ArticlePDF Available

Crustal heat production and estimate of terrestrial heat flow in central East Antarctica, with implications for thermal input to the East Antarctic ice sheet

Authors:

Abstract and Figures

Terrestrial heat flow is a critical first-order factor governing the thermal condition and, therefore, mechanical stability of Antarctic ice sheets, yet heat flow across Antarctica is poorly known. Previous estimates of terrestrial heat flow come from inversion of seismic and magnetic geophysical data, by modeling temperature profiles in ice boreholes, and by calculation from heat production values reported for exposed bedrock. Although accurate estimates of surface heat flow are important as an input parameter for ice-sheet growth and stability models, there are no direct measurements of terrestrial heat flow in East Antarctica coupled to either subglacial sediment or bedrock. Bedrock outcrop is limited to coastal margins and rare inland exposures, yet valuable estimates of heat flow in central East Antarctica can be extrapolated from heat production determined by the geochemical composition of glacial rock clasts eroded from the continental interior. In this study, U, Th and K concentrations in a suite of Proterozoic (1.2–2.0 Ga) granitoids sourced within the Byrd and Nimrod glacial drainages of central East Antarctica indicate average upper crustal heat production (Ho) of about 2.6 ± 1.9 μW m⁻³. Assuming typical mantle and lower crustal heat flux for stable continental shields, and a length scale for the distribution of heat production in the upper crust, the heat production values determined for individual samples yield estimates of surface heat flow (qo) ranging from 33–84 mW m⁻² and an average of 48.0 ± 13.6 mW m⁻². Estimates of heat production obtained for this suite of glacially-sourced granitoids therefore indicate that the interior of the East Antarctic ice sheet is underlain in part by Proterozoic continental lithosphere with average surface heat flow, providing constraints on both geodynamic history and ice-sheet stability. The ages and geothermal characteristics of the granites indicate that crust in central East Antarctica resembles that in the Proterozoic Arunta and Tenant Creek inliers of Australia, but is dissimilar to other areas characterized by anomalously high heat flow in the Central Australian Heat Flow Province. Age variation within the sample suite indicates that central East Antarctic lithosphere is heterogeneous, yet the average heat production and heat flow of four age subgroups cluster around the group mean, indicating minor variation in thermal contribution to the overlying ice sheet from upper crustal heat production. Despite their minor differences, ice-sheet models may favor a geologically realistic model of crustal heat flow represented by such a distribution of ages and geothermal characteristics.
Content may be subject to copyright.
1
Crustal heat production and estimate of terrestrial heat flow in central East
Antarctica, with implications for thermal input to the East Antarctic ice sheet
John W. Goodge1
1Department of Earth and Environmental Sciences, University of Minnesota, Duluth, MN 55812 USA
Correspondence to: John Goodge (jgoodge@d.umn.edu) 5
Abstract. Terrestrial heat flow is a critical first-order factor governing the thermal condition and, therefore, mechanical
stability of Antarctic ice sheets, yet heat flow across Antarctica is poorly known. Previous estimates of terrestrial heat flow
come from inversion of seismic and magnetic geophysical data, by modeling temperature profiles in ice boreholes, and by
calculation from heat production values reported for exposed bedrock. Although accurate estimates of surface heat flow are
important as an input parameter for ice-sheet growth and stability models, there are no direct measurements of terrestrial heat 10
flow in East Antarctica coupled to either subglacial sediment or bedrock. Bedrock outcrop is limited to coastal margins and
rare inland exposures, yet valuable estimates of heat flow in central East Antarctica can be extrapolated from heat production
determined by the geochemical composition of glacial rock clasts eroded from the continental interior. In this study, U, Th
and K concentrations in a suite of Proterozoic (1.2-2.0 Ga) granitoids sourced within the Byrd and Nimrod glacial drainages
of central East Antarctica indicate average upper crustal heat production (Ho) of about 2.6 ± 1.9 µW m-3. Assuming typical 15
mantle and lower crustal heat flux for stable continental shields, and a length scale for the distribution of heat production in
the upper crust, the heat production values determined for individual samples yield estimates of surface heat flow (qo)
ranging from 33-84 mW m-2 and an average of 48.0 ± 13.6 mW m-2. Estimates of heat production obtained for this suite of
glacially-sourced granitoids therefore indicate that the interior of the East Antarctic ice sheet is underlain in part by
Proterozoic continental lithosphere with average surface heat flow, providing constraints on both geodynamic history and 20
ice-sheet stability. The ages and geothermal characteristics of the granites indicate that crust in central East Antarctica
resembles that in the Proterozoic Arunta and Tenant Creek inliers of Australia, but is dissimilar to other areas characterized
by anomalously high heat flow in the Central Australian Heat Flow Province. Age variation within the sample suite indicates
that central East Antarctic lithosphere is heterogeneous, yet the average heat production and heat flow of four age subgroups
cluster around the group mean, indicating minor variation in thermal contribution to the overlying ice sheet from upper 25
crustal heat production. Despite their minor differences, ice-sheet models may favor a geologically realistic model of crustal
heat flow represented by such a distribution of ages and geothermal characteristics.
Copyright statement. The author agrees to the copyright terms stated by the journal.
30
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
2
1 Introduction
Heat production and heat flow are fundamental characteristics of continental crust (Rudnick and Fountain, 1995).
Together they provide important constraints on the thermal and petrogenetic history of cratonic lithosphere, and heat flow is
an indicator of modern geodynamic environments. Antarctic lithosphere is uniquely important because it underlies Earth’s
largest ice caps, including numerous subglacial lakes, and it is critical in governing the thermal state and mechanical stability 5
of overlying ice (Pollard et al., 2005; Jamieson and Sugden, 2008; Van Liefferinge and Pattyn, 2013; Schroeder et al., 2014).
Terrestrial heat flow in Antarctica has a strong influence on basal ice temperatures, amount of basal ice at its pressure
melting point, and the formation of liquid water, all of which affect basal ice conditions, mechanical properties of glacial bed
material, degree of basal sliding, erosional effectiveness, and the distribution of hydrologic networks and subglacial lakes
(e.g., Siegert, 2000; Pollard et al., 2005; Pollard and DeConto, 2009). Despite its importance in governing ice-sheet mass 10
balanceand therefore as an input parameter for ice-sheet growth and stability modelsonly a few estimates of conductive
heat flow are available from temperature profiles in Antarctic ice (e.g., Fischer et al., 2013) and none from direct subglacial
measurement. This is particularly problematic for East Antarctica, where the ice cap exceeds 4 km in many areas. In order to
develop accurate models of past ice-sheet behavior and forward models of ice-sheet stability, it is therefore crucial to have
good estimates of terrestrial heat flow from East Antarctica. 15
Continent-wide models for terrestrial heat flow come from both seismological and satellite magnetic data. To address a
lack of direct heat flow measurements in Antarctica, Shapiro and Ritzwoller (2004) modeled surface heat flow by first
correlating seismic velocity data from crust and upper mantle in regions of known heat flow, and then extrapolating these
results to a seismic model of Antarctic lithosphere. Over a broad region of East Antarctica they estimated surface heat flow
to be notably low, uniform, and similar to other old cratons (mostly 35-60 mW m-2). Similarly, Fox Maule et al. (2005) used 20
satellite magnetic data to model the depth to Curie temperature, and then inverted the resulting thermal profile to generate a
distribution of heat flow (see also Purucker, 2012); this modeling likewise predicted heat flow in East Antarctica to be
similar to the results obtained from seismology. In order to evaluate areas that may preserve very old ice, generally requiring
thick, slow-moving ice under cold conditions with low basal heat flux, Van Liefferinge and Pattyn (2013) derived an average
distribution of heat flow from a simple mean of existing geophysical models (Fig. 1); this continent-wide synopsis highlights 25
a relatively uniform pattern of low heat flow in East Antarctica (mostly <55 mW m-2). Using a thermal model that assumes
basal ice temperatures above Antarctic subglacial lakes are equal to the pressure-melting value, Siegert (2000) estimated
geothermal heat flow to vary between 37 and 65 mW m-2, although most estimates for East Antarctica are <60 mW m-2. In
general terms, these different models, despite coarse kernel size, are consistent with one another and indicate heat flow in
most of East Antarctica between about 35-60 mW m-2. 30
In addition to models based on remote geophysical observations, there are also some field-based estimates. Using a
measured temperature profile in the EPICA ice borehole at Dome C, Fischer et al. (2013) derived a geothermal heat flux of
about 54 mW m-2 by fitting a model of heat flow and basal ice melting to the thermal profile. A geological approach was
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
3
taken by Carson et al. (2014), who modeled heat flow using values for heat production estimated from the abundances of
radioactive elements in crustal rocks sampled from outcrop near the Amery Ice Shelf (compiled by Carson and Pittard,
2012). From a coastal transect across rock exposures at Prydz Bay, the resulting profile indicates that heat flow in this area
of Archean to Proterozoic igneous and metamorphic crust is highly variable over a distance of about 200 km, ranging from
an average of 31 mW m-2 in the Vestfold Hills, to 44 mW m-2 in the Rauer Islands, and 55-70 mW m-2 in the area of southern 5
Prydz Bay. Locally, high heat-producing Cambrian granitoids indicate heat flow values as high as ~85 mW m-2. Their model
results thus show variable heat flow governed to first order by the age and type of crust represented, and punctuated by heat
production spikes contributed from Th-rich granitoids. Although much of the area underlying the Rauer Islands and Vestfold
Hills have low heat flow (<50 mW m-2) typical of Proterozoic and Archean crust, Carson et al. (2014) emphasized that some
early Paleozoic granites with anomalously high heat production can cause local elevation of heat flow (>80 mW m-2), as 10
observed in the Central Australian Heat Flow Province (McLaren et al., 2003). Thus, spatially coarse models of heat flow
based on geophysical data across East Antarctica indicate relatively typical continental values ranging from about 35-60 mW
m-2, yet there are some indications of locally elevated heat flow in areas underlain by high heat-producing granites.
