Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-06-20T10:48:02.611Z Has data issue: false hasContentIssue false

A multi-century ice-core perspective on 20th-century climate change with new contributions from high-Arctic and Greenland (PARCA) cores

Published online by Cambridge University Press:  14 September 2017

Ellen Mosley-Thompson
Affiliation:
Byrd Polar Research Center, The Ohio State University, 1090 Carmack Road, Columbus, OH 43210-1002, USA E-mail: thompson.4@osu.edu Department of Geography, The Ohio State University, 1036 Derby Hall, 154 North Oval Mall, Columbus, OH 43210-1361, USA
Lonnie G. Thompson
Affiliation:
Byrd Polar Research Center, The Ohio State University, 1090 Carmack Road, Columbus, OH 43210-1002, USA E-mail: thompson.4@osu.edu Department of Geological Sciences, The Ohio State University, 275 Mendelhall Laboratory, 125 South Oval Mall, Columbus, OH 43210-1308, USA
Ping-Nan Lin
Affiliation:
Byrd Polar Research Center, The Ohio State University, 1090 Carmack Road, Columbus, OH 43210-1002, USA E-mail: thompson.4@osu.edu
Rights & Permissions [Opens in a new window]

Abstract

A global collection of high temporally resolved ice-core-derived δ18O records is examined to assess whether the proxy records are consistent with contemporaneous observed temperature variations in their respective regions. This is prerequisite to using the older parts of the proxy (δ18O) records to assess whether 20th-century temperatures remain within the range of longer-term natural variability. Excluding the high plateaus in East and West Antarctica where 20th-century temperatures show modest cooling, the ice-core records from other regions suggest modest to strong 20th-century warming. The recent warming over Greenland has been modest and spatially variable. The 20th-century warming over both the Barents Sea region and the Tibetan Plateau now falls well outside the range of prior longer-term temperature variability. Similarly, over the South American Andes and the Antarctic Peninsula the recent warming exceeds the long-term mean for the last 1000 and 500 years, respectively. The ice fields in these regions are in danger of being compromised or lost if the present warming trend in these regions persists.

Type
Research Article
Copyright
Copyright © The Author(s) [year] 2006

Introduction

Since the late 1960s when the first long cores were recovered from Greenland and Antarctica, ice cores have provided unique details about the nature of the Earth’s climate system with emphasis on glacial–interglacial cycles. Recently, greater attention has been focused on reconstructing high-resolution (optimally annual) proxy histories for the Holocene, with emphasis on the last few thousand years. This attention results, in part, from the need to assess the well-documented 20th-century warming of the Earth’s surface temperatures from a longer-term perspective (Reference Mann, Bradley and HughesMann and others, 1999; Reference Levitus, Antonov, Boyer and StephensLevitus and others, 2000; Reference HansenHansen and others, 2001; Reference Jones and MobergJones and Moberg, 2003; Reference Moberg, Sonechkin, Holmgren, Datsenko and KarlénMoberg and others, 2005). Restricted to very cold and/or high locations, ice-core-derived proxy histories provide a partial picture of past climatic and environmental conditions. These frozen archives are powerful contributors to multi-proxy reconstructions, providing multiple lines of evidence (e.g. changes in atmospheric dustiness and chemistry, in mass accumulation and in temperature) from remote locations where other proxy data are unavailable.

Over the last few decades, the Ice Core Paleoclimate Research Group at The Ohio State University (OSU) has collected a suite of tropical and subtropical cores from the South American Andes and the Tibetan Plateau, as well as polar cores from Antarctica, Greenland and Franz Josef Land (FJL), Russian Arctic. The analyses of these ice cores have emphasized the highest time resolution possible. More recently, a number of new annually resolved ice-core records from Greenland have been added to the OSU collection. The cores were acquired between 1995 and 1999 as part of the NASA–NSF (US National Science Foundation) PARCA (Program for Arctic Regional Climate Assessment) project (Reference Thomas and InvestigatorsThomas and others, 2001). These cores are particularly valuable as they come from sites that are widely distributed along the western side of Greenland. Previously, most available high-resolution Greenland ice-core histories that extend back more than a century were drilled in central and north-central Greenland or in southern Greenland near the Dye 2 and Dye 3 radar stations (now abandoned).

