Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-30T04:20:48.144Z Has data issue: false hasContentIssue false

Mechanism for the subglacial formation of cryogenic brines

Published online by Cambridge University Press:  08 May 2023

Sarah U. Neuhaus*
Affiliation:
Earth and Planetary Sciences, University of California Santa Cruz, Santa Cruz, CA 95064, USA
Slawek M. Tulaczyk
Affiliation:
Earth and Planetary Sciences, University of California Santa Cruz, Santa Cruz, CA 95064, USA
*
Corresponding author: Sarah U. Neuhaus, E-mail: suneuhau@ucsc.edu
Rights & Permissions [Opens in a new window]

Abstract

Cryogenic brines are under-studied, despite the fact that they may contain information about past ice-sheet behavior. Cryogenic brines form through cryoconcentration of seawater, although the specific setting and mechanism of formation have been debated. Previous conceptual models of brine formation require seawater isolation from the ocean in a closed basin experiencing freezing. We propose instead that they may form in pore spaces of marine sediments subjected to repeat cycles of ice-sheet advance and retreat. During periods of basal freezing, cryoconcentration produces hypersaline brines which experience downward flow driven by unstable density stratification. Our advection-diffusion model of porewater chemistry evolution successfully recreates the porewater chemistry of two deep Antarctic cores containing cryogenic brines (AND-1B and AND-2A), suggesting that cryogenic brines can be formed through the repeated isolation and cryoconcentration of marine waters within subglacial sediment pore spaces of modern and past ice sheets.

Type
Letter
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The International Glaciological Society

1. Introduction

Cryogenic brines are common in formerly glaciated regions in the Northern Hemisphere, having been found in deep boreholes across the Canadian Shield (Frape and Fritz, Reference Frape and Fritz1982; Bottomley and others, Reference Bottomley, Conrad Gregoire and Raven1994) and Fennoscandia (Starinsky and Katz, Reference Starinsky and Katz2003). Fewer cryogenic brines have been found in Antarctica, however this may simply be due to the difficulty of obtaining deep porewater samples from Antarctica. In fact, subsurface brines may be widespread around the edges of the Antarctic continent where basal freezing dominates (Foley and others, Reference Foley2019; Frank and others, Reference Frank, Haacker, Fielding and Yang2022; Piccione and others, Reference Piccione2022). We examined cryogenic brines collected from the pore spaces of two deep boreholes drilled into the seafloor in McMurdo Sound, Antarctica (Pompilio and others, Reference Pompilio2007; Frank and others, Reference Frank and Gui2010).

Chemical signatures in the cryogenic brines indicate that they are derived from the freezing (cryoconcentration) of seawater (Starinsky and Katz, Reference Starinsky and Katz2003; Frank and others, Reference Frank and Gui2010, Reference Frank, Haacker, Fielding and Yang2022; Gardner and Lyons, Reference Gardner and Lyons2019; Lyons and others, Reference Lyons2019), but their water isotopes are strongly depleted. The two prevailing conceptual models for the formation of cryogenic brines require the isolation of seawater in a basin which freezes over (Starinsky and Katz, Reference Starinsky and Katz2003; Grasby and others, Reference Grasby, Rod Smith, Bell and Forbes2013; Frank and others, Reference Frank, Haacker, Fielding and Yang2022). During freezing, seawater solutes are preferentially retained in the remaining liquid, resulting in the formation of a dense brine which seeps into the underlying sediments. Although the idea of an isolated marine basin could be used to explain the formation of cryogenic brines found in the Northern Hemisphere, it is difficult to transplant that idea to the Ross Sea Sector of Antarctica due to the lack of evidence for bathymetric highs large enough to have isolated marine waters from the ocean in the past (Tinto and others, Reference Tinto2019). Here, we propose a new model in which brines form in the sediment pore spaces below a marine-based ice sheet, without the need for an isolated marine basin.

2. Summary of methods and results

We modeled the concentration of Cl and δ18O in sediment porewaters that have experienced multiple cycles of glacial retreat and advance and compared our results to brines found in cores AND-1B and AND-2A (Pompilio and others, Reference Pompilio2007; Frank and others, Reference Frank and Gui2010), recovered from drill sites located beneath the McMurdo Ice Shelf (Fig. 1). We modeled the formation and vertical dispersal of cryogenic brines using a finite-difference code solving the 1-D vertical advection-diffusion equation:

(1)$$\displaystyle{{\partial C} \over {\partial t}} = D_{{\rm sed}}\displaystyle{{\partial ^2C} \over {\partial z^2}} + \displaystyle{{\partial D_{{\rm sed}}} \over {\partial z}}\displaystyle{{\partial C} \over {\partial z}} + K\displaystyle{{\Delta \rho } \over \rho }\displaystyle{{\partial C} \over {\partial z}}$$

where C is chemical concentration, t is time, D sed is the diffusion coefficient of the chemical parameter through sediments, z is depth below the seafloor, K is the hydraulic conductivity of the sediments and ρ is the density of the porewater. Derivation of this equation and the variables D sed, K and ρ is shown in the Supplemental materials.