Although there are no measurements of terrestrial heat flow obtained directly from subglacial sediment or rock in East
Antarctica, rock clasts eroded from the continental interior and transported to the margin can provide insight into the 15
subglacial geology and, therefore, heat production. In this paper, the concentrations of heat-producing elements in glacial
igneous rock clasts provide a unique opportunity to assess heat flow in the deep continental interior. The major ice drainages
in East Antarctica are marked by nearly radial flow away from central ice divides and domes toward the continental margin
(Fig. 2), providing natural proxy samples of the continental interior by bedrock erosion during glacial flow (Peucat et al.,
2002; Goodge et al., 2008, 2010, 2017). Unique among the major drainages, glacial ice in Byrd Glacier and related smaller 20
drainages moves non-radially from the main ice divides because it is obstructed by the high-standing Transantarctic
Mountains (peak elevations >4000 m). Ice flows through the mountains via channelized outlet glaciers, but it also ablates in
areas where it ramps up against the mountain range (Whillans and Cassidy, 1983). Glacial moraines are formed both along
the margins of the outlet glaciers and where ice is ablating, forming lag deposits adjacent to the mountains. As part of a study
of East Antarctic crustal history, glacial moraines were sampled at sites between the Byrd and Beardmore glaciers (Fig. 2). 25
Ice-velocity fields show that material transported in the greater Byrd system may have been eroded from a broad area of
central East Antarctica, potentially from near the upstream boundary along the major ice divide connecting Dome A and
Dome C.
As proxies for subglacial geology, igneous clasts eroded from central East Antarctica and collected from moraines
adjacent to the central Transantarctic Mountains were dated and analyzed geochemically for major and trace elements30
including the major heat-producing elements U, Th and Kyielding values of radioactive heat production. Surface heat flow
was then estimated by assuming mantle and lower crustal heat-flux contributions and a length scale for reduction in upper-
crustal radiogenic heat production. The results indicate that heat production among most of the samples in this group is
within the range expected for average continental crust and that terrestrial heat flow for a large region of central East
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
4
Antarctica is like that commonly observed in Precambrian shield areas. Estimates from this analysis corroborate previous
models of heat flow in East Antarctica and can be used as first-order constraints in ice-sheet and lithospheric thermal models.
2 Glacial igneous clasts
Igneous rock samples were collected at five sitesArgo Glacier, Lonewolf Nunataks, Milan Ridge, Mt. Sirius, and
Turret Nunatakthat yielded pre-Ross Orogen crystallization ages (>600 Ma), thereby providing evidence regarding 5
composition and age of the Precambrian East Antarctic shield (Fig. 2). At Lonewolf Nunataks, elongate bands of distributed
moraine and ice-matrix debris follow narrow flow lines related to ice movement along the southern margin of Byrd Glacier.
Sites at Milan Ridge and Argo Glacier, directly adjacent to exposed Precambrian basement in the Miller Range, comprise
thin, distributed morainal deposits dominated by a rich variety of crystalline rock types. Sites at Turret Nunatak and Mt.
Sirius are dominated by Gondwanide debris, but they also yielded a small number of distinctive crystalline clasts. Together, 10
these sites yielded a suite of pre-Ross igneous samples that were analyzed for their geochemical composition, age, and
isotopic composition. The igneous clasts consist mainly of intermediate to felsic igneous rocks that represent magmatic
components of the ice-covered East Antarctic craton (Goodge et al., 2017). They are granitic to granodioritic in composition
and contain hornblende, biotite and/or muscovite; some samples are two-mica granitoids of peraluminous composition.
Zircon U-Pb ages from this suite of glacially-transported granitoid clasts show that the crust in central East Antarctica 15
was formed by a series of magmatic events at ~2.01, 1.88-1.85, ~1.79, ~1.57, 1.50-1.41, and 1.20-1.06 Ga (Goodge et al.,
2008, 2010, 2012, 2017). The dominant granitoid populations are ca. 1.85, 1.45 and 1.20-1.06 Ga. None of these igneous
ages are known from the limited outcrop in the region. Samples of metamorphic rock clasts from the same moraines have
similar Proterozoic ages ranging from about 1.1-1.9 Ga (Goodge et al., 2010; Nissen et al., 2013). By comparison with
nearby mountain outcrops, the types and ages of these samples indicate that the crust of central East Antarctica comprises 20
plutonic and metamorphic rocks unlike those seen in the central Transantarctic Mountains (Nimrod Complex ages of ca. 3.1,
2.5, and 1.7 Ga; Goodge et al. (2001); Goodge and Fanning (2016)). Likewise, they are different from igneous and
metamorphic rocks exposed at the Terre Adélie coast (ages of ca. 2.4 and 1.7 Ga; Oliver and Fanning (1997)), although one
population is similar in age to ~1.6 Ga glacial clasts sampled by Peucat et al. (2002). As shown in Figure 2, the glacially-
eroded igneous clasts discussed here may also sample the Gamburtsev Subglacial Mountains, which is thought to have 25
nucleated growth of the East Antarctic Ice Sheet (DeConto and Pollard, 2003; Bo et al., 2009; Rose et al., 2013).
3 Analytical methods
Bulk-rock X-ray fluorescence (XRF) and inductively-coupled plasma mass spectrometry (ICPMS) analyses of major
and trace element compositions were completed in the GeoAnalytical Lab at Washington State University (Goodge et al.,
2017). Prior to analysis, fresh chips of each sample were hand-picked and a standard amount (approximately 28 g) was 30
ground in a swing mill with tungsten carbide surfaces for 2 minutes. For XRF analysis of major elements, 3.5 g of sample
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
5
powder was weighed into a plastic mixing jar with 7 g of spec pure dilithium tetraborate (Li2B4O7). The mixed powders were
emptied into graphite crucibles and loaded into a muffle furnace for fusion at 1000 °C. After removing from the oven to
cool, each bead was reground in the swing mill and the resulting glass powders were replaced in the graphite crucibles and
refused for 5 minutes, then cooled to form a glass bead. Their lower flat surfaces were then ground on 600 silicon carbide
grit and finished briefly on a glass plate to remove any metal from the grinding wheel. The concentrations of 29 elements in 5
the unknown samples were measured on a ThermoARL Advant’XP+ sequential X-ray fluorescence spectrometer by
comparing the X-ray intensity for each element with the intensity obtained from USGS standard samples (PCC-1, BCR-1,
BIR-1, DNC-1, W-2, AGV-1, GSP-1, G-2, and STM-1, using the values recommended by Govindaraju (1994) and beads of
pure vein quartz used as blanks for all elements except Si. Twenty standard beads are routinely run and used for recalibration
approximately once every three weeks or after the analysis of about 300 unknowns. The intensities for all elements were 10
corrected for line interference and absorption effects due to all the other elements using the fundamental parameter method.
For trace elements, powdered samples were mixed with 2 g of Li2B4O7 flux, placed in a carbon crucible and fused at
1000 °C in a muffle furnace for 30 minutes. After cooling, the resultant fusion bead was briefly ground in a carbon-steel ring
mill and a 250 mg portion was weighed into a 30 ml, screw-top Teflon PFA vial for dissolution in water, HNO3, H2O2, and
HF and warmed on a hot plate until a clear solution was obtained. Samples were then diluted to a final weight of 60 g with 15
de-ionized water. Solutions were analyzed for 27 elements on an Agilent model 4500 ICPMS and were diluted an additional
10x at the time of analysis using Agilent’s Integrated Sample Introduction System (ISIS). This yielded a final dilution factor
of 1:4800 relative to the amount of sample fused. Instrumental drift was corrected using Ru, In, and Re as internal standards,
and applying a linear interpolation between In and Re to compensate for mass-dependent differences in the rate and degree
of instrumental drift. Isobaric interferences of rare-earth and other oxides were optimized with correction factors using 20
mixed-element solutions. Standardization was accomplished by analyzing duplicates of three in-house rock standards
interspersed within each batch of 18 unknowns.
4 Heat production and estimated heat flow
4.1 Sample geochemical characteristics
Major, trace and rare-earth element geochemical data show that the granitoid samples are Si-rich and have >65 wt% 25
SiO2, and many have SiO2 = 70-75 wt%. Trace-element compositions show that the samples have enriched light rare-earth
element (LREE) and depleted heavy rare-earth element (HREE) abundances relative to chondrites, and they are enriched in
large ion lithophile (LIL) elements and slightly depleted in high field-strength (HFS) elements relative to mid-ocean ridge
basalt (MORB). Their trace and rare-earth element signatures are quite similar to modern continental-margin magmatic arc
systems (e.g., Cascades, Andes) or evolved volcanic arcs, and they show very similar patterns and abundances as magmas 30
interacting with thick crust (e.g., Davidson et al., 1990, 1991; Wörner et al., 1994). Some of the samples resemble Si-rich,
peraluminous leucogranites found in regions of over-thickened continental crust (Frost et al., 2001). In broad terms, then, the
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
6
trace-element compositions indicate that the melts that produced these igneous rocks interacted with thick, evolved
continental crust, but that they are dissimilar generally from intraplate granitoids.
4.2 Heat-producing elements
The concentrations of the heat-producing elements U, Th and K in 18 granitoid samples are listed in Table 1. The
concentration of U is generally low, ranging up to about 6 ppm (mean = 2.0). Thorium ranges widely, from 1-98 ppm (mean 5
= 23.7), and K similarly varies from 0.5-8 wt% K2O (mean = 4.34).
4.3 Heat production
The geochemical compositions of igneous rocks can be used to determine crustal heat production based on their
concentrations of radioactive elements. Heat production (Ho) was calculated for these clast samples based on rock density
and concentrations of U, Th and K by applying two different algorithms: 10
Ho = 10-2 * ρ * (9.67CU + 2.63CTh + 3.48CK) (1)
Ho = ρ * 0.9928CU*H(238U) + 0.0071CU*H(235U) + CTh*H(Th) + 1.19x10-4CK40*H(40K) (2)
where Ho is surface heat production (µW m-3), ρ is density (kg m
3), CU is the concentration of U (ppm), CTh is the
concentration of Th (ppm), CK is the concentration of K2O (wt%), H(238U) is the heat production from the isotope 238U
(9.37x10-5 W kg-1), H(235U) is the heat production from the isotope 235U (5.69x10-4 W kg-1), H(Th) is the heat production 15
from the isotope 232Th (2.69x10-5 W kg-1), and H(40K) is the heat production from the isotope 40K (2.79x10-5 W kg-1). Method
1 was calculated from the formula of Rybach (1976, 1988), using values from Hasterok and Chapman (2011). Method 2 uses
the formulation of Turcotte and Schubert (2014). Density (ρ) was assumed to be 2.7 x 103 kg m-3 in all cases. Using Method
1, the igneous clast compositions yield estimates of heat production ranging from 0.25-7.49 µW m-3, with an average of
about 2.6 µW m-3 and 1σ standard deviation of 1.9 µW m-3 (Table 1). Most of the variation observed in these samples comes 20
from variations in concentrations of U and Th. Method 2 gives quite similar results.
