ArticlePDF Available

A lithofacies analysis of a South Polar glaciation in the Early Permian: Pagoda Formation, Shackleton Glacier region, Antarctica

Authors:

Abstract and Figures

The currently favored hypothesis for Late Paleozoic Ice Age glaciations is that multiple ice centers were distributed across Gondwana and that these ice centers grew and shank asynchronously. Recent work has suggested that the Transantarctic Basin has glaciogenic deposits and erosional features from two different ice centers, one centered on the Antarctic Craton and another located over Marie Byrd Land. To work towards an understanding of LPIA glaciation that can be tied to global trends, these successions must be understood on a local level before they can be correlated to basinal, regional, or global patterns. This study evaluates the sedimentology, stratigraphy, and flow directions of the glaciogenic, Asselian–Sakmarian (Early Permian) Pagoda Formation from four localities in the Shackleton Glacier region of the Transantarctic Basin to characterize Late Paleozoic Ice Age glaciation in a South Polar, basin-marginal setting. These analyses show that the massive, sandy, clast-poor diamictites of the Pagoda Fm were deposited in a basin-marginal subaqueous setting through a variety of glaciogenic and glacially influenced mechanisms in a depositional environment with depths below normal wave base. Current-transported sands and stratified diamictites that occur at the top of the Pagoda Fm were deposited as part of grounding-line fan systems. Up to at least 100 m of topographic relief on the erosional surface underlying the Pagoda Fm strongly influenced the thickness and transport directions in the Pagoda Fm. Uniform subglacial striae orientations across 100 m of paleotopographic relief suggest that the glacier was significantly thick to “overtop” the paleotopography in the Shackleton Glacier region. This pattern suggests that the glacier was likely not alpine, but rather an ice cap or ice sheet. The greater part of the Pagoda Fm in the Shackleton Glacier region was deposited during a single retreat phase. This retreat phase is represented by a single glacial depositional sequence that is characteristic of a glacier with a temperate or mild subpolar thermal regime and significant meltwater discharge. The position of the glacier margin likely experienced minor fluctuations (readvances) during this retreat. Though the sediment in the Shackleton Glacier region was deposited during a single glacier retreat phase, evidence from this study does not preclude earlier or later glacier advance–retreat cycles preserved elsewhere in the basin. Ice flow directions indicate that the glacier responsible for this sedimentation was likely flowing off of an upland on the side of the Transantarctic Basin closer to the Panthalassan–Gondwanide margin (Marie Byrd Land), which supports the hypothesis that two different ice centers contributed glaciogenic sediments to the Transantarctic Basin. Together, these observations and interpretations provide a detailed local description of Asselian–Sakmarian glaciation in a South Polar setting that can be used to understand larger-scale patterns of regional and global climate change during the Late Paleozoic Ice Age.
Content may be subject to copyright.
Journal of Sedimentary Research, 2021, v. 91, 611–635
Research Article
DOI: 10.2110/jsr.2021.004
A LITHOFACIES ANALYSIS OF A SOUTH POLAR GLACIATION IN THE EARLY PERMIAN:
PAGODA FORMATION, SHACKLETON GLACIER REGION, ANTARCTICA
LIBBY R.W. IVES AND JOHN L. ISBELL
University of Wisconsin–Milwaukee, Department of Geosciences, Lapham Hall, 3209 North Maryland Avenue, Milwaukee, Wisconsin 53211, U.S.A.
e-mail: woodfor5@uwm.edu
ABSTRACT: The currently favored hypothesis for Late Paleozoic Ice Age glaciations is that multiple ice centers
were distributed across Gondwana and that these ice centers grew and shank asynchronously. Recent work has
suggested that the Transantarctic Basin has glaciogenic deposits and erosional features from two different ice
centers, one centered on the Antarctic Craton and another located over Marie Byrd Land. To work towards an
understanding of LPIA glaciation that can be tied to global trends, these successions must be understood on a local
level before they can be correlated to basinal, regional, or global patterns. This study evaluates the sedimentology,
stratigraphy, and flow directions of the glaciogenic, Asselian–Sakmarian (Early Permian) Pagoda Formation from
four localities in the Shackleton Glacier region of the Transantarctic Basin to characterize Late Paleozoic Ice Age
glaciation in a South Polar, basin-marginal setting. These analyses show that the massive, sandy, clast-poor
diamictites of the Pagoda Fm were deposited in a basin-marginal subaqueous setting through a variety of
glaciogenic and glacially influenced mechanisms in a depositional environment with depths below normal wave
base. Current-transported sands and stratified diamictites that occur at the top of the Pagoda Fm were deposited as
part of grounding-line fan systems. Up to at least 100 m of topographic relief on the erosional surface underlying
the Pagoda Fm strongly influenced the thickness and transport directions in the Pagoda Fm. Uniform subglacial
striae orientations across 100 m of paleotopographic relief suggest that the glacier was significantly thick to
‘‘overtop’’ the paleotopography in the Shackleton Glacier region. This pattern suggests that the glacier was likely
not alpine, but rather an ice cap or ice sheet. The greater part of the Pagoda Fm in the Shackleton Glacier region
was deposited during a single retreat phase. This retreat phase is represented by a single glacial depositional
sequence that is characteristic of a glacier with a temperate or mild subpolar thermal regime and significant
meltwater discharge. The position of the glacier margin likely experienced minor fluctuations (readvances) during
this retreat. Though the sediment in the Shackleton Glacier region was deposited during a single glacier retreat
phase, evidence from this study does not preclude earlier or later glacier advance–retreat cycles preserved
elsewhere in the basin. Ice flow directions indicate that the glacier responsible for this sedimentation was likely
flowing off of an upland on the side of the Transantarctic Basin closer to the Panthalassan–Gondwanide margin
(Marie Byrd Land), which supports the hypothesis that two different ice centers contributed glaciogenic sediments
to the Transantarctic Basin. Together, these observations and interpretations provide a detailed local description of
Asselian–Sakmarian glaciation in a South Polar setting that can be used to understand larger-scale patterns of
regional and global climate change during the Late Paleozoic Ice Age.
INTRODUCTION
Strata of the Transantarctic Basin (TAB) contain a complete South Polar
sedimentary record of the global ‘‘ icehouse’’ to ‘‘ greenhouse’’ transition
during the Early Permian (Collinson et al. 1994, 2006; Isbell et al. 2008b).
Sedimentation in the TAB was dominated by glaciogenic processes during
the Asselian–Sakmarian (Isbell et al. 2008c). This interval was part of the
Late Paleozoic Ice Age (LPIA, ~374–256 Ma) (Fielding et al. 2008c;
Monta˜
nez and Poulsen 2013). Widespread glaciation across Gondwana
characterized the LPIA, as did low pCO
2
, high pO
2
, generally low eustatic
levels with large magnitude fluctuations, low solar luminosity, and
increased d
18
O and d
13
C values relative to the rest of the Phanerozoic
(Gastaldo et al. 1996; Raymond and Metz 2004; Monta˜
nez and Soreghan
2006; Fielding et al. 2008d; Rygel et al. 2008; Monta˜
nez and Poulsen
2013).
The currently favored hypothesis for LPIA glaciations is that multiple
ice centers (ice sheets or ice caps) were distributed across Gondwana, and
that these ice centers grew and shank asynchronously over the LPIA’s ~80
Myr duration (Fielding et al. 2008c; Isbell et al. 2012; Monta˜
nez and
Poulsen 2013; L´opez-Gamund´ı et al. 2021; Rosa and Isbell 2021). The
character, distribution, and resulting sedimentary records of these glaciers
would have been driven by global, regional, and local climatic and
geologic influences (Isbell et al. 2012; Monta˜
nez and Poulsen 2013;
opez-Gamund´ı et al. 2021). The potential for local and regional
heterogeneity of LPIA glaciogenic strata is therefore extremely high. To
work towards an understanding of LPIA glaciation that can be tied to
Published Online: June 2021
Copyright Ó2021, SEPM (Society for Sedimentary Geology) 1527-1404/21/091-611
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
global trends, these successions must be understood on a local level before
they can be correlated to basinal, regional, or global patterns.
In this paper, we evaluate the sedimentology, stratigraphy, and flow
directions of four glaciogenic (Pagoda Fm) successions in the Shackleton
Glacier region of the TAB (Fig. 1B). The Pagoda Fm in the Shackleton
Glacier region has not previously been described and analyzed at the level
of detail reported in this study. The Shackleton Glacier region can offer a
different perspective to better-studied areas of the TAB (e.g., the
Beardmore Glacier region) because it was located in a basin marginal
position on the non-cratonic (or, ‘‘Panthalassan proximal’’) side of the
basin during the deposition of the Pagoda Fm (Fig. 1).
GEOLOGICAL SETTING
Asselian–Sakmarian glaciogenic strata of the Transantarctic Basin
(TAB) occur in discontinuous outcrops along the margin of the East
Antarctic Craton from Victoria Land, near Australia, to Dronning Maud
Land, near southern Africa (Frakes et al. 1971; Collinson et al. 1994; Isbell
et al. 2008c) (Fig. 2A). During the lower Cisuralian, the TAB was a narrow
(~100–200km-wide), trough-shaped basin that formed parallel and
proximal to the Gondwanide margin of the East Antarctica Craton, but
inboard of the Panthalassic margin (Fig. 3B) (Collinson et al. 1994; Elliot
2013; Isbell 2015; Elliot et al. 2017). In the central Transantarctic
Mountains and Victoria Land, glaciogenic strata occur in four sub-basins:
the Ohio Range to the Scott Glacier (Horlick Sub-basin), the Amundsen
Glacier to the Darwin Glacier area (Beardmore Sub-basin), south Victoria
Land (SVL), and north Victoria Land (NVL) (Figs. 1B, 3) (Frakes et al.
1966; Isbell et al. 2008c; Isbell 2010; Cornamusini et al. 2017). The
Shackleton Glacier region is located near the southern edge of the
Beardmore Sub-basin.
The origin and nature of the TAB during the Lower Permian is not well
understood. Hypotheses include intracratonic and extensional settings
(Collinson et al. 1994; Isbell 2015; Elliot et al. 2017). Regardless of what
processes drove basin formation at that time, the TAB was a narrow,
trough-shaped basin, with Proterozoic and early Paleozoic basement
shoulders, that paralleled the Panthalassan margin of the East Antarctic
Craton during deposition of the Pagoda Fm (Fig. 3) (Isbell et al. 1997b).
Though there is no evidence for upper Carboniferous to Sakmarian
orogenic activity in the central Transantarctic Mountains or adjacent Marie
Byrd Land, volcanic arcs and tectonic compression were occurring
elsewhere along Gondwana’s Panthalassan margin during that time,
including in eastern Australia, the Andean margin of South America and
Patagonia, the Ellsworth Mountains, Thurston Island, and the Antarctic
Peninsula (Fielding et al. 2001; Elliot 2013; Viza
´n et al. 2017). This same
margin was extremely active and subject to repeated, complex accretion
events throughout the Paleozoic (Veevers et al. 1994; Domeier and Torsvik
2014; Goodge 2020). As a result of this activity, the TAB evolved into a
foreland basin later in the Permian (Collinson et al. 1994; Elliot et al.
2017). During the Lower Jurassic, strata in the central Transantarctic
Mountains were pervasively intruded by sills associated with Ferrar Group
volcanism and the breakup of Gondwana (Elliot 1992).
SEDIMENTOLOGY AND STRATIGRAPHY OF THE PAGODA FORMATION
The Pagoda Fm is the basal unit in the Permian–Early Jurassic Victoria
Group (upper Beacon Supergroup) in the Beardmore Sub-basin of the
TAB. Rare palynomorphs and conchostracans suggest that the Pagoda and
Mackellar fms are Asselian–Sakmarian (Masood et al. 1994; Askin 1998;
Babcock et al. 2002). The Pagoda Fm overlies both the Kukri and Maya
regional erosional surfaces (Figs. 1B, 3) (Collinson et al. 1994; Isbell 1999;
Elliot 2013). The Maya Erosional Surface is a disconformity that separates
Devonian(?) clastics of the lower Beacon Supergroup from the Victoria
Group (Isbell 1999). The Kukri Erosional Surface separates the Beacon
Supergroup from underlying Ross Orogeny intrusions and associated
metasediments. Significant relief of at least 150 m occurs on these
unconformities (Fig. 4) (Isbell et al. 1997a, 2008c; Isbell 1999). Both the
Pagoda Fm, and the overlying, post-glacial Mackellar Fm, lap onto the
erosional surfaces, indicating that the Pagoda Fm and its equivalents in
other sub-basins often did not overtop the relief (Isbell et al. 1997a). The
lower Beacon Supergroup units are not present in the Shackleton Glacier
region, and the erosional surface underlying the Pagoda Fm is merged
Maya and Kukri surface (Isbell et al. 2008c). At all sites in this study, the
Pagoda Fm overlies Ross Orogeny granites.
Since their discovery, the Pagoda Fm and its equivalents throughout the
Transantarctic Mountains have been unanimously interpreted as glacio-
genic or glacially influenced because their predominant lithologies are
massive and laminated, sandy and silty diamictites (Long 1964a; Lindsay
1970a; Coates 1985; Barrett et al. 1986; Collinson et al. 1994; Isbell et al.
2008c). Minor lithologies of the Pagoda Fm include conglomeratic
sandstones, sandstones, mudrocks, and lonestone-bearing mudrocks (Isbell
et al. 2008c). Besides diamictites and lonestones, evidence for a glacial
origin for the Pagoda Fm includes striated and polished basement surfaces,
the prevalence of striated and faceted clasts, and a clear relationship
between local basement composition and lithologies of large clasts in the
diamictites (Lindsay 1969; Coates 1985). Detailed interpretations of
depositional environments have been made for a few Pagoda Fm localities
(Lindsay 1970a; Waugh 1988; Miller 1989; Isbell et al. 2001; Lenaker
2002; Long et al. 2008–2009; Koch 2010; Koch and Isbell 2013) and its
equivalents in Victoria Land (Askin et al. 1971; Barrett 1972; Barrett and
McKelvey 1981; Isbell 2010; Cornamusini et al. 2017), Horlick Mountains
(Frakes et al. 1966; Aitchison et al. 1988), and Ellsworth Mountains
(Ojakangas and Matsch 1981; Matsch and Ojakangas 1991). With few
exceptions, these analyses have invoked subaqueous, glacial-proximal
depositional settings. This is in contrast to early surveys that interpreted
diamictites as subglacially deposited ‘‘tillites’’ (Lindsay 1970a; Coates
1985; Miller 1989; Isbell et al. 1997b).
Isbell et al. (2008c) separated the Permian glaciogenic units in the
Transantarctic Mountains into basin-margin and basinal facies associa-
tions. Basin-margin successions are predicted to occur near basement highs
and along basin margins, are relatively thin (,100 m), contain evidence
for subglacial deformation and erosion, have deformation resulting from
proglacial glaciotectonism, and small (m-scale) gravity-driven deposition.
Basinal successions are thicker (100–500 m), have little-to-no evidence for
subglacial processes, and are more likely to contain stratified diamictites,
lonestone-bearing mudrocks, mudrocks, and larger (up to tens of meters)
mass-transport deposits. Evidence of grounded ice and grounding-line
processes have been identified in both basinal (e.g., Koch and Isbell
(2013)) and basin-margin (e.g., Isbell (2010)) facies associations. Based on
its paleogeographic position and Pagoda Fm thickness, the Shackleton
Glacier area is here predicted to contain the basin-margin facies.
In the Shackleton Glacier region, the glaciogenic facies of the Pagoda
Fm are underlain by a non-glacial, lacustrine facies association at a single
site on Mt. Butters (site MB-17). This facies and its depositional
environment are described in detail by Isbell et al. (2001). Below this
contact, the fine-grained lacustrine facies are pervasively sheared, likely
subglacially (Isbell et al. 2001). Lonestones, interpreted to be iceberg-
rafted debris, occur in the lower post-glacial Mackellar Fm in the
Shackleton Glacier region (Seegers 1996; Seegers-Szablewski and Isbell
1998). This suggests that glaciers were still present in the Transantarctic
Basin even after glaciogenic sedimentation was no longer dominant.
STUDY AREA AND METHODS
The sedimentary sections described in this paper were examined as part
of the U.S. Antarctic Program’s helicopter-supported Shackleton Glacier
Deep-Field Camp during the 2017–2018 austral summer (Table 1; Fig.
L.R.W. IVES AND J.L. ISBELL612 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 1.—Generalized geologic maps of study area. Maps are South Polar projections. A) Geologic map of the Central Transantarctic Mountains and Victoria Land, with
relevant outlet glaciers and mountain ranges labeled. Modified after Elliot (2013), Goodge and Fanning (2016), and Estrada et al. (2016). Box on inset map indicates extent of
this geologic map. Red box on geologic map indicates the extent of ‘‘map B.’’ B) Regional geologic map of the Shackleton Glacier area, noting the locations of sections
described in this study. MM-17 is Mt. Munson, MB-17 is Mt. Butters 1, MBSE-17 is Mt. Butters 2, and RS-18 is Reid Spur. Geology adapted from McGregor and Wade
(1969), and Mirsky (1969), aerial photos from LIMA Landsat imagery (Bindschadler et al. 2008).
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 613
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 2.—Paleogeographic reconstructions of Gondwana near the Carboniferous–Permian Boundary. All maps are south-polar projections. Star indicates the approximate
location of the Shackleton Glacier area. Continent distributions and paleolatitudes are based on Lawver et al. (2011) and copied from Isbell et al. 2012). Note that there are
differences in the positioning of some crustal blocks (e.g., Patagonia and New Zealand) between this reconstruction and Figure 3, which is modified after Elliot (2013). A)
Yellow regions indicate the modern extent of sedimentary basins containing Late Paleozoic Ice Age strata. Abbreviations include: Falkland Islands–Malvinas (FI), Ellsworth
Mountain block (EM), Antarctic Peninsula (AP), Thurston Island (TI), Marie Byrd Land (MBL), and the Challenger Plateau–western New Zealand (ChP). Basins are adapted
L.R.W. IVES AND J.L. ISBELL614 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
1B). These sections are located on the Mt. Butters Massif (MB-17, MBSE-
17) on the west side of the Shackleton Glacier, on the east face of Reid
Spur (RS-18) of the Ramsey Glacier, and Mount Munson (MM-17) at the
head of Barrett Glacier. The two sections at Mt. Butters are separated by
approximately 2 km.
Sites were selected for section descriptions based on accessibility,
continuity of exposure, and preliminary observations from previous
expeditions to the area. At MB-17 several shorter sections were measured
to capture lateral variability. Sedimentological data (texture, grain shape,
sedimentary structures, etc.) as well as paleotransport indicators (including
cross stratification, primary current lineations, striae, slickenslides, fold
hinge lines, and thrust-plane orientation) were logged in each section.
Measured sections were placed in context using outcrop-scale photographs
taken from helicopters (Fig. 4).
Structural and sedimentary orientations were measured using a Brunton
compass, with the azimuth set to 0008. Measurements were corrected for
magnetic declination using the NOAA Magnetic Declination Estimated
Value Calculator (NOAA 2019). Orientations were corrected for structural
dip and aggregate orientations calculated using Stereonet (v. 11) software
(Allmendinger et al. 2012; Cardozo and Allmendinger 2013). Some
measurements from Mt. Munson were collected by JLI during the 1997–
1998 field campaign to the Shackleton Glacier region. New flow-direction
measurements are in Table 2 and site descriptions are in Appendix A.
FACIES ASSOCIATIONS
The following descriptions are of glaciogenic facies associations (FA)
that constitute the Pagoda Fm in the Shackleton Glacier region (Table 3).
We use the term glaciogenic to refer to sedimentary systems whose
components are dominantly derived from glacial erosion and/or transport,
and whose depositional processes are glacier-driven. For example, a
succession that is the result of a plume from a subglacial jet would be
glaciogenic, but largely non-glacially-derived deep-water sediments with
the occasional outsized clast would instead be considered ‘‘glacially
influenced.’’ The characteristics of distinct sediment grain sizes are
consistent throughout these successions. Very-fi ne-grained sandstones and
shales are black in color. Fine- to medium-grained sandstones are generally
quartz-rich with some lithic and potassium feldspar grains. Cobble- and
boulder-size clasts are sourced from local basement lithologies (Fig. 1B),
and include predominantly phaneritic granitoids, with some gneiss,
quartzite, and gray, fine-grained metasedimentary rocks. All sand-size
and coarser-grained material in the Pagoda Fm occurs in all categories of
particle roundness (angular to well rounded), although finer-grained sands
are typically better-rounded than medium- and coarse-grained sands.
Striations occur on large clasts throughout the Pagoda Fm but are not
common. The lack of striations is possibly due to the hardness of
individual clasts, which are primarily composed of granite, quartz, and
feldspar (Dowdeswell et al. 1985; Bennett et al. 1997).
Massive Sandy Diamictite Facies Association (MSD)
MSD Description.—This diamictite is the dominant FA of the Pagoda
Fm. Similar lithologies occur throughout the Transantarctic Basin. In the
Shackleton Glacier region, the thickness of this FA ranges from 3 m at Mt.
Munson (MM-17) to 73 m at Mt. Butters (MB-17C). Since this diamictite
is almost wholly massive, there is no clear way to further subdivide these
successions. Where exposed, the lower contact overlies either a striated and
polished unconformity with the Queen Maud Batholith (MM-17 and
MBSE-17) or subglacially deformed lacustrine sediments (Isbell et al.
2001; MB-17C). The upper contact of this FA is sharp or erosional with
current-transported facies, including the Cross Bedded Sandstone (CBS)
facies association, the Heterogenous Sandy (HS) facies association, and the
Mackellar Fm. The contact between this FA and the HS facies association
is also gradational where facies HS1 (stratifi ed diamictite) is present above
the contact. The upper part of this FA are intercalated with turbidites (see
LS facies interpretation), and may interfinger with stratified diamictites and
mass-transport deposits (see HS facies association interpretation).
Approximately 90% of this FA is clast-poor to clast-rich diamictite
(Hambrey and Glasser 2003; Hambrey and Glasser 2012) with minor
amounts of sorted sands and gravels (Fig. 5). Clast abundances fluctuate
throughout the succession. Some intervals are sufficiently clast-poor that
they could be classified as muddy sandstones with dispersed clasts (,1%
clasts) (Fig. 5F) (Moncreiff 1989; Hambrey 1994; Hambrey and Glasser
2012). Most clast-rich parts of this diamictite contain 10–15% clasts (Fig.
5A, E), but some very limited areas contain up to ~30% clasts (Fig. 5C).
