Holocene climate change in East Antarctica (EA)

From ACCE
Revision as of 16:39, 22 August 2014 by Tonyp (Talk | contribs)

(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)
Jump to: navigation, search
This page is part of the topic Regional patterns of holocene climate change in Antarctica

Deglaciation history and the Pleistocene-Holocene transition

3.25 Selected Holocene environmental changes – East Antarctica

The widespread Antarctic early Holocene optimum between 11.5 and 9 ka BP is observed in all ice cores from coastal and continental sites (Steig et al., 2000[1]; Masson-Delmotte et al., 2004[2]) and coincided with biogenic sedimentation commencing in lakes along the East Antarctic margin and the occupation of ice-free land by biota between c. 13.5 and 10 ka BP (Figure 3.25). The early Holocene climate evolution seems to be consistent in the regions studied so far (i.e., Amery Oasis (70°40’S-68°00’E), the Larsemann Hills (69°20’S-76°50’E), and the Vestfold Hills (68°30’78°00’E)). In the Larsemann Hills some areas escaped glaciation during the LGM, whereas other areas became gradually ice-free between c. 13.5 and 4 ka BP (Hodgson et al., 2001[3]). Diatoms and pigment data point to the establishment of seasonally melting lake ice and snow cover and the development of microbial mats at c. 10.8 ka BP (Hodgson et al., 2005[4]), with evidence for relatively wet conditions between c. 11.5 and 9.5 ka BP in a lake on one of the northern islands (Verleyen et al., 2004b[5]). This is consistent with palaeolimnological evidence from the nearby Vestfold Hills, which were probably more fully glaciated during the LGM, but where lakes became ice-free and diatoms and rotifers inhabited the lakes from c. 11.4 14C ka BP (c. 13.2-13.4 ka BP); (Roberts and McMinn, 1999a[6]; Cromer et al., 2005[7]). At least parts of Amery Oasis were covered by locally expanded glaciers during the Late Pleistocene (Hambrey et al., 2007[8]), and deglaciation in some areas started around c. 11 ka BP (Fink et al., 2006[9]), whereas biogenic sediments started to accumulate in lakes in other parts of the region at c. 12.5 ka BP (Wagner et al., 2004[10]; Wagner et al., 2007[11]), broadly coincident with deglaciation of the Larsemann and Vestfold Hills. Deglaciation was followed by the establishment of a diatom community in one of the lakes, likely related to increased nutrient inputs and a reduction in ice and snow cover at c. 10.2 ka BP, marking the start of relatively warm conditions (Wagner et al., 2004[10]).

In Wilkes Land, parts of the Bunger Hills (66°10’S-101°00’E) remained ice-free during the LGM (Gore et al., 2001[12]), whereas the Windmill Islands (66°20’S-110°30’E) were probably glaciated (Goodwin, 1993[13]). Minimum ages for deglaciation in the Windmill Islands are also slightly younger than those from the oases near the Lambert Glacier; post-glacial lake sediments accumulated at c. 10.2 ka BP (Roberts et al., 2004[14]), biogenic sedimentation in the marine bays started around c. 10.5 ka BP (Cremer et al., 2003[15]; Hodgson et al., 2003[16]), and penguins occupied the region from at least c. 9 ka BP (Emslie and Woehler, 2005[17]). Relatively cool summer conditions near the Windmill Islands probably prevailed during the early Holocene, as reflected by the microfossil record in coastal marine sediment cores (Cremer et al., 2003[15]). The Bunger Hills were occupied by snow petrels from at least 10 ka (Verkulich and Hiller, 1994[18]), and organic sediments started to accumulate in the lakes there at the Pleistocene-Holocene boundary in association with extensive and relatively rapid ice melting, which similarly points to an early Holocene warm period at c. 9 +/- 0.5 ka BP (Verkulich et al., 2002[19]), and is followed by a marine optimum (see below, Kulbe et al., 2001[20]). Radiocarbon evidence suggests that large parts of the Southern Bunger Hills were rapidly deglaciated prior to 8 ka BP (Melles et al., 1997[21]).