Estimates of heat production versus age are plotted in Fig. 3. Compared to an average value for surface heat production
in stable continental shield regions of ~2 µW m-3 (Jaupart et al., 2016), most of the Antarctic clast samples are of similar
magnitude (mean of 2.61 µW m-3, with 11 falling between 1-4 µW m-3), although they are notably higher than values
reported for some cratonic areas (e.g., Canadian Shield and Grenville Orogen; Mareschal and Jaupart, 2013; Jaupart et al., 25
2016). As a group, they are quite similar to the global average granitic heat production of 2.5 µW m3 (Rybach, 1976; Haenel
et al., 1988), which in general is expected to be higher than the bulk upper crustal average of about 1.6 µW m3 (Jaupart et
al., 2016). Heat production from the clasts is comparable to estimates obtained from Archean and Paleoproterozoic bedrock
exposed in the coastal region of southern Prydz Bay (2.4-2.6 µW m-3; Carson and Pittard, 2012; Carson et al., 2013). Four of
the clasts give high values between 4.0-7.5 µW m-3, which are similar to global occurrences of crust characterized by high 30
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
7
heat production (Mareschal and Jaupart, 2013; Jaupart et al., 2016) and exemplified by the Central Australian Heat Flow
Province (CAHFP; Sandiford and McLaren, 2002; McLaren et al., 2003). Nonetheless, all but two of the samples in this
suite have heat production less than the mean for the CAHFP (4.6 µW m-3).
The variability in heat production shown by the data presented here resembles that observed in regions comprised by
Precambrian shields or granitic batholiths and likely represents real heterogeneities in the source region. Although the 5
precise distribution of heat-producing rocks in the source area from which these clast samples were eroded is not known, this
group may collectively provide a qualitatively-random sample that provides a means to assess average heat production for a
broad region of the continental interior. Compared to global examples, the Proterozoic igneous rocks in this study indicate
that heat production in central East Antarctica is like that of typical continental shield areas and demonstrably different from
the anomalously warm region represented by the CAHFP. Therefore, despite plausible geological and geophysical 10
correlations between cratonic rocks in southern Australia (Gawler craton) and the Wilkes Land region of East Antarctica
(e.g., Oliver and Fanning, 1997; Aitken et al., 2014; Goodge and Finn, 2010; Boger, 2011; Goodge and Fanning, 2016),
extrapolation of the high values reported for the CAHFP into East Antarctica may be unwarranted.
4.4 Heat flow
Geothermal heat flow can be estimated from the empirical relation with crustal heat production (Lachenbruch, 1968; 15
Roy et al., 1968). In the absence of direct terrestrial heat flow measurements, as is the case for Antarctica, it is possible to
calculate heat flow from heat production by assuming a thickness of the upper crustal heat-producing layer (Sandiford and
McLaren, 2002; Turcotte and Schubert, 2014). This thickness, hr, is the length scale for decrease in Ho with depth in the
upper crust (where most heat-producing elements are concentrated) and is determined from the slope of the q-H function.
Although Ho is thought to decrease exponentially with depth (Lachenbruch, 1968), a first-order estimate of terrestrial heat 20
flow can be obtained from:
qo = qm + qr + (Ho * hr) (3)
where qo is the surface heat flow (mW m-2), qm is the mantle heat flow, qr is the 'reduced' heat flow contributed by heat
production in the middle and lower crust, and other terms are as defined above. For stable Precambrian continental crust,
average values for qm are about 14 mW m-2 and qr is about 15 mW m-2 (Sandiford and McLaren, 2002; Perry et al., 2006; 25
Levy et al., 2010; Mareschal and Jaupart, 2013; Jaupert et al., 2016). Based on similarities in age and thickness to the
Canadian and Scandinavian shields, a value of 7.3 km for hr is used here. Using the relationship above and heat production
results, the surface terrestrial heat flow is estimated from the igneous clast population to range from about 31-84 mW m-2
(Table 1), with an average of 48.0 ± 13.6 mW m-2 (1σ standard deviation; Fig. 4). The average value may be regarded as an
integrated estimate of heat flow across the area of erosion within the catchment, but it is probably a maximum because it is 30
derived from values of heat production that are biased to crustal granites.
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
8
4.5 Uncertainties
Because estimates of heat flow are used in ice-sheet models, it is important to consider uncertainties in the values used
as input parameters. Here I consider uncertainties in the estimates of heat production and heat flow provided above.
4.5.1 Uncertainties in Ho
Laboratory precision on elemental analyses is very high (instrumental precision within 0.2% for K2O by XRF and within 5
2% for U and Th by ICPMS), density is assumed, and constants of heat-production for various elements are assumed.
Therefore, individual uncertainties for Ho were not calculated because they are expected to be very low relative to other
parameters involved in calculation of heat flow.
4.5.2 Uncertainties in qo
Uncertainties in the linear relationship used to calculate surface heat flow (qo) can be modeled using the following 10
expression:
qo = qm + qr + (Ho*hr) + (hr*Ho) (4)
where qo is the sum of uncertainties represented by the variables included in equation (3). Determining reasonable values
for most of the terms is problematical because the corresponding terms in the heat-flow equation are either based on
model-derived values or are simply poorly constrained by limited empirical data. Because geological and seismological data 15
indicate that East Antarctica is a stable craton, we can use typical cratonic values for qm and qr as a basis for evaluating
uncertainty in these terms. For this analysis, qm is taken to be ±2.5 mW m-2 based on a compilation of estimates worldwide
for stable continental shield areas that range mostly from 12-17 mW m-2 (Mareschal and Jaupart, 2013; Jaupart et al., 2016).
Uncertainty in the lower-crustal term, qr, is taken to be 3.0 mW m- 2, assumed as a general variance 20%) around a
representative value of 15 mW m-2 for lower-crustal heat flow. A mean value of Ho = 2.6 µW m-3 is used from the data 20
reported here and the representative average value of hr = 7.3 km that was used to calculate heat flow is assumed here.
Uncertainty in heat production, Ho, is taken as a 1σ standard deviation of the calculated values (1.86 mW m-2), and
uncertainty in the length scale, hr, is assumed to be 1500 m 20%). Based on these inputs, we can derive a general
uncertainty for the surface heat flow term (qo) of about 23 mW m-2 (Fig. 4). This is a large value compared to the nominal
mean value of 48 mW m-2 obtained here, and it reflects large natural variability in lithosphere properties as well as few direct 25
constraints on mantle heat flow, lower crustal heat flow, and the vertical distribution of heat-producing elements in
continental crust. Of this estimated uncertainty, 24% is contributed by the qm and qr terms, and 76% is attributed to the
multiplying effects of the thickness and uncertainty of the upper-crustal heat-producing layer (hr and hr). Only 8% is
contributed by Ho itself. Together, the large combined uncertainty is therefore contributed mainly by mantle heat flow,
lower crustal heat flow, and the vertical distribution of heat-producing elements; conversely, estimates of upper crustal heat 30
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
9
production from the glacial clast samples are not an important source of uncertainty. Nonetheless, the overall range in
surface heat flow covered by this uncertainty is consistent with the range of values reported for other cratons, lending
support to the idea that the recovered glacial clasts are indeed representative of heat flow known from typical Archean and
Proterozoic shield areas. Despite the inherent large uncertainties, the first-order results can help to inform future ice-sheet
modeling. 5
5 Discussion
The glacial igneous clasts sampled for this study indicate that upper crustal heat production for at least a part of central
East Antarctica is in the range of 0.3-7.5 µW m-3, with average value of 2.6 ± 1.9 µW m-3. Assuming typical values of mantle
heat flux, lower-crustal heat flux, and an upper-crustal length factor appropriate for stable continental cratons, the derived
heat production corresponds to an average surface heat flux of 48 mW m-2. Although this approach assumes cratonic values 10
for mantle and lower-crustal contributions, it is reasonable given what is known about East Antarctic lithosphere. The net
upper crustal contribution to surface heat flow is therefore about 19 mW m-2. The total surface heat flux is quite similar to
the average heat flux of 53 mW m-2 from 13 cratonic shield provinces globally (Jaupart et al., 2016). Likewise, Nyblade and
Pollack (1993) found average surface heat flow values of 42 mW m-2 for Archean provinces and 47 mW m-2 for
Paleoproterozoic provinces, which represents a general depletion of heat-producing elements in continental crust with 15
increasing age. The heat flow results obtained here are also similar to earlier estimates for East Antarctica determined by
geophysical modeling and inversion of ice borehole temperature profiles, which indicate a broad region with low to
moderate values of 50-60 mW m-2.
Taken together, the heat production and surface heat flow values estimated for the glacial igneous clasts discussed here
appear to be representative of typical Archean-Proterozoic cratonic lithosphere. As a group they are distinctly different from 20
the regional pattern shown by anomalously warm Proterozoic crust in central Australia with average qo = 80 mW m-2
(McLaren et al., 2003), which has been suggested to extend across the Wilkes Land margin of Antarctica based on
Gondwana supercontinent reconstructions (Carson et al., 2014; Aitken et al., 2014). Despite general age similarities among
some of the clast population with parts of the Gawler Craton, and basement age correlations that indicate continuity of
Mawson-type crust into the Wilkes sector of East Antarctica (Goodge and Fanning, 2016), the proxy heat production 25
determinations and heat flow estimates provided here suggest that central portions of the East Antarctic ice sheet are
underlain by stable continental crust with quite normal thermal properties represented by average values of heat production
of about 2.5 µW m-3 and heat flow of about 50 mW m-2.
