Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-19T03:47:26.869Z Has data issue: false hasContentIssue false

Cryostratigraphy of mid-Miocene permafrost at Friis Hills, McMurdo Dry Valleys of Antarctica

Published online by Cambridge University Press:  16 December 2020

Marjolaine Verret*
Affiliation:
Antarctic Research Centre, Victoria University of Wellington, Wellington, New Zealand
Warren Dickinson
Affiliation:
Antarctic Research Centre, Victoria University of Wellington, Wellington, New Zealand
Denis Lacelle
Affiliation:
Department of Geography, Environment and Geomatics, University of Ottawa, Ottawa, Canada
David Fisher
Affiliation:
Department of Earth Sciences, University of Ottawa, Ottawa, Canada
Kevin Norton
Affiliation:
School of Geography, Environment and Earth Sciences, Victoria University of Wellington, Wellington, New Zealand
Hannah Chorley
Affiliation:
Antarctic Research Centre, Victoria University of Wellington, Wellington, New Zealand
Richard Levy
Affiliation:
Antarctic Research Centre, Victoria University of Wellington, Wellington, New Zealand GNS Science, Lower Hutt, New Zealand
Tim Naish
Affiliation:
Antarctic Research Centre, Victoria University of Wellington, Wellington, New Zealand

Abstract

The origin and stability of ground ice in the stable uplands of the McMurdo Dry Valleys remains poorly understood, with most studies focusing on the near-surface permafrost. The 2016 Friis Hills Drilling Project retrieved five cores reaching 50 m depth in mid-Miocene permafrost, a period when Antarctica transitioned to a hyper-arid environment. This study characterizes the cryostratigraphy of arguably the oldest permafrost on Earth and assesses 15 Myr of ground ice evolution using the REGO model. Four cryostratigraphic units were identified: 1) surficial dry permafrost (0–30 cm), 2) ice-rich to ice-poor permafrost (0.3–5.0 m) with high solute load and δ18O values (-16.2 ± 1.8‰) and low D-excess values (-65.6 ± 4.3‰), 3) near-dry permafrost (5–20 m) and 4) ice-poor to ice-rich permafrost (20–50 m) containing ice lenses with low solute load and δ18O values (-34.6 ± 1.2‰) and D-excess of 6.9 ± 2.6‰. The near-surface δ18O profile of ground ice is comparable to other sites in the stable uplands, suggesting that this ice is actively responding to changing surface environmental conditions and challenging the assumption that the surface has remained frozen for 13.8 Myr. The deep ice lenses probably originate from the freezing of meteoric water during the mid-Miocene, and their δ18O composition suggests mean annual air temperatures ~7–11°C warmer than today.