Clasts range from pebble-size to 4 m in diameter. Clast compositions
includes granite, feldspar, vein quartz, gneiss, and fine-grained metasand-
stone. Clast shape ranges from rounded to angular. Faceted clasts are
common, but bullet-shaped and striated clasts are rare. The matrix is very
poorly sorted, with sizes ranging from muds through granule-size grains.
Matrix grain-size distributions remain constant within and between
outcrops of this facies, though mean matrix grain size increases slightly
in clast-rich sections relative to clast-poor sections.
The diamictite in this FA is massive, with very rare exceptions.
However, broadly defined zones of clast-poor or clast-rich diamictites do
from Isbell et al. (2012). B) Proposed positions of glacial centers during the Early Permian based on flow directions and position of basins and ‘‘ highlands.’’ Illustrated ice
centers are not meant to represent the whole possible extent of each proposed glacier, but where proposed glaciers were likely to be nucleated. The arrows reflect field
measurements of flow directions reported in the studies cited for each ice center. However, flow directions of glaciers are highly variable, both spatially and temporally, and the
true flow paths of these ancient ice centers were likely much more variable than the arrows on this map. Confidence is based on abundance of available lithologic data, and
both relative and absolute ages. Ice centers are as follows: MBL, the proposed Marie Byrd Land ice center, discussed in this study as the most likely source for the glaciogenic
sediments of the Pagoda Fm in the Shackleton Glacier region (Isbell et al. 1997b; Isbell 2010), a) Uruguay (Crowell and Frakes 1975; Assine et al. 2018; Fedorchuk et al.
2019), b) Asunsci´on (Frakes and Crowell 1969; Fran¸ca and Potter 1988; Limarino et al. 2014), c) Windhoek–Koakoveld Highlands (Martin 1981; Visser 1987; Fran¸ca et al.
1996; Rosa et al. 2016; Tedesco et al. 2016; Assine et al. 2018; Dietrich et al. 2019; Fallgatter and Paim 2019), d) Cargonian Highlands (Crowell and Frakes 1972; Visser
1997; Isbell et al. 2008a; Dietrich et al. 2019), e) Cape–Ventana Fold Belt (Visser 1997; Isbell et al. 2008a; Wopfner 2012), f) East African Thermal Rise (Rust 1975; Wopfner
2012), g) Patagonian Western Magmatic Arc (Pauls 2014; Survis 2015; Marcos et al. 2018), h) Zimbabwe (Wopfner 2012; Dietrich et al. 2019), i) Madagascar–SW India
(Veevers and Tewari 1995; Isbell et al. 2012), j) Chotanagpur and Chhattisgarh (Veevers and Tewari 1995; Dasgupta 2006; Isbell et al. 2012), k) Pilabra–Yilgarn, l) Kimberly
(see Mory et al. (2008); Martin et al. (2019), and references therein, m) Arunta–Musgrave (Mory et al. 2008 and Martinet al. 2019) and references therein, n) Bowen–
Gunnedah–Sydney (Fielding et al. 2008a, 2008b, 2010), o) Galilee (Fielding et al. 2008a, 2008b, 2010; Isbell et al. 2012), p) Wilson (Hand 1993; Rocchi et al. 2011; Jordan et
al. 2013), r) East Antarctic (Isbell et al. 1997a; Isbell 2010), s) Ellsworth (Frakes et al. 1971; Ojakangas and Matsch 1981; Matsch and Ojakangas 1992; Visser 1997).
TABLE 1.—Names and locations of sedimentary sections described in this
paper.
Location
Name
Section
Name Geographic Coordinates
Thickness of
Pagoda Fm
Mt. Butters 1 MB-17 S84851.0290W177825.216090 m
Mt. Butters 2 MBSE-17 S84853.0030W177822.354077 m
Reid Spur RS-18 S84847.035 0E178846.6800.62 m
Mt. Munson MM-17 S84845.3590E173841.11805m
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 615
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 3.—Regional geologic context and tectonic setting of the Transantarctic Basin during the Asselian–Sakmarian. A) Stratigraphy of the Beacon Supergroup across
different regions of the Transantarctic Basin. Adapted from Elliot (2013), Cornamusini et al. (2017), and Elliot et al. (2017). B) Tectonic setting of southern Gondwana during
the Permian, adapted from Elliot (2013). Regions of sedimentary deposition are shaded yellow. Note that there are differences in the positioning of some crustal blocks (e.g.,
Patagonia and New Zealand) between this reconstruction and Figure 2. C) Modern extent and isopach map of the Asselian–Sakmarian glaciogenic facies (Pagoda Fm and
equivalents) in the Horlick Sub-basin and Beardmore Sub-basin in the Transantarctic Mountains. Gray areas are outcrops (nunatuks). Lines show isopachs of the Pagoda Fm
(Beardmore Sub-basin) and Buckeye Fm (Horlick Sub-basin) from Isbell et al. (2008c).
L.R.W. IVES AND J.L. ISBELL616 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
occur. These zones are anywhere from 1 to 30 m thick. While we tracked
these changes in clast abundance vertically through the measured sections,
it was not our impression that these ‘‘zones’’ have any sort of horizontal
organization as might be implied by the terms ‘‘layers’’ or ‘‘ horizons,’’ and
that there are no distinct bounding surfaces between these zones. The
transitions between these zones are gradational. These gradational
transitions occur on the decimeter to meter scale. There was no clear
relationship between transition thickness and any other property of these
rocks. Zones of diamictites with similar clast concentrations cannot be
correlated between outcrops; the distribution of clast-poor and clast-rich
diamictites appears to be unique to each locality.
Crude and chaotic bedding occur very occasionally in the otherwise
massive diamictite. Where bedding can be discerned, the beds are 2–10 cm
thick, laterally discontinuous, and internally massive. Rare sedimentary
structures include ruck structures beneath large clasts (Fig. 5D) and
thinning of beds over large clasts (Fig. 5H). Rare beds of boulders and
cobbles occur in the lower part of this facies at both Mt. Butters sites (Fig.
5G). In all ‘‘boulder beds’’ the clasts were not striated, polished, or
uniformly oriented.
Sandstone and/or conglomerate bodies are also rare within this FA.
These bodies occur in the diamictite and are poorly to moderately sorted.
The grain compositions in these bodies are similar to sand- to gravel-size
grains in the diamictite. Most sand bodies are massive but can contain
layers of discrete grains sizes. However, these layers are often highly
deformed. Bodies of sorted sand and/or gravel occur in two distinct sizes.
Small bodies consist of sorted sediments that most often occur as
irregularly shaped, horizontally elongate lenses, ‘‘whisps,’’ and boudinage-
like bodies that are up to 20 cm thick and 100 cm long (Figs. 5A, B, E, F).
Such structures may occur alone or, more frequently, in bands and zones up
to 1 m thick. Occasionally, diapir-like structures made of sand and/or
gravel, 1 to 2 m thick, are also present. These structures project upward
into the diamict but are not associated with other sand or gravel bodies.
Larger bodies of well-sorted sand or gravel range in thickness from 1 to 3
m and are laterally discontinuous. These bodies have the same grain-size
distributions as the smaller bodies. Large sand bodies are typically
massive, but where stratification does exist in the sands and/or gravels the
beds are largely deformed and display water-escape structures. The large
sand bodies thin laterally and have overturned folds on their thicker ends.
In one instance, the main sand body was accompanied by smaller sand
bodies that trailed off away from its thick end (comet-like structure) (Fig.
5F; Sandstone with dispersed clasts). The lower contact between the
comet-like bodies and the surrounding diamictite has a slope of 308to 358
above horizontal, and the underlying diamictite sometimes has a fissile
structure, indicating shearing. Occasionally such contacts overlie smaller
sand lenses, sheath folds, ‘‘whisps,’’ and boudins.
MSD Interpretation.—This FA is most likely glaciogenic or glacially
influenced. Evidence for glacial transport of sediment in this FA includes
striations and polish on basement granites where diamictites rest directly
on granite (MBSE-17 and MM-17), the very poor sorting of sediment in
the system, and the presence of angular to rounded grains of all sizes,
faceted and striated clasts, and large boulders composed of local basement
lithologies. In glacial settings, massive diamictites, like the facies described
here, may be the result of subglacial till deposition (Evans et al. 2006),
settling from suspension of a meltwater plume (Visser 1994), settling from
suspension and rain-out from icebergs and iceberg scouring (Dowdeswell
et al. 1994; Lisitzin 2002), mass-transport deposits (Rodrigues et al. 2019),
or debris flows (Powell and Molnia 1989). Glacial depositional
environments that include these processes are subglacial, proglacial
proximal (but outside the influence of the grounding zone), and
grounding-zone-environments including ground-zone wedges (Batchelor
and Dowdeswell 2015; Demet et al. 2019; Dietrich and Hoffmann 2019),
morainal banks (Eidam et al. 2020), and ice-contact fans (Powell 1990;
FIG. 4.—Photographs of Mt. Butters outcrops showing the stratigraphy of the
Pagoda Fm and postglacial Mackellar Fm, and well as the relief of the Maya
Erosional Surface, in this area. Formations are labelled, and the Maya Erosional
Surface is marked by a green, dashed line. A) A photograph of section MB-17 taken
from a helicopter; view to the south. The saddle in the background of this photo is
approximately halfway between section MB-17 and MBSE-17. B) An outcrop that is
part of the Mt. Butters Massif, and occurs on a spur located southeast of section MB-
17. View is toward the southeast from MB-17. This site has never been visited on
foot, so the scale is not certain. Note the dip of the granitic basement towards the
southwest, which is consistent with basement dip measurements at site MB-17 and
MBSE-17 in this study. This exposure is the same as Figure 4 in (Isbell et al. 2001).
C) Photograph of section MBSE-17, view toward the SW. Purple oval shows
helicopter’s shadow.
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 617
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
Benn 1996). The most likely depositional processes and environments for
the MSD facies can be inferred by considering the unique features of this
facies.
The subglacial deposition of this FA is not likely to have been the
dominant process, but cannot be wholly ruled out. A tillite interpretation
for the MSD FA is supported by the massive, poorly sorted nature of the
diamictites. However, there is no strong evidence for glacier grounding (for
example, striated pavements or continuous erosional surfaces) above the
base of this facies. The boulder and cobble beds in this FA do not contain
uniform oriented, bulleted, or striated clasts. This suggests that the boulder
beds are lags; that they formed due to winnowing (Eyles 1988), not
subglacial processes.
This FA was more likely deposited subaqueously than subaerially. In
most of the sections measured in this study, the upper contact of the
diamictite is conformable with, and in some cases, gradationally transitions
into, various subaqueous facies (HS facies association). At section RSP-18,
TABLE 2.—Summary of paleo-transport measurements from this study (mean is Fisher mean 6a99). Note that due to the proximity of the study
locations to the south pole that the orientation of compass directions are not the same across the basin. Mack, Mackellar Fm; L. Pag, lower Pagoda
lacustrine facies (Isbell et al. 2001); T, trend; P, plunge; D, dip angle; DD, dip direction; S, strike; *, measured during 1997 field season.
Section (Height) Facies/Fm Feature Measurement Orientation N
Site MW-18
10 m above basement Mack Asymmetrical ripples Transport T: 3228, 3128, 00283
Site MM-17
0 m - Striae on basement Lineation T: 18081
*0 m - Striae on basement Lineation T: 095863.788
*4 m above base of Mackellar Mack Asymmetrical ripples Transport T: 1578628.587
13–17.5 m Mack Highly deformed slump features Vergence T: 1098, 1048, 11483
*4–48 m of Mackellar Fm Mack Asymmetrical ripples Transport T: 1098620.0810
Site MB-17
0 m - Dip and dip direction of basement Mean pole to plane T: 26285
P: 798612
Plane from mean S: 3528
pole D: 118
DD: 2628
MB-17A/B L. Pag Symmetrical ripple crest axes Mean lineation T: 346866.0 817
MB-17B L. Pag Slickenside lineation Mean T: 006 81
MB-17C (12 m) MSD Fold axes and small thrust faults Mean Vergence T: 22086 3187
MB-17C (20 m) MSD Plane of thrust faults Vergence T: 2568, 25182
MB-17C (58 m) MSD Slide surface Planes DD: 2438, 19182
D: 158,358
Plane from mean S: 1178
pole D: 238
DD: 206 8
MB-17C/D HA Asymmetrical ripples Mean transport T: 325830
(73–81m) Spread 2368- 0558
MB-17C/D HA Cross beds Transport T: 3468, 3568, 35685
2268, 2218
MB-17D HA Climbing Ripples Transport T: 2218, 25682
MB-17C (84 m) HA Grooves Direction of T: 2538624 85
(iceberg keel marks?) shallowing
MB-17C
(18 m above Pagoda) Mack Asymmetrical ripples Mean Transport T: 2198639 87
Site MBSE-17
0 m - Striae on Basement Lineation T: 3118, 31682
MBSE-17 (29 m) MSD Sheath fold hinge Orientation T: 0168,P:2081
Vergence 286 8??
MBSE-17 (31–32 m) MSD Thrust faults Mean pole to plane T: 26387
P: 698
Plane from mean S: 3538
pole D: 218
DD: 0838
Vergence: T: 2638
MBSE-17 (54 m) HA Crenulations on slide sufrace Lineation T: 2418, 25182
MBSE-17 (61–68 m) CBS Cross beds and asymm. ripples Mean transport T: 20886 29 87
MBSE-17 (21–26 m above Pagoda) Mack Asymmetrical ripples Mean transport T: 29886 6823
Site RSP-18
24–48 m LS Cross beds, asymm. ripples, and PCL Mean transport T: 17686 29 818
68–70 m Mack Asymmetrical ripples Transport T: 2618, 2418, 21184
2818
L.R.W. IVES AND J.L. ISBELL618 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
TABLE 3.—Description of facies and facies associations.
Facies
Association
Facies
Name Thickness Lithology Structures Formative Process
Depositional
Environment
Massive Sandy
Diamictite
(MSD)
5–75 m
(at least)
Clast-poor to clast-rich,
sandy diamictite; muddy
sand matrix; minor
amounts of
discontinuous sand and
gravel bodies
Diamictites are massive to
crudely bedded and
ungraded, sands are
generally massive, and
sometime laminated, but
always highly deformed
Glaciogenic, subaqueous
processes; likely a
combination of mass
transport, iceberg
rainout, iceberg
scouring, plume
sedimentation, and
subglacial till deposition
Glacier-proximal to
glacier-intermediate,
continental shelf
Laminated
Sands (LS)
~17 m Coarse to fine-grained,
well-sorted sandstones;
quartz-rich; subangular
to rounded
Fine- to medium-grained sands
that occur in thin, planar
beds with primary current
lineations; coarse-grained
sandstones are trough cross-
bedded; fine- to very-fine-
grain sandstones are
laminated, or thin-bedded
with unidirectional ripples
High-density turbidites
and/or a transitional
concentrated density
flow (Mulder and
Alexander 2001)
Distal or medial
portion of an ice-
contact fan or delta
Heterogenous
Sandy (HS)
HS1: Bedded
diamictite
1–10 m Clast-rich, sandy
diamictite; matrix is
moderately well-sorted
3–7 cm beds that are massive
with sharp, planar, and
laterally discontinuous
contacts; soft sediment
deformation associated with
facies HA2
Glaciogenic, subaqueous
processes; likely a
combination of iceberg-
rain-out and plume
sedimentation
Subaqueous glacier-
proximal and
grounding-line fan
system
HS2: Chaotic
sandstones
0–15 m Conglomerates to very
fine-grained sandstones
Bedding is usually massive,
soft-sediment deformation is
pervasive; few primary
sedimentary structures
preserved; secondary
structures include fold
noses, boudinage, faulting,
shear structures above and
below contacts, and ruck
structures
Mass-transport, gravity-
driven processes in the
form of slides, slumps,
and/or mass-transport
deposits
HS3: Stratified
sandstones
0–15 m Coarse- to very f ine-
grained sandstones
Medium- to coarse-grained
sandstones are thickly
laminated to bedded with
planar cross beds, trough
cross beds, climbing ripples,
3D ripples that are
asymmetrical or climbing,
hummocky and swaly cross
stratification, and
symmetrical ripples with
bundled upbuilding; very-
fine and fine sandstones are
laminated or thin-bedded
include unidirectional
ripples, some flaser ripples
and climbing ripples;
occasional, small-scale soft-
sediment deformation occurs
in all lithologies
Current- dominated
transport and deposition,
with some slumping;
unconfined flow; poorly
sorted sediment source;
large variations in
current velocities;
occasional wave
reworking
— Cross-bedded
Sandstone
(CBS)
10–30 m Well-sorted, medium- to
very coarse-grained
quartz arenite sandstone;
rare pebbles, cobbles,
and conglomerates
Low-angle and trough
crossbeds; crossbed sets
range in thicknesses from
~15 cm to ~1.5 m; rare
thin beds with asymmetrical
ripples; occur within
amalgamated channels in
multi-storied, multi-lateral
sand sheet ,2 km wide
Strong tractive flow
confined to a series of
amalgamated channels;
subaqueous
Unconfined,
distributive flow;
likely glacial-
proximal subaqueous
fan/ grounding-line
fan
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 619
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 5.—Photographs of the Massive Sandy Diamictite (MSD) facies association. Rulers are 50 cm long when folded in half and 1 m long when unfolded. Marks on rulers
are in cm. The weathered color of the diamictite is gray to beige and tends to be redder when the matrix has a higher percentage of sand. Small sand bodies are highlighted by
blue where present, and orange line indicates important bedding planes. A) A characteristic clast-poor section of the MSD facies association with a sand ‘‘whisp’’ (MB-17C).
B) A diamictite section of this facies that is very clast poor, yet contains several large boulders (MB-17C). C) A clast-rich section of the MSD diamictite (MB-17C). D) A ruck
structure in the diamictite made by a boulder (MM-17). E) Examples of small sand bodies in the MSD diamictite that experienced simple shear strain, resulting in sheath fold,
‘‘stringers,’’ boudinage, and en echelon structures. (MBSE-17). F) Example of small sand-body bands in an otherwise massive, clast-poor diamictite. (MB-17C). G) A boulder
and cobble bed in an otherwise massive diamictite (MBSE-17). H) Diamictite beds onlapping on to a large, 4 m boulder in the diamictite. Inset picture shows the whole, ~4-
m-diameter boulder in outcrop, with person for scale.
L.R.W. IVES AND J.L. ISBELL620 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
the massive diamictite is interstratified with turbidites, a subaqueous
process (see LS facies interpretation). Additionally, there are no
characteristics in this FA that suggest subaerial exposure (e.g., paleosols,
desiccation features, wind-transported sediments).
The lack of clear or laterally continuous stratification, as well as a wide
range of grain sizes in the matrix throughout this diamictite, indicate that
the deposition of these sediments was not principally controlled by
‘‘sorting’’ processes such as currents and/or low-density flows. The
massive nature of beds may be a primary depositional feature, as is the
case in subglacial deposits, supraglacial debris, or subaqueous plume
sedimentation. Alternatively, this facies may also be the result of
secondary processes, such as redeposition by mass-transport and high-
density gravity-driven processes–flows (Wright and Anderson 1982;
Vesely et al. 2018), or homogenization due to iceberg scour (Dowdeswell
et al. 1994), dewatering (Collinson and Mountney 2019), or bioturbation
(Svendsen and Mangerud 1997; Murray et al. 2013). Where crude
stratification does occur, there are indicators that settling from suspension
may have been a key depositional process. Ruck structures beneath large
clasts suggest those clasts were dropped into the surrounding diamictite
from ice-rafted debris (Fig. 5D) (Thomas and Connell 1985). The
thinning of diamictite beds over large clasts suggests that there was some
component of settling from suspension in their depositional process (Fig.
5H). In glaciomarine settings, this most often occurs as meltwater plume
sedimentation.
Small, deformed, sorted sediment bodies occur in many depositional
settings alongside diamictites, including in till (Kessler et al. 2012) and
proximal glaciomarine sediments (Domack 1983; Sheppard et al. 2000).
Depending on their depositional context, small sand bodies are proposed to
be sourced from iceberg dumps, winnowing due to dewatering, or from
incorporating subglacial sediment into till through freeze-on. The small
sand and gravel bodies all indicate pervasive simple shear (whisps,
stringers, boudinage, and sheath folds) or loading (diapir structures). In at
least one case, shearing was associated with an overlying, thicker
sandstone body (Fig. 5F). If the diamictite experienced similar strain
conditions, representative structures would likely be impossible to observe
in outcrop, due to the homogeneous natures of the diamictite. The large
sand bodies are most likely mass-transport deposits that underwent
slumping or non-turbulent flow (Posamentier and Martinsen 2011;
Rodrigues et al. 2019).
The combination of processes (subglacial till deposition, iceberg
rain-out, plume sedimentation, iceberg scouring, and mass transport)
that potentially contributed to the deposition of the massive diamictite
and related facies are most likely to occur in a glacier-proximal to
glacier-intermediate (not distal or proximal) setting, with water depths
below storm wave base (i.e., offshore–transition to offshore) (Licht et
al. 1999; Powell and Cooper 2002). In the Cenozoic, comparable
depositional models have been proposed for similar successions in the
Yakataga Fm, Alaska (Eyles and Lagoe 1990), Weddell Sea, Ross Sea,
and George V regions of Antarctica (Anderson et al. 1980; McKay et al.
2009), as well as St. George’s Bay, Newfoundland (Sheppard et al.
2000). Of these examples, St. George’s Bay is likely the most analogous
to the Shackleton Glacier region during the Permian, since it is not an
open-shelf setting, but an embayment whose topography is controlled
by much older basement rocks (Batterson and Sheppard 2000; Shaw
2016).
Plumes emitted from subglacial and englacial meltwater jets were
likely the primary sources of sediments in this system. Plume
sedimentation is not likely to occur where the glacier meltwater is
denser than the ambient water in the depositional environment, which is a
condition associated with lacustrine conditions and resulting hyperpycnal
flows. Therefore the deposition of this facies most likely occurred in
marine or estuarine conditions (Powell 1990). Variations in iceberg
calving, fluctuating glacial hydraulic systems, and minor movement of
the ice margin may explain the variation of matrix grain size and clast
abundance throughout the facies. Glacial hydraulic systems and icebergs
capable of producing sufficient sediment to create this FA are
characteristic of temperate to ‘‘mild’’ subpolar glaciers (Matsch and
Ojakangas 1991; Hambrey and Glasser 2012; Dowdeswell et al. 2016;
Kurjanski et al. 2020).
