barrett, P., Deconto, R., Dunbar, R., Cofaigh, C.O., Passchier, S., Naish, T., 2008.
Recent advances in understanding Antarctic climate evolution. Antarctic Science 20, 313
325.
Shipboard Scientific Party, 1989b. Site 739, in: Barron, J., Larsen, B. (Eds.), Proc. ODP, Init.
Repts. (Ocean Drilling Program) 119, 289-344.
Sleep, N. H., 2006. Mantle plumes from top to bottom. Earth Science Reviews 77, 231–271.
Solheim, A., Forsberg, C.F., Pittenger, A., 1991. Stepwise consolidation of glacigenic sediments
related to the glacial history of Prydz Bay, East Antarctica, in: Barron, J., Larsen, B.
(Eds.), Proceedings of the Ocean Drilling Program, Scientific Results 119, pp. 169-182.
Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a two
stage model. Earth and Planetary Science Letters 26, 207-221.
Stock, J.D., Montgomery, D.R., 1996. Estimating paleorelief from detrital mineral age ranges.
Basin Research 8, 317-327.
Sugden, D.E., 1996. The East Antarctic Ice Sheet: Unstable ice or unstable ideas? Transactions
of the Institute of British Geographers 21, 443-454.
Sugden, D.E., Marchant, D.R., Denton, G.H., 1993. The case for a stable East Antarctic Ice
Sheet: The background. Geografiska Annaler 75, 151-154.
!
26!
Tagami, T., O’Sullivan, P.B., 2005. Fundamentals of fission-track thermochronology.
Reviews in Mineralogy and Geochemistry 58, 19-47.
Taylor, J., Siegert, M.J., Payne, A.J., Hambrey, M.J., O’Brien, P.E., Cooper, A.K., Leitchenkov,
G., 2004. Topographic controls on post-Oligocene changes in ice-sheet dynamics, Prydz
Bay region, East Antarctica. Geology 32, 197-200.
Thomson, S.N., Gehrels, G.E., Ruiz, J., Buchwaldt, R., 2012. Routine low-damage U-Pb dating
of apatite using laser ablation-multicollector-ICPMS. Geochemistry, Geophysics,
Geosystems 13, Q0AA21, 23 pp
Thomson, S.N., Reiners, P.W., Tochilin, C.J., Hemming, S.R., Gehrels, G.E., 2011. The
Cenozoic history of East Antarctica subglacial erosion and sediment flux from the
offshore detrital Thermochronometric record. Eos Transactions AGU, Abstract PP33B
1934.
van de Flierdt, T., Hemming, S.R., Goldstein, S.L., Gehrels, G.E., Cox, S.E., 2008. Evidence
against a young volcanic origin of the Gamburtsev Subglacial Mountains, Antarctica.
Geophysical Research Letters 35, L21303, doi:10.1029/2008GL035564.
Volpi, V., Rebesco, M., Diviacco, P., 2009. New insights in the evolution of Antarctic glaciation
from depth conversion of well-log calibrated seismic section of Prydz Bay. International
Journal of Earth Sciences 98, 1991-2007.
Veevers, J.J., 2003. Pan-African is Pan-Gondwanaland: Oblique convergence drives rotation
during 650-500 Ma assembly. Geology 31, 501-504.
Veevers, J.J., Saeed, A., 2007. Gamburtsev Subglacial Mountains provenance of Permian
Triassic sandstones in the Prince Charles Mountains and offshore Prydz Bay: Integrated
U-Pb and TDM ages and host-rock affinity from detrital zircons. Gondwana Research 14,
316-342.
Veevers, J.J., Saeed, A., O’Brien, P.E., 2008. Provenance of the Gamburtsev Subglacial
Mountains from U-Pb and Hf analysis of detrital zircons in Cretaceous to Quaternary
sediments in Prydz Bay and beneath the Amery Ice Shelf. Sedimentary Geology 211, 12
32.
Williams, T., van de Flierdt, T., Chung, E., Roy, M., Hemming, S.R., Goldstein, S.L.,
Abrahamowicz, M., Tremblay, B., 2007. U.S. Geological Survey and The National
Academies 10th International Symposium on Antarctic Earth Sciences, extended abstract
037.