Although terrestrial climate archives are present in the ice-free regions in Dronning Maud Land (Matsumoto et al., 2006[22]), surprisingly little information is available about the deglaciation and post-glacial climate evolution there. The Untersee Oasis (71°S-13°E) was probably ice-free during the LGM as shown by 14C dating of organic deposits from snow petrels, which indicates an occupation during at least the past 34 ka (Hiller et al., 1988[23]). Some islands in the Lützow-Holm Bay near Syowa Station (69°00'S-39°35'E) are believed to have been ice-free for at least 40 ka and probably longer, as evidenced by AMS 14C dates of individual in situ marine fossils from raised beach deposits (Miura et al., 1998[24]).

In summary, parts of some East Antarctic oases escaped glaciation during the LGM, whereas others were probably glaciated and became gradually ice-free at the Pleistocene-Holocene boundary, with some regional differences in the timing of deglaciation and colonization by biota. The early Holocene climate optimum is detected in terrestrial and coastal marine records between c. 11.5-9.5 ka BP, centred on c. 10 ka BP, when most glaciated regions became ice-free and organic deposits started to accumulate, and in ice cores between 11.5 and 9 ka BP (Masson et al., 2000[25]).

After the early Holocene

All eastern Antarctic sites show a weak climate optimum between 6 and 3 ka but in general the period after deglaciation shows complex and less consistent patterns than those observed at the Holocene-Pleistocene boundary. In the oases near the Lambert Glacier, relatively dry conditions occurred on land between c. 9.5 and 7.4 ka BP, and in the Larsemann Hills, lake levels dropped below their present position (Verleyen et al., 2004b[5]). Marine sediments in isolation basins at c. 7.4 and 5.2 ka BP are consistent with a marine climate optimum, which is not clearly evident in the terrestrial sediments. In contrast, warm conditions prevailed on land in the Amery Oasis in the early Holocene (c. 10.2 ka BP), which lasted until c. 6.7 ka BP and with a clear optimum between c. 8.6 and 8.4 ka BP (Cremer et al., 2007[26]), whereas cold conditions prevailed from c. 6.7 ka BP onwards until c. 3.7 ka BP (Wagner et al., 2004[10]). In the Vestfold Hills, isostatic rebound and the emergence of isolation lakes from the sea resulted in a major ecosystem change, which hampers detailed palaeoclimatological inferences from being made for the period after the early Holocene optimum, particularly in lower altitude lakes (Fulford-Smith and Sikes 1996[27]; Roberts and McMinn, 1999b[28]).

In Wilkes Land, open water conditions were inferred from the marine sediments of an isolation basin in the Windmill Island between c. 8 and 4.8 14C ka BP (c. 9–4.5 ka BP), but the dating uncertainty is large because 14C dates are few and there is a variable reservoir effect throughout the sediment (Roberts et al., 2004[14]). Relatively cool summer conditions were observed in a nearby marine bay with a combination of winter sea ice and seasonal open water conditions between c. 10.5-4 ka BP (Cremer et al., 2003[15]; Hodgson et al., 2003[16]). The peak of this (marine) cooling period was pinpointed at between c. 7 ka and 5 ka BP, when penguin colonies where abandoned on one of the peninsulas (Emslie and Woehler, 2005[17]). In the Bunger Hills cold and dry conditions prevailed between c. 9 and 5.5 ka BP, with a low input of glacial meltwater in the lakes and a permanent lake-ice cover (Verkulich et al., 2002[19]), coincident with the extensive occupation of snow petrels between c. 8 and 6 ka BP (Verkulich and Hiller, 1994[18]) probably as a result of more distal glaciers and snow fields. In contrast, a marine optimum was identified in coastal sediments between c. 9.4 to 7.6 ka BP (Kulbe et al., 2001[20]), followed by cold marine conditions between c. 7.6 and 4.5 ka BP deduced from low organic carbon accumulation rates.

In summary, the period following deglaciation shows complex patterns with a marine climate optimum in some areas apparently out of phase with a terrestrial optimum or coincident with cool and dry conditions on land. Dating uncertainties prevent an in depth correlation between the different anomalies. This might reflect the fact that the organic fraction in marine sediments records spring to autumn conditions (including sea ice blooms), whereas lacustrine biotic assemblages largely reflect summer conditions when the lakes are ice-free and primary production peaks, and in some cases spring (under-ice) blooms (Hodgson and Smol, 2008[29]).