Figure 1. Location of AND-1B and AND-2A boreholes. Ross Sea is indicated in black. The MODIS Mosaic of Antarctica (Haran and others, Reference Haran, Bohlander, Scambos, Painter and Fahnestock2021) and ice velocity data (Rignot and others, Reference Rignot, Mouginot and Scheuchl2017) are plotted using Antarctic Mapping Tools (Greene and others, Reference Greene, Gwyther and Blankenship2017).

We modeled porewater concentrations of Cl and δ18O for a 2 km sediment column (seafloor to 2000 mbsf) which has experienced 100 000-year cycles of ice-sheet retreat and re-advance. We chose to model the concentrations of Cl and δ18O because we expect that they did not interact chemically with the sediments at temperatures prevailing in the shallow subsurface (Morin and others, Reference Morin2010), and thus are solely indicators of freezing- and transport-related processes in the sedimentary column. In our model runs, we exposed the simulated sedimentary column to three different upper boundary conditions (Fig. 2), representing the different parts of a simplified glacial cycle: (i) seawater exposure representing deglaciations during interglacials (e.g. modern conditions), (ii) basal melting when the ice sheet is overriding the study site and climactic (e.g. surface temperature) as well as glaciological conditions (e.g. large ice thickness) lead to a positive thermal energy balance at the ice base, and (iii) basal freezing when the ice sheet is overriding the study site and climactic and glaciologic conditions lead to a negative thermal energy balance at the ice base. A description of how we altered the boundary conditions is shown in the Supplementary materials.

Figure 2. Schematic of the three different periods examined during the 100 000-year model runs. During seawater periods, the porewater in the topmost element reflect seawater concentrations. During freezing periods, Cl concentration increases, and δ18O decreases slightly. During melting periods, Cl decreases, and δ18O decreases significantly.

Because we were interested in determining whether cryogenic brines could form in sediment pore spaces, we altered the free parameters to allow the model outputs to fit the observations from AND-1B and AND-2A, respectively. In doing so, we were able to fit our model simulations to the observed concentrations of Cl and δ18O (Fig. 3) using reasonable model parameters and simplified glacial cycles, thus demonstrating the feasibility of our mechanism. The R 2 values indicating model fit are shown in Figures 3c and f. A list of all free parameters and the corresponding range of values over which we examined the free parameters is outlined in Table S1. Our model of the AND-1B core produced Cl concentrations that were 1.8 times more concentrated than the local seawater, and our model of the AND-2A core produced Cl concentrations that were five times more concentrated than the local seawater.

Figure 3. Model results compared to observed brine concentrations in the AND-1B core (a–c) and in the AND-2A core (d–f). (a, d) Cl concentration. (b, e) δ18O. (c, f) R 2 values indicating model fit after each glacial cycle. Light gray shading denotes cycles over which the R 2 value is significant for Cl and the dark gray shading denotes the cycles over which R 2 was significant for both Cl and δ18O. The dark lines in (a), (b), (d) and (e) denote the model results from the glacial cycles that best fit the observations for both Cl and δ18O based on R 2 values shown in (c) and (f). The results from ten glacial cycles are shown in (a) and (b). The results from 20 glacial cycles are shown in (d) and (e).

3. Conclusions and future research directions

The model was able to create cryogenic brines of similar Cl and δ18O composition to the brines observed in the pore spaces of the AND-1B and AND-2A cores. We thus conclude that cryogenic brines may form in sediment pore spaces just as well as in an isolated marine basin. The mechanism proposed here is a more plausible explanation, at least in the Antarctic setting considered here (McMurdo Sound). Given the current lack of a forebulge in the presence of the Antarctic Ice Sheet or a moraine large enough to cut off access to the ocean, it is unlikely that these locations were isolated from the ocean except when overridden by ice. However, we do not rule out other methods of cryogenic brine formation in other Antarctic settings – namely in the McMurdo Dry Valleys where the topographic and glaciologic history may permit brine formation in an isolated marine basin (Frank and others, Reference Frank, Haacker, Fielding and Yang2022).