Webb, P.-N., 1990. Review: the Cenozoic history of Antarctica and its global impact. Antarctic
Science 2, 3-21.
!
27!
Whitehead, J.M., Quilty, P.G., McKelvey, B.C., O’Brien, P.E., 2006. A review of the Cenozoic
stratigraphy and glacial history of the Lambert Graben – Prydz Bay region, East
Antarctica. Antarctic Science 18, 83-99.
Young, D.A., Wright, A.P., Roberts, J.L., Warner, R.C., Young, N.W., Greenbaum, J.S.,
Schroeder, D.M., Holt, J.W., Sugden, D.E., Blankenship, D.D., van Ommen, T.D.,
Siegert, M.J., 2011. A dynamic early East Antarctic Ice Sheet suggested by ice-covered
fjord landscapes. Nature 474, 72-75.
Zachos, J.C., Breza, J.R., Wise, S.W., 1992. Early Oligocene ice-sheet expansion on Antarctica:
Stable isotope and sedimentological evidence from Kerguelen Plateau, southern Indian
Ocean. Geology 20, 569-573.
!
28!
Figure 1. Isostatically rebounded subglacial topography (80 times vertically exaggerated) of the
Lambert Graben/Prydz Bay region of East Antarctica showing the location of core 739C used in
this study and the modeled pre-glacial drainage into the Lambert Graben (Jamieson and Sugden,
2008; after Cox et al., 2010).
!
29!
Figure 2. Bedrock geology map of the area surrounding the Lambert Graben and Prydz Bay
showing location of core 739C (after van de Flierdt et al., 2008; Fitzsimons, 2003).
!
30!
Figure 3. Stratigraphy of ODP core 739C. Analyzed segments of the core are indicated with
arrows along the left side of the column (Volpi et al., 2009; Shipboard Scientific Party, 1989).
!
31!
Figure 4. (a) Zircon U-Pb and (b) zircon fission track probability density plots. Samples from
each age division of core 739C are combined in each plot. Note the dominant ~500 Ma peak in
all of the U-Pb spectra and the increasing occurrence of Precambrian grains in the Oligocene
samples. Also note the two populations of FT ages (~250-300 Ma and ~120 Ma). (n=number of
analyses)
!
32!
Figure 5. Apatite U-Pb probability density plots. Samples from each age division of core 739C
are combined in each plot. Note the dominant ~500 Ma peak in every age of sediments.
(n=number of analyses)
!
33!
Figure 6. Apatite fission track probability density plots. The data for each sample is shown
separately and correlated with its location in core 739C. Note the decreasing trend in mean
cooling age throughout the Early Oligocene, corresponding with a change in lag time from ~240180 My, compared with a change in lag time during the late Miocene through Quaternary from
~170 Ma-150 Ma. (n=number of analyses)
!
34!
Figure 7. Apatite (U-Th)/He probability density plots. All samples for each age distribution of
core 739C are combined in each plot. Raw ages are plotted in the histograms. Red PDF curves
represent raw ages. Green PDF curves represent plot ages, or the average between the raw age
and the fully corrected age of a grain. Note the shift to younger ages (<100 Ma) after the Early
Oligocene. (n=number of analyses)
!
35!
Figure 8. Plot of stratigraphic age versus central apatite fission track age for all samples. Black
lines represent constant lag time. Note the decrease in lag time in the Early Oligocene samples
from ~250 My to ~180 My, compared with the nearly constant lag time of ~160 My (nearly
steady-state erosion) in the late Miocene through Quaternary samples. Error bars are ±1σ for
central fission track ages and ±1 Ma uncertainty for depositional ages.
!
36!
Figure 9. HeFTy time-temperature models with histograms showing apatite fission track (blue)
and raw apatite (U-Th)/He (green) age spectra for late Miocene samples. Youngest fission track
age (~80 Ma) corresponds to a modeled incision depth of ~2.8 km by the late Miocene. Youngest
(U-Th)/He age (~25 Ma) corresponds to an incision depth of ~3 km by the late Miocene.
!
37!
Download