Mid Holocene warm period – Hypsithermal

A Mid Holocene Hypsithermal (MHH) is present in various ice, lake and marine core records from Antarctica (see Hodgson et al., 2004a[30]) for a review) including ice-free oases near the Lambert Glacier (note: the timing of the MHH differs from the marine/sea ice inferred ‘hypsithermal’ in some of the records discussed in Changes in sea ice extent through the Holocene - suggesting that sea ice responds to different forcing). In the Larsemann Hills the MHH is dated between c. 4 and 2 ka BP. There, relatively wet conditions occurred on land, and predate the coastal marine optimum observed in isolation basins (Verleyen et al., 2004a[31]; Verleyen et al., 2004b[5]). A short return to dry conditions and low water levels is present in one of the lake records at c. 3.2 ka BP (Verleyen et al., 2004b[5]). The relatively wet MHH is coincident with the restart of biogenic sedimentation in Progress Lake at 3.5 ka BP, after at least 40 ka of permanent lake ice cover (Hodgson et al., 2006a[32]), and the formation of proglacial lakes occupying Stornes, the eastern of the two main peninsulas in the Larsemann Hills between c. 3.8 and 1.4 14C ka BP (c. 4.4 - 4.1 and 1.3 ka BP, Hodgson et al., 2001[3]). In the Amery Oasis, the relatively warm conditions of the MHH between c. 3.2 and 2.3 ka BP are inferred from abundant organic matter deposition in Lake Terrasovoje (Wagner et al., 2004[10]). In the Vestfold Hills a decline in lake salinity could be inferred between c. 4.2 and 2.2 ka BP, but dates are uncertain (Björck et al., 1996[33]; Roberts and McMinn, 1996[34], 1999a[6]). This period of low salinity is however broadly consistent with the warm and humid conditions between c. 4.7 and 3 ka BP proposed by Björck et al. (1996[33]) after reinterpretation of previously published results (Pickard et al., 1986[35]). In contrast, Bronge (1992[36]) inferred relatively cold conditions between c. 5 and 3 ka BP and a short but marked cooling event between c. 2.3 and 2 ka BP.

In Wilkes Land, enhanced biological production, probably reflecting more open water conditions and a climate optimum, occurred between c. 4 and 1 ka BP (Kirkup et al., 2002[37]). This coincided with a local marine optimum characterised by open water and stratified conditions caused by enhanced meltwater input (Cremer et al., 2003[15]). In this area the MHH coincided with the readvance of the Law Dome ice margin after c. 4 ka BP, in response to an increase in precipitation (Goodwin, 1996[38]).

In the Bunger Hills a stepwise increase in primary production was reported in the lakes between c. 4.7 and 2 ka BP (Melles et al., 1997[21]). The start of the MHH slightly postdates the start of warm conditions between c. 5.5 and 2 ka BP, interpreted by Verkulich et al. (2002[19]) from the pattern of draining of ice-dammed lakes. A marine optimum occurred in this area between c. 3.5 and 2.5 ka BP, which was preceded by a gradual warming from c. 4.5 ka BP onwards (Kulbe et al., 2001[20]).

In the Lützow-Holm Bay region a rapid isostatic rebound (6 m in c. 1,000 years) occurred between c. 4.7 and 3 ka BP, which was linked to the rapid removal of part of the regional ice mass, most likely as a result of melting caused by the MHH (Okuno et al., 2007[39]).

In summary, there is evidence for a MHH in EA, but dating uncertainties are still high in some areas. Because the MHH acts as one of the historical analogues for the present day warming climate, there is a pressing need for well-dated lake sediment records to study this past climate anomaly and its influence on ecosystem functioning. There is a disappointing lack of long-term high-resolution records from ice-free areas in the Dronning Maud Land region.

The past 2,000 years - Neoglacial cooling, the Little Ice Age and recent climate change

Much attention has been paid to the fluctuations in climate that gave rise in the Northern Hemisphere to the well-documented Medieval Warm Period (900-1300 AD / 1100-700 yr BP), and Little Ice Age (LIA) consisting of cool intervals beginning about 1650, 1770, and 1850 AD, each separated by slight warming intervals / 300-100 yr BP), the earlier part of the latter coinciding with the Maunder Minimum, a period of minimal sunspot activity and low solar output from 1645-1715 AD / 355-285 yr BP). The search in Antarctica for short-term climate signals like these that are apparent in the Northern Hemisphere is an important element in understanding how the Earth’s climate system works.