The time frame in which the deposition of this facies occurred is
difficult to infer, especially without any evidence in the facies for glacier
grounding above its lower contact. Sedimentation rates in glaciomarine
settings are highly variable, even for the same glacier, and strongly depend
on glacier conditions and proximity to the ice front (Hallet et al. 1996).
Rates of accumulation will also depend on physiography of the
depositional area (e.g., fjord vs. open shelf). Accumulating ~100 m of
glaciomarine diamictite could take anywhere from a few years (Cowan and
Powell 1991) to a few millennia (Partin and Sadler 2016; Domack and
Powell 2018). The lack of non-glaciogenic deposits (see LS and HS facies
descriptions for glacial interpretations) interstratified with the MSD facies
suggests that the deposition of this facies occurred on the shorter end of
this time scale.
Laminated Sands Facies Association (LS)
LS Description.—This is a sandstone FA that occurs only at site RS-18
and is interstratified with facies MSD. This succession is ~17 m thick and
laterally continuous across the outcrop. Its lower contact is erosional above
MSD, and its upper contact was covered. Internally, this FA consists of
fining-upward packages 3–5 m thick (Fig. 6). This FA consists of coarse-
to fine-grained, well-sorted sandstones. The dominant lithology in this FA
is fine- to medium-grained sands that occur in thin, planar-laminated beds
with primary current lineations. Coarse-grained sandstones at the bases of
some packages are trough cross-bedded. Fine- to very-fine grained
sandstones are laminated, or thin-bedded with unidirectional cross-laminae
and/or ripples. The sandstone is quartz-rich, and grains are subangular to
rounded. Pebbles up to 8 cm in diameter occur at the bases of some cross-
beds. The uppermost parts of fining-upward packages sometimes include
fine- to very-fine-grained black-colored sandstone.
LS Interpretation.—This FA is most likely the result of a series of
noncohesive density-flow events, in the form of high-density turbidites
and/or a transitional concentrated density flows (Mulder and Alexander
2001). Fining sequences such as these could also be formed in a fluvial or
shallow marine setting, where bedload-dominated currents are common.
However, a fluvial setting is inconsistent with the lack of channelization or
indication of surface exposure in the RSP-18 succession. Similarly, a
shoreface setting is unlikely because there is no evidence of emergence or
wave action in the succession. The current directions measured in this unit
are unidirectional, which is also not indicative of a shoreface. Since these
turbidites are interstratified with facies MSD, they are most likely the distal
or medial part of an ice-contact fan or delta (Lønne 1995; Dowdeswell et
al. 2015).
Heterogenous Sandy Facies Association (HS) HS Stratigraphy.—
This FA (Fig. 7) occurs at both Mt. Butters sites above a gradational
contact with the massive diamictite (MSD), and below a sharp, erosional
contact with the cross-bedded sands facies (CBS) at MBSE-17 and a sharp,
planar contact with the Mackellar Fm at MB-17 (Fig. 8). The lower part of
the facies association begins as interbedded, discontinuous bodies of
deformed, sorted sands and gravels (facies HS2) in stratified diamictites
(facies HS1). Undeformed, moderately sorted, stratified sandstone bodies
with a range of grain sizes (facies HS3) occur in the middle of the
succession and eventually become the dominant facies near the top of the
succession. This FA ranges in thickness from 1 to 15 m. Lithologies in this
facies are interstratified.
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 621
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
Facies HS1: Stratified diamictite
Description.—This facies is a clast-rich, sandy diamictite (Figs. 7D, G,
I, 8). This diamictite facies is similar to the massive diamictite facies
(MSD), but is consistently stratified, is more clast-rich, and the matrix is
better sorted. The matrix in this facies is moderately to well sorted. Matrix
grain sizes range from medium to very fine sand. The mean matrix grain
size varies between beds. Most beds have a mean matrix grain size of
medium sand, but some beds have a dominantly very fine grained sand
matrix. Clasts in this facies are angular to subrounded, and have a size
range similar to MSD, granule to 1 m boulders. The beds in this facies are
3–7 cm thick, are planar, laterally discontinuous, and have sharp contacts.
The distribution of clasts is random and unrelated to bedding planes.
Larger clasts often punctuate bedding planes. This facies has a gradational
lower contact overlying the massive diamictite (MSD) facies. Vertical and
lateral contacts with other facies in this FA (HS3 and HS2) are most often
sharp, sometimes loaded, deformed, or erosional. This bedded diamictite is
frequently interbedded with facies HS2. Strata in this facies adjacent to
contacts with facies HS2 often display soft-sediment deformation,
including load structures and sheared contacts.
Interpretation.—Similar to facies MSD, this facies was likely
deposited in a glacier-proximal glaciomarine setting, but was dominated
by plume sedimentation and iceberg rainout (Eyles 1987; Licht et al. 1999;
Powell and Domack 2002; McKay et al. 2009). The consistency of the
stratification in this facies compared to the MSD facies suggests that the
sediment in this facies was not as frequently subject to remobilization,
either through gravity-driven transport or iceberg scour. The better sorting
of the matrix in this facies relative to the massive diamictite facies indicate
that sediment-sorting processes were more active than during the
deposition of MSD. In a glacial-proximal setting, this likely means that
turbulence kept fine-grained sediment suspended in the water column and
did not allow it to settle out. The clast-rich composition and the thin nature
of the beds also suggest that depositional processes were relatively
constant, compared to the high variability of clast contents in facies MSD.
Loaded, deformed, and erosional contacts in this facies, and with other
facies in this FA, indicate that this facies was rheologically ‘‘ weak,’’ or
experienced ductile deformation at low strain magnitudes. Therefore,
during and shortly following deposition, facies HS1 was likely water-
saturated. This also suggests that that sedimentation was rapid, that this
facies was subject to gravity-driven processes, and that these sediments
were deformed by gravity-driven deposits (Figs. 7D, F, 8) (Visser 1994).
Facies HS2: Chaotic Sandstones
Description.—This facies consists of very fine- to coarse-grained
sandstones (Fig. 7A–C). Beds in this facies are laterally discontinuous
(Figs. 7G, 8). The thicknesses of sandstone bodies are laterally inconsistent
and range in thickness from 0.5 m to 2 m. Widths of sandstone bodies
range from ~1 m to outcrop scale. Sandstone beds may be interbedded
with one another, but are dominantly interbedded with the surrounding
either massive (MSD FA) of stratifi ed (HS1) diamictite facies. Sandstone
bodies are irregularly shaped, but generally have planar to lenticular
shapes. Lower and lateral contacts are deformed, sharp, or erosional, and
often show evidence of soft-sediment deformation (Fig. 7D, F). Upper
contacts are sharp and conformable. Contacts between sandstone bodies
are erosional or deformed.
Beds in this facies are often internally massive, and soft-sediment
deformation is pervasive (Fig. 7E, F). Primary sedimentary structures are
sometimes preserved, but this is rare and only occurs in a small area of any
given bed. Secondary structures in this facies include fold noses (Fig. 7E),
boudins, faults, and other simple shear structures above and below contacts
(Fig. 7D–G), and ruck structures associated with rare outsized clasts (Fig.
7I). Grain size in this facies ranges from conglomerate to very-fi ne-grained
sandstone. Fine- to medium-grained sandstones tend to be well sorted,
while coarse-grained sandstones and conglomerates are poorly sorted. The
medium- and coarse-grained sandstones occur more frequently than finer-
grained lithologies.
Interpretation.—The sandstone bodies in this facies are most likely the
result of mass-transport, gravity-driven processes (Posamentier and
Martinsen 2011; Sobiesiak et al. 2018; Rodrigues et al. 2019). Preserved
primary sediment structures in some of these bodies indicate that sediment
sorting due to current transport likely occurred before the remobilization
and final deposition of these sediments. Irregular lateral contacts between
this facies and the two diamictites facies (HS1 and MSD) indicate that
mass-transported bodies were sandstone rafts deposited into pre-existing
diamictites by gravity driven processes. The deposition of mass-transport
FIG. 6.—Photograph of the Laminated Sands
(LS) facies association at site RS-18 (Reid Spur).
Triangles show fining-upward packages separated
by erosional surfaces. Reddish lithologies are
medium-grained sandstone to coarse-grained
sandstone, and have planar lamination and low-
angle cross-stratification. Black lithologies are
fine-grained sandstones.
L.R.W. IVES AND J.L. ISBELL622 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
deposits (MTDs) into the diamictite indicates that diamictite deposition
was contemporaneous with MTD emplacement. This also suggests that the
diamictite was weak (highly susceptible to deformation), likely due to both
high pore-water pressures and a lack of consolidation.
Facies HS3: Stratified Sandstones
Description.—This facies consists of these facies very fine- to coarse-
grained sandstones (Fig. 7A–C). Medium- to coarse-grained sandstones
are thickly laminated to bedded. Common sedimentary structures in
medium- to coarse-grained sandstones include planar cross beds, trough
cross beds, climbing ripples, and 3D ripples that are asymmetric or
climbing (Fig. 7A). Rare sedimentary structures include hummocky and
swaly cross stratification and symmetrical ripples with bundled upbuilding
(Fig. 7B, C). Lateral and vertical variations in sedimentary structures
within a unit of the same lithology or grain-size is common. Coarser sands
are occasionally massive or contain laterally discontinuous sand or gravel
lenses. Trough cross-beds occasionally have pebbles at the bases of
FIG. 7.—Photographs of the Heterogenous Sandy (HS) facies association at Mt. Butters in section MB-17 and MBSE-17. Marks on all rulers are in cm. Rulers are50cm
when folded in half and 1 m long when unfolded. White dashed lines have been used to highlight contacts and important bedding surfaces. A) Medium-grained sandstone
layer in facies HS3. The sandstone body is composed mostly of climbing dunes (cross beds) and capped by asymmetrical ripples. Note sharp contact with black-colored,
bedded diamictite (facies HS1) in lower part of image (MB-17). B) Fine- to medium-grained sandstone in facies HS3 with up-building symmetrical ripples (MB-17). C)
Swaly and hummocky cross-stratification in sandstone layer of facies HS3 (MB-17). D) Loaded, possibly boudinage, contact between a massive sandstone (HS2) and bedded
diamictite (HS3) (MB-17). E) Sandstone with a soft-sediment recumbent fold nose in facies HS2, likely at the front of a slump (MB-17). F) Deformed contact between facies
HS2 and HS3. Jacob’s staff for scale; marking is every 10 cm. (MB-17). G) Outcrop showing interfi ngering between facies in this facies association. Black-colored sediments
are bedded diamictites (facies HS1), other lithologies show sand and gravel of facies HS2. (MB-17). H) Groove structures on top of facies HS1. View approximately toward
north (down the Shackleton Glacier). I) Granitic outsized clast punctuating interlaminated sandstone beds in the bedded diamictite facies HS1 (MB-17).
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 623
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 8.—Photo mosaics (A, B, and C) and interpretive sketches (B0and C0) of the upper part of section MBSE-17 at Mt. Butters. Mosaics A and B were taken from a
helicopter, and mosaic C from the ground. This is the opposite side of the outcrop shown in Figure 4C. View is to the north, and the ridge runs roughly east–west. Part A
shows lateral variations in the architecture of the cross-bedded sandstone (CBS) facies. Figures B and B0highlight the stratigraphic relationships between the Massive
Diamictite (MSD) facies association, the Heterogenous Sandy (HS) facies association, and the CBS facies association. Black lines in CBS denote channel erosional surfaces.
Black lines in HS3 indicate soft-sediment deformation. The part of this outcrop highlighted by C and C 0is located within the red box on Part B. Part C shows a part of the
outcrop that is characteristic of the HS facies association, and was selected to illustrate the pervasive nature of soft-sediment deformation in this facies association.
L.R.W. IVES AND J.L. ISBELL624 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
troughs. Very fi ne- and fine-grained sandstones are laminated or thin-
bedded. Sedimentary structures in fine-grained and very fine-grained
sandstones include unidirectional cross-laminae, ripples, planar and wavy
laminae, some flaser-bedded cross-laminated units, climbing-ripple
laminae, and rare outsized clasts with ruck structures. Lithologies of all
grain sizes also have minor amounts of soft-sediment deformation,
including dewatering structures, minor folds, and loading.
At the upper contact of the Heterogenous Sandy FA with the Mackellar
Fm in section MB-17, these sandstones had at least five large, shallow,
east–west-oriented grooves (Table 2) that occur in a massive, well-sorted
sandstone (Fig. 7H). All grooves were ~1–2 m wide and 10 cm deep.
Berms ~10–20 cm high bound the grooves on their long sides. All of the
grooves gradually shallow towards the west (2158), and two of the grooves
had prow-like berms at their eastern terminus.
Strata in this facies are laterally continuous across the outcrop.
Sandstone bodies are generally wedge-shaped and thicken in the direction
of flow. Erosional surfaces are common in the facies. This facies has an
erosional lower contact with both diamictite facies (MSD and HS1), and a
sharp, conformable contact with the overlying UFG facies association.
Interpretation.—This facies was likely deposited rapidly in an
unconfined, subaqueous setting with limited reworking by waves and
icebergs. The sedimentary structures in this facies indicate that current-
dominated transport and deposition occurred, followed by syndepositional
or postdepositional slumping of some deposits. The wide range of grain
sizes and sedimentary structures indicate a sediment source with a wide
range of grain sizes and huge variations in current velocities during
deposition. Common sedimentary structures, such as planar cross beds,
trough cross beds, and asymmetrical ripples, suggest unidirectional,
relatively high-velocity currents. Climbing ripples suggest decrease in flow
velocity downcurrent, which is characteristic of unconfined flows, which is
supported by the lack of channelized deposits. Flaser bedding, as well as
abrupt changes in grain size and sedimentary structures within and
between beds, indicates that current velocities were highly variable and
fluctuated. Rare soft-sediment deformation suggests high pore-water
pressure during rapid deposition. Rare hummocky and swaly cross-
stratification and symmetrical ripples form under oscillatory flow
conditions created by surface waves, possibly during reworking by storms
(Reineck and Singh 1980; Dumas and Arnott 2006; Collinson and
Mountney 2019). Their rare occurrence and interstratification with fine-
grained sediments suggest that these wave features formed below normal
wave base. The shape of the grooves on the upper contact of this FA at site
MB-17, their position in the upper contact of a massive (homogenized)
sandstone, and immediately below the contact with the dropstone-bearing
lower Mackellar Fm suggest that these features are iceberg keel marks
(Dowdeswell et al. 1994; Vesely and Assine 2014). These massive
sandstones were likely deposited in a similar way to other sandstones in
this facies, but were homogenized by iceberg actions.
HS Depositional Environment.—The three facies in this FA represent
a complex depositional environment that is characteristic of subaqueous,
glacier-intermediate to -proximal settings in front of the terminus of
temperate to ‘‘mild’’ subpolar glaciers. Evidence for the glaciogenic origin
of this FA includes pebbles with ruck structures (representing ice-rafted
debris), iceberg keel marks (facies HS3), the very poor sorting of sediment
in the system, and the wide range of sedimentary grain shapes in the
Pagoda Fm sandstones, which are described at the beginning of the facies
section.
The stratified diamictite (facies HS1) was deposited primarily through
plume sedimentation in a glacier-proximal setting, and is the dominant, or
‘‘background,’’ sedimentation type in this FA. The gradational contact
separating MSD (massive diamictite) and facies HS1 suggests a gradual
shift in depositional environments between the two. Sediment composition
is consistent between the two diamictite facies, suggesting that the
sediment source did not change, but that the depositional environment
shifted from glacier-intermediate to glacier-proximal. The deposition of
both diamictites was likely controlled by the same processes (i.e., plume
sedimentation, iceberg rain-out, iceberg scouring, and mass transport) but
to different degrees. This shift from glacier-intermediate to glacier-
proximal was most likely driven by a minor readvance of the glacier
margin, but it may have also been an apparent effect caused by the
progradation of the overlying grounding-line fan system (HS2 and HS3).
The sandstone facies in this FA (facies HS2 and HS3) most likely
represent the medial part of a subaqueous grounding-line fan(s) system
(Powell 1990; Lønne 1995; Dowdeswell et al. 2015). In facies HS3, the
high-velocity, unidirectional current transport combined with abrupt
changes in grain size (i.e. current velocity), unconfined flow, and
interstratification with the bedded diamictite (facies HS1) are characteristic
of grounding-line fans (Powell 1991). The gravity-driven transport of
facies HS2 sandstone bodies were likely derived from deposits similar to
(or the same as) facies HS3. ‘‘Shedding’’ of sediments is characteristic of
the rapid sedimentation in grounding-line fan systems (Benn 1996; Powell
and Alley 1997; Lønne et al. 2001). Intense, ductile deformation and
loading along contacts throughout this FA indicate that all facies were
water-saturated, unconsolidated, and generally had the consistency of soup,
suggesting rapid deposition and that they were therefore prone to
resedimentation (Fig. 8B).
The wave reworking of some sandstone beds, as indicated by
hummocky and swaly cross-stratification and wave-ripple stratification
(symmetrical ripples) in facies HS3 suggest that this depositional
environment was occasionally subjected to surface-wave activity (below
normal wave base). These features indicate that there was not perennial ice
cover during the deposition of this FA. This evidence for wave reworking
suggests a similar, wave-winnowing origin for the boulder beds in facies
MSD.
Where this FA is well developed in outcrops at Mt. Butters (sites MB-17
and MBSE-17), the succession has a general coarsening and increase in
sorting trends upward. This trend indicates the increase in the proximity of
the energy and sediment source, either through progradation of the
grounding-line system and/or advance of the glacial front.
Cross Bedded Sandstone Facies (CBS)
CBS Description.—This facies occurs at Mt. Butters section MBSE-17,
and consists of an erosionally-based, laterally extensive, channel-form
sandstone body 10–30 m thick and several hundred meters wide that cuts
into and through a laterally continuous thick sandstone sheet at the top of a
coarsening-upward succession of the HS (1–3) facies association (Figs. 8,
9). The sandstone body is laterally continuous across outcrop MBSE-17
but is not present at section MB-17, which is ~2 km north (Fig. 1B). The
basal CBS erosional surface has a relief of up to 10 m, and lower contacts
with all HS facies and the MSD facies (Fig. 10). The upper contact of this
facies with the overlying Mackellar Fm is sharp and horizontal.
This facies occurs in multistoried, multilateral sand-filled channel-form
bodies displaying nonsequential, lateral compensational stacking patterns
(Fig. 9). Individual channels are meter-scale thick and tens of meters wide,
trough-shaped in cross section, and are filled by either vertical or
downstream accretion dipping to the east. Channels are truncated by the
bases of overlying channel bodies. Channel stacking is nonsequential and
disorganized, with some aggradation. This facies is composed of well-
sorted, medium- to very coarse-grained quartz sandstone, with minor
occurrences of conglomerate lenses and beds (Fig. 9). Mudrocks were not
observed in the sandstone bodies. Very rare pebble- and small-cobble-size
clasts occur throughout the sandstones. Those clasts have lithologies
similar to clasts observed in both diamictite facies (HS1 and MSD).
Sedimentary structures are almost exclusively 0.15–1.5-m-thick sets of
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 625
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
low-angle stratification and trough cross-beds (Fig. 9A). Thin beds with
asymmetrical ripples also occur but are rare.
Adjacent to the described section (MBSE-17), the edge of the channel-
form body appears to extend across the top of the underlying strata as a
wing-like extension (Fig. 8A). Channel-form sandstone bodies occur in the
wing. These bodies appear to transition laterally into other thick sandstones
with channelized bases. Most notably, the contact between the channelized
sandstones in the wing and the underlying HS facies appears to be sharp
CBS Interpretation.—The CBS facies in the Pagoda Fm was deposited
by strong, tractive, confined flow as indicated by the occurrence of the
basal erosion surface and the internal channel bodies filled by downstream-
accreting bar forms and cross-stratification. The trough shape of the
internal sandstone-filled channels and their multistoried and multilateral
characteristics suggest that the channels were stationary during flow in the
channels and that they did not migrate until channel switching occurred
and new channels formed as older channels filled and were abandoned
(Friend 1983). The occurrence of low-angle stratification and meter-scale
trough cross beds organized into downstream-accreting bodies with
massive bedded to lenticular gravels suggest high flow velocities. Such
features are characteristic of highly dynamic systems where aggradation in
channels likely forced channel switching to adjacent areas on the
depositional surface. The occurrence of this facies association on top of
unconfined coarsening-upward HS (1–3) facies association and the
occurrence of wings that appear as a continuation of the HS (1–3)
coarsening-upward succession suggest that the CBS facies formed as part
of a HS–CBS larger-scale dispersal system. The presence of wings also
suggests that parts of the CBS system were unconfined and represent
‘‘overbank’’ deposition on surfaces in areas between channels. Together,
these patterns are most characteristic of an unconfined, distributive setting
(Funk et al. 2012). This unit is similar to some grounding-line fan systems
that authors have called subaqueous outwash fans (Visser et al. 1987;
Thomas and Chiverrell 2006; Rose 2018).
Whether this facies was deposited subaerially or subaqueously is
unclear. However, the CBS sandstone body does not contain evidence for
shallow-water wave reworking, pedogenesis, or subaerial exposure,
whereas facies both below and above this facies represent subaqueous
deposition below normal wave base, and likely below storm wave base.
Therefore, a subaqueous setting seems likely.
DEPOSITIONAL MODEL
Most of the Pagoda Fm in the Shackleton Glacier Area is composed of
massive diamictites (facies MSD) that likely formed in glacier-proximal to
glacier-intermediate environments, at depths largely below normal wave
base, through a variety of glaciogenic and glacially influenced processes.
These massive diamictites are also conformably overlain by, and
interstratified with, grounding-line fan deposits (facies associations LS,
HS, and CBS). These glacially derived lithologies are conformably succeed
by prodeltaic, fine-grained facies of the Mackellar Fm.
Evidence for the grounded advance of a glacier(s) in the Shackleton
Glacier region is present at base of the MSD facies at both Mt. Butters sites
(MB-17 and MBSE-17) and at Mt. Munson (MM-17). The lower contact
of the MSD facies is not exposed at Reid Spur, so similar inferences cannot
be made for that locality. No conclusive evidence for subglacial
deformation or erosion was observed higher in the Pagoda Fm at any
site examined, though a subglacial origin for the MSD facies cannot be
wholly ruled out. This advance was likely made by a glacier whose
thickness exceed 100 m and flowed from north to south across the
Shackleton Glacier region. The advance would have come from the
direction of the present Ross Sea and crossing the TAB’s margins
perpendicular to the elongate trend of TAB (See discussion in prior
section; Fig. 11D). This observation, along with other data from the TAB,
strongly suggest that there were at least two ice centers in Antarctica during
the Permian, one located on the East Antarctica craton and one in present
day West Antarctica (Isbell 2010).