Estimates of terrestrial heat flow such as those provided here can also be used to assess the effect of heat flow on ice-
sheet mass balance. For example, Pollard et al. (2005) evaluated the effect of varying heat flow regimes on ice-sheet 30
behavior by modeling changes in Antarctic ice volume, ice-sheet surface elevation, and area of the base at its pressure-
melting point as a function of differing heat-flow regimes. Their models used three different geothermal heat flow
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
10
distributions: (a) uniform heat flow of 37.7 mW m-2, representing typical values of Archean cratons; (b) uniform at 75.4 mW
m-2, to mimic Proterozoic lithosphere characterized by high crustal heat production; and (c) spatially varying heat flow based
on the distributions of different crustal provinces extrapolated from craton-margin geology, and including values of 41 and
55 mW m-2 across most of East Antarctica. The values of heat production and heat flow estimated for central East Antarctica
in this study are most consistent with their third approach; the average heat flow value of the Proterozoic granitoid samples is 5
higher than in the case of uniform Archean lithosphere, yet lower than that assumed for Proterozoic lithosphere with high
crustal heat production. Because the modeling of Pollard et al. (2005) shows a large effect of heat flow on the area of the ice-
sheet base at its pressure-melting point, inputting appropriate values of crustal heat flow is vitally important for predicting,
for example, the thermal and physical conditions of the basal ice-sheet regime.
To provide a simple model for the distribution of heat flow across the catchment area sampled in this study, mean heat 10
flow values were calculated in two ways (Table 2). First, the set of 18 samples was divided into equal quintiles representing
ranges of 10 mW m-2 each. Average heat flow values were calculated for each quintile, as was a percentage of the
measurements falling in that range (Fig. 5a). Each quintile thus represents a proportionally-based average heat flow value
that could be used as an input for ice sheet models. Assuming that the igneous and metamorphic crust beneath the EAIS is
heterogeneous in age and composition, this proportional distribution of heat flow values may better reflect the complexities 15
of crustal geothermal input as a function of subglacial area compared to a simple average. Second, the samples were grouped
by age and average heat flow values calculated for each of four groups (Fig. 5b). This approach provides a reasonable
estimate of heat flow potentially contributed by igneous crust proportionally represented by different age groups. Although
the sample values were divided arbitrarily into five groups using the first method, this approach shows that about 61% of the
sample results are <50 mW m-2 (also indicated by the skewed distribution of values in Figure 4), indicating that the bulk of 20
crust underlying the EAIS has relatively low long-range average heat flow. The second approach, perhaps more useful from
a modeling perspective because it groups samples by age, illustrates that for individual age groups the values are also quite
modest, ranging from about 42-55 mW m-2 and similar to the total group average. It is noteworthy that this range is nearly
identical to the heterogeneous heat-flow model adopted by Pollard et al. (2005), appearing to validate the earlier study.
Future ice-sheet stability modeling combined with the estimates of low to intermediate sub-glacial heat flow found in this 25
study may thus help to further refine predictions of ice-sheet behavior.
6 Conclusions
Based on geochemical analysis of a suite of glacially-eroded granitic rock clasts, average heat production from an
inferred large Proterozoic igneous crustal province in central East Antarctica is estimated to be about 2.5 µW m-3, and the
corresponding average surface heat flow is about 48 mW m-2. These geothermal properties are quite similar to average 30
Archean and Proterozoic cratonic shields globally, despite being biased here to granitic compositions. Although the source of
the granite clasts is not precisely known, they were likely derived from a region extending into central East Antarctica from
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
11
near the inlet to Byrd Glacier. This region contrasts with other areas marked by high heat flow, such as the Central Australia
Heat Flow Province and some parts of East Antarctica near Prydz Bay, indicating that crust in those areas does not extend
into central regions of the continental interior.
Heat flow estimated in this study is valuable for several reasons. First, the values are very similar to an estimate of heat
flow derived by modeling of a borehole temperature profile near Dome C (54 mW m-2; Fischer et al., 2013), helping to 5
validate the earlier model finding. Likewise, they are consistent with the general range of values indicated by inversion of
geophysical data from cratonic East Antarctica (e.g., Shapiro and Ritzwoller, 2004; Fox Maule et al., 2005). In detail, the
values obtained here show a similar range to those indicated in the model derived from magnetic data (Fox Maule et al.,
2005), both of which indicate that lithologic and, therefore, geothermal variations are real. Second, the new data provide a
unique estimate of heat production and terrestrial heat flow that can be used as an input to ice-sheet stability models. In 10
particular, they validate the general approach by Pollard et al. (2005) in which basal heat flow is varied by area depending on
age and character of the subglacial geology. There is similar variability within this sample group that probably reflects the
lithologic heterogeneity to be expected in continental shields. Third, although the data presented here provide a good
approximation of both heat production and heat flow in an otherwise inaccessible region of East Antarctica, the existing
uncertainties associated with extrapolating heat flow from heat production illustrate the critical need for precise in-situ 15
measurement of terrestrial heat flow from the subglacial environment. One attempt to do so beneath the Whillans Ice Stream
in West Antarctica (Fisher et al., 2015) measured a heat flux of 285 mW m-2. This extraordinarily high value, even greater
than that observed on modern ocean ridges (typically ~100-250 mW m-2 near the ridge axis and one third of that for oceanic
crust >50 Ma; Stein, 1995), likely reflects a large component of advective heat transport by flowing water and is therefore
not representative of terrestrial heat flow in West Antarctica. Despite the difficulty in obtaining reliable heat flow data from 20
the subglacial environment, it should be a high research priority that can be addressed by drilling through the ice sheets at as
many sites as possible in order to assess crustal heterogeneity. Last, these estimates of low to moderate crustal heat flow
indicate that some large regions of the interior East Antarctic ice sheet may be expected to be frozen at the bed, which is of
use to future drilling projects that plan to intersect the glacial bed.
Data availability. Data supporting the conclusions are listed in Table 1. 25
Sample availability. Samples referred to in this study are housed at the University of Minnesota Duluth and available on
request to the author.
Competing interests. The author declares that he has no conflict of interest.
Acknowledgments. Field and analytical portions of this project were supported by the National Science Foundation (award
0944645). Jacqueline Halpin and Jean-Claude Mareschal provided helpful feedback on the approach to estimating heat 30
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
12
production and heat flow, and Jeff Severinghaus kindly reviewed an earlier draft manuscript. John Swenson kindly provided
insight into the treatment of uncertainties.
References
Aitken, A.R.A., Young, D.A., Ferraccioli, F., Betts, P.G., Greenbaum, J.S., Richter, T.G., Roberts, J.L., Blankenship, D.D.,
Siegert, M.J.: The subglacial geology of Wilkes Land, East Antarctica, Geophysical Research Letters, 41, 5
doi:10.1002/2014GL059405, 2014.
Bo, S., Siegert, M.J., Mudd, S.M., Sugden, D., Fujita, S., Xiangbin, C., Yunyun, J., Xueyuan, T., Yuansheng, L.: The
Gamburtsev Mountains and the origin and early evolution of the Antarctic Ice Sheet, Nature, 459, 690693,
doi:10.1038/nature08024, 2009.
Boger, S.D.: Antarctica Before and after Gondwana, Gondwana Research, 19, 335-371, doi: 10.1016/j.gr.2010.09.003, 10
2011.
Carson, C.J., McLaren, S., Roberts, J.L., Boger, S.D., Blankenship, D.D.: Hot rocks in a cold place: high sub-glacial heat
flow in East Antarctica, J. Geol. Soc. Lond., doi:10.1144/jgs2013-030, 2013.
Carson, C.J., Pittard, M.: A reconnaissance crustal heat production assessment of the Australian Antarctic Territory (AAT),
Geoscience Australia Record, Report 2012-63, 2012. 15
Davidson, J.P., McMillan, N.J.M., Moorbath, S., Wörner, G., Harmon, R.S., López-Escobar, L.: The Nevados de Payachata
volcanic region (18S/69W, N. Chile) II. Evidence for widespread crustal involvement in Andean magmatism,
Contributions to Mineralogy and Petrology, 105, 412432, 1990.
Davidson, J.P., Harmon, R.S., Wörner, G.: The source of central Andean magmas: Some considerations. In Harmon, R.S.,
and Rapela, C.W. (eds.), Andean Magmatism and Its Tectonic Setting, Geological Society of America Special Paper, 20
265, 233–244, 1991.
DeConto, R.M., Pollard, D.: Rapid Cenozoic glaciation of Antarctica induced by declining atmospheric CO2, Nature, 421,
245249, doi:10.1038 /nature01290, 2003.
Fischer, H., Severinghaus, J., Brook, E., and 27 others: Where to find 1.5 million yr old ice for the IPICS “Oldest-Ice” ice
core, Clim. Past, 9, 2489-2505, doi:10.5194/cp-9-2489-2013, 2013. 25
Fisher, A.T., Mankoff, K.D., Tulaczyk, S.M., Tyler, S.W., Foley, N., and the WISSARD Science Team: High geothermal
heat flux measured below the West Antarctic Ice Sheet, Science Advances, 1(6), 1-9, doi:10.1126/sciadv.1500093,
2015.
Fox Maule, C., Purucker, M.E., Olsen, N., Mosegaard, K.: Heat flux in Antarctica revealed from satellite magnetic data,
Science, 309, 464-467, doi:10.1126/science.1106888, 2005. 30
Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis D.J., Frost, C.D.: A geochemical classification for granitic rocks,
J. Petrology, 42, 2033-2048, doi:10.1093/petrology/42.11.2033, 2001.
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
13
Goodge, J.W., Fanning, C.M., Bennett, V.C.: U-Pb evidence of ~1.7 Ga crustal tectonism during the Nimrod Orogeny in the
Transantarctic Mountains, Antarctica: implications for Proterozoic plate reconstructions, Precambrian Research, 112,
261-288, doi:10.1016/S0301-9268(01)00193-0, 2001.