In the Lambert Glacier region there is some evidence of Neoglacial cooling in the Larsemann Hills (Hodgson et al., 2005[4]) leading to dry conditions around 2000 yr BP, 700 yr BP and between c. 300 and 150 yr BP in some of the lakes (Verleyen et al., 2004b[5]). The lake evidence parallels declines in sea bird populations during the past 2,000 years (Liu et al., 2007[40]). In the Amery Oasis, Neoglacial cooling followed the MHH from c. 2.3 ka BP onwards, with a short return to a relatively warmer climate between c. 1.5 and 1 ka BP (Wagner et al., 2004[10]). In the Vestfold Hills an increase in fast ice extent is observed from c. 1.7 ka BP (McMinn, 2000[41]), broadly coincident with the Chelnock Glaciation on land (Adamson and Pickard, 1986[42]). Cold conditions were inferred between c. 1.3 ka BP and 250 yr BP (Bronge, 1992[36]) and low precipitation has tentatively been inferred from c. 1.3-1.5 ka BP, but dating uncertainty is high (Roberts and McMinn, 1999a[6]). Meltwater input into the lakes gradually decreased from 3 ka BP onwards (Fulford-Smith and Sikes, 1996[27]). A palaeohydrological model derived from the reconstructed changes in salinity and water level throughout sediment cores suggested that there was no significant change in evaporation for the last c. 700 years, but that a lower evaporation period is evident between c. 150 - 200 yr BP, suggestive of a mild LIA-like event in the Vestfold Hills (Roberts et al., 2001[43]).

In Wilkes Land, Neoglacial cooling and persistent sea ice cover were observed in the marine bays near the Windmill Islands (Kirkup et al., 2002[37]; Cremer et al., 2003[15]). A slow decrease in lake water salinity was observed on land during the late Holocene, and there is no evidence for a LIA-like event there (Hodgson et al., 2006c[44]). Instead, we see a very rapid salinity rise during the past few decades that has been brought about either by increased evaporation rates or a decrease in precipitation. In the Bunger Hills a rapid growth of the petrel population after c. 2 ka BP was reported by Verkulich and Hiller (1994[18]), and coincides with climate cooling (Melles et al., 1997[21]; Verkulich et al., 2002[19]) and a glacier readvance during recent centuries (Adamson and Colhoun, 1992[45]). After the cooling event at c. 2 ka BP relatively warm conditions, yet colder than during the Hypsithermal, prevailed with an additional cooling trend that started recently (Verkulich et al., 2002[19]).

In the Lützow-Holm Bay region lake sediment cores are starting to provide insights in past climate variability during the late Holocene. An increase in the total organic carbon to total nitrogen ratio was linked to an increase in the aquatic moss vegetation from c. 1.1 ka BP onwards (Matsumoto et al., 2006[22]). However, detailed palaeoclimatic records are still lacking for the region.

In summary, most East Antarctic areas experienced Neoglacial cooling, with some markedly cooler or drier events in places. The apparent differences in the dating of these events from one part of the region to another might be related to dating uncertainties or to the lack of high-resolution records in some areas. With the exception of some ice cores (such as Law Dome where CO2 mixing ratios were in the range 275-284 ppm with the lower levels during 1550-1800 AD, probably as a result of colder climate (Etheridge et al., 1996[46])) there is neither convincing evidence for a Little Ice Age event, nor for anything corresponding convincingly, and region-wide, to the Medieval Warm Period of the Northern Hemisphere.