Ice-Core-Derived δ18O Records

Figure 1 presents the decadally averaged δ18O records organized geographically (north to south), along with regional maps showing their locations. The records from Tibet, the Andes and Antarctica have been published separately, but not collectively with the newer δ18O histories from Greenland and FJL. This synthesis is limited to the last millennium, as many of the records are relatively short, extending back just a few centuries to one half-millennium. Six of the δ18O records extend to earlier millennia, four into the Last Glacial Stage. The longer records include those from the Dunde and Guliya ice cores from the Tibetan Plateau (Reference ThompsonThompson and others, 1989, Reference Thompson1997, respectively), the Quelccaya and Huascarán cores from Peru (Reference Thompson, Mosley-Thompson, Dansgaard and GrootesThompson and others, 1986, Reference Thompson1995, respectively), the Sajama ice cap in Bolivia (Reference ThompsonThompson and others, 1998) and the Plateau Remote site from East Antarctica (Reference Mosley-Thompson, Jones, Bradley and JouzelMosley-Thompson, 1996). Table 1 provides basic information about each core. Details of the dating procedures are in the references.

Fig. 1. Decadally averaged δ18O histories are shown from north to south (top to bottom). Horizontal bars in each record show the average δ18O values for the pre- and post-1900 time intervals. The 1000 year Northern Hemisphere reconstruction is from Reference Mann, Bradley and HughesMann and others (1999) and updated by Reference Mann and JonesMann and Jones (2003). Superimposed is the observed near-surface temperature record (Reference Jones and MobergJones and Moberg, 2003). Maps show locations for the cores.

Table 1. Information for all ice cores discussed in the text Timescale under revision.

The δ18O histories from two different drilling projects at the Summit Site in central Greenland are included for comparison with the newer Greenland records (Fig. 1). The Greenland Ice Sheet Project 2 (GISP2) record (Reference Grootes, Stuiver, White, Johnsen and JouzelGrootes and others, 1993) is shown along with a composite from two cores, called Site T cores, drilled 4 km apart at the Summit in 1989 by OSU (Reference Mosley-Thompson, Thompson, Dai, Davis and LinMosley-Thompson and others, 1993). The δ18O history from the Windy Dome Ice Field on Graham Bell Island, FJL, (Reference HendersonHenderson, 2002) is included along with the five new Greenland δ18O histories: GITS, NASA-U, D2, D3, Raven (Fig. 1), collected by the PARCA initiative. With the exception of the GISP2 core, all records discussed here were analyzed and dated by the authors using consistent procedures. Moreover, unlike many of the earlier Greenland records, the timescales for the PARCA cores were constructed using three seasonally varying parameters: the insoluble dust concentrations and oxygen isotopic ratios (δ18O), both measured at OSU, and hydrogen peroxide (H2O2) measured at the University of Arizona (Reference McConnellMcConnell and others, 2001). The timescales were further confirmed with beta radioactivity horizons from thermonuclear bomb tests and by identification of known volcanic eruptions (Reference Mosley-ThompsonMosley-Thompson and others, 2001, Reference Thompson, Mosley-Thompson, Davis, Lin, Henderson and Mashiotta2003).

To facilitate their comparison, each record has been normalized with respect to its entire length (for cores shorter than one millennium) or with respect to the last millennium for the longer cores. The annual Z score is the standardized deviation from its respective mean (annual Z score = (annual value-record mean)/record standard deviation). The annual Z scores are shown (Fig. 1) as unweighted 10 year (decadal) averages plotted at the midpoint of the decade. The goal here is to examine a global (pole-to-pole) array of ice-core-derived (δ18O) proxy temperature histories to assess 20th-century climate changes within a longer-term perspective. Thus, the averages for the pre-1901 period (1900 to the oldest year in the record or AD 1000) are compared with those from the post-1900 part of the record (1901 to the most recent year in the record). It would be ideal if all the records were of equivalent temporal length; however, the time, expense and logistics required to collect ice cores dictates the depth to which cores are recovered at a given site.