Type
Earth Sciences
Copyright
Copyright © Antarctic Science Ltd 2020

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Billups, K. & Schrag, D.P. 2002. Paleotemperatures and ice volume of the past 27 Myr revisited with paired Mg/Ca and 18O/16O measurements on benthic foraminifera. Paleoceanography, 17, 3–1.CrossRefGoogle Scholar
Coulombe, S., Fortier, D., Lacelle, D., Kanevskiy, M. & Shur, Y. 2019. Origin, burial and preservation of late Pleistocene-age glacier ice in Arctic permafrost (Bylot Island, NU, Canada). The Cryosphere, 13, 10.5194/tc-13-97-2019.CrossRefGoogle Scholar
Craig, H. 1965. The measurement of oxygen isotope paleotemperatures. In Tongiorgi, E., ed. Stable isotopes in oceanographic studies and paleotemperatures. Pisa: V. Lischi, 161182.Google Scholar
Craig, H., Gordon, L. & Horibe, Y. 1963. Isotopic exchange effects in the evaporation of water: 1. Low-temperature experimental results. Journal of Geophysical Research, 68, 10.1029/JZ068i017p05079.Google Scholar
Dahl-Jensen, D., Albert, M., Aldahan, A., Azuma, N., Balslev-Clausen, D., Baumgartner, M., Berggren, A.-M., et al. 2013. Eemian interglacial reconstructed from a Greenland folded ice core. Nature, 493, 10.1038/nature11789.Google Scholar
Dansgaard, W. 1964. Stable isotopes in precipitation. Tellus, 16, 436468.CrossRefGoogle Scholar
Dickinson, W.W. & Rosen, M.R. 2003. Antarctic permafrost: an analogue for water and diagenetic minerals on Mars. Geology, 31, 10.1130/0091-7613(2003)031<0199:APAAFW>2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Dickinson, W.W., Schiller, M., Ditchburn, B.G., Graham, I.J. & Zondervan, A. 2012. Meteoric Be-10 from Sirius Group suggests high elevation McMurdo Dry Valleys permanently frozen since 6 Ma. Earth and Planetary Science Letters, 355, 10.6073/pasta/9dda244e95ce8dfcdbd9b94d2289ef1d.Google Scholar
Doran, P. & Fountain, A. 2019. McMurdo Dry Valleys LTER: Daily measurement summaries from Friis Hills Meteorological Station (FRSM) in Taylor Valley, Antarctica from 1993 to present ver 6. Environmental Data Initiative, 10.6073/pasta/4c19209bd0cc73f5b30d6b98eca47bc9.Google Scholar
Fisher, D., Lacelle, D. & Pollard, W. 2020a. A model of unfrozen water content and its transport in icy permafrost soils: effects on ground ice content and permafrost stability. Permafrost and Periglacial Processes, 31, 10.1002/ppp.2031.CrossRefGoogle Scholar
Fisher, D.A., Lacelle, D., Pollard, W. & Faucher, B. 2020b. A model for stable isotopes of residual liquid water and ground ice in permafrost soils using arbitrary water chemistries and soil-specific empirical residual water functions. Permafrost and Periglacial Processes, 10.1002/ppp.2079.CrossRefGoogle Scholar
Fisher, D., Lacelle, D., Pollard, W., Davila, A. & McKay, C.P. 2016. Ground surface temperature and humidity, ground temperature cycles and the ice table depths in University Valley, McMurdo Dry Valleys of Antarctica. Journal of Geophysical Research - Earth Surface, 121, 10.1002/2016JF004054.CrossRefGoogle Scholar
Fitzgerald, P.G. 1992. The Transantarctic Mountains of southern Victoria Land: the application of apatite fission track analysis to a rift shoulder uplift. Tectonics, 11, 10.1029/91TC02495.CrossRefGoogle Scholar
Flower, B.P. & Kennett, J.P. 1994. The middle Miocene climatic transition: East Antarctic ice sheet development, deep ocean circulation and global carbon cycling. Palaeogeography, Palaeoclimatology, Palaeoecology, 108, 10.1016/0031-0182(94)90251-8.CrossRefGoogle Scholar
Fountain, A., Fernandez-Diaz, J.C., Obryk, M.K., Levy, J., Gooseff, M.N., van Horn, D.J., et al. 2017. High-resolution elevation mapping of the McMurdo Dry Valleys, Antarctica, and surrounding regions. Earth System Science Data, 9, 10.5194/essd-9-435-2017.CrossRefGoogle Scholar