Goodge, J.W., Fanning, C.M., Brecke, D.M., Licht, K.J., Palmer, E.F.: Continuation of the Laurentian Grenville province in
western East Antarctica, Journal of Geology, 118, 601-619, doi:10.1086/656385, 2010. 5
Goodge, J.W., Fanning, C.M., Vervoort, J.D., Radakovich, A.L.: More SWEAT: Discovery of Mesoproterozoic and
Paleoproterozoic igneous crust in East Antarctica strengthens the case for Laurentia-Antarctica connections in Rodinia,
Geological Society of America Abstracts with Programs, 44(7), 599, 2012.
Goodge, J.W., Vervoort, J.D., Fanning, C.M., Brecke, D.M., Farmer, G.L., Williams, I.S., Myrow, P.M., DePaolo, D.J.: A
positive test of East Antarctica-Laurentia juxtaposition within the Rodinia supercontinent, Science, 321, 235-240; 10
doi:10.1126/science.1159189, 2008.
Goodge, J.W., Fanning, C.M.: Mesoarchean and Paleoproterozoic history of the Nimrod Complex, central Transantarctic
Mountains, Antarctica: Stratigraphic revisions and relation to the Mawson Continent in East Gondwana, Precambrian
Research, 285, 242-271, doi:10.1016/j.precamres.2016.09.001, 2016.
Goodge, J.W., Fanning, C.M., Fisher, C.M., Vervoort, J.D.: Proterozoic crustal evolution of central East Antarctica: Age and 15
isotopic evidence from glacial igneous clasts, and links with Australia and Laurentia, Precambrian Research, in press,
2017.
Govindaraju, K.: Compilation of working values and sample description for 383 geostandards, Geostandards Newsletter, 18
(Special Issue), 1-158, 1994.
Haenel, R., Rybach, L., Stegena L., eds.: Handbook of Terrestrial Heat-Flow Density Determination: Dordrecht, Kluwer 20
Academic Publishers, 486 pp, 1988.
Hasterok, D., Chapman, D.S.: Heat production and geotherms for the continental lithosphere, Earth and Planetary Science
Letters, 307, 5970, doi: 10.1016/j.epsl.2011.04.034, 2011.
Jamieson, S.S.R., Sugden, D.E.: Landscape Evolution of Antarctica. In Cooper, A.K., Barrett, P.J., Stagg, H., Storey, B.,
Stump, E., Wise, W., eds., Antarctica: A Keystone in a Changing World (Proceedings of the 10th International 25
Symposium on Antarctic Earth Sciences). Washington, DC, The National Academies Press, 2008.
Jaupart, C., Mareschal, J.-C., Iarotsky, L.: Radiogenic heat production in the continental crust, Lithos, 262, 398-427,
10.1016/j.lithos.2016.07.017, 2016.
Lachenbruch, A.H.: Preliminary geothermal model of the Sierra Nevada, Journal of Geophysical Research, 73, 69776989,
1968. 30
Lévy, F., Jaupart, C., Mareschal, J., Bienfait, G., Limare, A.: Low heat flux and large variations of lithospheric thickness in
the Canadian Shield, Journal of Geophysical Research, 115, 6404, 2010.
Mareschal, J.-C., Jaupart, C.: Radiogenic heat production, thermal regime and evolution of continental crust,
Tectonophysics, 609, 524534, doi:10.1029/2009JB006470, 2013.
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
14
McLaren, S., Sandiford, M., Hand, M., Neumann, N., Wyborn, L., Bastrakova, I.: The hot southern continent: heat flow and
heat production in Australian Proterozoic terranes. In Hillis, R.R., Müller, R.D. (eds.), Evolution and Dynamics of the
Australian Plate, Geological Society of Australia Special Publications, 12(22), 151161, 2003.
Nissen, C.I., Fanning, C.M., and Goodge, J.W.: New evidence of Proterozoic metamorphic events in East Antarctica from
in-situ U-Pb age dating of monazite in metamorphic glacial clasts, central Transantarctic Mountains, Antarctica, 5
Geological Society of America Abstracts with Programs, 45, 798, 2013.
Nyblade, A.A.: Heat flow and the structure of Precambrian lithosphere, Lithos, 48, 8191, 1999.
Nyblade, A.A., Pollack, H.N.: A global analysis of heat flow from Precambrian terrains: implications for the thermal
structure of Archean and Proterozoic lithosphere, Journal of Geophysical Research, 98, 1220712218, 1993.
Oliver, R.L., Fanning, M.: Antarctica: precise correlation of Paleoproterozoic terrains, In Ricci C. A., ed., The Antarctic 10
Region: Geological Evolution and Processes, Siena, Terra Antarctica Publications, 163172, 1997.
Perry, H., Rosieanu, C., Mareschal, J.C., Jaupart, C.: Thermal regime of the lithosphere in Canada, Canadian Journal of
Earth Sciences, 47, 389408, doi:10.1139/E09-059, 2010.
Peucat, J.-J., Capdevila, R., Fanning, C.M., Ménot, R.-P., Pécora, L., Testut, L.: 1.60 Ga felsic volcanic blocks in the
moraines of the Terre Adélie craton, Antarctica: Comparisons with the Gawler Range volcanics, South Australia, 15
Australian Journal of Earth Sciences, 49, 831-845, doi:10.1046/ j.1440-0952.2002.00956, 2002.
Pollard, D., DeConto R.M., Nyblade, A.A.: Sensitivity of Cenozoic Antarctic ice sheet variations to geothermal heat flux,
Glob. Planet. Change, 49, 63-74, doi:10.1016/j.gloplacha.2005.05.003, 2005.
Purucker, M.E.: Antarctica Basal Heat Flux, http://websrv.cs.umt.edu/isis/index.php/Antarctica_Basal_Heat_Flux, 2012.
Rignot, E., Mouginot, J., Scheuchl, B.: Ice flow of the Antarctic ice sheet, Science, 333, 1427-1430; 20
doi:10.1126/science.1208336, 2011.
Rose, K.C., Ferraccioli, F., Jamieson, S.S.R., Bell, R.E., Corr, H., Creyts, T.T., Braaten, D., Jordan, T.A., Fretwell, P.T.,
Damaske, D.: Early East Antarctic Ice Sheet growth recorded in the landscape of the Gamburtsev Subglacial Mountains,
Earth and Planetary Science Letters, 375, 112, doi:10.1016/j.epsl.2013 .03.053, 2013.
Roy, R.F., Blackwell, D.D., Birch, F.: Heat generation of plutonic rocks and continental heat flow provinces, Earth and 25
Planetary Science Letters, 5, 112, 1968.
Rudnick, R.L., Fountain, D.M.: Nature and composition of the continental crust: a lower crustal perspective, Reviews of
Geophysics, 33, 267309, 1995.
Rybach, L.: Radioactive heat production in rocks and its relation to other petrophysical parameters, Pure and Applied
Geophysics, 114, 309317, 1976. 30
Rybach, L.: Determination of heat production rate, In R. Haenel, L. Rybach, L. Stegena, (eds.), Handbook of Terrestrial
Heat-Flow Density Determination, Kluwer Academic Publishers, Dordrecht, pp. 125-142, 1988.
Sandiford, M., McLaren, S.: Tectonic feedback and the ordering of heat producing elements within the continental
lithosphere, Earth and Planetary Science Letters, 204, 133150, doi:10.1016/S0012-821X(02)00958-5, 2002.
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
15
Schroeder, D.M., Blankenship, D.D., Young, D.A., Quartini, E.: Evidence for elevated and spatially variable geothermal flux
beneath the West Antarctic Ice Sheet, Proc. Natl. Acad. Sci. U. S. A., 111(25), 90709072, 2014.
Shapiro, N.M., Ritzwoller, M.H.: Inferring surface heat flux distributions guided by a global seismic model: particular
application to Antarctica, Earth Planet. Sci. Lett., 223, 213224, 2004.
Siegert, M.J.: Antarctic subglacial lakes, Earth-Science Reviews, 50, 2950, 2000. 5
Stein, C.: Heat flow of the Earth, In Global Earth Physics: A Handbook of Physical Constants, American Geophysical
Union, Washington, D.C., Reference Shelf 1, 144-158, 1995.
Turcotte, D.L., Schubert, G.: Geodynamics, Cambridge, Cambridge University Press, 623 p, 2014.
Van Liefferinge, B., Pattyn, F.: Using ice-flow models to evaluate potential sites of million year-old ice in Antarctica, Clim.
Past, 9, 28592887, doi:10.5194/cpd-9-2859-2013, 2013. 10
Whillans, I.M., Cassidy, W.A.: Catch a falling star: Meteorites and old ice, Science, 222, 55-57, 1983.
Wörner, G., Moorbath, S., Horn, S., Entenmann, J., Harmon, R.S., Davidson, J., López-Escobar, L.: Large- and fine-scale
geochemical variations along the Andean arc of northern Chile (17. 522S), In K.J. Reutter, E. Scheuber, and P.J.
Wigger, eds., Tectonics of the Southern Central Andes: Structural Evolution of an Active Continental Margin, Springer-
Verlag, Berlin, 7792, 1994. 15
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
16
Figure 1: Terrestrial heat flow in Antarctica, from the mean geothermal heat flux model of Van Liefferinge and Pattyn (2013),
which averages heat flow determined from multiple geophysical datasets. Inset white box shows area of Figure 2, including glacial
drainage sourcing bedrock igneous rock clasts (white outline).
65°S
75°S
180°
90°E90°W
1000 km
120
Heat flow (mW m-2)
110
100
90
80
70
60
50
Fig. 2.