Before comparing the δ18O records and drawing general conclusions, the advantages and limitations of using δ18O as a proxy for regional air temperature are reviewed briefly. With respect to ice cores, the δ18O–air-temperature (T a) relationship has been most extensively investigated in Greenland and Antarctica. In Antarctica, a strong linear relationship between δ18O and Ta has been demonstrated (Reference Aldaz and DeutschAldaz and Deutsch, 1967; Reference Mosley-Thompson, Bradley and JonesMosley-Thompson, 1992; Reference Peel, Bradley and JonesPeel, 1992; Reference JouzelJouzel, 1999), with slope (a) values ranging from 0.76 to 0.92%˚C−1. In Greenland the slope of the linear relationship ranges from 0.62 to 0.67%˚C−1 (Reference Dansgaard, Johnsen, Clausen and GundestrupDansgaard and others 1973; Reference Johnsen, Dansgaard and WhiteJohnsen and others, 1989). More recently, using central Greenland borehole temperatures, Reference Cuffey, Clow, Alley, Stuiver, Waddington and SaltusCuffey and others (1995) calibrated the paleothermometer, δ18O = αTa + β. They concluded that δ18O provides a ‘faithful’ proxy for long-term average temperature at that site (GISP2); however, they caution that it is inappropriate to use a single set of constants and β) to infer past climate changes, as they depend upon factors that change with time. Thus, in keeping with the general practice in ice-core paleothermometry (Reference JouzelJouzel and others, 1997; Reference JouzelJouzel, 1999), we adopt the convention for the polar cores such that more negative δ18O values reflect cooler air temperatures and less negative δ18O values reflect warmer air temperatures at the time of condensation.

There has been much less investigation of the δ18O–T a relationship in non-polar ice cores, in part because records have been limited. Reference Rozanski, Araguás-Araguás and GonfiantiniRozanski and others (1992, Reference Rozanski, Araguás-Araguás, Gonfiantini, Swart, Lohmann, McKenzie and Savin1993) reviewed an extensive set of in situ data collected over three decades, mainly from mid-latitude, lower-elevation sites. They suggested that as a first approximation, temperature controls the isotopic composition of precipitation at high and mid-latitudes while the amount of precipitation (the amount effect) controls the isotopic composition in tropical regions (Reference Rozanski, Johnsen, Schotterer and ThompsonRozanski and others, 1997). They also concluded that the relationship between longer-term changes in δ18O and T a for a given location is more appropriate for paleoclimate investigations (vs short-term applications).

Additional studies have addressed the roles of air temperature and the precipitation ‘amount effect’ in determining the δ18O values in precipitation (snowfall) from a regional perspective. Reference Yao, Thompson, Mosley-Thompson, Zhihong, Xingping and LinYao and others (1996) conducted the first study of the δ18O–T a relationship in discrete precipitation events at three sites on the Tibetan Plateau. They reported that averaging δ18O and air temperatures for discrete precipitation events over longer periods (months to years) removed the synoptic effects and pointed to air temperature as the dominant control on longer-term δ18O variations at individual sites.

Reference Thompson, Mosley-Thompson and HendersonThompson and others (2000) address the conundrum with Andean precipitation whereby the seasonal δ18O–T a relationship is opposite that in the polar regions. In the Andes, summer snow has more negative δ18O values than winter snow, and the seasonal δ18O range can be large, up to 20%, although the seasonal temperature range is small. Reference Grootes, Stuiver, Thompson and Mosley-ThompsonGrootes and others (1989) used a three-step model, employing Rayleigh fractionation in each step, to explain the δ18O values measured on the Quelccaya ice cap. They concluded that air circulation and air mass stability, rather than temperature, determine the seasonal δ18O cycle over Quelccaya. They did not address the longer-term relationship between δ18O and Ta in the tropics that appears consistent with that in the higher latitudes (more negative δ18O implies colder T a at condensation). Reference Thompson, Mosley-Thompson and HendersonThompson and others (2000) suggest that seasonal changes in the cycle of deep tropical convection modulate the height of the mean condensation level (hence temperature of condensation). Deep summer convection produces precipitation that is more isotopically depleted than winter precipitation that condenses at lower levels in the atmosphere. Reference Henderson, Thompson and LinHenderson and others (1999) and Reference VuilleVuille and others (2003) have investigated the controls on recent snowfall over Sajama, Quelccaya and Huascaran. Reference Henderson, Thompson and LinHenderson and others (1999) note that the spatial distribution of temperature anomalies in the western tropical Atlantic influences atmospheric circulation at 500 hPa and thereby isotopic fractionation over the Amazon Basin, the primary source of precipitation for Huascaran and Quelccaya. Reference VuilleVuille and others (2003) note that their modeled δ18O values depend strongly on precipitation amount, but conclude that the δ18O signal in precipitation is influenced by a combination of mechanisms (precipitation amount, temperature, variability of the moisture source and changes in atmospheric circulation). Reference Bradley, Vuille, Hardy and ThompsonBradley and others (2003) report strong linkages between sea surface temperatures across the equatorial Pacific Ocean and δ18O in ice cores from the tropical Andes as well as the Himalaya (Dasuopu glacier). They, like Reference Henderson, Thompson and LinHenderson and others (1999), report a strong link between δ18O and El Nino-Southern Oscillation (ENSO) variability, which is not surprising as ENSO integrates all the mechanisms mentioned above. The δ18O– T a relationship on longer timescales (decades to centuries) is of greater importance here. Reference Thompson, Mosley-Thompson and HendersonThompson and others (2000, Reference Thompson, Mosley-Thompson, Davis, Lin, Henderson and Mashiotta2003) present evidence that on longer timescales, δ18O variations in Andean precipitation more strongly reflect variations in air temperature than precipitation. Nevertheless, this issue is ripe for further investigation, as the dearth of records and their short duration present challenges to fully addressing the controls on the δ18O– T a relationship at high-elevation sites in the low latitudes. Here the trends in δ18O, when averaged over decadal and longer timescales, are assumed to reflect primarily trends in air temperature rather than in precipitation.