This mechanism of cryogenic brine formation is exciting because it allows for the possibility of interpreting past ice-sheet behavior. Given the number of free parameters tuned in our study to allow the model outputs to fit the observations from the AND-1B and AND-2A cores, it is unlikely that our results are unique. However, future studies may be able to perform a more robust determination of the free parameters for a specific field site to produce unique model results. These results would allow for the determination of past ice-sheet dynamics, namely the timing of past grounding line advance and retreat as well as conditions at the base of the ice (i.e. freezing or melting). This could be a useful tool to help pinpoint the timing of grounding line retreat following the Last Glacial Maximum (LGM) (e.g. Neuhaus and others, Reference Neuhaus2021). Dating the timing of post-LGM grounding line retreat in the Ross Sea has been notoriously difficult (Anderson and others, Reference Anderson2014) due to the paucity of datable material in the Ross Sea sediments. Examining cryoconcentrated porewater samples could provide another avenue of investigation.

Additionally, cryogenic brines may be of interest to planetary scientists as they represent analogs of the types of fluids hypothesized to exist on other planetary bodies (namely Mars, Europa and Enceladus) and represent the most likely environment for finding extraterrestrial life in our own Solar System. For instance, there is evidence hinting at the presence of liquid water below the ice cap on Mars' south pole (Orosei and others, Reference Orosei2018; Lauro and others, Reference Lauro2021). Given the cold temperatures at the base of the Martian ice caps, the liquid water is assumed to contain high solute concentrations. These brines could have formed in the sediment pore spaces in a manner analogous to the mechanism proposed here.

Despite containing information about past ice-sheet dynamics, cryoconcentrated brines are under-sampled largely because sampling of subglacial waters has been biased toward fast-moving ice (Kamb, Reference Kamb2001; Tulaczyk and others, Reference Tulaczyk2014), where the ice sheet is not frozen to the bed and basal melting is prevalent. Because cryoconcentration can occur in sediment pore spaces, it is likely that there are brines hidden in the sediments where the Antarctic Ice Sheet is frozen to the bed, namely along most of the ice-sheet margin in Antarctica (Foley and others, Reference Foley2019). If we were to sample deep sediment cores at those locations we would find cryogenic brines, which could inform us of the past ice-sheet dynamics at that site.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/aog.2023.28.

Acknowledgements

This material is based upon work supported by the National Science Foundation under grant No. 1644187 and 1745124.

Author contributions

S. U. N. and S. M. T. co-designed this research. S. U. N. performed the analysis and wrote the manuscript with input from S. M. T.