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
17
Figure 2: Map showing potential source areas for dated glacial igneous clasts, superimposed on the Bedmap2 subglacial
topography of East Antarctica (Fretwell et al., 2013). Principal features are ice-sheet catchment areas (marked by thin blue
drainage divides), ice flow directions in the broad Byrd Glacier drainage (arrows, Rignot et al. (2011), and areas of Precambrian
basement exposure (pink). Composite source area (outlined by heavy white line) was determined from the ice flow-fields that 5
contribute ice to each of the five sample sites (white circles). Because transport distance is not known for any of the individual
clasts, possible bedrock sources could lie anywhere between the sample sites and the top of the ice-shed overlapping the
Gamburtsev Subglacial Mountains (GSM). Other abbreviations: ASB, Aurora Subglacial Basin; GM, Grove Mountains; LT,
Lambert trough; LV, Lake Vostok; MR, Miller Range; PCM, Prince Charles Mountains; WSB, Wilkes Subglacial Basin. Outlet
glaciers: Bd, Beardmore; By, Byrd; Ni, Nimrod; Sc, Scott; Sh, Shackleton. 10
SOURCE
AREA
120°E
150°E
150°W
120°W
90°W
60°W
30°W
90°E
70°S75°S80°S85°S
60°E30°E
180°
MR
By
Bd
Ni
Sc
Sh
LT
PCM
GM
LV
ASB
GSM
Dome A
Dome C
WSB
Ross Sea
0 500km
Byrd Gl
Lambert Gl
Recovery
Ice
Stream
Support Force
Ice Stream
MSA
AGA
MRA
TNA LWA
drainage divides
ice flow
exposed Precambrian basement
EA
WA
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
18
Figure 3. Plot of surface heat production (Ho) vs. age for igneous glacial clasts. Data listed in Table 1. Range of heat production
values in Ross Orogen granites from unpublished data. For comparison, values are shown from the Canadian Shield (Mareschal
and Jaupart, 2013; Jaupart et al., 2016), areas of high heat production in stable continental provinces (‘hot crust’; Mareschal and
Jaupart, 2013; Jaupart et al., 2016), East Antarctica (Carson and Pittard, 2012; Carson et al., 2014), and the Central Australian 5
Heat Flow Province (CAHFP; McLaren et al., 2003).
Ross Orogen
granitoids
(~500 Ma)
avg granites
avg continental crust
0
5
10
15
1000 1500 2000 2500 3000 3500
Heat production (Ho, μW m-3)
Age (Ma)
Heat production in glacial igneous clasts
Igneous clasts
East Antarctica
CAHFP (Australia Proterozoic)
Canadian Shield
Hot crust
Average crust
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
19
Figure 4. Histogram of heat flow values estimated from heat production in the glacial clasts. Mean value (magenta) is 48.0 mW m-2
(n = 18) with a 1σ standard deviation of 13.6 mW m
-2. Consideration of uncertainties in calculation of heat flow indicates an
overall uncertainty of about ±21 mW m-2 (see text). Range of heat flow modeled for East Antarctica shown for comparison (light
pink, Van Liefferinge and Pattyn, 2013). Global average values for Archean cratons and Proterozoic lithosphere shown by ruled 5
bars (Nyblade et al., 1999).
±23
mean =
48.0
East Antarctic
heat flow models
0
1
2
3
4
5
20 30 4 0 50 6 0 70 80 90
Heat flow (mW m-2)
Number
Heat flow
Archean
cratons
Proterozoic
lithosphere
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
20
Figure 5: Summary pie diagrams showing distribution of heat flow estimates from this study. (a) Distribution of heat flow by
quintiles between 30-80 mW m-2. Quintile averages shown, with the highest quintile represented by one sample with calculated
heat flow of about 84 mW m-2. (b) Distribution of average heat flow by age groups. Values shown in Table 2.
5
1209 Ma: 55.4 mW m-2
1455 Ma: 41.8 mW m-2
1807 Ma: 49.6 mW m-2
2013 Ma: 47.5 mW m-2
12%
41%35%
12%
b. Heat flow distribution by age group
33%
28%
22%
11%
6%
a. Heat flow distribution by quintile
Quintile 1: 35.5 mW m-2
Quintile 2: 43.5 mW m-2
Quintile 3: 56.0 mW m-2
Quintile 4: 63.3 mW m-2
Quintile 5: 83.6 mW m-2
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
21
Table 1. Estimates of Ho and qo for igneous clast samples, central East Antarctica.
Sample'
Age'
Th'
K2O'
Ho'(Method'1)'a'
Ho'(Method'2)'a'
qo'b'
''
(Ma)'
(ppm)'
(wt%)'
(µW'm>3)'
(µW'm>3)'
(mW'm>2)'
10LWA>13.1'
1204'
27.7'
3.80'
3.81'
3.79'
56.8'
10LWA>11.1'
1213'
32.9'
4.72'
3.43'
3.40'
54.1'
10MSA>2.3'
1410'
38.0'
7.26'
3.43'
3.35'
54.1'
10TNA>1.1'
1430'
18.5'
4.39'
2.04'
1.98'
43.9'
10LWA>6.5'
1432'
15.8'
8.16'
2.59'
2.46'
47.9'
10LWB>4.3'
1448'
11.4'
5.76'
1.66'
1.57'
41.2'
10LWB>3.8'
1470'
5.8'
4.19'
1.12'
1.05'
37.2'
10LWA>20.1'
1486'
8.6'
2.35'
1.14'
1.11'
37.4'
10MSA>3.5'
1508'
1.5'
1.57'
0.25'
0.23'
30.9'
10LWA>6.4'
1570'
26.9'
5.65'
4.11'
4.05'
59.0'
10LWA>14.1'
1786'
41.9'
5.77'
4.93'
4.89'
65.0'
10LWB>4.5'
1848'
13.4'
3.89'
1.45'
1.39'
39.6'
10LWA>6.3'
1850'
98.3'
5.38'
7.49'
7.54'
83.6'
10LWA>7.1'
1854'
18.2'
6.83'
2.09'
1.99'
44.3'
10LWB>4.1'
1865'
10.4'
4.92'
1.51'
1.44'
40.1'
10MSA>3.3'
1876'
1.0'
2.29'
0.78'
0.74'
34.7'
10LWA>10.1'
2010'
51.3'
0.73'
4.47'
4.54'
61.6'
10LWA>8.1'
2015'
4.9'
0.53'
0.61'
0.61'
33.4'
Mean'
'
23.7'
4.34'
2.61'
2.56'
48.0'
Std.'dev.'
'
24.0'
2.18'
1.86'
1.89'
13.6'
a'Heat'production'(Ho)'was'calculated'from'geochemical'analysis'in'two'ways.'Method'1' uses'Equation'(1),'using'values'after'
Rybach'(1988)'and'Hasterok'and'Chapman'(2011).'Method'2'uses'Equation'(2).'Both'assume'an'average'density'(ρ)'for'
granitic'rocks'of'2.7'x'103'kg'm>3.'
b'Surface'heat'flow'(qo)'determined'from'Equation'(3)'(Turcotte'and'Schubert,'2014).'5
Table 2: Heat flow estimates in proportions based on quintile ranges and age.
Binned'by'quintile'
'
Binned'by'age'
Quintile'
Heat'flow'
(mW'm>2)'
No.'
%'
'
Age'(Ma)'
Heat'flow'
(mW'm>2)'
No.'
%'
1'
35.5'
6'
0.33'
'
1209'
55.4'
2'
0.12'
2'
43.5'
5'
0.28'
'
1455'
41.8'
7'
0.41'
3'
56.0'
4'
0.22'
'
1770'
49.6'
6'
0.35'
4'
63.3'
2'
0.11'
'
2013'
47.5'
2'
0.12'
5'
83.6'
1'
0.06'
''
''
''
''
''
'
The Cryosphere Discuss., https://doi.org/10.5194/tc-2017-134
Manuscript under review for journal The Cryosphere
Discussion started: 14 August 2017
c
Author(s) 2017. CC BY 4.0 License.
... Exploring the internal structure of the lithosphere and its geodynamics is one of the frontier topics in Earth Science (Artemieva, 2006(Artemieva, , 2018Ashchepkov et al., 2012;Björk & Winsosr, 2006;Dalton, Bao, & Ma, 2017;Goodge, 2018;Hu, He, & Wang, 2001;Majorowicz, Grad, & Polkowski, 2019;Mareschal & Jaupart, 2004;Qiu, Zuo, Chang, & Li, 2014;Wang, 2015). In 1914, following the early work of Barrel (1914) on the concept of the lithosphere relative to the asthenosphere, different approaches have been used to define the lithosphere based on different physical and chemical boundary conditions, such as thermal lithosphere (Griffin et al., 1999), petrological lithosphere (Rudnick, McDonough, & O'Connell, 1998), seismic-thermal lithosphere (Goes, Govers, & Vacher, 2000), mechanical lithosphere (Schubert, Turcotte, & Olson, 2001), seismological lithosphere (Anderson, 1995), elastic lithosphere (Watts, 2001), chemical lithosphere (Lee, Lenardic, Cooper, Niu, & Levandar, 2005), and electric lithosphere (Clowes, 1997). ...
... The thermal lithospheric structure refers to the proportion of the heat flow in the crust and the mantle, and their competency and structural relationship (Blackwell, 1971;Wang, 2015). The crust-mantle heat flow distribution affects the current crustal and mantle activity, deep temperature conditions, and the formation of basins and mountains (Artemieva, 2018;Artemieva & Mooney, 2001;Dalton et al., 2017;Goodge, 2018;Majorowicz et al., 2019;Norden, Förster, & Balling, 2008;Rudnick & Nyblade, 1999;Wang, 2015), and is an important window to understand the tectonic evolution of the lithosphere and the deep dynamics of basins (Hu et al., 2001;Mareschal & Jaupart, 2004;Björk & Winsosr, 2006;Artemieva, 2006;Zuo et al., 2013;Zuo, Li, Li, & Hao, 2015;Zuo, Song, Li, Wang, & Yang, 2017;Gong, Zhang, & Ye, 2011;Qiu et al., 2014;Li, Zuo, Qiu, & Gao, 2017;Liu, Zhang, Zhang, & He, 2016;Jiang et al., 2019). Pollackm & Chapman (1977) first computed the "thermal" lithospheric thickness in North America and Australia. ...
... to 850°C, with an average value of 816°C ± 36°C (Table 7), and the lithospheric thermal structure is "hot crust and cold mantle" type. Terrestrial heat flow is an objective reflection of the dynamics of the sedimentary basin and the thermal evolution of the lithospheric structure (Artemieva, 2018;Goodge, 2018;Majorowicz et al., 2019;Wang, 2015). Sedimentary basins of different ages and different origins show obvious differences in their present-day thermal states (Artemieva, 2006;Björk & Winsosr, 2006;Gong et al., 2011;Hu et al., 2001;Li et al., 2017;Liu et al., 2016;Zuo et al., 2011). ...