Discussion

Placing the well-documented 20th-century warming of the Earth’s globally averaged temperature within a temporal perspective longer than existing meteorological (Reference Jones and MobergJones and Moberg, 2003) and oceanographic (Reference Levitus, Antonov, Boyer and StephensLevitus and others, 2000) observations (∼1860 to the present) requires proxy records such as ice-core-derived δ18O histories. Figure 1 reveals that the averaged δ18O for post-1900 precipitation is enriched (to differing degrees) relative to that for pre-1901 precipitation in all cores except those from Siple Station, West Antarctica, and Plateau Remote, East Antarctica. The observed regional trends in temperatures and the associated ice-core-derived δ18O records are discussed below. If the 20th-century δ18O trends among the ice cores are broadly consistent with the contemporaneous temperature trends for their respective regions, then the earlier portions (pre-20th-century) of the δ18O records should also reflect regional temperature trends and provide a viable longer-term perspective for the recent warming.

Large regional differences exist among the Greenland δ18O histories, even between the GISP2 and Summit 1989 records that are ∼ 4 km apart. The 20th-century enrichment appears larger in Summit 1989 than in GISP2, but this does not result from the different record lengths (∼500 vs ∼1000years, respectively). To facilitate comparison, the average δ18O value was calculated for the GISP2 segment that overlaps the Summit 1989 record, and, as Figure 1 reveals, the 500 year δ18O average is only slightly less than that for the 1000 year segment. Thus, the difference cannot be resolved with data in hand, and likely reflects some combination of differences in dating approaches, sampling schema and real spatial differences among the cores.

The 5 year unweighted running means of the annual δ18O values (Fig. 2) highlight the regional variability that exists over Greenland on annual to decadal timescales. This variability arises from (1) differences in regional climate forcing, (2) more localized conditions (e.g. an isotopic record is only produced when snow is falling), and (3) post-depositional modification by deflation, drifting and redeposition of surface snow. Despite this regional variability, the coherency of the Greenland δ18O records (Fig. 2) provides strong evidence that δ18O of the precipitation (snow) does reflect large-scale, multi-year variations in air temperature. The vertical bars in Figure 2 highlight three periods of marked isotopic depletion (cooling) that are widely recorded along the west side of the ice sheet from the northwest (GITS) to the southwest (Raven). The two closely spaced shaded bars highlight two decades of well-documented volcanically induced cooling: (1) 1810−20 with the eruptions of Unknown in 1809 and Tambora, Indonesia, in 1815 (Reference Dai, Mosley-Thompson and ThompsonDai and others, 1991; Reference Thompson, Mosley-Thompson, Davis, Lin, Henderson and MashiottaMosley-Thompson and others, 2003) and (2) 1830−40 with the eruptions of an unidentified volcano in ∼1832 and Coseguina, Nicaragua, in 1835 (Reference Cole-Dai, Mosley-Thompson and ThompsonCole-Dai and others, 1997). The third bar highlights a brief, but large and spatially coherent, cooling from 1917 to 1919 that is documented in Arctic temperature composites (see e.g. Reference Jones and MobergJones and Moberg, 2003, fig. 2).