When the glacier margin retreated from the Mt. Butters, Mt. Munson,
and Reid Spur sites, the deposition of glacier-proximal deposits was
FIG. 9.—Photographs of the Cross Bedded Sandstone (CBS) facies at section
MBSE-17 on Mt. Butters. A) An example of low-angle and trough cross-bed sets in
this facies. Measuring stick is 1 m long. B) Photograph of facies in outcrop, noting
occurrences of minor lithologies.
L.R.W. IVES AND J.L. ISBELL626 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 10.—Sedimentary logs and paleotransport directions from sites described in this study. Paleotransport directions are plotted to align with the maps of modern
Transantarctic Mountains presented in this study, so that north is toward the bottom of the page and varies by geographic position (see inset map). Details of paleotransport
measurements are available in Table 2. MM-17 is Mt. Munson, MB-17(A–C) is Mt. Butters 1, MBSE-17 is Mt. Butters 2, and RS-18 is Reid Spur. Lithologies are grouped by
their interpreted facies or facies association. Colored bars next to each log are used to indicate the distribution of facies associations in these sections, and colored areas in
between sections show interpretation of each facies extent outside the section. Dark green corresponds to the localized lacustrine facies of the Pagoda Fm at Mt. Butters
described by Isbell et al. (2001). Bright green represents the Massive Sandy Diamictite (MSD) facies association. Different shades of purple represent grouping of the
Heterogenous Sandy facies association (HS1, stratified diamictite; HS2, chaotic sandstones; HS3, stratified sandstones). The inset map shows the location of each section and
isopachs of the Pagoda Fm in the Beardmore Sub-basin from Isbell et al. (2008c). Light blue represents the Cross Bedded Sandstone facies (CBS). Dark blue represents the
Laminated Sandstone facies association (LS). Red represents the Mackellar Fm. The datum for these columns were chosen using the last evidence for glaciogenic sediments,
either the uppermost outsized clast or diamictite, which is a marker that also serves as an upper sequence boundary.
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 627
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
FIG. 11.—Box diagrams showing progressive phases of the depositional model for the Pagoda Fm and lowermost, glacially influenced Mackellar Fm in the Shackleton Glacier region, alongside maps showing the modern
locations of the sites described in this study with transport orientations related to each part of the depositional model. See Table 3 for flow directions. A) Map of the modern central Transantarctic Mountains. Gray areas are
approximate locations of nunatuk regions. Solid black lines are isopachs of the Pagoda Fm, copied from Isbell et al. (2008c). Red square indicates map area in Parts B–F. Note that the smaller map areas are rotated relative to
the larger ‘‘map A.’’ Yellow stars show site locations described in this study. B) Paleotopography of the Shackleton Glacier Area before deposition of the Pagoda Fm. Map shows strike and dip of granite surface underlying
the Pagoda Fm at Mt. Butters site MB-17, corrected for modern structural conditions. This map also includes the Isbell et al. (2008c) isopach lines and derived basin-axis orientation for reference. C) Proposed depositional
conditions for the lacustrine facies at the base of the Pagoda Fm at site MB-17 (Isbell et al. 2001). The green, double-sided arrow shown on the map shows the orientation of symmetrical (wave) ripple crests, which parallels
the strike of the underlying basement. D) Proposed depositional conditions for Massive Sandy Diamictite (MSD) facies and Laminated Sands (LS) facies during retreat of the glacier out of the Shackleton Glacier area. On
the map, purple wedge indicates range of flow direction in the LS facies, and green wedges indicate down-slope transport direction in the MSD facies. Blue double-headed arrows show glacier flow directions measured in
this study. E) Proposed conditions during the deposition of the grounding-line fan represented by facies associations MSD, Heterogenous Sandy (HS), Cross-bedded Sands (CBS), and Mackellar Fm. The red wedge shows
range of flow directions in the Mackellar Fm, blue and purple show the range of flow direction for the CBS and HS flow directions, respectively, and the blue arrow indicates the mean flow direction of the CBS facies at site
MBSE-17. F) Proposed conditions during the deposition of the lower Mackellar Fm.
L.R.W. IVES AND J.L. ISBELL628 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
initiated. The massive diamictite (MSD) facies that dominated the Pagoda
Fm at Mt. Butters and Reid Spur (RSP-18) was likely deposited through a
combination of glaciogenic depositional processes characteristic of (sensu
lato) glaciomarine settings, a combination of settling from suspension of
neritic sediments, plume sedimentation from subglacial and englacial jets,
as well as iceberg sedimentation and mixing. The prevalence of plume
sedimentation and lack of hyperpycnal (under-) flow indicate that the water
in the TAB was likely either marine or brackish. It should be noted that the
TAB did not have a shelf, in the formal sense, as it was a trough-shaped
basin that was not directly connected to the open ocean.
The Shackleton Glacier region during the Cisuralian was not an open
shelf, but a near-coast setting with ample topographic relief, and water
depths below normal wave base, such as St. George’s Bay, Newfoundland
(Sheppard et al. 2000). Ultimately, many of these sediments were likely
remobilized by gravity-driven slides, slumps, and flows. This is especially
true at the Mt. Butters site, where the vergence of soft-sediment
deformation features in the MSD facies follow the local paleotopographic
slope (Fig. 11B, E). Soft-sediment deformation caused by mass-transport
deposits (facies HS) in bedded diamictites at Mt. Butters shows that the
diamictites were unconsolidated, water-saturated, and subject to resedi-
mentation, suggesting relatively rapid deposition.
At Mt. Butters and Reid Spur, massive diamictites are interstratified
with and overlain by grounding-line fan deposits in the form of the LS, HS,
and CBS facies associations. The stacked density flows (facies LS) at Reid
Spur represent the medial to distal part of a fan, likely in relatively deep
water. Flow direction in these sediments are towards the basin axis (Fig.
11D), suggesting that the basin geometry controlled topography in this
area. On the other hand, the fan deposits at Mt. Butters are more proximal
to the glacier margin. The laminated diamictites, reactivation surfaces, and
storm-wave deposits in the fan at the top of the Mt. Butters section (MB-17
and MBSE-17) suggest that this fan was built gradually along a relatively
stable margin and not during a single, catastrophic drainage event
(Dowdeswell et al. 2016). The successions at Mt. Butters from the HS
facies association through the CBS facies association represents either the
progradation of one of these fans, a minor readvance associated with the
deposition of the fan, or a combination of the two (Fig. 11E). The
dispersed flow directions in the HS facies are characteristic of a fan, though
their orientation broadly toward the north is away from the basin axis and
opposed to glacier flow directions. Flow directions in the CBS facies are
well clustered toward the southwest (Table 2, Fig. 11E). Influence of the
paleotopographic slope at Mt. Butters and the radial nature of fan
geometries are the most likely cause from these seemingly antagonistic
flow directions (Fig. 11E). The orientation of the ice front would have also
likely been influenced by this topography.
STRATIGRAPHIC FRAMEWORK
Basin-Margin vs. Basinal Facies Associations
Isbell et al. (2008c) described two generalized facies association of the
Pagoda Fm: one that occurs in basinal settings and another that occurs
along the basin margins. The sites described in this study from Mt.
Munson and Mt. Butters in the Pagoda Fm are characteristic of the Basin
Margin FA, because they are relatively thin successions (,100 m) and
have evidence for subglacial erosion in the form of polished and striated
bedrock (MBSE-17 and MM-17), and subglacially sheared lacustrine
sediments (MB-17; Isbell et al. 2001). The site at Reid Spur (RSP-18) is
also most characteristic of the Basin Margin FA, because diamictite facies
there are thick and unstratified and has a poorly sorted matrix, suggesting
plume sedimentation and gravity-driven redeposition. However, this
section also likely represents a transition between the two facies
associations. This is indicated by the LS facies at this site, which is
attributable to the mid- to distal part of a grounding-line fan, and by the
inferred thickness of the Pagoda Fm at Reid Spur (~100 m), which is the
same as the transition thickness between Isbell et al. (2008c)’s two FAs.
Effects of Topography
The transport directions in these successions strongly suggest that
paleotopography (relief on the Maya Erosional Surface) played a
significant role in the deposition of the Pagoda in the Shackleton Glacier
region. The influence of topography is particularly evident in the section
MB-17 (Fig. 11). The surface of the basement at MB-17 dips toward the
west at 118(Figs. 4, 11B; Appendix A). Wave-ripple crests in facies BFG
are parallel to the strike of the basement surface at MB-17 (Fig. 11C),
suggesting that the paleotopography created by this surface was sufficient
to affect and orient wave action. The transport directions of slumping and
other gravity-driven processes in facies MSD at both sites MB-17 and
MBSE-17 are also generally towards the west (Figs. 10, 11D), following
the same slope. Flow directions in the grounding-line fan facies
associations (HS and CBS) have a wide spread ranging from the southwest
toward the east. However, in the HS facies association at MBSE-17,
gravity-driven transport is still towards the west (Figs. 10, 11E). The south-
southeast flow directions in the turbidite facies (LS) at site RSP-18 do not
align with the flow direction at Mt. Butters, suggesting that those facies
have a separate origin than the Mt. Butter’s grounding line fan(s) (Figs. 10,
11D).
Paleotopographic control on the deposition of Pagoda Fm and Mackellar
Fm has been noted by authors throughout the TAB (Lindsay 1970b; Barrett
1972; Isbell et al. 1997a, 2008c; Cornamusini et al. 2017). In previous
studies, ice-flow directions (usually striae on bedrock or clast pavements)
have often been combined with other transport directions in the Pagoda Fm
to infer a generalized transport direction. However, recent work in modern,
high-relief, glaciated landscapes has shown that the relationship between
glacier flow directions, other transport directions, and topography can be
used to infer the thickness of the glacier relative to the magnitude of relief
on the landscape (Landvik et al. 2014). In other words, whether or not a
glacier ‘‘follows’’ the underlying topography is a function of the glacier’s
thickness. Therefore, indicators of glacier flow should be considered
separately from other transport directions.
The ice-flow directions below the Pagoda Fm in the Shackleton Glacier
region are oriented generally northwest to southeast at Mt. Butters and Mt.
Munson (Fig. 11D). Though none of the striae observed during this study
had unidirectional indicators, previous workers in this area have found
glacially carved features in the basement underlying the Pagoda Fm that
show glacier flow was basin-ward (toward the south) or along the basin
axis (toward the southeast) (Appendix A). These uniform flow directions
on both a paleotopographic high (MM-17) and paleotopographic low (MB-
17 and MBSE-17) suggest that the glacier, when it created these striae, was
sufficiently thick to ‘‘ overtop’’ the pre-existing topography in the
Shackleton Glacier Area. Based on the difference in Pagoda Fm thickness,
the local relief between site MB-17 (Mt. Butters) and MM-17 (Mt.
Munson) was at least 85 m, and the onlapping of the Mackellar Fm onto
basement in this area suggests that localized relief may have exceeded 100
m (Seegers 1996; Isbell et al. 1997a; Seegers-Szablewski and Isbell 1998).
This scale of relief is on the scale of large hills. The topographic
prominence of subglacial features on the scale of 100 m is considered
negligible in studies of modern ice-sheet margins (e.g., Lindb ¨ack and
Pettersson 2015; Cooper et al. 2019), but would likely perturb or redirect
the flow of relatively thin glaciers.
This discussion is all to say that the thickness of the glacier that created
these striae more likely than not greatly exceeded the thickness of local
topographic relief (~100 m), and that flow was most likely toward the
center of the TAB. This inference suggests that the glacier was more likely
an ice cap or ice sheet than an alpine glacier.
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 629
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
Glacial Systems Tracts
In this study, the sequence stratigraphy of the Pagoda Fm in the
Shackleton Glacier region can be considered only at Mt. Butters, because
that is the only location where the authors were able to measure complete
sections. Since the Mt. Butters locations are basin-marginal successions
(Fig. 11), this analysis of glacial sequence stratigraphy should not be
considered applicable to basinal successions of the Pagoda Fm. The
Pagoda Fm at Mt. Butters is unique in the TAB because it contains only a
single glacial sequence as defined by Powell and Cooper (2002) and
Rosenblume and Powell (2019) (Fig. 12). The sequence described in this
paper is bounded at its base by a surface of glacial erosion (defined by
striae on bedrock and the deformation of underlying lacustrine sediments
(see Isbell et al. 2001) and at its top by an iceberg termination surface
(defined by the final outsized clast in the section). Since the Pagoda Fm
in this location is overlain by the nonglacial Mackellar Fm, there is no
true maximum retreat surface beyond the iceberg termination surface
(i.e., the last dropstone in the lower Mackellar Fm). Most of the
succession likely represents a glacial-retreat systems tract, though there is
likely some fraction of the massive diamictite facies above the erosional
surface that is more likely to have been subglacially deposited and would
therefore represent a glacial-maximum systems tract. The transition
FIG. 12.—Glacial sequence stratigrapahy of the
Mt. Butters sections, after Powell and Cooper
(2002) and Rosenblume and Powell (2019). The
depositional systems are defined as N, nonglacial;
D, glacier distal; P, glacier proximal; and I, ice
contact.
L.R.W. IVES AND J.L. ISBELL630 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
between those two systems tracts would be defined by a grounding-line
retreat surface.
This sequence is most consistent with Rosenblume and Powell (2019)’s
Type I ‘‘idealized glacial sequence,’’ which is a model developed to reflect
a sedimentation sequence deposited during the retreat of a relatively warm
subpolar glacier with a glacial erosion surface as a lower sequence
boundary and sufficient meltwater for the development of grounding-line
fans. This sequence model was developed based on upper Miocene
sediments from the Ross Sea region of Antarctica (Rosenblume and Powell
2019), whose climatic and geology were likely reasonably similar to the
TAB during the Permian. The Type I idealized sequence is interpreted to
represent a dynamic climatic glacial system with very high erosion rates
and debris fluxes, which is consistent with the depositional model
presented for the Pagoda Fm in this paper.
DISCUSSION
This study finds that most of sediments in the Pagoda Fm in the
Shackleton Glacier region were deposited during the retreat of a temperate
to ‘‘mild’’ subpolar glacier. The key indicators for this retreat include the
presence of grounding-line fan systems, ample evidence for plume
sedimentation and rapid deposition, as well as abundant glacially
transported clasts with a wide size range made of local basement
lithologies. This glacier had a subaqueous terminus during the deposition
of the Pagoda Fm in the Shackleton Glacier area, either in a marine or a
brackish setting.
One of the ultimate goals of the study of LPIA glaciogenic sediments is
to infer the ‘‘type’’ and distribution of glaciation experienced in any given
basin. Glacier ‘‘type’’ typically refers to the glacier’s thermal regime and its
size (i.e., ice sheet, ice cap, or ice field). Such characteristics of glaciers are
controlled by many factors but generally tie back into climate and geologic
setting. Glaciogenic sedimentary deposits are often used to infer the
thermal regime of their parent glacier (Dowdeswell et al. 2016; Kurjanski
et al. 2020). Recent studies all agree that glaciogenic deposits in the TAB
are most likely the result of transport and deposition by temperate or
‘‘mild’’ subpolar glaciers, largely because grounding-line fans are common
in the Pagoda Fm and its equivalents, which suggest that the TAB glaciers
released an abundance of meltwater (Isbell et al. 2008c; Isbell 2010; Koch
and Isbell 2013; Cornamusini et al. 2017).
The presence of subaqueous fan deposits interspersed in the massive
diamictite facies of the Pagoda Fm in the Shackleton Glacier area suggests
that the glacier had an active, persistent, and organized subglacial
hydrologic system, and that during its retreat the glacial margin was
stationary for at least a few years at a time to create grounding-line fans
(Cowan and Powell 1991; Hunter et al. 1996; Dowdeswell et al. 2015,
2016). These inferences are also supported by the succession’s stratigraphy,
which is characteristic of a ‘‘ Type I’’ (‘‘mild’’ subpolar glacier with
grounding-line-fan development) glacial systems tract (Rosenblume and
Powell 2019). Whether the glacier responsible for the deposition of the
Pagoda Fm in the Shackleton Glacier Region had more of a ‘‘ mild
subpolar’’ thermal regime, similar to modern glaciers in eastern Svalbard,
or more of a truly ‘‘temperate’’ thermal regime, like modern glaciers of
southern Alaska, is difficult to discern. To clarify likely thermal regime,
additional glaciogenic successions and nonglacial paleoclimate indicators
in the TAB should be examined. In either case, glaciers with mild thermal
regimes and developed subglacial hydrological systems are far and away
the more prolific producers and transporters of sediments, and that the
sedimentary record is therefore biased towards them. The presence of
deposits from temperate or mild subpolar glaciers does not mean the
thermal regime of the glaciers were never cold-based or that there was not
lateral variation in glacier thermal regime.
The unidirectional orientation of subglacial flow indicators, along with
evidence to topographic relief exceeding ~100 m in the Shackleton
Glacier region, suggest that glacier thickness also exceeded 100 m during
its maximum. Such observations do not allow an inference of maximum
possible ice thickness. However, we can infer that the glacier was more
likely part of an ice sheet or ice cap than an ice field or alpine glacier. The
glacier also may have thinned substantially during its retreat and
subsequent deposition of the Pagoda Fm in this area.
Throughout the TAB, Asselian–Sakmarian glaciogenic depositional
environments were locally and regionally variable in part due to the
inherent complexity of glacial processes and preservation potential, but
also due to the topography of the pre-existing landscape. Most studies of
glaciogenic rocks in the TAB note up to several hundred meters of
topographic relief on the underlying basement that is not wholly ‘‘filled’’
by glacial sediments and continued to influence postglacial sediment
deposition (Isbell et al. 1997a). Studies of both modern (e.g., Lawson
(1979)) and ancient (e.g., Cornamusini et al. (2017)) glacial sedimentary
systems have observed how topographic effects can result in seemingly
contradictory flow directions. This study is another example. Transport
directions at all of the sites in this study appear contradictory, unless
paleotopography and processes that created each feature are considered
(Fig. 11). For example, at Mt. Butters the flow directions of Permian
glaciers in the TAB appears to be from north to south, while the transport
directions of gravity-driven deposits and flow directions in the grounding-
line fan are perpendicular to that (Fig. 11D, E). If averaged together, these
flow directions would imply general transport toward the southwest, when
the most likely scenario was that the mass- and current-transported deposits
followed a paleotopographic slope and the glacier did not.
The Pagoda Fm in the Shackleton Glacier region likely represents a
single glacial–interglacial cycle, stratigraphically represented by a single
glacial sequence. In this context, the phrase ‘‘glacial–interglacial cycle’’
refers to the advance of a glacier into the basin and its whole retreat out
of the basin, which is stratigraphically defined by a surface of glacial
retreat. This is not to say that the position of the glacier margin did not
fluctuate during that cycle, but that only one grounded erosional surface
is present in the study area and only one surface of glacier retreat (Powell
and Cooper 2002; Rosenblume and Powell 2019). No instances of a
grounded readvance over any of the sections examined in this study were
observed beyond the basal erosional surfaces. As previously discussed,
the processes and environments responsible for the extensive deposition
of the massive, glaciogenic diamictite facies were likely diverse and
largely subaqueous. The preservation of such sediments is most probable
if they were deposited during the final retreat phase with no subsequent
advances over the area (Kurjanski et al. 2020). This is especially true in a
basin-marginal position in a basin like the TAB that was trough-shaped
and not exposed to open-marine conditions. Evidence for glacier
readvance above the basal erosional surface of the Pagoda Fm does
exist at other Pagoda Fm outcrops in the TAB, including in the
Beardmore Sub-basin (Lindsay 1970a; Miller 1989; Koch and Isbell
2013). This evidence typically occurs in thicker, basinal facies
association which are likely areas with higher accommodation than the
basin margins.
The flow directions and facies analyses from this study strongly support
the hypothesis that an ice center was positioned inboard of the
Panthalassan margin of Antarctica (an area that is now Marie Byrd Land)
during the Asselian–Sakmarian (Fig. 2B, ice center ‘‘q’’ ). The presence of
such an ice sheet is a relatively new hypothesis that was first proposed by
Isbell (2010) and Isbell et al. (2008c) based on transport directions in
South Victoria Land. In recent publications, this proposed ice center on the
Panthalassan side of the TAB has been inconsistently included (Fielding et
al. 2008c; Isbell et al. 2012; Monta˜
nez and Poulsen 2013) and excluded
(Fielding et al. 2010; Craddock et al. 2019) from LPIA ice-center
reconstructions. Evidence from this study and Isbell (2010) shows that an
ice center should be included on the non-cratonic edge of the TAB in
reconstructions including the Gzhelian–Sakmarian, an interval also
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 631
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
referred to as ‘‘Event 5’’ (L ´opez-Gamund´ı et al. 2021), and Australian ‘‘P1’’
(Fielding et al. 2008b). Ice-flow directions elsewhere in the TAB clearly
indicate that glaciers also flowed into the TAB off of the East Antarctic
Craton and along the TAB’s basin axis toward the Wisconsin and Ohio
ranges (Fig. 2B, ice center ‘‘r’’ ). The multiple ice centers contributing to
sedimentation in the TAB may not have been synchronous in their
advances and retreats throughout the LPIA. This additional evidence for an
ice center over Marie Byrd Land is important because it helps to explicate
the hypothesis that glaciation during the LPIA consisted of asynchronous,
discrete ice centers and not a single, large ice sheet centered over
Antarctica (Isbell et al. 2012; Monta ˜
nez and Poulsen 2013) (Fig. 2B).
Inferences made from glaciogenic and glacially influenced sediments
(‘‘near-field’’ records) can be tied to global and ‘‘far-field’’ records of
climate change from the ~80 Myr of the LPIA to approach a holistic
understanding of the effects that the onset, duration, and ultimate collapse
of a global icehouse-influenced Earth, both geologically and biologically
(Monta˜
nez et al. 2007; Rygel et al. 2008; Soreghan et al. 2019).
CONCLUSIONS
The Pagoda Fm in the Shackleton Glacier region is glaciogenic and
was deposited in a basin-marginal subaqueous setting, by a glacier with a
temperate or mild subpolar thermal regime. The dominant lithology in the
Pagoda Fm here is massive, sandy, clast-poor diamictite. The depositional
processes governing these diamictites were subaqueous glacial processes
in a marine or brackish setting; likely a combination of mass transport,
iceberg rain-out, iceberg scouring, plume sedimentation, and subglacial
till deposition. Current-transported sands and stratified diamictites in the
Pagoda Fm were deposited as part of grounding-line fan systems. In the
Shackleton Glacier region, all glaciogenic sediments in the Pagoda Fm
were likely deposited during the retreat phase of a single glacial
sequence. The transport directions and thicknesses of strata along the
TAB margin were strongly controlled by topographic relief on the
underlying erosional surface. Glacier flow directions (towards the south
and southeast) and trends in Pagoda Fm thicknesses in the Shackleton
Glacier Area support the hypothesis that an ice center was present toward
the Panthalassan margin of East Antarctica (Marie Byrd Land) during the
LPIA.