References

Anderson, JB and 10 others (2014) Ross Sea paleo-ice sheet drainage and deglacial history during and since the LGM. Quaternary Science Reviews 100, 3154. doi: 10.1016/j.quascirev.2013.08.020Google Scholar
Bottomley, DJ, Conrad Gregoire, D and Raven, KG (1994) Saline ground waters and brines in the Canadian Shield: geochemical and isotopic evidence for a residual evaporite brine component. Geochimica et Cosmochimica Acta 58(5), 14831498. doi: 10.1016/0016-7037(94)90551-7Google Scholar
Foley, N and 9 others (2019) Evidence for pathways of concentrated submarine groundwater discharge in east Antarctica from helicopter-borne electrical resistivity measurements. Hydrology 6(2), 115. doi: 10.3390/HYDROLOGY6020054Google Scholar
Frank, TD, Gui, Z and ANDRILL SMS Science Team (2010) Cryogenic origin for brine in the subsurface of southern McMurdo Sound, Antarctica. Geology 38(7), 587590. doi: 10.1130/G30849.1Google Scholar
Frank, TD, Haacker, EMK, Fielding, CR and Yang, M (2022) Origin, distribution, and significance of brine in the subsurface of Antarctica. Earth-Science Reviews 234(August), 104204. doi: 10.1016/j.earscirev.2022.104204Google Scholar
Frape, SK and Fritz, P (1982) The chemistry and isotopic composition of saline groundwaters from the Sudbury Basin, Ontario. Canadian Journal of Earth Sciences 19(4), 645661. doi: 10.1139/e82-053Google Scholar
Gardner, CB and Lyons, WB (2019) Modelled composition of cryogenically produced subglacial brines, Antarctica. Antarctic Science 31(3), 165166. doi: 10.1017/S095410201900004XGoogle Scholar
Grasby, SE, Rod Smith, I, Bell, T and Forbes, DL (2013) Cryogenic formation of brine and sedimentary mirabilite in submergent coastal lake basins, Canadian Arctic. Geochimica et Cosmochimica Acta 110, 1328. doi: 10.1016/j.gca.2013.02.014Google Scholar
Greene, CA, Gwyther, DE and Blankenship, DD (2017) Antarctic mapping tools for Matlab. Computers & Geosciences 104, 151157. doi: 10.1016/j.cageo.2016.08.003Google Scholar
Haran, T, Bohlander, J, Scambos, T, Painter, T and Fahnestock, M (2021) MODIS Mosaic of Antarctica 2003–2004 (MOA2004) Image Map, Version 2. NASA National Snow and Ice Data Center DAAC. doi: 10.5067/68TBT0CGJSOJGoogle Scholar
Kamb, B (2001) Basal zone of the West Antarctic ice streams and its role in lubrication of their rapid motion. Antarctic Research Series 77, 157199.Google Scholar
Lauro, SE and 12 others (2021) Multiple subglacial water bodies below the south pole of Mars unveiled by new MARSIS data. Nature Astronomy 5(1), 6370. doi: 10.1038/s41550-020-1200-6Google Scholar
Lyons, WB and 9 others (2019) The geochemistry of englacial brine from Taylor Glacier, Antarctica. Journal of Geophysical Research: Biogeosciences 124(3), 633648. doi: 10.1029/2018JG004411Google Scholar
Morin, RH and 5 others (2010) Heat flow and hydrologic characteristics at the AND-1B borehole, ANDRILL McMurdo Ice Shelf project, Antarctica. Geosphere 6(4), 370378. doi: 10.1130/GES00512.1Google Scholar
Neuhaus, SU and 6 others (2021) Did Holocene climate changes drive West Antarctic grounding line retreat and readvance? Cryosphere 15(10), 46554673. doi: 10.5194/tc-15-4655-2021Google Scholar
Orosei, R and 21 others (2018) Radar evidence of subglacial liquid water on Mars. Science 361(6401), 490493. doi: 10.1126/science.aar7268Google Scholar
Piccione, G and 10 others (2022) Subglacial precipitates record Antarctic ice sheet response to late Pleistocene millennial climate cycles. Nature Communications 13(1), 5428. doi: 10.1038/s41467-022-33009-1Google Scholar
Pompilio, M and 11 others (2007) Petrology and geochemistry of the AND-1B core, ANDRILL McMurdo Ice Shelf Project, Antarctica. Terra Antarctica 14(3), 255288.Google Scholar
Rignot, E, Mouginot, J and Scheuchl, B (2017) MEaSUREs InSAR-Based Antarctica Ice Velocity Map, Version 2. [antarctica_ice_velocity_450m_v2]. Boulder, Colorado, USA: NASA National Snow and Ice Data Center Distributed Active Archive Center. doi: 10.5067/D7GK8F5J8M8R (accessed May 17, 2021).Google Scholar
Starinsky, A and Katz, A (2003) The formation of natural cryogenic brines. Geochimica et Cosmochimica Acta 67(8), 14751484. doi: 10.1016/S0016-7037(02)01295-4Google Scholar
Tinto, KJ and 30 others (2019) Ross Ice Shelf response to climate driven by the tectonic imprint on seafloor bathymetry. Nature Geoscience 12(6), 441449. doi: 10.1038/s41561-019-0370-2Google Scholar
Tulaczyk, S and 16 others (2014) WISSARD at Subglacial Lake Whillans, West Antarctica: scientific operations and initial observations. Annals of Glaciology 55(65), 5158. doi: 10.3189/2014aog65a009Google Scholar
Figure 0

Figure 1. Location of AND-1B and AND-2A boreholes. Ross Sea is indicated in black. The MODIS Mosaic of Antarctica (Haran and others, 2021) and ice velocity data (Rignot and others, 2017) are plotted using Antarctic Mapping Tools (Greene and others, 2017).

Figure 1

Figure 2. Schematic of the three different periods examined during the 100 000-year model runs. During seawater periods, the porewater in the topmost element reflect seawater concentrations. During freezing periods, Cl concentration increases, and δ18O decreases slightly. During melting periods, Cl decreases, and δ18O decreases significantly.

Figure 2

Figure 3. Model results compared to observed brine concentrations in the AND-1B core (a–c) and in the AND-2A core (d–f). (a, d) Cl concentration. (b, e) δ18O. (c, f) R2 values indicating model fit after each glacial cycle. Light gray shading denotes cycles over which the R2 value is significant for Cl and the dark gray shading denotes the cycles over which R2 was significant for both Cl and δ18O. The dark lines in (a), (b), (d) and (e) denote the model results from the glacial cycles that best fit the observations for both Cl and δ18O based on R2 values shown in (c) and (f). The results from ten glacial cycles are shown in (a) and (b). The results from 20 glacial cycles are shown in (d) and (e).

Supplementary material: File

Neuhaus and Tulaczyk supplementary material

Neuhaus and Tulaczyk supplementary material

Download Neuhaus and Tulaczyk supplementary material(File)
File 272.8 KB