References

  1. Steig, E.J., Morse, D.L., Waddington, E.D., Stuiver, M., Grootes, P.M., Mayewski, P.A., Twickler, M.S. and Whitlow, S.I. 2000. Wisconsinan and Holocene climate history from an ice core at Taylor Dome, western Ross Embayment, Antarctica, Geografiska Annaler Series a-Physical Geography, 82A, 213-235.
  2. Masson-Delmotte, V., Stenni, B. and Jouzel, J. 2004. Common millennial-scale variability of Antarctic and Southern Ocean temperatures during the past 5000 years reconstructed from the EPICA Dome C ice core, The Holocene, 14, 145-151.
  3. 3.0 3.1 Hodgson, D.A., Noon, P.E., Vyverman, W., Bryant, C.L., Gore, D.B., Appleby, P., Gilmour, M., Verleyen, E., Sabbe, K., Jones, V.J., Ellis-Evans, J.C. and Wood, P.B. 2001. Were the Larsemann Hills ice-free through the Last Glacial Maximum?, Antarctic Science, 13, 440-454.
  4. 4.0 4.1 Hodgson, D.A., Verleyen, E., Sabbe, K., Squier, A.H., Keely, B.J., Leng, M.J., Saunders, K.M. and Vyverman, W. 2005. Late Quaternary climate-driven environmental change in the Larsemann Hills, East Antarctica, multi-proxy evidence from a lake sediment core, Quaternary Research, 64, 83-99.
  5. 5.0 5.1 5.2 5.3 5.4 Verleyen, E., Hodgson, D.A., Sabbe, K. and Vyverman, W. 2004b. Late Quaternary deglaciation and climate history of the Larsemann Hills (East Antarctica), Journal of Quaternary Science, 19, 361-375.
  6. 6.0 6.1 6.2 Roberts, D. and McMinn, A. 1999a. A diatom-based palaeosalinity history of Ace Lake, Vestfold Hills, Antarctica, The Holocene, 9, 401-408.
  7. Cromer, L., Gibson, J.A.E., Swadling, K. and Ritz, D.A. 2005. Faunal microfossils: Indicators of Holocene ecological change in a saline Antarctic lake, Palaeogeography, Palaeoclimatology, Palaeoecology, 221, 83-97.
  8. Hambrey, M.J., Glasser, N.F., McKelvey, B.C., Sugden, D.E. and Fink, D. 2007. Cenozoic Landscape Evolution of an East Antarctic Oasis (Radok Lake Area, Northern Prince Charles Mountains), and Its Implications for the Glacial and Climatic History of Antarctica, Quaternary Science Reviews, 26, 598-626.
  9. Fink, D., McKelvey, B., Hambrey, M.J., Fabel, D. and Brown, R. 2006. Pleistocene Deglaciation Chronology of the Amery Oasis and Radok Lake, Northern Prince Charles Mountains, Antarctica. Earth and Planetary Science Letters, 243, 229-243.
  10. 10.0 10.1 10.2 10.3 10.4 Wagner, B., Cremer, H., Hultzsch, N., Gore, D.B. and Melles, M. 2004. Late Pleistocene and Holocene history of Lake Terrasovoje, Amery Oasis, East Antarctica, and its climatic and environmental implications, Journal of Paleolimnology, 32, 321-339.
  11. Wagner, B., Hultzsch, N., Melles, M. and Gore, D.B. 2007. Indications of Holocene sea-level rise in Beaver Lake, East Antarctica, Cambridge Journals Online, 125-128.
  12. Gore, D.B., Rhodes, E.J., Augustinus, P.C., Leishman, M.R., Colhoun, E.A. and Rees-Jones, J. 2001. Bunger Hills, East Antarctica: Ice free at the Last Glacial Maximum, Geology, 29, 1103-1106.
  13. Goodwin, I.D. 1993. Holocene deglaciation, sea-level change, and the emergence of the Windmill Islands, Budd Coast, Antarctica, Quaternary Research, 40, 70-80.
  14. 14.0 14.1 Roberts, D., McMinn, A., Cremer, H., Gore, D. and Melles, M. 2004. The Holocene evolution and palaeosalinity history of Beall Lake, Windmill Islands (East Antarctica) using an expanded diatom-based weighted averaging model, Palaeogeography Palaeoclimatology Palaeoecology, 208, 121-140.
  15. 15.0 15.1 15.2 15.3 15.4 Cremer, H., Gore, D., Melles, M., Roberts, D. 2003. Palaeoclimatic significance of late Quaternary diatom assemblages from southern Windmill Islands, East Antarctica, Palaeogeography Palaeoclimatology Palaeoecology, 195, 261-280.