Fig. 2. The 5 year running mean of the annual δ18O histories from five new Greenland ice cores are shown along with the 1989 Site T record collected at the Summit (GISP2) site. Vertical shaded bars highlight specific time intervals discussed in the text. Horizontal bars in each record show the average δ18O value before and after the large cooling in 1917–19 that was followed by a rapid, widespread warming over much of the Arctic.

This brief cold event preceded an abrupt warming in the high Arctic (60−90˚ N) from ∼1920 to ∼1940 (Reference RogersRogers, 1985). The warming was strongest in the Kara and Barents Sea region and only slightly weaker in western Baffin Bay. It is hypothesized to have been linked to the retreat of sea ice in response to enhanced westerly flow (Reference Bengtsson, Semenov and JohannessenBengtsson and others, 2004). Over the ice sheet the post-1920 warming is recorded by slightly enriched δ18O values at all sites except NASA-U (Fig. 2). Note that the δ18O-inferred warming is most pronounced at the two southernmost sites (Raven and D3). The new Greenland δ18O records are relatively short and provide a limited temporal perspective, but when coupled with the longer records from Summit they suggest modest 20th-century warming over Greenland with alternating multi-decadal periods of warming and cooling, consistent with the limited available meteorological observations (for a summary see Reference BoxBox, 2002).

An ice-core record from the Windy Dome ice cap in FJL yields quite a different picture (Reference HendersonHenderson, 2002). The decadally averaged δ18O record (Fig. 1) suggests a large warming that began ∼1910 and was sustained until mid-century, consistent with observations in the Barents Sea region as discussed above. Longer ice cores from this region are needed to assess the 20th-century warming relative to that associated with the Medieval Warm Period (MWP).

The available high-resolution δ18O records from Northern Hemisphere high-latitude ice cores are generally consistent with 20th-century observations. Specifically, they record (1) the short-term cooling in response to explosive tropical volcanic eruptions that injected sulfate aerosols into the tropical stratosphere where they were globally distributed, (2) the brief large-scale cooling from 1917 to 1919, and (3) the subsequent rapid (few years), large-scale warming in the Arctic (Reference RogersRogers, 1985; Reference OverpeckOverpeck and others, 1997; Reference Jones and MobergJones and Moberg, 2003) that persisted for several decades. Thus it is realistic to assume that δ18O variations in the pre-20th-century part of the ice-core records also reflect regional temperature trends. The collection of annually dated δ18O records discussed here suggests that 20th-century warming has been spatially variable over Greenland but, on balance, temperatures have been modestly warmer than in previous centuries, particularly over the southern part of the ice sheet. With the exception of the two southern sites (D3 and Raven), the magnitude of the warming over the 20th century remains smaller than the decadal-scale variability. Eastward in the Russian Arctic, the FJL δ18O record suggests that 20th-century warming in that high-Arctic region is well outside the norm for the last half-millennium.

Ice cores from tropical and subtropical regions are limited, but new cores are being slowly added to the archive. Two low-latitude ice-core δ18O composite records have been previously published (Reference Thompson, Mosley-Thompson, Davis, Lin, Henderson and MashiottaThompson and others, 2003) and are included in Figure 1. The composite for the South American Andes is derived from the Quelccaya, Huascaran and Sajama cores, while the Tibetan Plateau composite is based on the Dunde, Guliya and Dasuopu cores. Note that these two composite δ18O histories reveal some notable regional differences over most of the last millennium. The Andean composite shows a clear Little Ice Age or recent neoglacial cool period (AD ∼1450−1880) as well as warmer conditions from AD∼1100 to 1350, correlative with the MWP. The Tibetan composite shows neither of these multi-century climate variations. The only consistent, century-scale feature between the two regional composites is the marked isotopic enrichment in 20th-century precipitation. Although high-resolution, well-dated tropical and subtropical ice-core records are scarce, they paint a consistent picture, specifically, that the 20th-century warming in the Andes and over the Tibetan Plateau is now outside the range of natural variability for the last millennium. The recent warming in lower-latitude, high-elevation regions is consistent with atmospheric temperatures (Reference Diaz, Eischeid, Duncan and BradleyDiaz and others, 2003; Reference Jones and MobergJones and Moberg, 2003), Intergovernmental Panel on Climate Change model predictions (Reference Cubasch and HoughtonCubasch and others, 2001) and the widespread recession of alpine glaciers (Reference Thompson, Mosley-Thompson, Davis, Lin, Henderson and MashiottaThompson and others, 2003; Reference OerlemansOerlemans, 2005).