Grootes, P.M., Steig, E.J., Stuiver, M., Waddington, E.D., Morse, D.L. & Nadeau, M.J. 2001. The Taylor Dome Antarctic 18O record and globally synchronous changes in climate. Quaternary Research, 56, 289298.CrossRefGoogle Scholar
Gasson, E., DeConto, R.M., Pollard, D. & Levy, R.H. 2016. Dynamic Antarctic ice sheet during the early to mid-Miocene. Proceedings of the National Academy of Sciences of the United States of America, 113, 10.1073/pnas.1516130113.Google ScholarPubMed
Gooseff, M.N., Lyons, W., McKnight, D.M., Vaughn, B.H., Fountain, A.G. & Dowling, C. 2006. A stable isotopic investigation of a polar desert hydrologic system, McMurdo Dry Valleys, Antarctica. Arctic, Antarctic, and Alpine Research, 38, 10.1657/1523-0430(2006)038[0060:ASIIOA]2.0.CO;2.CrossRefGoogle Scholar
Hagedorn, B., Sletten, R.S., Hallet, B., McTigue, D.F. & Steig, E.J. 2010. Ground ice recharge via brine transport in frozen soils of Victoria Valley, Antarctica: insights from modeling δ18O and δD profiles. Geochimica et Cosmochimica Acta, 74, 10.1016/j.gca.2009.10.021.CrossRefGoogle Scholar
Herold, N., Huber, M., Müller, R. & Seton, M. 2012. Modeling the Miocene Climatic Optimum: ocean circulation. Paleoceanography, 27, 10.1029/2010PA002041.CrossRefGoogle Scholar
Jouzel, J., Masson-Delmotte, V., Cattani, O., Dreyfus, G., Falourd, S., Hoffmann, G., et al. 2007. Orbital and millennial Antarctic climate variability over the past 800,000 years. Science, 317, 10.1126/science.1141038.CrossRefGoogle ScholarPubMed
Lacelle, D., Davila, A.F., Pollard, W.H., Andersen, D., Heldmann, J., Marinova, M. & McKay, C.P. 2011. Stability of massive ground ice bodies in University Valley, McMurdo Dry Valleys of Antarctica: using stable O-H isotope as tracers of sublimation in hyper-arid regions. Earth and Planetary Science Letters, 301, 10.1016/j.epsl.2010.11.028.CrossRefGoogle Scholar
Lacelle, D., Lapalme, C., Davila, A.F., Pollard, W., Marinova, M., Heldmann, J. & McKay, C.P. 2016. Solar radiation and air and ground temperature relations in the cold and hyper-arid Quartermain Mountains, McMurdo Dry Valleys of Antarctica. Permafrost and Periglacial Processes, 27, 10.1002/ppp.1859.CrossRefGoogle Scholar
Lacelle, D., Davila, A.F., Fisher, D., Pollard, W.H., DeWitt, R., Heldmann, J., et al. 2013. Excess ground ice of condensation-diffusion origin in University Valley, Dry Valleys of Antarctica: evidence from isotope geochemistry and numerical modeling. Geochimica et Cosmochimica Acta, 120, 10.1016/j.gca.2013.06.032.CrossRefGoogle Scholar
Lapalme, C., Lacelle, D., Pollard, W., Fisher, D., Davila, A. & McKay, C.P. 2017. Distribution and origin of ground ice in University Valley, McMurdo Dry Valleys, Antarctica. Antarctic Science, 29, 10.1017/S0954102016000572.CrossRefGoogle Scholar
Lewis, A.R. & Ashworth, A.C. 2015. An early to middle Miocene record of ice-sheet and landscape evolution from the Friis Hills, Antarctica. Bulletin, 128, 10.1130/B31319.1.Google Scholar
Lyons, W., Welch, K.A., Fountain, A.G., Dana, G.L., Vaughn, B.H. & McKnight, D.M. 2003. Surface glaciochemistry of Taylor Valley, southern Victoria Land, Antarctica and its relationship to stream chemistry. Hydrological Processes, 17, 10.1002/hyp.1205.Google Scholar
Lyons, W., Welch, K.A., Snyder, G., Olesik, J., Graham, E.Y., Marion, G.M. & Poreda, R.J. 2005. Halogen geochemistry of the McMurdo Dry Valleys lakes, Antarctica: clues to the origin of solutes and lake evolution. Geochimica et Cosmochimica Acta, 69, 305323.CrossRefGoogle Scholar
Mackay, J. 1990. Seasonal growth bands in pingo ice. Canadian Journal of Earth Sciences, 27, 10.1139/e90-116.CrossRefGoogle Scholar
Marchant, D. & Head III, J.W. 2007. Antarctic dry valleys: microclimate zonation, variable geomorphic processes, and implications for assessing climate change on Mars. Icarus, 192, 10.1016/j.icarus.2007.06.018.CrossRefGoogle Scholar