  16. 16.0 16.1 Hodgson, D.A., McMinn, A., Kirkup, H., Cremer, H., Gore, D., Melles, M., Roberts, D. and Montiel, P. 2003. Colonization, succession, and extinction of marine floras during a glacial cycle: A case study from the Windmill Islands (east Antarctica) using biomarkers, Paleoceanography, 18(3), 1067, doi:1010.1029/2002PA000775.
  17. 17.0 17.1 Emslie, S.D., Woehler, E.J. 2005. A 9000-year record of Adélie penguin occupation and diet in the Windmill Islands, East Antarctica, Antarctic Science, 17, 57-66.
  18. 18.0 18.1 18.2 Verkulich, S.R. and Hiller, A. 1994. Holocene Deglaciation of the Bunger Hills Revealed by C-14 Measurements on Stomach Oil Deposits in Snow Petrel Colonies, Antarctic Science, 6, 395-399.
  19. 19.0 19.1 19.2 19.3 19.4 Verkulich, S.R., Melles, M., Pushina, Z.V. and Hubberten, H-W. 2002. Holocene Environmental changes and development of Figurnoye Lake in the southern Bunger Hills, East Antarctica, Journal of Paleolimnology, 28, 253-267.
  20. 20.0 20.1 20.2 Kulbe, T., Melles, M., Verkulich, S.R. and Pushina, Z.V. 2001. East Antarctic climate and environmental variability over the last 9400 years inferred from marine sediments of the Bunger Oasis, Arctic Antarctic and Alpine Research, 33, 223-230.
  21. 21.0 21.1 21.2 Melles, M., Kulbe, T., Verkulich, S.R., Pushina, Z.V. and Hubberten, H-W. 1997. Late Pleistocene and Holocene environmental history of Bunger Hills, East Antarctica, as revealed by fresh-water and epishelf lake sediments. In: Ricci CA (ed) The Antarctic Region: Geological Evolution and Processes. Proceedings of the VII International Symposium on Antarctic Earth Sciences, Siena, 1995. Terra Antarctica, Sienna, 809-820.
  22. 22.0 22.1 Matsumoto, G.I., Komori, K., Enomoto, A., Imura, S., Takemura, T., Ohyama, Y. and Kanda, H. 2006. Environmental changes in Syowa Station area of Antarctica during the last 2300 years inferred from organic components in lake sediment cores, Polar Biosciences, 19, 52-61.
  23. Hiller, A., Wand, U., Kampf, H. and Stackebrandt, W. 1988. Occupation of the Antarctic continent by petrels during the past 35 000 years: inferences from a 14C study of stomach oil deposits, Polar Biology, 9, 69-77.
  24. Miura, H., Maemoku, H., Igarishi, A. and Moriwaki, K. 1998. Late Quaternary raised beach deposits and radiocarbon dates of marine fossils around Lützow-Holm Bay, Special map series of National Institute of Polar Research, 6, Tokyo.
  25. Masson, V., Vimeux, F., Jouzel, J., Morgan, V., Delmotte, M., Ciais, P., Hammer, C., Johnsen, S., Lipenkov, V.Y., Mosley-Thompson, E., Petit, J.R., Steig, E.J., Stievenard, M. and Vaikmae, R. 2000. Holocene climate variability in Antarctica based on 11 ice-core isotope records, Quaternary Research, 54, 348-358.
  26. Cremer, H., Heiri, O., Wagner, B. and Wagner-Cremer, F. 2007. Abrupt climate warming in East Antarctica during the early Holocene, Quaternary Science Reviews, 26, 2012-2018.
  27. 27.0 27.1 Fulford-Smith, S.P. and Sikes, E.L. 1996. The evolution of Ace Lake, Antarctica, determined from sedimentary diatom assemblages, Palaeogeography Palaeoclimatology Palaeoecology, 124, 73-86.
  28. Roberts, D. and McMinn, A. 1999b. Palaeohydrological modelling of Ace Lake, Vestfold Hills, Antarctica, The Holocene, 9, 515-520.
  29. Hodgson, D.A. and Smol, J.P. 2008. High latitude palaeolimnology. In Polar Lakes and Rivers - Arctic and Antarctic Aquatic Ecosystems, (W. F. Vincent, and J. Laybourn-Parry, Eds.). Oxford University Press, Oxford, UK.