The number of annually resolved and carefully dated δ18O records available from Antarctica is also limited. Three δ18O records, each representing a different part of the continent, are shown in Figure 1. The annually dated Dyer Plateau (DP) core (Reference ThompsonThompson and others, 1994) is located along the spine of the Antarctic Peninsula, a region where temperatures have increased strongly since the 1950s (Reference Marshall, Lagun and Lachlan-CopeMarshall and others, 2002; Reference King, Turner, Marshall, Connolley, Lachlan-Cope, Domack, Burnett, Leventer, Conley, Kirby and BindschadlerKing and others, 2003). Temperature records here are short, with observations extending back to the late 1940s. The decadally averaged 500 year δ18O history suggests cooler temperatures from AD∼1700 to 1950, followed by 18O enrichment that is contemporaneous with the observed warming in the region (Reference Mosley-Thompson, Thompson, Domack, Burnett and LeventerMosley-Thompson and Thompson, 2003). The 20th-century average δ18O value exceeds that for the previous 400 years, and the isotopically inferred warming of the last few decades lies well outside the range of natural variability for the last half-millennium.

On the high dry polar plateaus of East and West Antarctica, δ18O records are available from Plateau Remote and Siple Stations. At Siple Station (SS), the high annual accumulation (461 mmw.e.) results in an excellently preserved and easily interpretable annual record (Reference Mosley-Thompson, Bradley and JonesMosley-Thompson, 1992). At Plateau Remote (PR), the low accumulation (∼40mmw.e.) makes annual resolution difficult, but well-known time-stratigraphic volcanic horizons provide excellent time constraints for the last millennium (Reference Mosley-Thompson, Jones, Bradley and JouzelMosley-Thompson, 1996; Reference Cole-Dai, Mosley-Thompson, Wight and ThompsonCole-Dai and others, 2000). The δ18O history at PR shows a very modest 18O depletion (cooling) in 20th-century precipitation relative to that over the last millennium, while the 18O depletion is stronger at SS. Both suggest a recent cooling relative to the earlier part of the record. Antarctic meteorological records are short, few in number and biased toward coastal stations and stations in the Peninsula (Reference TurnerTurner and others, 2005). Two long records exist for the interior of the continent (Vostok and Amundsen-Scott South Pole Station) and both show long-term cooling trends. At South Pole (SP), average annual near-surface temperatures have declined slowly by ∼0.5˚C since record keeping began in 1958 although interannual variability is high (data provided to E.M.-T. by SP meteorological personnel in July 2005).

Thus, recent δ18O trends in Antarctic precipitation appear regionally consistent with the observed temperature trends. The recent 18O enrichment in the DP core from the spine of the Peninsula and the modest 18O depletion in the SS and PR cores on the high polar plateau are consistent with the warming and cooling trends observed in those regions, respectively (Reference Vaughan, Marshall, Connolley, King and MulvaneyVaughan and others, 2001 and references therein; Reference Thompson and SolomonThompson and Solomon, 2002; Reference TurnerTurner and others, 2005). Assuming that the earlier portions of these δ18O records also provide credible temperature proxies, the recent warming in the Peninsula region is unusual within the longer (500years) perspective available, and the slight cooling in East and West Antarctica lies within the range of past variability (1000years for PR and 500 years for SS).

Conclusions

High temporally resolved ice-core-derived δ18O records have been examined to assess whether they provide reliable proxies for examining regional 20th-century temperature trends within a longer-term perspective. The largest isotopically inferred warming is found in the Russian Arctic. Warmer 20th-century temperatures characterize the Tibetan Plateau and the South American Andes and extend into the Antarctic Peninsula. The 20th-century warming in these regions lies outside the range of natural variability as discerned from the earlier part of their respective δ18O records. Over Greenland a modest warming that is most pronounced over the southwestern part of the ice sheet has ensued since ∼1920. Modest cooling has dominated 20th-century temperatures over the high plateaus of East and West Antarctica. The unique climate histories preserved in the glaciers and ice caps from the Russian Arctic through the tropics to the Antarctic Peninsula may soon be degraded and/or lost if the current warming in those regions persists.