SUPPLEMENTAL MATERIAL
Supplemental appendices are available from SEPM’s Data Archive: https://
www.sepm.org/supplemental-materials.
ACKNOWLEDGMENTS
This work would have been impossible without the hard work of all the
people who made the 2017–2018 Shackleton Deep Field Camp such a success,
including the talented people of the National Science Foundation, Antarctic
Support Contract, Ken Borek Air, Petroleum Helicopters, Inc., New York Air
National Guard, and the U.S. Air Force. Special thanks are owed to Edith Taylor
and Rudolf Serbet for aiding in the field planning, Danny Uhlmann and Ted
Grosgebauer for keeping us from tumbling off of cliffs, and to Patty Ryberg,
Rudolf Serbert, Brian Atkinson, and Erik Gulbranson for their companionship
and cooking in the deep field. Thanks also to Kate Pauls and Eduardo Rosa for
their feedback on the manuscript. Funding for this research came from National
Science Foundation OPP-1443557, EAR-1729219, and OISE-1559231 grants,
the University of Wisconsin–Milwaukee Graduate Fellowships programs, the
P.E.O. Scholar Awards Program, The American Federation of Mineralogical
Societies, The Wisconsin Geological Society, University of Wisconsin–
Milwaukee (RGI grant), and the University of Wisconsin–Milwaukee
Department of Geosciences. And finally, thanks are due to Fernando Vesely
and an anonymous reviewer for their thoughtful comments and suggestions that
improved this manuscript.
REFERENCES
AITCHISON, J.C., BRADSHAW, M.A., AND NEWMANN, L., 1988, Lithofacies and origin of the
Buckeye Formation: late Paleozoic glacial and glaciomarine sediments, Ohio Range,
Transantarctic Mountains, Antarctica: Palaeogeography, Palaeoclimatology, Palaeoecol-
ogy, v. 64, p. 93–104.
ALLMENDINGER ,R.W.,CARDOZO, N., AND FISHER, D.M., 2012, Structural Geology
Algorithms: Vectors and Tensors in Structural Geology: Cambridge University Press,
302 p.
ANDERSON,J.B.,KURTZ, D.D., DOMACK, E.W., AND BALSHAW, K.M., 1980, Glacial and glacial
marine seidment of the Antarctic continental shelf: Journal of Geology, v. 88, p. 399–
414.
ASKIN, R.A., 1998, Floral trends in the Gondwana high latitudes: palynological evidence
from the Transantarctic Mountains: Journal of African Earth Sciences, v. 27, p. 12–13.
ASKIN, R.A., BARRETT, P.J., KOHN, B.P., AND MCPHERSON, J.G., 1971, Stratigraphic sections
of the Beacon Supergroup (Devonian and older(?) to Jurassic) in south Victoria Land, in
Barrett, P.J., ed., Antarctic Data Series No. 2: Wellington, Victoria University.
ASSINE, M.L., DESANTA ANA, H., VEROSLAVSKY, G., AND VESELY, F.F., 2018, Exhumed
subglacial landscape in Uruguay: erosional landforms, depositional environments, and
paleo-ice flow in the context of the late Paleozoic Gondwanan glaciation: Sedimentary
Geology, v. 369, p. 1–12.
BABCOCK, L.E., ISBELL, J.L., MILLER, M.F., AND HASIOTIS, S.T., 2002, New late Paleozoic
conchostracan (Crustacea: Branchiopoda) from the Shackleton Glacier Area, Antarctica:
Age and paleoenvironmental implications: Journal of Paleontology, v. 76, p. 70–75.
BARRETT, P.J., 1972, Late Paleozoic glacial valley at Alligator Peak, Southern Victoria Land,
Antarctica: New Zealand Journal of Geology and Geophysics, v. 15, p. 262–268.
BARRETT, P.J., AND MCKELVEY, B.C., 1981, Permian tillites of southern Victoria Land,
Antarctica, in Hambrey, M.J., and Harland, W.B., eds., Earth’s pre-Pleistocene Glacial
Record: Cambridge, Cambridge University Press, p. 233–236.
BARRETT, P.J., ELLIOT, D.H., AND LINDSAY, J.F., 1986, The Beacon Supergroup (Devonian–
Triassic) and Ferrar Group (Jurassic) in the Beardmore Glacier area, Antarctica, in
Turner, M.D., and Splettstoesser, J.F., eds., Geology of the central Transantarctic
Mountains: American Geophysical Union, Antarctic Research Series, p. 339–428.
BATCHELOR, C.L., AND DOWDESWELL, J.A., 2015, Ice-sheet ground-zone wedges (GZWs) on
high-latitude continental margines: Marine Geology, v. 363, p. 65–92.
BATTERSON, M., AND SHEPPARD, K., 2000, Deglacial history of northern St. George’s Bay,
western Newfoundland: Newfoundland Department of Mines and Energy, Geological
Survey of Newfoundland, Current Research, v. 1, p. 33–47.
BENN, D.I., 1996, Subglacial and subaqueous processes near a glacier grounding line:
sedimentological evidence from a former ice-dammed lake, Achnasheen Scotland:
Boreas, v. 25, p. 23–36.
BENNETT, M.R., HAMBREY, M.J., AND HUDDART, D., 1997, Modification of clast shape in high
artic glacial environments: Journal of Sedimentary Research, v. 67, p. 550–559.
BINDSCHADLER, R., VORNBERGER, V., FLEMING, A., FOX, A., MULLINS, J., BINNIE, D., PAULSEN,
S.J., GRANNEMAN,B.,AND GORODESTZKY, D., 2008, The Landsat image mosaic of
Antarctica: Remote Sensing of Environment, v. 112, p. 4214–4226.
CARDOZO,N.,AND ALLMENDINGER, R.W., 2013, Spherical projections with OSXStereonet:
Computers & Geosciences, v. 51, p. 193–205.
COATES, D.A., 1985, Late Paleozoic glacial patterns in the central Transantarctic
Mountains, Antarctica, in Turner, M.D., and Splettstoesser, J.F., eds., Geology of the
Central Transantarctic Mountains: American Geophysical Union, Antarctic Research
Series, p. 325–338.
COLLINSON, J.D., AND MOUNTNEY, N., 2019, Sedimentary Structures: Edinburgh, Academic
Press, 340 p.
COLLINSON, J.W., ISBELL, J.L., ELLIOT, D.H., MILLER, M.F., AND MILLER, J.M.G., 1994,
Permian–Triassic Transantarctic basin, in Veevers, J.J., and Powell, C.M., eds., Permian–
Triassic Pangean basins and Foldbelts along the Panthalassan Margin of Gondwanaland:
Geological Society of America, Memoir 184, p. 173–222.
COLLINSON, J.W., HAMMER, W.R., ASKIN, R.A., AND ELLIOT, D.H., 2006, Permian–Triassic
boundary in the central Transantarctic Mountains, Antarctica: Geological Society of
America, Bulletin, v. 118, p. 747–763.
COOPER, M.A., JORDAN, T.M., SCHROEDER, D.M., SIEGERT, M.J., WILLIAMS, C.N., AND
BAMBER, J.L., 2019, Subglacial roughness of the Greenland Ice Sheet: relationship with
contemporary ice velocity and geology: The Cryosphere, v. 13, p. 3093–3115.
CORNAMUSINI, G., TALARICO, R.M., SIMONETTE, C., SPINE, A., OLIVETTI,V.,AND WOO,J.,
2017, Upper Paleozoic glacigenic deposits of Gondwana: stratigraphy and paleoenvir-
onmental significance of a tillite succession in Northern Victoria Land (Antarctica):
Sedimentary Geology, v. 358, p. 51–69.
COWAN, E.A., AND POWELL, R.D., 1991, Ice-proximal sediment accumulation rates in a
temperate glacial fjord, south-eastern Alaska, in Anderson, J.B., and Ashley, G.M., eds.,
Glacial Marine Sedimentation: Paleoclimatic Significance: Ge ological Society of
America, Special Paper 261, p. 61–73.
CRADDOCK, J.P., OJAKANGAS, R.W., MALONE, D.H., KONSTANTINOU, A., MORY, A.J., BAUER,
W. , T HOMAS, R.J., AFFINATI, S.C., PAULS, K.N., ZIMMERMAN, U., BOTHA, G., ROCHAS-
CAMPOS, A., DOS SANTOS, P.R., TOHVER, E., RICCOMINI, C., MARTIN, J., REDFERN,J.,
HORSTWOOD,M.,AND GEHRELS, G., 2019, Detrital Zircon Provenance of Permo-
Carboniferous Glacial Diamictites across Gondwana: Earth-Science Reviews, v. 192, p.
285–316.
L.R.W. IVES AND J.L. ISBELL632 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
CROWELL, J.C., AND FRAKES, L.A., 1972, Late Paleozoic glaciation: Part V, Karoo Basin,
South Africa: Geological Society of America, Bulletin, v. 83, p. 2887–2919.
CROWELL, J.C., AND FRAKES, L.A., 1975, Late Paleozoic glaciaition, in Campbell, K.S.W.,
ed., Gondwana Geology: Papers Presented at the Third Gondwana Symposium
Canberra, Australia, 1973: Canberra, Australia, Australian National University Press,
p. 313–331.
DASGUPTA, P., 2006, Facies characteristics of Talchir Formation, Jharia Basin, India:
implications for initiation of Gondwana sedimentation: Sedimentary Geology, v. 185, p.
59–78.
DEMET, B.P., NITTROUER, J.A., ANDERSON, J.B., AND SIMKINS, L.M., 2019, Sedimentary
processes at ice sheet grounding-zone wedges revealed by outcrops, Washington State
(USA): Earth Surface Processes and Landforms, v. 44, p. 1209–1220.
DIETRICH, P., FRANCHI, F., SETLHABI, L., PREVEC, R., AND BAMFORD, M., 2019, The nonglacial
diamictite of Toutswemogala Hill (lower Karoo Supergroup, central Botswana):
implications of the extent of the Late Paleozoic Ice Age in the Kalahari–Karoo Basin:
Journal of Sedimentary Research, v. 89, p. 1–15.
DIETRICH,P.,AND HOFFMANN, A., 2019, Ice-margin fluctuation sequences and grounding
zone wedges: the record of the Late Palaeozoic Ice Age in the eastern Karoo Basin
(Dwyka Group, South Africa): The Depositional Record, v. 5, no. 2, p. 247–271.
DOMACK, E.W., 1983, Facies of Late Pleistocene glacial–marine sediments on Whidbey
Island, Washington: an isostatic glacial–marine sequence, in Molina, B.F., ed., Glacial–
Marine Sedimentation: Boston, Springer, p. 535–570.
DOMACK, E.W., AND POWELL, C.M., 2018, Modern Glaciomarine Environments and
Sediments, in Menzies, J., and van der Meer, J.J.M., eds., Past Glacial Environments:
Elseveir, p. 105–158.
DOMEIER, M., AND TORSVIK, T.H., 2014, Plate tectonics in the late Paleozoic: Geoscience
Frontiers, v. 5, p. 303–350.
DOWDESWELL, J.A., HAMBREY, M.J., AND WU, R., 1985, A comparison of clast fabric and
shape in Late Precambrian and modern glacigenic sediments: Jour nal of Sedimentary
Petrology, v. 55, p. 691–704.
DOWDESWELL, J.A., WHITTINGTON, R.J., AND MARIENFELD, P., 1994, The origin of massive
diamicton facies by iceberg rafting and scouring, Scoresby Sund, East Greenland:
Sedimentology, v. 41, p. 21–35.
DOWDESWELL, J.A., HOGAN, K.A., ARNOLD, N.S., MUGFORD, R.I., WELLS, M., HIRST, J.P.P.,
AND DECALF, C., 2015, Sediment-rich meltwater plumes and ice-proximal fans at
themargins of modern and ancient tidewater glaciers: Observations and modeling:
Sedimentology, v. 62, p. 1665–1692.
DOWDESWELL, J.A., CANALS, M., JAKOBSSON, B.J., TODD, B.J., DOWDESWELL, E.K., AND
HOGAN, K.A., 2016, The variety and distribution of submarine glacial landforms and
implications for ice-sheet reconstruction, in Dowdeswell, J.A., Canals, M., Jakobsson,
B.J., Todd, B.J., Dowdeswell, E.K., and Hogan, K.A., eds., Atlas of Submarine Glacial
Landforms: Modern, Quaternary and Ancient: Geological Society of London, Memoir
46, p. 519–552.
DUMAS,S.,AND ARNOTT, R.W.C., 2006, Origin of hummocky and swaley cross-
stratification: the controlling influence of unidirectional current strength and aggradation
rate: Geology, v. 34, p. 1073–1076.
EIDAM, E.F., SUTHERLAND, D.A., DUNCAN, D., KIENHOLZ, C., AMUNDSON, J.M., AND MOTYKA,
R.J., 2020, Morainal bank evolution and impact on terminus dynamics during a tidewater
glacier stillstand: Journal of Geophysical Research: Earth Surface, v. 125, no. 11,
e2019JF005359.
ELLIOT, D.H., 1992, Jurassic magmatism and tectonism associated with Gondwanaland
break-up: an Antarctic perspective, in Storey, B.C., Alabaster, T., and Pankhurst, R.J.,
eds., Magmatism and the Causes of Continental Break-Up: Geological Society of
London, Special Publication 68, p. 165–184.
ELLIOT, D.H., 2013, The geological and tectonic evolution of the Transanatarctic
Mountains: a review, in Hambrey, M.J., Barker, P.F., Barrett, P.J., Bowman, V., Davies,
B., Smellie, J.L., and Tranter, M., eds., Antarctic Palaeoenvironments and Earth-Surface
Processes: Geological Society of London, Special Publication 381, p. 7–35.
ELLIOT, D.H., FANNING, C.M., ISBELL, J.L., AND HULETT, S.R.W., 2017, The Permo-Triassic
Gondwana sequence, central Transantarctic Mountains, Antarctica: zircon geochronol-
ogy, provenance, and basin evolution: Geosphere, v. 13, p. 155–178.
ESTRADA, S., LA
¨UFER, A., ECKELMANN, K., HOFMANN, M., GA
¨RTNE R, A., AND LINNEMANN,U.,
2016, Continuous Neoproterozoic to Ordovician sedimentation at the East Gondwana
margin: implications from detrital zircons of the Ross Orogen in northern Victoria Land,
Antarctica: Gondwana Research, v. 37, p. 426–448.
EVANS, D.J.A., PHILLIPS, E.R., HIEMSTRA, J.F., AND AUTON, C.A., 2006, Subglacial till:
formation, sedimentary characteristics and classifi cation: Earth-Science Reviews, v. 78,
p. 115–176.
EYLES, C.H., 1987, Glacially influenced submarine-channel sedimentation in the Yakataga
Formation, Middleton Island, Alaska: Journal of Sedimentary Petrology, v. 57, p. 1004–
1017.
EYLES, C.H., 1988, A model for striated boulder pavement formation on glaciated, shallow-
marine shelves: an example from the Yakataga Formation, Alaska: Journal of
Sedimentary Petrology, v. 58, p. 62–71.
EYLES, C.H., AND LAGOE, M.B., 1990, Sedimentation patterns and facies geometry on a
temperate glacially-influenced continental shelf: the Yakataga Formation, Middleton
Island, Alaska, in Dowdeswell, J.A., and Scourse, J.D., eds., Glacimarine Environments:
Processes and Sediments: Geological Society of London, Special Publication 53, p. 363–
386.
FALLGATTER, C., AND PAIM, P.S.G., 2019, On the origin of the Itarar´e Group basal
nonconformity and its implications for the late Paleozoic glaciation in the Paran´a Basin,
Brazil: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 531, 108225.
FEDORCHUK, N.D., ISBELL, J.L., GRIFFIS, N.P., MONTAN
˜EZ, I.P., VESELY, F.F., IANNUZZI, R.,
MUNDIL, R., YIN, Q.-Z., PAULS, K.N., AND ROSA, E.L., 2019, Origin of paleovalleys on the
Rio Grande do Sul Shield (Brazil): implications for the extent of late Paleozoic glaciation
in west-central Gondwana: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 531,
108738.
FIELDING, C.R., SLIWA, R., HOLCOMBE, R.J., AND JONES, A.T., 2001, A new palaeogeographic
synthesis for the Bowen, Gunnedah and Sydney basins of eastern Australia, in Hill, K.C.,
and Bernecker, T., eds., Eastern Australasian Basins Symposium: A Refocused Energy
Perspective for the Future: Victoria, Australia, Petroleum Exploration Society of
Australia, Special Publication, p. 269–278.
FIELDING, C.R., FRANK, T.D., BIRGENHEIER, L.P., RYGEL, M.C., JONES, A.T., AND ROBERTS,J.,
2008a, Stratigraphic imprint of the late Palaeozoic ice age in eastern Australia: a record
of alternating glacial and nonglacial climate regime: Geological Society of London, v.
165, p. 129–140.
FIELDING, C.R., FRANK, T.D., BIRGENHEIER, L.P., RYGEL, M.C., JONES, A.T., AND ROBERTS,J.,
2008b, Stratigraphic record and facies associations of the late Paleozoic ice age in
eastern Australia (New South Wales and Queensland), in Fielding, C.R., Frank, T., and
Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and Space: Geological
Society of America, Special Paper 441, p. 41–57.
FIELDING, C.R., FRANK, T.D., AND ISBELL, J.L., 2008c, The Late Paleozoic Ice Age: a review
of current understanding and synthesis of global climate patter ns, in Fielding, C.R.,
Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and
Space: Geological Society of America, Special Publication 441, p. 343–354.
FIELDING, C.R., FRANK, T.D., AND ISBELL, J.L., eds., 2008d, Resolving the Late Paleozoic Ice
Age in Time and Space: Geological Society of America, Special Publication 441, 354 p.
FIELDING, C.R., FRANK, T.D., ISBELL, J.L., HENRY, L.C., AND DOMACK, E.W., 2010,
Stratigraphic signature of the late Palaeozoic Ice Age in the Parmeener Supergroup of
Tasmania, SE Australia, and inter-regional comparisons: Palaeogeography, Palae-
oclimatology, Palaeoecology, v. 298, p. 70–90.
FRAKES, L.A., AND CROWELL, J.C., 1969, Late Paleozoic glaciation: I, South America:
Geological Society of America, Bulletin, v. 80, p. 1007–1042.
FRAKES, L., MATTHEWS, J., NEDER, I., AND CROWELL, J., 1966, Movement directions in late
Paleozoic glacial rocks of the Horlick and Pensacola Mountains, Antarctica: Science, v.
153, p. 746–749.
FRAKES, L.A., MATTHEWS, J.L., AND CROWELL, J.C., 1971, Late Paleozoic glaciation: Part III,
Antarctica: Geological Society of America, Bulletin, v. 82, p. 1581–1604.
FRANC¸A, A.B., AND POTTER, P.E., 1988, Estratigrafi a, ambiente deposicional e analise de
reservatorio do Grupo Itarare (Permocarbonifero), Bacia do Parana (Parte 1):
PETROBRAS, Boletim de Geociencias, v. 2, p. 147–191.
FRANC¸A, A.B., WINTER, W.R., AND ASSINE, M.L., 1996, Arenitos Lapa–Vila Velha; um
modelo de trato de sistemas subaquosos canal–lobos sob influencia glacial, Grupo Itarare
(C-P), Bacia do Parana: Revista Brasileira de Geociencias, v. 26, p. 43–56.
FRIEND, P.F., 1983, Towards the fi eld classification of alluvial architecture or sequence, in
Collinson, J.D., and Lewin, J., eds., Modern and Ancient Fluvial Systems, International
Association of Sedimentologists, Special Publication 6, p. 345–354.
FUNK, J.E., SLATT, R.M., AND PYLES, D.R., 2012, Quantification of static connectivity
between deep-water channels and stratigraphically adjacent architectural elements using
outcrop analogs: American Association of Petroleum Geologists, Bulletin, v. 96, p. 277–
300.
GASTALDO, R.A., DIMICHELE, W.A., AND PFEFFERKORN, H.W., 1996, Out of the icehouse into
the greenhouse; a late Paleozoic analog for modern global vegetational change: GSA
Today, v. 6, no. 10, p. 1–7.
GOODGE, J.W., 2020, Geological and tectonic evolution of the Transantarctic Mountains,
from ancient craton to recent enigma: Gondwana Research, v. 80, p. 50–122.
GOODGE, J.W., AND FANNING, M.C., 2016, Mesoarchean and Paleoproterozoic history of the
Nimrod Complex, central Transantarctic Mountains, Antarctica: stratigraphic revisions
andrelation to the Mawson Continent in East Gondwana: Precambrian Research, v. 285,
p. 242–271.
HALLET, B., HUNTER, L., AND BOGEN, J., 1996, Rates of erosion and sediment evacuation by
glaciers: a review of field data and their implications: Global and Planetary Change, v.
12, p. 213–235.
HAMBREY, M.J., 1994, Glacial Environments: Vancouver, University of British Columbia
Press, 296 p.
HAMBREY, M.J., AND GLASSER, N.F., 2003, Glacial sediments: processes, environments and
facies, in Middleton, G.V., ed., Encyclopedia of Sediments and Sedimentary Rocks:
Dordrecht, Kluwer Academic Publishers, p. 316–331.
HAMBREY, M.J., AND GLASSER, N.F., 2012, Discriminating glacier thermal and dynamic
regimes in the sedimentary record: Sedimentary Geology, v. 251–252, p. 1–33.
HAND, S.J., 1993, Palaeogeography of Tasmania’s Permo-Carboniferous glacige nic
sediments, in Findlay, R.H., Unrug, R., Banks, M.R., and Veevers, J.J., eds., Gondwana
Eight: Assembly, Evolution and Dispersal: Rotterdam, A.A. Balkema, p. 459–469.
HUNTER, L.E., POWELL, R.D., AND LAWSON, D.E., 1996, Flux of debris transported by ice at
three Alaskan tidewater glaciers: Journal of Glaciology, v. 42, p. 123–135.
ISBELL, J.L., 1999, The Kukri Erosion Surface: a reassessment of its relationship to rocks of
the Beacon Supergroup in the central Transantarctic Mountains, Antarctica: Antarctic
Science, v. 11, p. 228–238.