Article
The Chagan Depression in the Yingen‐Ejinaqi Basin, located at the intersection of the Paleo‐Asian Ocean and the Tethys Ocean domains is an important region to gain insights on terrestrial heat flow, lithospheric thermal structure and deep geodynamic processes. Here we compute terrestrial heat flow values in the Chagan Depression using a large set of system steady‐state temperature data from four representative wells and rock thermal conductivity. We also estimate the “thermal” lithospheric thickness, mantle heat flow, ratio of mantle heat flow to surface heat flow and Moho temperature to evaluate the regional tectonic framework and deep dynamics. The results show that the heat flow in the Chagan Depression ranges from 66.5 to 69.8 mW/m2, with an average value of 68.3 ± 1.2 mW/m2. The Chagan Depression is characterized by a thin “thermal” lithosphere, high mantle heat flow and high Moho temperature, corresponding to the lithospheric thermal structure of “cold mantle and hot crust” type. We correlate the formation of the Yingen‐Ejinaqi Basin to the Early Cretaceous and Cenozoic subduction of the western Pacific Plate and the Cenozoic multiple extrusions. Our results provide new insights into the thermal structure and dynamics of the lithospheric evolution in central China.
... where the surface geology is unknown, the lithologies and probable heat production is best constrained by 900 determining the probable heat production of each drainage basin based on its detrital clasts (e.g. Goodge, 2018). 901 ...
Preprint
Full-text available
Abstract. Antarctic geothermal heat flow (GHF) affects the temperature of the ice sheet, determining its ability to slide and internally deform, as well as the behaviour of the continental crust. However, GHF remains poorly constrained, with few and sparse local, borehole-derived estimates, and large discrepancies in the magnitude and distribution of existing continent-scale estimates from geophysical models. We review the methods to extract GHF, compile borehole and probe-derived estimates from measured temperature profiles, and recommend the following future directions: 1) Obtain more borehole-derived estimates from the subglacial bedrock and englacial temperature profiles. 2) Estimate GHF beneath the interior of the East Antarctic Ice Sheet (the region most sensitive to GHF variation) via long-wavelength microwave emissivity. 3) Estimate GHF from inverse glaciological modelling, constrained by evidence for basal melting. 4) Revise geophysically-derived GHF estimates using a combination of Curie depth, seismic, and thermal isostasy models. 5) Integrate in these geophysical approaches a more accurate model of the structure and distribution of heat production elements within the crust, and considering heterogeneities in the underlying mantle. And 6) continue international interdisciplinary communication and data access.
... The averaged shear-wave velocity within the NELVZ (from ~10 to 25 km depth) is consistent with shear wave velocities for granite-gneiss to granite-granodiorite rock types in this depth range (Christensen, 1996 - Table 2). This hypothesis provides a common source for the gravity low observed in both NE Greenland and NN and could explain regional measurements of elevated heat flow if we consider the relative thickness of this granitic layer ( Pascal et al., 2007) and the presence of high heat-producing granitoid rocks in NE Greenland as found in other Proterozoic terranes (McLaren et al., 2003;Carson et al., 2014;Goodge, 2018). ...
Article
Full-text available
We present a high-resolution shear-wave velocity model of Greenland’s lithosphere from regional and teleseismic Rayleigh-waves recorded by the Greenland Ice Sheet Monitoring Network supplemented with observations from several temporary seismic deployments. To construct Rayleigh-wave group-velocity maps, we integrated signals from regional and teleseismic earthquakes with several years of ambient seismic noise and used the dispersion to constrain crustal and upper-mantle seismic shear-wave velocity structure. Specifically, we used a Markov Chain Monte Carlo technique to estimate 3-D shear-wave velocities beneath Greenland to a depth of 200 km. Our model reveals four prominent anomalies: a deep high-velocity feature extending from southwestern to northwestern Greenland that may be the signature of a thick cratonic keel; a corridor of relatively low upper-mantle velocity across central Greenland that could be associated with lithospheric modification from the passage of the Iceland plume beneath Greenland or interpreted as a tectonic boundary between cratonic blocks; an upper-crustal southwest-northeast trending boundary separating Greenland into two regions of contrasting tectonic and crustal properties; and a mid-crustal low-velocity anomaly beneath northeastern Greenland. The nature of this mid-crustal anomaly is of particular interest given that it underlies the onset of the Northeast Greenland Ice Stream, and raises interesting questions regarding how deeper processes may impact the ice stream dynamics and the evolution of the Greenland ice sheet.
Article
Geothermal heat flow (GHF) is an elusive physical property, yet it can reveal past and present plate tectonic processes. In Antarctica, GHF has further consequences in predicting the response of ice sheets to climate change. In this Review, we discuss variations in Antarctic GHF models based on geophysical methods and draw insights into tectonics and GHF model usage for ice sheet modelling. The inferred GHF at continental scale for West Antarctica (up to 119 mW m−2, 95th percentile) points to numerous contributing influences, including non-steady state neotectonic processes. Combined influences cause especially high values in the vicinity of the Thwaites Glacier, a location critical for the accurate prediction of accelerated loss of Antarctic ice mass. The inferred variations across East Antarctica are more subtle (up to 66 mW m−2, 95th percentile), where slightly elevated values in some locations correspond to the influence of thinned lithosphere and tectonic units with concentrations of heat-producing elements. Fine-scale anomalies owing to heat-producing elements and horizontal components of heat flow are important for regional modelling. GHF maps comprising central values with these fine-scale anomalies captured within uncertainty bounds can thus enable improved ensemble-based ice sheet model predictions of Antarctic ice loss. Geothermal heat flow, a driver of tectonic processes, must be inferred by indirect means in Antarctica. This Review probes variations in Antarctic heat flow maps, discusses the corresponding tectonic insights and provides recommendations for the use of such maps in ice sheet modelling. Differences between geothermal heat flow maps for Antarctica that are derived using alternative approaches provide greater insight into its tectonic evolution than anomalies that are constrained from one model alone.Non-steady state processes and heat-producing elements in the upper crust contribute markedly to the spatial distribution of anomalously high geothermal heat flow values (>60%).High geothermal heat flow anomalies in West Antarctica are a consequence of multiple contributing sources, such as neotectonic rifting, volcanism and a mantle heat anomaly.The stable lithosphere of East Antarctica has relatively subtle geothermal heat flow anomalies, many of which are difficult to separate from model uncertainties and currently remain unresolved.Fine-scale geothermal heat flow variations can be accounted for, through low and high bounds to possible geothermal heat flow in the form of uncertainty maps, to provide robust inputs to predictive modelling of Antarctic ice sheet evolution.Geothermal heat flow is a boundary condition for modelling ice loss. In particular, the fast-changing Thwaites Glacier of West Antarctica, and the outlet glaciers of the Wilkes and Aurora Basins of East Antarctica, are locations of great concern. Differences between geothermal heat flow maps for Antarctica that are derived using alternative approaches provide greater insight into its tectonic evolution than anomalies that are constrained from one model alone. Non-steady state processes and heat-producing elements in the upper crust contribute markedly to the spatial distribution of anomalously high geothermal heat flow values (>60%). High geothermal heat flow anomalies in West Antarctica are a consequence of multiple contributing sources, such as neotectonic rifting, volcanism and a mantle heat anomaly. The stable lithosphere of East Antarctica has relatively subtle geothermal heat flow anomalies, many of which are difficult to separate from model uncertainties and currently remain unresolved. Fine-scale geothermal heat flow variations can be accounted for, through low and high bounds to possible geothermal heat flow in the form of uncertainty maps, to provide robust inputs to predictive modelling of Antarctic ice sheet evolution. Geothermal heat flow is a boundary condition for modelling ice loss. In particular, the fast-changing Thwaites Glacier of West Antarctica, and the outlet glaciers of the Wilkes and Aurora Basins of East Antarctica, are locations of great concern.
Article
Full-text available
Abstract High-quality maps of Geothermal heat flow (GHF) are crucial when modeling ice dynamics, shape, and mass loss of the Antarctic Ice Sheet, which is one of the largest potential contributors to sea level rise. The determination of GHF remains challenging, as in situ data are sparse and geophysical models exhibit large discrepancies in amplitude and resolution, especially on regional scales. Using a novel approach implementing a joint inversion of gravity and seismic tomography data with various geophysical and mineral physics information, we estimate the 3D thermal lithospheric structure and present a new GHF map. The resulting surface heat flow correlates with the location of subglacial volcanism and can represent a boundary condition for accurate ice dynamics models that can explain observed acceleration in the ongoing ice mass loss. Absolute values are within the range of other seismology-based methods and are much lower than those obtained using for example, magnetic data. High uncertainties remain in the parametrization of the upper crustal structure and thermal parameters.
Article
Full-text available
We present PetroChron Antarctica, a new relational database including petrological, geochemical and geochronological data sets along with computed rock properties from geological samples across Antarctica. The database contains whole-rock geochemistry with major/trace element and isotope analyses, geochronology from multiple isotopic systems and minerals for given samples, as well as an internally consistent rock classification based on chemical analysis and derived rock properties (i.e., chemical indices, density, p-velocity, and heat production). A broad range of meta-information such as geographic location, petrology, mineralogy, age statistics and significance are also included and can be used to filter and assess the quality of the data. Currently, the database contains 11,559 entries representing 10,056 unique samples with varying amounts of geochemical and geochronological data. The distribution of rock types is dominated by mafic (36%) and felsic (33%) compositions, followed by intermediate (22%) and ultramafic (9%) compositions. Maps of age distribution and isotopic composition highlight major episodes of tectonic and thermal activity that define well known crustal heterogeneities across the continent, with the oldest rocks preserved in East Antarctica and more juvenile lithosphere characterizing West Antarctica. PetroChron Antarctica allows spatial and temporal variations in geology to be explored at the continental scale and integrated with other Earth-cryosphere-biosphere-ocean data sets. As such, it provides a powerful resource ready for diverse applications including plate tectonic reconstructions, geological/geophysical maps, geothermal heat flow models, lithospheric and glacial isostasy, geomorphology, ice sheet reconstructions, biodiversity evolution, and oceanography.