  30. Hodgson, D.A., Doran, P.T., Roberts, D. and McMinn, A. 2004a. Paleolimnological studies from the Antarctic and subantarctic islands. In: Pienitz R, Douglas MSV, Smol JP (eds) Developments in Palaoenvironmental Research. Long-term Environmental Change in Arctic and Antarctic Lakes, 8, Springer, Dordrecht, 419-474.
  31. Verleyen, E., Hodgson, D.A., Sabbe, K., Vanhoutte, K. and Vyverman, W. 2004a. Coastal oceanographic conditions in the Prydz Bay region (East Antarctica) during the Holocene recorded in an isolation basin, The Holocene, 14, 246-257.
  32. Hodgson, D.A., Verleyen, E., Squier, A.H., Sabbe, K., Keely, B.J., Saunders, K.M. and Vyverman, W. 2006a. Interglacial environments of coastal east Antarctica: comparison of MIS 1 (Holocene) and MIS 5e (Last Interglacial) lake-sediment records, Quaternary Science Reviews, 25, 179-197.
  33. 33.0 33.1 Björck, S., Olsson, S., Ellis-Evans, J.C., Häkansson H., Humlum O., De Lirio, J.M. 1996. Late Holocene palaeoclimatic records from lake sediments on James Ross Island, Antarctica, Palaeogeography Palaeoclimatology Palaeoecology, 121, 195-220.
  34. Roberts, D. and McMinn, A. 1996. Palaeosalinity reconstruction from saline lake diatom assemblages in the Vestfold Hills, Antarctica. In: Lyons WB, Howard-Williams C, Hawes I (eds) Ecosystem Processes in Antarctic Ice-Free Landscapes: Polar Desert Ecosystems Conference Proceedings, Christchurch, New Zealand, 207-220,
  35. Pickard, J., Adamson, D.A., Harwood, D.M., Miller, G.H., Quilty, P.G., Dell, R.K. 1986. Early Pliocene marine sediments in the Vestfold Hills, East Antarctica: implications for coastline, ice sheet and climate, South African Journal of Science, 82, 520-521.
  36. 36.0 36.1 Bronge, C. 1992. Holocene climatic record from lacustrine sediments in a freshwater lake in the Vestfold Hills, Antarctica, Geografiska Annaler, 74A, 47-58.
  37. 37.0 37.1 Kirkup, H., Melles, M. and Gore, D.B. 2002. Late Quaternary environment of Southern Windmill Islands, East Antarctica, Antarctic Science, 14, 385-394.
  38. Goodwin, I.D. 1996. A mid to late Holocene re advance of the Law Dome ice margin, Budd Coast, East Antarctica, Antarctic Science, 8, 395-406.
  39. Okuno, J., Miura, H. and Maemoku, H. 2007. The possibility of the rapid melting of ice sheet in Holocene around the Lutzow-Holm Bay, Antarctica, Geophysical Research Abstracts, 7.
  40. Liu, X., Sun, L., Xie, Z., Yin, X., Zhu, R. and Wang, Y. 2007. A preliminary record of the historical seabird population in the Larsemann Hills, East Antarctica, from geochemical analyses of Mochou Lake sediments, Boreas, 36, 182-197.
  41. McMinn, A. 2000. Late Holocene increase in sea ice extent in fjords of the Vestfold Hills, eastern Antarctica, Antarctic Science, 12, 80-88.
  42. Adamson, D.A., Pickard, J. 1986. Physiography and geomorphology of the Vestfold Hills. In: Pickard J (ed) Antarctic Oasis: terrestrial environments and history of the Vestfold Hills. Academic Press, Sydney, 99-139.
  43. Roberts, D., Van Ommen, T.D., McMinn, A., Morgan, V. and Roberts, J.L. 2001. Late Holocene East Antarctic climatic trends from ice core and lake sediment proxies, The Holocene, 11, 117-120.
  44. Hodgson, D.A., Roberts, D., McMinn, A., Verleyen, E., Terry, B., Corbett, C. and Vyverman, W. 2006c. Recent rapid salinity rise in three East Antarctic lakes, Journal of Paleolimnology, 36, 385-406.
  45. Adamson, D.A., Colhoun, E.A. 1992. Late Quaternary Glaciation and Deglaciation of the Bunger Hills, Antarctica, Antarctic Science, 4, 435-446.
  46. Etheridge, D.M., Steele, L.P., Langenfields, R.J. Francey, J.M. Barnola and Morgan, V.I. 1996. Natural and anthropogenic changes in atmospheric CO2 over the last 1000 years from air in Antarctic firn and ice, J. Geophys. Res., 101, 4115-4128.