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 633
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
ISBELL, J.L., 2010, Environmental and paleogeographic implications of glaciotectonic
deformation of glaciomarine deposits within Permian strata of the Metschel Tillite,
southern Victoria Land, Antarctica, in opez-Gamund´ı, O.R., and Buatois, L.A., eds.,
Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological
Society of America, Special Publication 468, p. 81–100.
ISBELL, J.L., 2015, Permian and Triassic sedimentation in the central Transantarctic
mountains, southern Victoria Land and Northern Victoria Land: a south polar view of
Gondwanan during the Paleozoic–Mesozoic transition, The 21st Internation Symposium
on Polar Science: Polar Region as a Key Observatory for the Changing Globe and
Beyond: Incheon, Republic of Korea, Korea Polar Institute.
ISBELL, J.L., GELHAR, G.A., AND SEEGERS, G.M., 1997a, Reconstruction of preglacial
topography using a postglacial flooding surface: Upper Paleozoic deposits, central
Transantarctic Mountains, Antarctica: Journal of Sedimentary Research, v. 67, p. 264–
273.
ISBELL, J.L., SEEGERS, G.M., AND GELHAR, G.A., 1997b, Upper Paleozoic glacial and
postglacial deposits, central Transantarctic Mountains, Antarctica, in Martini, I.P., ed.,
Late Glacial and Postglacial Environmental Changes: Quaternary, Carboniferous–
Permian, and Proterozoic: Oxford, U.K., Oxford University Press, p. 230–242.
ISBELL, J.L., MILLER, M.F., BABCOCK, L.E., AND HASIOTIS, S.T., 2001, Ice-marginal
environment and ecosystem prior to initial advance of the late Paleozoic ice sheet in the
Mount Butters area of the central Transantarctic Mountains, Antarctica: Sedimentology,
v. 48, p. 953–970.
ISBELL, J.L., COLE, D.I., AND CATUNEANU, O., 2008a, Carboniferous–Permian glaciation in
the main Karoo Basin, South Africa: stratigraphy, depositional controls, and glacial
dynamics, in Fielding, C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late
Paleozoic Ice Age in Time and Space: Geological Society of America, Special
Publication 441, p. 71–82.
ISBELL, J.L., FRAISER, M.L., AND HENRY, L.C., 2008b, Examining the complexity of
environmental change during the late Paleozoic and early Mesozoic: Palaios, v. 23, p.
267–269.
ISBELL, J.L., KOCH, Z.J., SZABLEWSKI, G.M., AND LENAKER, P.A., 2008c, Permian glacigenic
deposits in the Transantarctic Mountains, Antarctica, in Fielding, C.R., Frank, T.D., and
Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and Space: Geological
Society of America, Special Publication 441, p. 59–70.
ISBELL, J.L., HENRY, L.C., GULBRANSON, E.L., LIMARINO, C.O., FRAISER, M.L., KOCH, Z.J.,
CICCIOLI, P.L., AND DINEEN, A.A., 2012, Glacial paradoxes during the late Paleozoic ice
age: evaluating the equilibrium line altitude as a control on glaciation: Gondwana
Research, v. 22, p. 1–19.
JORDAN, T.A., FERRACCIOLI, F., ARMADILLO, E., AND BOZZO, E., 2013, Crustal architecture of
the Wilkes Subglacial Basin in East Antarctica, as revealed fromairborne gravity data:
Tectonophysics, v. 585, p. 196–206.
KESSLER, T.C., KLINT, K.E.S., NILSSON,B.,AND BJERG, P.L., 2012, Characterization of sand
lenses embedded in tills: Quaternary Science Reviews, v. 53, p. 55–71.
KOCH, Z.J., 2010, Reevaluation of the late Paleozoicglacigenic deposits of the Pagoda
Formation at Tillite Glacier, central Transanatarctic Mountains, Antarctica [Ph.D.
Thesis]: University of Wisconsin–Milwaukee, Milwaukee, 145 p.
KOCH, Z.J., AND ISBELL, J.L., 2013, Processes and products of grounding-line fans from the
Permian Pagoda Formation, Antarctica: insight into glacigenic conditions in polar
Gondwana: Gondwana Research, v. 24, p. 161–172.
KURJANSKI, B., REA, B.R., SPAGNOLO, M., CORNWELL, D.G., HOWELL,J.,AND ARCHER, S.,
2020, A conceptual model for glaciogenic reservoirs: from landsystems to reser voir
architecture: Marine and Petroleum Geology, v. 115, 104205.
LANDVIK, J.Y., ALEXANDERSON, H., HENRIKSEN, M., AND INGO
´LFSSON,´
O., 2014, Landscape
imprints of changing glacial regimes during ice-sheetbuild-up and decay: a conceptual
model from Svalbard: Quaternary Science Reviews, v. 92, p. 258–268.
LAWSON, D.E., 1979, Sedimentological analysis of the western terminus region of the
Matanuska Glacier, Alaska: Hanover, New Hampshire, U.S. Army, Cold Regions
Research and Engineering Laboratory, 122 p.
LAWVER, L.A., DALZIEL, I.W.D., NORTON, I.O., AND GAHAGAN, L.M., 2011, The Plates 2011
Atlas of Plate Reconstructions (500 Ma to Present Day): University of Texas, Technical
Report 198, 189 p.
LENAKER, P.A., 2002, Sedimentology of Permian glacial deposits in the Darwin Glacier
region, Antarctica [M.S. Thesis]: University of Wisconsin–Milwaukee, 173 p.
LICHT, K.J., DUNBAR, N.W., ANDREWS, J.T., AND JENNINGS, A.E., 1999, Distinguishing
subglacial till and glacial marine diamictons in the western Ross Sea, Antarctica:
implications for a last glacial maximum grounding line: Geological Society of America,
Bulletin, v. 111, p. 91–103.
LIMARINO, C.O., ALONSO-MURUAGA, P.J., CICCIOLI, P.L., LOINAZE, V.S.P., AND CE
´SARI, S,N.,
2014, Stratigraphy and palynology of a late Paleozoic glacial paleovalley in the Andean
Precordillera, Argentina: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 412, p.
223–240.
LINDBA
¨CK, K., AND PETTERSSON, R., 2015, Spectral roughness and glacial erosion of a land-
terminating section of the Greenland Ice Sheet: Geomor phology, v. 238, p. 149–159.
LINDSAY, J.F., 1969, Stratigraphy and sedimentation of Lower Beacon rocks in the central
Transantarctic Mountains, Antarctica: Institute of Polar Studies, Report 33: Research
Foundation and the Institute of Polar Studies, The Ohio State University, 58 p.
LINDSAY, J.F., 1970a, Depositional environment of Paleozoic glacial rocks in the central
Transantarctic Mountains: Geological Society of America, Bulletin, v. 81, p. 1149–1172.
LINDSAY, J.F., 1970b, Paleozoic cave deposit in the central Transantarctic Mountains: New
Zealand Journal of Geology and Geophysics, v. 13, p. 1018–1023.
LISITZIN, A.P., 2002, Sea-Ice and Iceberg Sedimentation in the Ocean: Berlin, Springer-
Verlag, 800 p.
LONG, W.E., 1964a, The stratigraphy of the Horlick Mountains, in Adie, R.J., ed.,
International Symposium on Antarctic Geology: Amsterdam, North Holand Publishing,
p. 352–363.
LONG, W.E., 1964b, The stratigraphy of the Ohio Range, Antarctica [Ph.D. Thesis]: The
Ohio State University, Columbus, Ohio, 340 p.
LONG, W.E., MCLELL, D., COLLINSON, J.W., AND ELLIOT, D.H., 2008–2009, Geology of the
Nilsen Plateau, Queen Maud Mountains, Transantarctic Mountains: Terra Antarctica, v.
15, p. 239–253.
LØNNE, I., 1995, Sedimentary facies and depositional architecture of ice-contact
glaciomarine systems: Sedimentary Geology, v. 98, p. 13–43.
LØNNE, I., NEMEC, W., BLIKRA, L.H., AND LAURITSEN, T., 2001, Sedimentary architecture and
dynamic stratigraphy of a marine ice-contact system: Journal of Sedimentary Research,
v. 71, p. 922–943.
LO
´PEZ-GAMUNDI
´, O., LIMARINO, C.O., ISBELL, J.L., PAULS, K.N., CE
´SARI, S.N., AND MURUAGA,
P.A., 2021, The Late Paleozoic Ice Age along the southwestern margin of Gondwana:
facies models, age constraints, correlation and sequence stratigraphic framework: Journal
of South American Earth Sciences, v. 107, 103056.
MARCOS, P., GREGORI, D.A., BENEDINI, L., BARROS, M., STRAZZERE, L., AND PIVETTE, C.P.,
2018, Pennsylvanian glacimarine sedimentation in the Cushamen Formation, western
North Patagonian Massif: Geoscience Frontiers, v. 9, p. 485–504.
MARTIN, H., 1981, The late Palaeozoic Gondwana glaciation: Geologische Rundschau, v.
70, p. 480–496.
MARTIN, J.R., REDFERN, J., HORSTWOOD, M.S.A., MORY, A.J., AND WILLIAMS, B.P.J., 2019,
Detrital zircon age and provenance constraints on late Paleozoic ice-sheet growth and
dynamics in Western and Central Australia: Australian Journal of Earth Sciences, v. 66,
p. 183–207.
MASOOD, K.R., TAYLO R , T.N., HORNER,T.,AND TAYLOR, E.L., 1994, Palynology of the
Mackellar Formation (Beacon Supergroup) of East Antarctica: Review of Palaeobotany
and Palynology, v. 83, p. 329–337.
MATSCH, C.L., AND OJAKANGAS, R.W., 1991, Comparison in depositional style of ‘‘polar’’
and ‘‘temperate’’ glacial ice: late Paleozoic Whiteout Conglomerate (West Antarctica)
and late Proterozoic Mineral Fork Formation (Utah), in Anderson, J.B., and Ashley,
G.M., eds., Glacial Marine Sedimentation: Paleoclimatic Signifi cance: Geological
Society of America, Special Publication 261, p. 191–206.
MATSCH, C.L., AND OJAKANGAS, R.W., 1992, Stratigraphy and sedimentology of the
Whiteout Conglome rate: a late Paleozoic glacigenic sequence in the Ellsworth
Mountains, West Antarctica, in Webers, G.F., Craddock, G.F., and Splettstoesser, J.F.,
eds., Geology of the Ellsworth Mountains, Antarctica: Geological Society of America,
Memoir 170, p. 37–62.
MCGREGOR, V.R., AND WADE, F.A., 1969, Geologic Map of Antarctica, Western Queen
Maud Mountains: American Geographical Society of New York, Sheet 16.
MCKAY, R., BROWNE, G., CARTER, L., COWAN, E., DUNBAR, G., KRISSEK, L., NAISH,T.,
POWELL, R., REED, J., TALARICO,F.,AND WILCH, T., 2009, The stratigraphic signature of the
late Cenozoic Antarctic Ice Sheets in the Ross Embayment: Geological Society of
America, Bulletin, v. 121, p. 1537–1561.
MILLER, J.M.G., 1989, Glacial advance and retreat sequences in a Permo-Carboniferous
section, central Transantarctic Mountains: Sedimentology, v. 36, p. 419–430.
MINSHEW, V.H., 1967, Geology of the Scott Glacier and Wisconsin Range areas, central
Transantarctic Mountains, Antarctica [Ph.D. Thesis]: The Ohio State University, 268 p.
MIRSKY, A., 1969, Geologic Map of Antarctica, Ohio Range to Liv Glacier: American
Geographical Society, Sheet 17.
MONCREIFF, A.C.M., 1989, Classifi cation of poorly sorted sediments: Sedimentary Geology,
v. 65, p. 191–194.
MONTAN
˜EZ, I.P., AND POULSEN, C.J., 2013, The late Paleozoic ice age: an evolving paradigm:
Annual Review of Earth and Planetary Sciences, v. 41, p. 1–28.
MONTAN
˜EZ, I., AND SOREGHAN, G.S., 2006, Earth’s fickle climate: lessons learned from deep-
time ice ages: Geotimes, v. 51, p. 24–27.
MONTAN
˜EZ, I.P., TABOR, N.J., NIEMEIER, D., DIMICHELE, W.A., FRANK, T.D., FIELDING, C.R.,
ISBELL, J.L., BIRGENHEIER, L.P., AND RYGEL, M.C., 2007, CO
2
-forced climate and
vegetation instability during late Paleozoic deglaciation: Science, v. 315, p. 87–91.
MORY, A.J., MARTIN, J.R., AND REDFERN, J., 2008, A review of Permian–Carboniferous
glacial deposits in Western Australia, in Fielding, C.R., Frank, T.D., and Isbell, J.L., eds.,
Resolving the Late Paleozoic Ice Age in Time and Space: Geological Society of
America, Special Publication 441, p. 29–40.
MULDER,T.,AND ALEXANDER, J., 2001, The physical character of subaqueous sedimentary
density flows and their deposits: Sedimentology, v. 48, p. 269–299.
MURRAY, K.T., MILLER, M.F., AND BOWSER, S.S., 2013, Depositional processes beneath
coastal multi-year sea ice: Sedimentology, v. 60, p. 391–410.
NOAA, 2019, NCEI Geomagnetic Calculators, National Geophysical Data Center,
National Oceanic and Atmospheric Administration, NOAA’s National Centers for
Environmental Information (NCEI), formerly the National Geophysical Data Center, and
the collocated World Data Service for Geophysics, operated by NOAA/NESDIS/NCEI.
OJAKANGAS, R.W., AND MATSCH, C.L., 1981, The late Paleozoic Whiteout Conglomerate: a
glacial and glaciomarine sequence in the Ellsworth Mountains, West Antarctica, in
L.R.W. IVES AND J.L. ISBELL634 JSR
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
Hambrey, M.J., and Harland, W.B., eds., Earth’s pre-Pleistocene Glacial Record:
Cambridge, Cambridge University Press, p. 241–244.
PARTIN, C.A., AND SADLER, P.M., 2016, Slow net accumulation sets snowball Earth apart
from all younger glacial episodes: Geology, v. 44, p. 1019–1022.
PAULS, K.N., 2014, Sedimentology and paleoecology of fossil-bearing, high-latitude mari ne
and glacially influenced desposits in the Tepuel Basin, Patagonia, Argentina [M.S.
Thesis]: University of Wisconsin–Milwaukee.
POSAMENTIER, H.W., AND MARTINSEN, O.J., 2011, The character and genesis of submarine
mass-transport deposits: insights from outcrop and 3D seismic data, in Shipp, R.C.,
Weimer, P., and Posamentier, H.W., eds., Mass-Transport Deposits in Deepwater
Settings: Society for Sedimentary Geology, Special Publication 96, p. 7–38.
POWELL, R.D., 1990, Glacimarine processes at grounding-line fans and their growth to ice
contact deltas, in Dowdeswell, J.A., and Scourse, J.D., eds., Glacimarine Environments:
Processes and Sediments: Geological Society of London, Special Publication 53, p. 53–
73.
POWELL, R.D., 1991, Grounding-line systems as second-order controls on fluctuations of
tidewater termini of temperate glaciers, in Anderson, J.B., and Ashley, G.M., eds.,
Glacial Marine Sedimentation: Paleoclimat ic Significance: Geological Society of
America, Special Publication 261, p. 75–93.
POWELL, R.D., AND ALLEY, R.B., 1997, Grounding-line systems: processes, glaciological
inferences and the stratigraphic record, in Barker, P.F., and Cooper, A.C., eds., Geology
and Siesmic Stratigraphy of the Antarctic Margin, 2: American Geophysical Union,
Antarctic Research Series 71, p. 169–187.
POWELL, R.D., AND COOPER, J.M., 2002, A glacial sequence stratigraphic model for
temperate, glaciated continental shelves: Geological Society, Special Publication 203, p.
215–244.
POWELL, R., AND DOMACK, E., 2002, Modern glaciomarine environments, in Menzies, J.,
ed., Modern and Past Glacial Environments: Oxford, Butterworth-Heinemann, p. 361–
389.
POWELL, R.D., AND MOLNIA, B.F., 1989, Glacimarine sedimentary processes, facies and
morphology of the south-southeast Alaska shelf and fjords: Marine Geology, v. 85, p.
359–390.
RAYMOND, A., AND METZ, C., 2004, Ice and its consequences: glaciation in the late
Ordovician, Late Devonian, Pennsylvanian–Permian, and Cenozoic compared: Journal
of Geology, v. 112, p. 655–670.
REINECK, H.E., AND SINGH, I.B., 1980, Depositional Sedimentary Environments: Berlin,
Springer-Verlag, 551 p.
ROCCHI, S., BRACCIALI, L., DIVINCENZO, G., GEMELLI, M., AND GHEZZO, C., 2011, Arc
accretion to the early Paleozoic Antarctic margin of Gondwana in Victoria Land:
Gondwana Research, v. 19, p. 594–607.
RODRIGUES, M.C.N., TRZAKOS, B., ALSOP, G.I., AND VESELY, F.F., 2019, Making a
homogenite: an outcrop perspective into the evolution of deformation within mass-
transport deposits: Marine and Petroleum Geology, v. 112, 104033.
ROSA, E.L.M., AND ISBELL, J.L., 2021, Late Paleozoic glaciation: Reference Module in Earth
Systems and Environmental Sciences, Elsevier, p. 534–545.
ROSA, E.L.M., VESELY, F.F., AND FRANC¸A, A.B., 2016, A review on late Paleozoic ice-related
erosional landforms in the Paran´a Basin: origin and paleogeographical implications:
Brazilian Journal of Geology, v. 46, p. 147–166.
ROSE, P., BYERLEY, G., VAUGHAN, O., CATER, J., REA, B.R., SPAGNOLO, M., AND ARCHER, S.,
2018, Aviat: a Lower Pleistocene shallow gas hazard developed as a fuel gas supply for
the Forties Field, in Bowman, M., and Levell, B., eds., Petroleum Geology of NW
Europe: 50 Years of Learning: Geological Society of London, Petroleum Geology
Conference Series 8, Proceedings, p. 485–505.
ROSENBLUME, J.A., AND POWELL, R.D., 2019, Glacial sequence stratig raphy of ANDRILL-
1B core reveals adynamic subpolar Antarctic Ice Sheet in Ross Sea during the late
Miocene: Sedimentology, v. 66, p. 2072–2097.
RUST, I.C., 1975, Tectonic and sedimentary framework of Gondwana basins in southern
Africa, in Campbell, K.S.W., ed., Gondwana Geology: Canberra, Australian National
University Press, p. 537–564.
RYGEL, M.C., FIELDING, C.R., FRANK, T.D., AND BIRGENHEIER, L.P., 2008, The magnitude of
late Paleozoic glacioeustatic fluctuations: a synthesis: Journal of Sedimentar y Research,
v. 78, p. 500–511.
SEEGERS, G.M., 1996, Sedimentology of the Permian Mackellar Formation, Central
Transantarctic Mountains, Antarctica [M.S. Thesis]: University of Wisconsin–Milwau-
kee, 106 p.
SEEGERS-SZABLEWSKI, G., AND ISBELL, J.L., 1998, Stratigraphy and depositional environ-
ments of Lower Permian post-glacial rocks exposed between the Byrd and Nimrod
Glaciers, Antarctica: Journal of African Earth Sciences, v. 27, p. 175–176.
SHAW, J., 2016, Paraglacial landscapes in St George’s Bay, Newfoundland, Canada, in
Dowdeswell, J.A., Canals, M., Jakobsson, B.J., Todd, B.J., Dowdeswell, E.K., and
Hogan, K.A., eds., Atlas of Submarine Glacial Landforms: Modern, Quaternary and
Ancient: Geological Society of London, Memoir 46, p. 97–98.
SHEPPARD, K., BELL,T.,AND LIVERMAN, D.G.E., 2000, Late Wisconsinan stratigraphy and
chronology at Highlands, southern St. George’s Bay, southwest Newfoundland:
Quaternary International, v. 68–71, p. 275–283.
SOBIESIAK, M.S., KNELLER, B., ALSOP, G.I., AND MILANA, J.P., 2018, Styles of basal
interaction beneath mass transport deposits: Marine and Petroleum Geology, v. 98, p.
629–639.
SOREGHAN, G.S., SOREGHAN, M.J., AND HEAVENS, N.G., 2019, Explosive volcanism as a key
driver of the late Paleozoic ice age: Geology, v. 47, p. 600–604.
SURVIS, S.R., 2015, Sedimentology and stratigraphy of high-latitude, glacigenic deposits
from the Late Paleozoic Ice Age in the Tepuel–Genoa Basin, Patagonia, Argentina [M.S.
Thesis]: University of Wisconsin–Milwaukee, 104 p.
SVENDSEN, J.I., AND MANGERUD, J., 1997, Holocene glacial and climatic variations on
Spitsbergen, Svalbard: The Holocene, v. 7, p. 45–57.
TEDESCO, J., CAGLIARI, J., COITINHO, J.D.R., LOPES, R.D.C., AND LAVINA, E.L.C., 2016, Late
Paleozoic paleofjord in the southernmost Parana Basin (Brazil): geomor phology and
sedimentary fill: Geomor phology, v. 269, p. 203–214.
THOMAS, G.S.P., AND CHIVERRELL, R.C., 2006, A model of subaqueous sedimentation at the
margin of the Late Midlandian Irish Ice Sheet, Connemara, Ireland, and its implications
for regionally high isostatic sea levels: Quaternary Science Reviews, v. 25, p. 2868–
2893.
THOMAS, G.S.P., AND CONNELL, R.J., 1985, Iceberg drop, dump, and grounding structures
from Pleistocene glacio-lacustrine sediments, Scotland: Journal of Sedimentary
Petrology, v. 55, p. 243–249.
VEEVERS, J.J., AND TEWARI, R.C., 1995, Gondwana master basin of Peninsular India between
Tethys and the interior of the Gondwanaland province of Pangea: Geological Society of
America, Memoir 187, 72 p.
VEEVERS, J.J., POWELL, C.M., COLLINSON, J.W., AND LO
´PEZ-GAMUNDI
´, O.R., 1994, Synthesis,
in Veevers, J.J., and Powell, C.M., eds., Permian–Triassic Pangean basins and foldbelts
along the Panthalassan Margin of Gondwanaland: Geological Society of America,
Memoir 184, p. 331–353.
VESELY,F.,AND ASSINE, M.L., 2014, Ice-keel scour marks in the geologic record: evidence
from Carboniferous soft-sediment striated surfaces in the Paran´a Basin, southern Brazil:
Journal of Sedimentary Research, v. 84, p. 26–39.