Marinova, M.M., McKay, C.P., Pollard, W.H., Heldmann, J.L., Davila, A.F., Andersen, D.T., et al. 2013. Distribution of depth to ice-cemented soils in the high-elevation Quartermain Mountains, McMurdo Dry Valleys, Antarctica. Antarctic Science, 25, 10.1017/S095410201200123X.CrossRefGoogle Scholar
McKay, C.P., Mellon, M.T. & Friedmann, E.I. 1998. Soil temperatures and stability of ice-cemented ground in the McMurdo Dry Valleys, Antarctica. Antarctic Science, 10, 3138.CrossRefGoogle ScholarPubMed
McKay, R., Naish, T., Carter, L., Riesselman, C., Dunbar, R., Sjunneskog, C., et al. 2012. Antarctic and Southern Ocean influences on late Pliocene global cooling. Proceedings of the National Academy of Sciences of the United States of America, 109, 10.1073/pnas.1112248109.Google ScholarPubMed
Naish, T., Wolff, E., Carter, L., McKay, R. & Powell, M. 2010. Antarctic climate evolution during the Quaternary (last 2.6 Ma) from continental margin, Southern Ocean and ice core records. Quaternary International, 219, 10.1016/j.quaint.2010.02.006.Google Scholar
Paxman, G.J., Jamieson, S.S., Hochmuth, K., Gohl, K., Bentley, M.J., Leitchenkov, G. & Ferraccioli, F. 2019. Reconstructions of Antarctic topography since the Eocene-Oligocene boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 535, 10.1016/j.palaeo.2019.109346.CrossRefGoogle Scholar
Porter, S. & Beget, J.E. 1981. Provenance and depositional environments of late Cenozoic sediments in permafrost cores from lower Taylor Valley, Antarctica. Dry Valley Drilling Project, 33, 10.1029/AR033p0351.CrossRefGoogle Scholar
Porter, T. & Opel, T. 2020. Recent advances in paleoclimatological studies of Arctic wedge- and pore-ice stable-water isotope records. Permafrost and Periglacial Processes, 31, 10.1002/ppp.2052.CrossRefGoogle Scholar
Roy, C. 2019. The origin of massive ground ice in raised marine sediments along the Eureka Sound Lowlands, Nunavut, Canada. PhD thesis, McGill University [Unpublished], 122 pp.Google Scholar
Sofer, Z. & Gat, J. 1975. The isotope composition of evaporating brines: effect of the isotopic activity ratio in saline solutions. Earth and Planetary Science Letters, 26, 10.1016/0012-821X(75)90085-0.CrossRefGoogle Scholar
Sturm, C., Zhang, Q. & Noone, D. 2010. An introduction to stable water isotopes in climate models: benefits of forward proxy modelling for paleoclimatology. Climate of the Past, 6, 10.5194/cp-6-115-2010.CrossRefGoogle Scholar
Swanger, K.M. 2017. Buried ice in Kennar Valley: a late Pleistocene remnant of Taylor Glacier. Antarctic Science, 29, 10.1017/S0954102016000687.CrossRefGoogle Scholar
Valletta, R.D., Willenbring, J.K., Lewis, A.R., Ashworth, A.C. & Caffee, M. 2015. Extreme decay of meteoric beryllium-10 as a proxy for persistent aridity. Scientific Reports, 5, 17813, 10.1038/srep17813.CrossRefGoogle ScholarPubMed
Van Everdingen, R.O. 1998. Multi-language glossary of permafrost and related ground-ice terms. Ottawa: International Permafrost Association, 159 pp.Google Scholar
Wang, Y., Hou, S., Masson-Delmotte, V. & Jouzel, J. 2009. A new spatial distribution map of δ18O in Antarctic surface snow. Geophysical Research Letters, 36, 10.1029/2008GL036939.CrossRefGoogle Scholar
Watanabe, K. & Mizoguchi, M. 2002. Amount of unfrozen water in frozen porous media saturated with solution. Cold Regions Science and Technology, 34, 103110.CrossRefGoogle Scholar
Wilch, T.I., Lux, D.R., Denton, G.H. & McIntosh, W.C. 1993. Minimal Pliocene-Pleistocene uplift of the dry valleys sector of the Transantarctic Mountains: a key parameter in ice-sheet reconstructions. Geology, 21, 10.1130/0091-7613(1994)022<0668:MPPUOT>2.3.CO;2.2.3.CO;2>CrossRefGoogle Scholar
Supplementary material: PDF

Verret et al. supplementary material

Verret et al. supplementary material

Download Verret et al. supplementary material(PDF)
PDF 825.8 KB