Article
Full-text available
Dome A is the summit of the East Antarctic Ice Sheet, underlain by the rugged Gamburtsev Subglacial Mountains. The rugged basal topography produces a complex hydrological system featuring basal melt, water transport and storage, and freeze‐on. In a companion study, we used an inverse model to infer the spatial distributions of geothermal heat flow (GHF) and accumulation rate that best fit a variety of observational constraints. Here, we present and analyze the best‐fit state of the ice sheet in detail. Our modeled result agrees well with the observed water bodies and freeze‐on structures, while also predicting a significant amount of unobserved water and suggesting a change in stratigraphic interpretation that reduces the volume of the freeze‐on units. Our modeled stratigraphy agrees well with observations, and we predict that there will be two distinct patches of ice up to 1.5 Ma suitable for ice coring underneath the divide. Past divide migration could have interrupted stratigraphic continuity at the old ice patches, but various indirect lines of evidence suggest that the divide has been stable for about the last one and a half glacial cycles, which is an encouraging but not definitive sign for stability in the longer term. Finally, our GHF estimate is higher than previous estimates for this region, but consistent with possible heterogeneity in crustal heat production.
Article
Full-text available
We present a refined map of geothermal heat flow for Antarctica, Aq1, based on multiple observables. The map is generated using a similarity detection approach by attributing observables from geophysics and geology to a large number of high‐quality heat flow values (N = 5792) from other continents. Observables from global, continental and regional datasets for Antarctica are used with a weighting function that allows the degree of similarity to increase with proximity and how similar the observables are. The similarity detection parameters are optimized through cross‐correlation. For each grid cell in Antarctica, a weighted average heat flow value and uncertainty metrics are calculated. The Aq1 model provides higher spatial resolution in comparison to previous results. High heat flow is shown in the Thwaites Glacier region, with local values over 150 mWm−2. We also map elevated values over 80 mWm−2 in Palmer Land, Marie Byrd Land, Victoria Land and Queen Mary Land. Very low heat flow is shown in the interior of Wilkes Land and Coats Land, with values under 40 mWm−2. We anticipate that the new geothermal heat flow map, Aq1, and its uncertainty bounds will find extended use in providing boundary conditions for ice sheet modeling and understanding the interactions between the cryosphere and solid Earth. The computational framework and open architecture allow for the model to be reproduced, adapted and updated with additional data, or model subsets to be output at higher resolution for regional studies.
Article
Full-text available
Key Points First heat flow measurements from the Coompana Province, southern Australia, indicate values of 40–70 mW/m ² (57 ± 3 mW/m ² average) A tectonically reconstructed heat flow map constrains the thermal regime of geological provinces formerly contiguous with East Antarctica Geophysical models of Antarctic heat flow are likely underestimating crustal sources of radiogenic heat that influence subglacial melting
Article
Full-text available
We create a 3-D density, temperature, and composition model of the Antarctic lithosphere using an integrative approach combining gravity, topography, and tomography data with mineral physics constraints and seismic data on crustal structures. The latter is used to create a new Moho and crustal density model. Temperature and thermal density variations are estimated based on S wave velocities from two independent tomography models (SL2013sv and AN1-S). Results of the Antarctic continent show the well-known distinction between East and West Antarctica in temperature and density to a depth of about 200 km. Incorporating compositional variations in the temperature calculations increases temperatures in depleted regions by up to 150 °C, giving improved insights into thermal structures. The thickness of the lithospheric root also varies strongly between these regions, with values below 100 km in the west and above 200 km in the east. Regions with negative compositional density variations (<−0.040 g/cm³ at 100 km), high depletion (Mg # > 91.5), and low temperatures (<800 °C; central Dronning Maud Land, along the east flank of the Transantarctic Mountains) are interpreted as Precambrian cratonic fragments. Nearly undepleted lithosphere is found in the Lambert Graben and the Aurora Subglacial Basin and is attributed to Mesozoic rifting activity that has caused lithospheric rejuvenation.
Article
Full-text available
Magmas erupted over the last 20 m.y. within the Central Volcanic Zone (CVZ) of the Andes are strongly differentiated and depleted in 143Nd and enriched in 87Sr and 18O (87Sr/86Sr >0.705, 143Nd/144Nd <0.5126, δ18O >+6.5%o) relative to intra-oceanic arc magmas. These characteristics may reflect closed-system differentiation from melts of an incompatible element enriched sub-arc mantle (ancient subcontinental lithosphere), or the ubiquitous crustal contamination of island arc-like primitive magmas, derived from an asthenospheric mantle source. It is not possible to choose unambiguously which of these models is applicable, since all of the observed lavas of the CVZ are differentiated relative to melts that could have been in equilibrium with the mantle. We contend that the evidence currently available favors the latter possibility. CVZ volcanic centers are located on exceptionally thick (c. 70 km) continental crust and parent magmas are likely to have ponded and differentiated extensively at one or more levels in the crust during ascent. Furthermore, some compositional characteristics (e.g., δ18O >+6.5%o-greater than the δ18O values expected from ultramafic mantle; and Pb isotope ratios that exhibit a correspondence with the Pb-isotope composition of the crust through which the magmas were erupted) are consistent with crustal contamination of mantle-derived magmas, rather than closed-system differentiation of mantle-derived magmas, as the process predominantly determining magma evolution in the central Andes.
Book
Full-text available
GeoRef, Copyright 2014, American Geological Institute.
Article
Full-text available
Wilkes Land is a key region for studying the configuration of Gondwana and for appreciating the role of tectonic boundary conditions on East Antarctic Ice Sheet (EAIS) behavior. Despite this importance, it remains one of the largest regions on Earth where we lack a basic knowledge of geology. New magnetic, gravity and subglacial topography data allow the region's first comprehensive geological interpretation. We map lithospheric domains and their bounding faults, including the suture between Indo-Antarctica and Australo-Antarctica. Furthermore, we image subglacial sedimentary basins, including the Aurora and Knox Subglacial Basins, and the previously unknown Sabrina Subglacial Basin. Commonality of structure in magnetic, gravity and topography data suggest that pre-EAIS tectonic features are a primary control on subglacial topography. The preservation of this relationship after glaciation suggests that these tectonic features provide topographic and basal boundary conditions that have strongly influenced the structure and evolution of the EAIS.
Article
Rock clasts entrained in glacial deposits sourced from the continental interior of Antarctica provide an innovative means to determine the age and composition of ice-covered crust. Zircon U-Pb ages from a suite of granitoid clasts collected in glacial catchments draining central East Antarctica through the Transantarctic Mountains show that crust in this region was formed by a series of magmatic events at ∼2.01, 1.88-1.85, ∼1.79, ∼1.57, 1.50-1.41, and 1.20-1.06 Ga. The dominant granitoid populations are ca. 1.85, 1.45 and 1.20-1.06 Ga. None of these igneous ages are known from limited outcrop in the region. In addition to defining a previously unrecognized geologic history, zircon O and Hf isotopic compositions from this suite have: (1) mantle-like δ¹⁸O signatures (4.0-4.5‰) and near-chondritic Hf-isotope compositions (εHf ∼ +1.5) for granitoids of ∼2.0 Ga age; (2) mostly crustal δ¹⁸O (6.0-8.5‰) and variable Hf-isotope compositions (εHf = -6 to +5) in rocks with ages of ∼1.88-1.85, ∼1.79 and ∼1.57 Ga, in which the ∼1.88-1.79 Ga granitoids require involvement of older crust; (3) mostly juvenile isotopic signatures with low, mantle-like δ¹⁸O (∼4-5‰) and radiogenic Hf-isotope signatures (εHf = +6 to +10) in rocks of 1.50-1.41 Ga age, with some showing crustal sources or evidence of alteration; and (4) mixed crustal and mantle δ¹⁸O signatures (6.0-7.5‰) and radiogenic Hf isotopes (εHf = +3 to +4) in rocks of ∼1.2 Ga age. Together, these age and isotopic data indicate the presence in cratonic East Antarctica of a large, composite igneous province that formed through a punctuated sequence of relatively juvenile Proterozoic magmatic events. Further, they provide direct support for geological correlation of crust in East Antarctica with both the Gawler Craton of present-day Australia and Proterozoic provinces in western Laurentia. Prominent clast ages of ∼2.0, 1.85, 1.57 and 1.45 Ga, together with sediment source linkages, provide evidence for the temporal and spatial association of these cratonic elements in the Columbia supercontinent. Abundant ∼1.2-1.1 Ga igneous and metamorphic clasts may sample crust underlying the Gamburtsev Subglacial Mountains, indicating the presence of a Mesoproterozoic orogenic belt in the interior of East Antarctica that formed during final assembly of Rodinia.
Article
Studies of heat flow from Precambrian terrians have demonstrated three empirical relationships; a temporal relationship between heat flow and tectonic age, a spatial pattern between heat flow and the proximity of Archean cratons, and a temporal relationship between heat flow and the age of lithosphere stabilization. In the first relationship, heat flow is inversely related to tectonic age. The second pattern is characterized by low heat flow from Archean cratons and Proterozoic terrains adjacent to cratonic margins (pericratonic terrains), and higher heat flow from Proterozoic terrains that are more than a few hundred kilometers from a craton. In the third pattern, heat flow decreases as the age of stabilization of the lithosphere increases. A number of interpretations of Precambrian heat flow have been offered to explain one or more of these relationships. The simple cooling of a thermal boundary layer predicts essentially no change in heat flow in terrains older than ≈1.5 Ga, and therefore does not likely provide a comprehensive framework for the interpretation of Precambrian heat flow. By contrast, two other interpretations, (1) thicker lithosphere beneath Archean terrains than beneath Proterozoic terrains, and (2) greater heat production in Proterozoic crust than in Archean crust, when combined with the special structural configuration of sutures, can both contribute to the spatial and temporal heat flow distributions. Xenolith thermobarometry constraints on lithospheric temperatures, however, limit the contribution of age-dependent crustal heat production, and therefore at least part of the heat flow distributions derive from variations in lithosphere thickness.