VESELY, F.F., RODRIGUES, M.C.N.L., DAROSA, E.L.M., AMATO, J.A., TRZASKOS, B., ISBELL,
J.L., AND FEDORCHUK, N.D., 2018, Recurrent emplacement of non-glacial diamictite
during the late Paleozoic ice age: Geology, v. 46, p. 615–618.
VISSER, J.N.J., 1987, The palaeogeography of part of southwestern Gondwana during the
Permo-Carboniferous glaciation: Palaeogeography, Palaeoclimatology, Palaeoecology, v.
61, p. 205–219.
VISSER, J.N.J., 1994, The interpretation of massive rain-out and debris-flow diamictites
from the glacial marine environment, in Deynoux, M., Miller, J.M.G., Domack, E.W.,
Eyles, N., Fairchild, I.J., and Young, G.M., eds., Earth’s Glacial Record: Cambridge,
Cambridge University Press, p. 83–94.
VISSER, J.N.J., 1997, A review of the Permo-Carboniferous glaciation in Africa, in Martini,
I.P., ed., Late Glacial and Postglacial Environmental Changes: Quaternary, Carbonifer-
ous–Permian, and Proterozoic: Oxford, U.K., Oxford University Press, p. 169–191.
VISSER, J.N.J., LOOCK, J.C., AND COLLISTON, W.P., 1987, Subaqueous outwash fan and esker
sandstones in the Permo-Carboniferous Dwyka Formation of South Africa: Journal of
Sedimentary Petrology, v. 57, p. 467–478.
VIZA
´N, H., PREZZI, C.B., GEUNA, S.E., JAPAS, M.S., RENDA, E.M., FRANZESE,J.,AND VAN
ZELE, M.A., 2017, Paleotethys slab pull, self-lubricated weak lithospheric zones, poloidal
and toroidal plate motions, and Gondwana tectonics: Geosphere, v. 13, 1541–1554.
WAUGH, B.J., 1988, Sedimentology and petrology of sandstones from the Permo-
Carboniferous Pagoda Formation, central Transantarctic Mountains, Antarctica [M.S.
Thesis]: Vanderbilt University, 146 p.
WOPFNER, H., 2012, Late Palaeozoic–Early Triassic deposition and climates between
Samfrau and Tethys: a review, in Ga
˛
siewicz, M., and Słowakiewicz, M., eds., Paleozoic
Climate Cycles: Their Evolutionary and Sedimentological Impact: Geological Society of
London, Special Publication 376, p. 5–32.
WRIGHT, R., AND ANDERSON, J.B., 1982, The importance of sediment gracity flow to
sediment transport and sorting in a glacial marine environment, Eastern Weddell Sea,
Antarctica: Geological Society of America, Bulletin, v. 93, p. 951–963.
Received 13 January 2021; accepted 13 April 2021.
EARLY PERMIAN SOUTH POLAR GLACIATIONJSR 635
Downloaded from http://pubs.geoscienceworld.org/sepm/jsedres/article-pdf/91/6/611/5335572/i1527-1404-91-6-611.pdf
by USGS Library user
on 18 June 2021
... Similar to the muddy diamictite facies (Dmm-m), the physical characteristics of clasts in the sandy diamictite facies (Dmm-s) suggest that it was deposited in a glaciogenic or glacially-influenced setting, which is an interpretation consistent with other studies of this FA in the Wynyard Fm (e.g., Powell, 1990;Fielding et al., 2010;Henry et al., 2012). The sandy, massive diamictite facies (Dmm-s) is characteristic of proximal glaciomarine successions created through a combination of subglacial deposition, plume sedimentation, iceberg rainout, and resedimentation through masstransport (e.g., Eyles and Lagoe, 1990;Powell and Cooper, 2002;McKay et al., 2009;Ives and Isbell, 2021). The sandstone facies (Scfmc) in this FA was most likely deposited from low density turbidity currents (Mulder and Alexander, 2001). ...
... The stratification of the bedded, sandy diamictite facies (Dms-s), and its inter-stratification with fine-grained laminae (facies Fs), suggests that the diamictite was deposited in the same manner as the massive, sandy diamictite (facies Dmm-s), but in a more glacier-distal setting. That is, through a combination of plume sedimentation, iceberg rainout, and resedimentation through mass-transport (e.g., Eyles and Lagoe, 1990;Powell and Cooper, 2002;McKay et al., 2009;Henry et al., 2012;Ives and Isbell, 2021). The internal normal grading of the laminated, fine-grained facies (facies Fs), as well as its variations in laminae thickness, and sharp contacts between laminae of different grain sizes and colors suggest that the introduction of sediment into the water column prior to settling from suspension was episodic and likely event-driven. ...
... Divided in this way, Sequence I is~245 m thick and Sequence II is at least 170 m thick. These thicknesses are more consistent with sequences deposited by warm-based glaciers, whose sequences are on the order to hundreds to thousands of meters thick (e.g., Zellers, 1995;Cowan et al., 2010;Henry et al., 2012), than polar or sub-polar glaciers, whose sequences are typically <100 m (e.g., Rosenblume and Powell, 2019;Ives and Isbell, 2021). Thick sequences at the margins of warm-based glaciers are generally thought to be the result of their large sediment fluxes (Cowan and Powell, 1991), which allow for rapid accumulation of sediments that may drive subsidence due to loading, and the creation of additional accommodation space (Zellers, 1995). ...
Article
Full-text available
Glaciation during the Late Paleozoic Ice Age climatic interval (~ 370–260 Ma) was likely dynamic; consisting of numerous ice centers that grew and shrank asynchronously through time. Improvements in understanding of the spatial and temporal depositional complexity of glaciomarine sedimentary systems have shown that in order to characterize the conditions of Late Paleozoic glaciation, glaciogenic depositional systems and their stratigraphy must fundamentally be understood at the local level. To that end, this study reexamines the physical sedimentology and sequence stratigraphy of the type-section of the Permo-Carboniferous, glaciomarine Wynyard Formation (Wynyard Tillite) of the Tasmanian Basin by describing a 415 m thick interval of the formation, beginning at its basal erosional unconformity with the Proterozoic Burnie (Oonah) Formation. The glacigenic nature of the Wynyard Formation stratigraphy is indicated by the characteristics of clasts throughout the succession (striated, faceted, angular, and variable lithologies and grain sizes). Three glacial depositional environments were interpreted: a grounding zone wedge deposited in an ice-contact setting, glacier-proximal portion of a grounding line fan and/or morainal bank, and cyclopelites deposited in a glacier-intermediate to glacier-distal setting. Mass transport and turbidite deposits are common throughout all facies associations, as is soft sediment deformation likely caused by slumping. Several boulder pavements occur throughout the succession, indicating periodic glacier grounding. Together, these facies associations indicate that this Wynyard Fm succession is composed of glacigenic sediments that were deposited in sub-aqueous, marine, predominantly proglacial environments. The sequence stratigraphic analysis indicates that this entire succession of the Wynyard Fm was likely deposited during the single retreat phase of the “Wynyard Glacier”, with one notable readvance over the area. The interpretations made by this work enhance the understanding of glaciation of the Tasmanian Basin during the LPIA through facies analyses and a sequence stratigraphic approach, which allowed for detailed subdivision of lithologies that enabled inferences regarding the type, timing, and extent of the “Wynyard glacier”. This Wynyard Fm succession was likely deposited during a single glacier retreat with some minor readvances that may have occurred on the order of decades to years. Additionally, the glacier's grounding line was likely never more than a few kilometers upglacier (south) of this location during deposition. These finding are significant, because the massive diamictites that comprise this succession had previously been considered homogeneous and therefore resisted detailed interpretations. Constraining the often-complex depositional histories of glacigenic strata in similar fashions across the Tasmanian Basin allow us to better understand how these glacigenic deposits fit into the global climate system of the Late Paleozoic Ice Age.
... Sandstones deposited on unstable slopes near the glacial front usually show soft sediment deformation and slumping structures (Boulton, 1990;Isbell, 2010;Koch and Isbell, 2013). Thin sandstone and conglomerate layers could also be the results of iceberg dumping in both proximal and distal settings (Domack, 1983;Thomas and Connel, 1985;Sheppard et al., 2000;Ives and Isbell, 2021). This facies association shows the typical assemblage of a morainal bank association in an ice-contact to ice-proximal setting (Rosenblume and Powell, 2019), which forms along the margin of glaciers grounded with marine termination. ...
... Diamictite, as shown for FA1, could be generated from various processes in different depositional environments. The presence of weak stratification, faint lamination, and mud wisps within the thick bodies of unstratified or weakly stratified diamictite reveals that deposition should be likely linked with high decantation rate from suspended sediments under the water column, rather than with mass transport processes (Bjørlykke et al., 1976;Visser, 1994Visser, , 1996Powell and Domack, 2002;Ives and Isbell, 2021), which cannot be completely ruled out and possibly may have sometimes played a minor role. Clasts long axis alignment could be the result of decantation from icebergs rainout, while it is also been recorded in subglacial tills (Benn, 1995;Hicock et al., 1996;Hooyer and Iverson, 2000;Carr and Rose, 2003;McKay et al., 2009). ...
... Changes of clast abundance within the diamictite reflect variations of iceberg calving or glacial hydraulic system . The laminated diamictite includes minor small, soft-sediment deformed sandstone lenses that could be sourced from winnowing due to dewatering or from iceberg dumps in an ice-proximal environment (Domack, 1983;Thomas and Connel, 1985;Sheppard et al., 2000;Ives and Isbell, 2021). The laminated diamictite was likely deposited in an ice-proximal, up to ice-contact glaciomarine environment dominated by sediment plumes and/or iceberg rainout. ...
Article
The Late Paleozoic Ice Age (LPIA) is one of the coldest periods in Earth history, and led to the diachronous development of widespread ice centers across Gondwana in the Carboniferous and Permian. In Tasmanian Basin, located between northern Victoria Land (Antarctica) and Australia, the lowermost part of the Parmeener Supergroup (Late Carboniferous to Triassic), consisting of the Wynyard Tillite and its correlatives (Truro and Stockers tillites), recorded LPIA glacial sedimentation linked with ice covers that developed in the region. We carried out a detailed facies analysis of two drillcores which recovered glaciogenic sequences deposited in the Tasmanian Basin. Facies associations vary from possibly sub-glacial or ice-contact to ice-distal. Diamictite is the most common facies and its deposition is driven both by gravity and sediment remobilization processes and suspension settling with ice rafted debris accumulation. Mudstone layers, with and without dropstones, are interposed between diamictite intervals, recording ice-distal to non-glacial conditions respectively. Facies associations are indicative of subaqueous deposition in glacimarine environment. The glacial sequence stratigraphy approach was applied and glacial system tracts and bounding surfaces, which define glacial sequences, were identified. The stacking pattern of the facies associations allow us to demonstrate that the glacial sequences in the Tasmanian Basin recorded different phases of advance and retreat of the glacial front into the basin, at about the end of the main glacial phase.
... Massive, sandy, thick glacigenic diamictites like this facies are most characteristic of sub-aqueous, proglacial, glacier-proximal sedimentation (Eyles and Lagoe 1990;McKay et al., 2009;Ives and Isbell, 2021). Such successions are created through a combination of subglacial deposition, plume sedimentation, iceberg rain-out, and resedimentation through mass-transport. ...
... Divided in this way, Sequence I is ~ 245 m thick and Sequence II is at least 170 m thick. These thicknesses are more consistent with sequences deposited by temperate glaciers, whose sequences are on the order to hundreds to thousands of meters thick (e.g., Zellers, 1995;Cowan et al., 2010;Henry et al., 2010), than polar or sub-polar glaciers, whose sequences are typically less than 100 m (e.g., Ives and Isbell, 2021). Thick sequences at the margins of temperate glaciers are generally thought to be the result of their large sediment fluxes (Powell and Cowan, 1986), which allow for rapid accumulation of sediments that may drive subsidence due to loading, and the creation of additional accommodation space (Zellers, 1995). ...
... An additional line of inference is the relationship between this diamictite (Dmm-m) and the sorted sediment bodies contained within it (facies Cm and Sc-vc). Mass-transport deposits (e.g.,Sobiesiak et al., 2016;Ives and Isbell, 2021), glacial-marine sediments, and subglacial tills(Ruszczyńska-Szenajch, 1987;Kessler et al., 2012) can all contain rafts, lenses, and stringers of sorted, stratified sediments. Many of the smaller sandstone and conglomerate bodies in this facies could have been deposited as part of a subglacial till, as iceberg dumps, or as part of a mass-transport deposit. ...
Thesis
Full-text available
The Late Paleozoic Ice Age (LPIA; ~ 374 – 256 Ma) is the longest Phanerozoic icehouse interval. this interval in Earth’s history was largely defined by extensive glaciation of the southern hemisphere at both polar and temperate latitudes. Glaciers are powerful climatic and geologic actors, especially during icehouse periods, and widespread glaciation can have a significant influence on both regional and global climate and geology. Therefore, constraining the characteristics of LPIA glaciers is essential to developing a global-scale understanding of this key climatic event in Earth’s history. The manuscripts in this dissertation examine the sedimentology, transport directions, stratigraphy, and detrital zircon provenance of the Pennsylvanian – Permian glacigenic succession from the LPIA at locations in the Transantarctic (Antarctica) and Tasmanian (Australia) basins. The Transantarctic and Tasmanian basins share many characteristics that make them interesting and important places to study LPIA glacigenic rocks. In both basins, sediments were deposited during a ~ 14 Myr icehouse interval spanning the Pennsylvanian-Permian boundary during which time glaciation is thought to have been the most extensive of the LPIA. During this interval, both basins were located at high (> 60˚) southern latitudes along the Panthalassan margin of southeastern Gondwana. The similarities in paleogeographic, geologic, and temporal contexts between the Transantarctic and Tasmanian basins mean that characterizing and comparing LPIA glaciations in both areas is critical to understanding the late Paleozoic glacial maximum at polar latitudes. The works presented in this dissertation demonstrate that building an accurate, nuanced understanding of global glaciations during the LPIA, requires beginning at the local scale and building outward. Chapter 2 examines the Pagoda Formation of the Transantarctic Basin at four locations in the Shackleton Glacier Region of Antarctica. The dominant lithology in the Pagoda Fm at those locations is a massive, sandy, clast-poor diamictite. Depositional processes governing these diamictites were proglacial, subaqueous glacial processes, likely a combination of mass transport, iceberg rain-out, iceberg scouring, plume sedimentation, and subglacial till deposition. Some of the deposits are part of grounding-line fan systems. All glacigenic sediments in the Pagoda Fm at these locations were likely deposited during the retreat phase of a single, up to 90 m thick glacial sequence. Flow directions from these successions support the hypothesis that an ice center was present toward the Panthalassan margin of East Antarctica (Marie Byrd Land) during the LPIA. Chapter 3 describes the basal 415 m of the type section of the Wynyard Formation of the Tasmanian Basin, which outcrops along the coast of northwestern Tasmania. Facies associations in this succession include muddy massive diamictite, sandy massive diamictite, and rhythmically laminated fine-grained facies. Respectively, these sediments were deposited as a grounding-zone wedge, proglacial, proximal grounding line fan or morainal bank, and proglacial, glacier-distal cyclopelites. In this succession, the basal Wynyard Fm was deposited in glacier-proximal to glacier-distal, marine environments on a continental shelf at water depths below storm wave base. All facies associations contain mass transport and turbidite deposits that could have been driven by slope instability due to rapid deposition. The “Wynyard Glacier” was most likely an outlet glacier or ice stream draining a large ice cap or ice sheet. Chapter 4 is a detrital zircon geochronology provenance study of sandstones from the Wynyard Formation. These data represent the first such measurements from the Wynyard Formation anywhere in the basin. With these data, and using a “local first” approach, we demonstrated that all measured detrital zircon dates from the Wynyard Fm can be attributed to zircon sources that occur within 33 km of the sample location along the glacier’s flow path. Therefore, while the detrital zircon provenance signature of the Wynyard Fm also supports the hypothesis that the Wynyard Glacier flowed from south to north, this information does not impart insight into where the ice center was nucleated.
... In contrast, more selective applications of proxies have determined that the Antarctic diamictites occurred in a wide variety of environmental settings representing erosion and deposition by subglacial (sliding, deforming, and lodgement), rockfall, glaciomarine (grounding line, morainal bank, meltwater plume, iceberg and sea/lake ice rafting; iceberg turbation), sediment gravity flow, and mass transport processes (e.g., Ojakangas, 1991, 1992;Lenaker, 2002;Isbell et al., 2008Isbell et al., , 2012Isbell, 2010;Koch and Isbell, 2011;Cornamusini et al., 2017;Ives and Isbell, 2021). In this scenario, grooved and striated surfaces resulted from subglacial ploughing/sliding, turbation by icebergs, and as mass transport glide planes (Isbell et al., 2008). ...
... Such studies contributed to the identification of multiple ice sheets, that supplied sediment and ice to the basin (Figs. 1C, 2; Isbell et al., 2008;Isbell, 2010;Elliot et al., 2016;Ives and Isbell, 2021;Zurli et al., 2021). Ice transversely entered the trough-shaped basin from ice sheets located on the East Antarctic Craton and in Marie Bryd Land with subglacial and ice proximal glaciomarine sedimentation occurring along basin margins while predominantly meltwater plume, iceberg rafted, sediment gravity flow and mass transport deposition occurred along the basin axis (Lenaker, 2002;Isbell et al., 2008;Isbell, 2010;Ives and Isbell, 2021). ...
... 1C, 2; Isbell et al., 2008;Isbell, 2010;Elliot et al., 2016;Ives and Isbell, 2021;Zurli et al., 2021). Ice transversely entered the trough-shaped basin from ice sheets located on the East Antarctic Craton and in Marie Bryd Land with subglacial and ice proximal glaciomarine sedimentation occurring along basin margins while predominantly meltwater plume, iceberg rafted, sediment gravity flow and mass transport deposition occurred along the basin axis (Lenaker, 2002;Isbell et al., 2008;Isbell, 2010;Ives and Isbell, 2021). This emerging view contributed to the idea that multiple ice sheets occurred scattered across Antarctica and Gondwana (Isbell et al., 2003(Isbell et al., , 2012Fielding et al., 2008a;Montañez and Poulsen, 2013;Rosa and Isbell, 2021). ...
Article
The late Paleozoic Ice Age (LPIA) was one of Earth's most important Phanerozoic climatic events lasting for over 100 Mys. Despite its importance, its history is controversial with two hypotheses that portray glaciation differently (Fig. 1). Traditional views characterize the LPIA as a continuous glacial event that lasted from the Middle Mississippian until the Late Permian with a massive ice sheet that covered Gondwana throughout this interval. This approach often uses only one or two proxies to define the glaciation. The other emerging hypothesis suggests that numerous ice sheets occurred in Gondwana with individual glacial events lasting up to 10 Mys alternating with glacial minima/non-glacial intervals of similar duration. Both views are still prevalent. Both near and far-field proxies are used to define the ice age. Near-field proxies include the occurrence/absence of diamictites, glaciotectonic deposits/landforms, striated clasts and clast pavements, outsized clasts (dropstones), rhythmites, cyclic diamictite-bearing successions, glendonites, grooved and striated surfaces, streamline landforms, and U-shaped paleovalleys. Detrital zircons and chemical index of alteration (CIA) studies help to delineate the occurrence, extent, and location of glaciation. Multiple complexities occur with the use of these proxies as different non-glacial processes and driving factors can produce similar features or results. Far-field proxies focus on identifying changes in eustacy. These include the occurrence of cyclic successions composed of alternating nonmarine and marine strata (cyclothems), depth of incised valleys, paleotopographic relief, phosphatic black shales, and changing oxygen isotope ratios. Like the near-field record, far-field proxies are complex indicators with varied nuances that make their application challenging. Here we discuss the limitations and use of these proxies and promote a multiproxy approach to investigating Earth's glacial intervals. We suggest that studies incorporate multiple proxies coupled with detailed environmental, paleoflow, and paleogeographic analyses to better constrain the occurrence, timing, and extent of glaciation and its influence on global systems. This approach will provide a robust view of the LPIA. We also consider the magnitude and nature of sea-level response to changing ice volumes by discussing ice-volume fluctuations, basin subsidence's modification of glacioeustacy, and sea-level's response to global isostatic adjustment (GIA). In considering these features, it becomes apparent that glacioeustacy is more complex than previously envisioned.
... Zurli et al., 2022). Central Transantarctic Mountains stratigraphy is modified from Elliot, 2013;Elliot et al., 2017;Ives and Isbell, 2021. Southern Victoria Land stratigraphy is modified from Elliot, 2013, Liberato et al., 2017, and Ives and Isbell, 2021. ...
... Central Transantarctic Mountains stratigraphy is modified from Elliot, 2013;Elliot et al., 2017;Ives and Isbell, 2021. Southern Victoria Land stratigraphy is modified from Elliot, 2013, Liberato et al., 2017, and Ives and Isbell, 2021. Northern Victoria Land stratigraphy is modified from Cornamusini et al., 2017 andBomfleur et al., 2021 (2020) are in accordance. ...
Article
Full-text available
Between Permian to Triassic, the Earth experienced climatic and biotic crises, included the greatest mass extinction at the Permian-Triassic boundary. These climatic and biological changes are reflected in both marine and terrestrial depositional systems. Over this time span, the Gondwana supercontinent hosted numerous large basins that may preserve the paleoenvironment response to global changes in the sedimentary record. This study provides a lithostratigraphic reappraisal of the latest Paleozoic-Mesozoic alluvial Beacon Supergroup at Allan Hills (Convoy Range), which is one of the most complete sedimentary sequences in Antarctica. Fieldwork stratigraphic-lithological observation, facies analysis, and petrographic characterization of sedimentary rocks allow the identification of six depositional units. The investigations point out for a conformable relationship between depositional and lithostratigraphic units, characterized by changes in the fluvial style. The reconnaissance of a “transitional interval” showing intermediate features between the Permian Weller Coal Measures and the Triassic Feather Conglomerate strengthen the conformable nature of the sequence across the Permian-Triassic boundary in this region. The lithological features of such interval strongly resemble those observed in the coeval deeply studied Eastern Australia successions crossing the Permian-Triassic boundary as well as the end-Permian extinction. More precisely, the uppermost coal occurrence, just above a glossopterid macroflora-bearing carbonaceous mudstone within the “transitional interval”, marks the disappearance of coal-peat forming Permian vegetation which corresponds with the terrestrial end-Permian extinction, thus representing one of the few end- Permian extinction records in Antarctica.
... In the Transantarctic Basin, Isbell et al. (2012) demonstrated the occurrence of ice centers in the eastern and western flanks of the basin draining toward the depocenter, even if the basin morphology was articulated and local highs could exist. In the central TAM, tillite, known as Pagoda Formation, deposited in a glaciomarine setting (Isbell et al., 2012;Ives & Isbell, 2021) as well as in SVL (Metschel Tillite;Isbell, 2010;Zurli et al., 2022b); on the contrary NVL recorded the glaciolacustrine deposition of the Lanterman Formation dated to Asselian (Cornamusini et al., 2017), while glaciomarine conditions occurred in Tasmania (Zurli et al., 2022a). ...
Article
The Late Paleozoic Ice Age (LPIA) was one of the most severe glacial phases of Phanerozoic, which led to the development of ice centers within Gondwana. The sedimentary record of the Beacon Supergroup preserves strata related to early Permian ice dynamics. Here we document outcrops across the northern and southern Victoria Land basins in the Transantarctic Mountains (Antarctica); differences come out from the two regions, emphasizing the articulated physiography of the basin and the typology and size of the ice centers. The most common facies are massive to crudely stratified diamictite, sometimes interlayered with mudstone strata. Tillites deposited in a subglacial to ice proximal setting within a glaciomarine to glacio-continental environment, recording climate pulsating waxing and waning ice extent. Diamictite were petrographically analyzed to characterize the mineralogical composition of coarse-grained fraction. Macroscopic and thin section analysis have been carried out on cobbles to determine their petrographic features. The latter reveal that most of them were sourced from the local crystalline basement and from the Devonian sedimentary strata. Clasts composition, together with other provenance tools and paleo-ice flow indicators, allow reconstructing the geological setting of the Permian Victoria Land Basin and the paleo-ice dynamics and ice extent during LPIA in Antarctica.
... Soliani, 1982Soliani, , 1997López-Gamundí et al., 2016;Le Heron et al., 2019Rosa et al., 2019;Isbell et al., 2021) versus by grounded or floating ice (iceberg and sea/lake ice keel scour marks; cf. Woodworth-Lynas and Dowdeswell, 1994;Vesely and Assine, 2014;Rosa et al., 2016;Isbell et al., 2021;Ives and Isbell, 2021) are not fully acknowledged. Such recognition of floating ice turbation in the LPIA record often results in notable reinterpretations of the glacial history for an area, including the local occurrence or absence of glaciers (Woodworth-Lynas and Dowdeswell, 1994;Vesely and Assine, 2014;Rosa et al., 2016;Vesely, 2020), thus dramatically reshaping our understanding of the paleogeographic and paleoclimatic history of Gondwana. ...
Article
Whalebacks, roche moutonnées, and S-forms carved on Ediacaran granitoids near Cerro de las Cuentas, Uruguay, along with overlying diamictites, siltstones, and sandstones displaying soft-sediment grooved and striated surfaces in the Pennsylvanian San Gregorio Formation, record the glacial to post-glacial transition in the linked Norte, southern Paraná and Chaco-Paraná basins of Uruguay, Brazil, and Argentina respectively. Early authors reported these features resulted from subglacial abrasion and deposition as lodgement tills and glaciotectonites. Our re-examination reveals a nuanced setting with changing ice thicknesses, subglacial kinematics, and ice proximal glaciomarine dynamics associated with advance and retreat of an ice stream, or multiple advances of the Uruguayan Ice lobe, during glaciation of the Late Paleozoic Ice Age (LPIA) in these basins. The preserved landforms indicate temperate glacial conditions. Whalebacks form under 1.6 to 2.5 km-thick ice and likely formed when the lobe extended across the Uruguayan and Rio Grande do Sul shields into the adjacent Paraná Basin. Previously unidentified m-scale roches moutonnées cut into one whaleback developed under thinner ice where reduced basal pressure allowed for the opening of air and water-filled cavities, thus facilitating quarrying on the lee side of basement bumps. S-forms provide additional evidence for the occurrence of subglacial waters, indicating that the basal ice was at or above its pressure melting point. The lower meter of the overlying strata consists of interstratified trace fossil-bearing, laminated siltstones; thin-bedded diamictites; and current-rippled sandstones. Trace fossils belonging to the Mermia ichnofacies within the basal siltstones, as well as acritarchs in the overlying siltstones, suggest that these sediments were deposited in ice-proximal subaqueous settings with contributions from meltwater discharge. Graded siltstone laminae suggest settling from suspension likely from meltwater plumes, while thin-bedded diamictites were deposited either as debris flows or as two-component sedimentation with fines settling from suspension and coarser particles introduce as iceberg-rafted dropstones. Current-rippled sandstones indicate the occurrence of underflow currents. Soft-sediment troughs, grooves, and striations cutting these sediments display curved and sinuous paths with some features oriented perpendicular, and one oriented opposite to the overall trend. They contain marginal and terminal berms typical of iceberg scour marks suggesting transit across the area by icebergs calving from a tidewater ice front located to the SE.
... Given the known relief of the Kukri and Maya erosion surfaces, which is as much as 700 m (Isbell 1999), it is here suggested that the occurrence of granite at Cascade Bluff is the result of palaeotopography on the combined erosion surface, relief which needs to have been no more than 200-300 m. In the central TAM (Figure 2), the Pagoda Formation at Mount Munson just north of Mount Wade ( Figure S3) (Ives and Isbell 2021) is only a few metres thick and overlies basement granite, is absent at Mount Bowers 100 km to the west (Figure 2) (Elliot et al. 2014), and near Clarkson Peak 200 km to the northwest (Figure 2) is only a veneer on Devonian sandstones (Elliot and Isbell 2021). ...
Article
Full-text available
Only at Cape Surprise, central Transantarctic Mountains, is there exposed stratigraphic evidence for major offset along the range front, which marks a major boundary in Antarctica. Several faults parallel to the range front have been identified in the Devonian to Triassic Gondwana strata in the hinterland. Analysis of the stratigraphy based on field observations and the United States Geological Survey (USGS) aerial photographs, in conjunction with USGS topographic sheets and satellite-derived elevation measurements, suggests an array of faults with varying orientations and displacements. Fault offsets range up to an estimated 850 metres. No additional range-parallel faults have been identified and no clear pattern of faulting is evident in the hinterland of the frontal escarpment. Faulting may date from the time of slow uplift during the Cretaceous as well as the more rapid Cenozoic uplift of the range. Only a few faults in the hinterland can be allied with the fontal fault system. Cenozoic uplift and associated denudation was accompanied by glaciation of Antarctica, which is documented by Sirius Group strata. These deposits, which pre-date today’s polar landscape, are older than mid Miocene, and in part may date from the earliest stages of warm-based glaciation in the early Oligocene.
Article
Full-text available
For more than 150 years, geologic characteristics claimed to be evidence for pre-Pleistocene glaciations have been debated. Advancements in recent decades, in understanding features generated by mainly glacial and mass flow processes, are here reviewed. Detailed studies of data offered in support of pre- Pleistocene glaciations have led to revisions that involve environments of mass movements. Similarities and differences between Quaternary glaciogenic and mass movement features are examined, to provide a more systematic methodology for analysing the origins of more ancient deposits. Analyses and evaluation of data are from a) Quaternary glaciogenic sediments, b) formations which have been assigned to pre-Pleistocene glaciations, and c) formations with comparable features associated with mass movements (and occasionally tectonics). Multiple proxies are assembled to develop correct interpretations of ancient strata. The aim is not per se to reinterpret specific formations and past climate changes, but to enable data to be evaluated using a broader and more inclusive conceptual framework. Regularly occurring pre-Pleistocene features interpreted to be glaciogenic, have often been shown to have few or no Quaternary glaciogenic equivalents. These same features commonly form by sediment gravity flows or other non-glacial processes, which may have led to misinterpretations of ancient deposits. These features include, for example, environmental affinity of fossils, grading, bedding, fabrics, size and appearance of erratics, polished and striated clasts and surfaces (“pavements”), dropstones, and surface microtextures. Recent decades of progress in research relating to glacial and sediment gravity flow processes have resulted in proposals by geologists, based on more detailed field data, more often of an origin by mass movements and tectonism than glaciation. The most coherent data of this review, i.e., appearances of features produced by glaciation, sediment gravity flows and a few other geological processes, are summarized in a Diamict Origin Table. Keywords Diamictite, Tillite, Sediment gravity flow (SGF), Striation, Groove, Dropstone, Paleoclimate, Fossil vegetation, Glaciogenic proxies, Surface microtexture, Late Paleozoic ice age
Article
Full-text available
Sedimentary processes are known to help facilitate tidewater glacier advance, but their role in modulating retreat is uncertain and poorly quantified. In this study we use repeated seafloor bathymetric surveys and satellite‐derived terminus positions from LeConte Glacier, Alaska, to evaluate the evolution of a morainal bank and related changes in terminus dynamics over a 17‐year period. The glacier experienced a rapid retreat between 1994 and 1999, before stabilizing at a constriction in the fjord. Since then, the glacier terminus has remained stabilized while constructing a morainal bank up to 140 m high in water depths of 240–260 m, with rates of sediment delivery of 3.3 × 10⁵ to 3.8 × 10⁵ m³ a⁻¹. Based on repeated interannual surveys between 2016 and 2018, the moraine is a dynamic feature characterized by push ridges, evidence of active gravity flows, and bulldozing by the glacier at rates of up to meters per day. Beginning in 2016, the summertime terminus has become increasingly retracted, revealing a newly emerging basin potentially signaling the onset of renewed retreat. Between 2000 and 2016, the growing moraine reduced the exposed submarine area of the terminus by up to 22%, altered the geometry of the terminus during seasonal advances, and altered the terminus stress balance. These feedbacks for calving, melting, and ice flow likely represent mechanisms whereby moraine growth may delay glacier retreat, in a system where readvance is unlikely.
Article
Full-text available
The Transantarctic Mountains (TAM) are one of Earth's great mountain belts and are a fundamental physiographic feature of Antarctica. They are continental-scale, traverse a wide range of latitudes, have high relief, contain a significant proportion of exposed rock on the continent, and represent a major arc of environmental and geological transition. Although the modern physiography is largely of Cenozoic origin, this major feature has persisted for hundreds of millions of years since the Neoproterozoic to the modern. Its mere existence as the planet's longest intraplate mountain belt at the transition between a thick stable craton in East Antarctica and a large extensional province in West Antarctica is a continuing enigma. The early and more cryptic tectonic evolution of the TAM includes Mesoarchean and Paleoproterozoic crust formation as part of the Columbia supercontinent, followed by Neoproterozoic rift separation from Laurentia during breakup of Rodinia. Development of an Andean-style Gondwana convergent margin resulted in a long-lived Ross orogenic cycle from the late Neoproterozoic to the early Paleozoic, succeeded by crustal stabilization and widespread denudation during early Gondwana time, and intra-cratonic and foreland-basin sedimentation during late Paleozoic and early Mesozoic development of Pangea. Voluminous mafic volcanism, sill emplacement, and layered igneous intrusion are a primary signature of hotspot-influenced Jurassic extension during Gondwana breakup. The most recent phase of TAM evolution involved tectonic uplift and exhumation related to Cenozoic extension at the inboard edge of the West Antarctic Rift System, accompanied by Neogene to modern glaciation and volcanism related to the McMurdo alkaline volcanic province. Despite the remote location and relative inaccessibility of the TAM, its underlying varied and diachronous geology provides important clues for reconstructing past supercontinents and influences the modern flow patterns of both ice and atmospheric circulation, signifying that the TAM have both continental and global importance through time.
Article
Full-text available
The subglacial environment of the Greenland Ice Sheet (GrIS) is poorly constrained both in its bulk properties, for example geology, the presence of sediment, and the presence of water, and interfacial conditions, such as roughness and bed rheology. There is, therefore, limited understanding of how spatially heterogeneous subglacial properties relate to ice-sheet motion. Here, via analysis of 2 decades of radio-echo sounding data, we present a new systematic analysis of subglacial roughness beneath the GrIS. We use two independent methods to quantify subglacial roughness: first, the variability in along-track topography – enabling an assessment of roughness anisotropy from pairs of orthogonal transects aligned perpendicular and parallel to ice flow and, second, from bed-echo scattering – enabling assessment of fine-scale bed characteristics. We establish the spatial distribution of subglacial roughness and quantify its relationship with ice flow speed and direction. Overall, the beds of fast-flowing regions are observed to be rougher than the slow-flowing interior. Topographic roughness exhibits an exponential scaling relationship with ice surface velocity parallel, but not perpendicular, to flow direction in fast-flowing regions, and the degree of anisotropy is correlated with ice surface speed. In many slow-flowing regions both roughness methods indicate spatially coherent regions of smooth beds, which, through combination with analyses of underlying geology, we conclude is likely due to the presence of a hard flat bed. Consequently, the study provides scope for a spatially variable hard- or soft-bed boundary constraint for ice-sheet models.
Article
Full-text available
Along the easternmost edge of the Karoo–Kalahari Basin (KKB) of Botswana, the Toutswemogala Hill succession exposes a 30–50-m-thick suite of siliciclastic deposits interpreted by some as glaciogenic in origin tied to the Permo-Carboniferous Late Paleozoic Ice Age (LPIA). Six facies associations (FA) were recognized in this succession, resting unconformably on a highly uneven Archean gneissic basement, and consisting from base to top of: 1) clast-supported breccia made up of angular cobbles and boulders ubiquitously derived from the underlying basement, 2) well-bedded siltstones sealing or locally interdigitated with the underlying breccia, and bearing abundant remnants of Glossopteris sp. leaves, 3) a chaotic to faintly laminated matrix-supported diamictite bearing angular and subrounded clasts and tree logs attributed to the genus Megaporoxylon, 4) cross-bedded conglomerate bearing well-rounded quartz and clasts, 5) planar-laminated to ripple-laminated, poorly sorted, muddy sandstones showcasing dispersed mud chips that grade upward into 6) poorly sorted, cross-bedded coarse-grained sandstones displaying convolute beds and abundant imprints of unidentifiable tree logs. No evidence of glaciogenic processes have been found in this succession, in the form of either pavement or clasts striations. The breccia and diamictite are interpreted as scree and mass-flow deposits, respectively. Along with the age of the deposits, inferred from the plant debris (upper Carboniferous to lower Permian), the stratigraphic position of this sedimentary succession resting on the Archean basement suggests that it corresponds to the Dukwi Formation, a stratigraphic equivalent of the Dwyka Group in the Main Karoo Basin. This would explain the resemblance of the facies to those recovered at the base of the central Kalahari–Karoo Basin and in the neighboring Tuli, Ellisras, and Tshipise basins. The absence of diagnostic criteria for glacial processes in the studied succession raises the question of the extent, in both time and space, of the LPIA-related ice masses over southern Africa and particularly in southeastern Botswana. It is suggested here that during this glacial epoch, spatially restricted ice masses were confined in bedrock valleys (valley glaciers) in an uplifted setting otherwise characterized by non-glaciogenic processes, further strengthening the scenario of fragmented ice masses over southern Gondwana.
Article
Full-text available
In the eastern part of the Karoo Basin of South Africa, the sedimentary record of the Late Palaeozoic Ice Age, the Dwyka Group, consists of an up to 200 m thick accumulation of massive to crudely‐stratified diamictite occasionally interstratified with siltstone, sandstone and conglomerate horizons. Three distinct sedimentary units, separated by intervening glacial erosion surfaces, are viewed as ice‐margin fluctuation sequences. The lowermost one, resting on highly uneven, glacially‐abraded Archaean basement, has been interpreted as a grounding zone wedge deposited after the retreat and stabilization of the ice margin after the inundation of the Karoo Basin. The grounding zone wedge interpretation is based on its thickness (up to 100 m), the dominance of diamictite, and its facies assemblage and inferred depositional processes (rain‐out of debris, dropstone dumping, mass and debris flow, till). Overlying the grounding zone wedge deposits are sedimentary units interpreted as glaciofluvial or ice‐contact delta and grounding zone wedges, respectively. By analogy with Quaternary sedimentary sequences, deposition of the Dwyka Group in the study area might have been very rapid (tens to hundreds of thousand years) and may hence correlate with the ultimate deglacial sequence of the Western Karoo Basin, as both successions are covered by the postglacial Ecca Group. Although commonly observed and imaged on modern, high‐latitude continental shelves, grounding zone wedges have never been interpreted in the ancient geological record. This paper therefore outlines a model defining criteria necessary for identifying grounding zone wedges. This article is protected by copyright. All rights reserved.
Article
The southwestern margin of South America offers a complete record of the Late Paleozoic Ice Age (LPAI) that affected the Gondwana supercontinent. The tripartite division of LPIA glacial episodes has been refined with the help of new radiometric dates and biostratigraphic (flora and fauna) zonations in recent years to five shorter-lived discrete events: 1. Latest Devonian-earliest Tournaisian, 2. Tournaisian, 3. Visean, 4. Serpukhovian – Early Bashkirian, 5. late Pennsylvanian-earliest Permian. The glacial events 1, 2 and 3, and 4 are capped by postglacial transgressive deposits with marine fauna. The unbalanced preservation potential of the glacial deposits, skewed toward the glaciomarine sediments, provides an uneven stratigraphic record with few cases of continental glacial sedimentation, confined to the Serpukhovian – Early Bashkirian event, and numerous examples of glacial sedimentation in marine environments. Glacial sedimentation in marine settings has been grouped in two main facies associations: a valley-glacier-retreat (fjord) facies association and a submarine-retreat (glaciomarine apron) facies association in open-marine areas. Transitional facies, correspondent to those formed by the flooding of valleys during postglacial transgressions, are widely distributed along the Protoprecordillera, where paleofjord successions are well exposed particularly in western Paganzo Basin and mapped in subsurface in the Tarija basin. A general paleofjord model includes (from base to top) the following stages: (i) Incision of paleovalley and deposition of subglacial diamictites in ice contact deltas, (ii) Early Transgressive stage characterized by resedimentation of subglacial material by subaqueous sediment gravity flows and slumps in proglacial settings, (iii) Maximum flooding (late transgressive stage) dominated by black shales or laminated mudstones related to a marine incursion that flooded valleys; normal marine or brackish conditions may dominate this stage and (iv) Highstand: progradation of a fluvial-deltaic system including in some cases Gilbert-type deltas. In glaciomarine apron environments, the basal facies includes massive clast-supported conglomerates, with few striated and polished clasts, followed by fining-upward successions including thinly bedded diamictites with ice-rafted debris (IRD) and locally contorted sandstone masses in diamictite beds, indicative of mass-emplacement mechanisms. The presence of inter- and intratill pavements suggests glacial advance/retreat fluctuations along the basin margins. Deglaciation sequences, reflecting deposition mainly during the retreat of ice sheets, ice caps and alpine glaciers and successive deglaciation, can be used as operational tools for the analysis of glacial successions in SW Gondwana. They are characterized as rather simple upward-thinning successions in open marine settings as exemplified in most of the Calingasta-Uspallata Basin, Sauce Grande (Ventana Foldbelt, VFB) and central portions of the Paraná and Karoo basins. In more proximal areas (i.e. paleofjords) this vertical trend is commonly punctuated with deltaic wedges fed by nearby provenance areas. The late Paleozoic glacial-related successions of southwestern Gondwana exhibit a common tripartite motif, equivalent to second-order sequences with estimated durations of 10–80 Myr. The lower section corresponds to glacial and glacially-influenced diamictites; the middle interval is initiated with postglacial transgressions. The lower and middle intervals correspond to the deglaciation sequence as described and identified in several basins of Gondwana. Finally, the upper term includes coastal progradation, followed in some places by continentalization, accompanied in many sectors by increasing aridization. Examples of second-order sequences can be identified in the thick late Paleozoic successions of the Paraná and Karoo basins and in the VFB. Thinner second order sequences can be identified in the Calingasta-Uspallata, Rio Blanco, Paganzo and San Rafael basins. In the Paganzo and San Rafael basins the middle interval is also punctuated by short lived marine ingressions. The basal sequence boundary is commonly an abrasion surface (glacial erosion surface, GES) developed on bedrock. Deglaciation sequences are assigned to third order sequences made up of, when present, of a thin lowstand system tract (LST) of subglacial deposits followed upward by thick glaciomarine and glacially influenced sediments. These facies are part of a thick transgressive systems tract (TST) that culminates with marine shales that reflect interglacial or postglacial conditions during ice retreat. Thus, the deglaciation sequences are proposed to be third order sequences made up of LST-TST or exclusively TST.
Chapter
Glacial deposits provide a long-term record of climate and sea level changes on Earth. Detailed study of sedimentary rocks deposited during and immediately after glacial episodes is paramount to accurate palaeoclimatic reconstructions and for our understanding of global climatic and eustatic changes. This book presents new information and interpretations of the ancient glacial record, looking in particular at the Late Proterozoic and Late Paleozoic eras. The influence of global tectonics on the origins and distribution of ice masses and the character of glacial deposits through geologic time is emphasised. Sequence stratigraphic techniques are applied to glaciogenic successions, and explanations for possible low-latitude glaciation during the Late Proterozoic era and the association of carbonate deposits with glaciogenic rocks are put forward. Early interglacial conditions, represented by dark grey mudrocks and ice keel scour features are discussed. These studies, from key workers in International Geological Correlation Program Project 260, will aid the understanding of the Earth's climatic history.
Article
Glaciogenic sediments are present in many hydrocarbon-producing basins across the globe but their complex nature makes it difficult to characterise the reservoir-quality sedimentary units. Despite this, Ordovician glacial deposits in North Africa, and Carboniferous-Permian glaciogenic sequences in the Middle East, have been proven to host significant, economical, hydrocarbon accumulations. Additionally, discoveries have been made in the shallow (<1000 m below seabed), glacial, Pleistocene sedimentary succession of the North Sea (e.g. Peon and Aviat). This paper provides a predictive exploration framework in the form of a conceptual model of glaciogenic sediment-landform distributions. The model is based on the extensive onshore glacial sedimentary record integrated with available offshore data. It synthesises the published knowledge, drawing heavily on glacial landsystem models, glacial geomorphology and sedimentology of glaciogenic deposits to provide a novel conceptual model allowing for the efficient description and interpretation of glacial sediments and landforms in the subsurface. Subsequently, land-terminating and water-terminating ice sheet depositional systems are described and discussed, with respect to ice advance and retreat cycles. This detailed description focuses on the macro-scale stratigraphic organisation of glacial sediments with relation to the ice margin, aiding the prediction of glaciogenic sediment distributions, and their likely geometry, architecture and connectivity as reservoirs.