Acknowledgements

We thank the numerous scientists, engineers and field assistants who participated in the ice-core drilling projects that recovered these cores and conducted the laboratory analyses to extract the proxy records contained therein. Two anonymous reviewers provided valuable suggestions. These projects were funded over the years by the NSF, NASA, the US National Oceanic and Atmospheric Administration and the US National Geographic Society. This is Byrd Polar Research Center contribution No. 1325.

References

Aldaz, L. and Deutsch, S.. 1967. On a relationship between air temperature and oxygen isotope ratio of snow and firn in the South Pole region. Earth Planet Sci. Lett, 3(3), 267−274.CrossRefGoogle Scholar
Bengtsson, L, Semenov, V.A. and Johannessen, O.M.. 2004. The early twentieth-century warming in the Arctic - A possible mechanism. J. Climate, 17, 4045−4057.2.0.CO;2>CrossRefGoogle Scholar
Box, J.E. 2002. Survey of Greenland instrumental temperature records: 1873−2001. Int. J. Climatol., 22(15), 1829−1847.CrossRefGoogle Scholar
Bradley, R.S., Vuille, M., Hardy, D. and Thompson, L.G.. 2003. Low latitude ice cores record Pacific sea surface temperatures. Geophys. Res. Lett., 30(4), 1174. (10.1029/2002GL016546.)CrossRefGoogle Scholar
Cole-Dai, J., Mosley-Thompson, E. and Thompson, L.G.. 1997. Annually resolved Southern Hemisphere volcanic history from two Antarctic ice cores. J. Geophys. Res., 102(D14), 16,76116,771.CrossRefGoogle Scholar
Cole-Dai, J., Mosley-Thompson, E., Wight, S.P. and Thompson, L.G.. 2000. A 4100-year record of explosive volcanism from an East Antarctic ice core. J. Geophys. Res., 105(D19), 24,43124,441.CrossRefGoogle Scholar
Cubasch, U. and 8 others. 2001. Projections of future climate change. In Houghton, J.T. and 7 others, eds. Climate change 2001: the scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, Cambridge University Press, 525582.Google Scholar
Cuffey, K.M., Clow, G.D., Alley, R.B., Stuiver, M., Waddington, E.D. and Saltus, R.W.. 1995. Large Arctic temperature change at the Wisconsin–Holocene glacial transition. Science, 270(5235), 455458.CrossRefGoogle Scholar
Dai, J., Mosley-Thompson, E. and Thompson, L.G.. 1991. Ice core evidence for an explosive tropical volcanic eruption 6 years preceding Tambora. J. Geophys. Res., 96(D9), 17,36117,366.CrossRefGoogle Scholar
Dansgaard, W., Johnsen, S.J., Clausen, H.B. and Gundestrup, N.. 1973. Stable isotope glaciology. Medd. Grønl., 197(2), 153.Google Scholar
Diaz, H.F., Eischeid, J.K., Duncan, C. and Bradley, R.S.. 2003. Variability of freezing levels, melting season indicators, and snow cover for selected high-elevation and continental regions in the last 50 years. Climatic Change, 59(1–2), 3352.CrossRefGoogle Scholar
Grootes, P.M., Stuiver, M., Thompson, L.G. and Mosley-Thompson, E.. 1989. Oxygen isotope changes in tropical ice, Quelccaya, Peru. J. Geophys. Res., 94(D1), 11871194.CrossRefGoogle Scholar
Grootes, P.M., Stuiver, M., White, J.W.C., Johnsen, S. and Jouzel, J.. 1993. Comparison of oxygen isotope records from the GISP2 and GRIP Greenland ice cores. Nature, 366(6455), 552554.CrossRefGoogle Scholar
Hansen, J.E. and 7 others. 2001. A closer look at United States and global surface temperature change. J. Geophys. Res., 106(D20), 23,94723,963.CrossRefGoogle Scholar
Henderson, K.A. 2002. An ice core paleoclimate study of Windy Dome, Franz Josef Land (Russia): development of a recent climate history for the Barents Sea. (PhD thesis, Ohio State University.)Google Scholar
Henderson, K.A., Thompson, L.G. and Lin, P.N.. 1999. Recording of El Niño in ice core δ18O records from Nevado Huascarán, Peru. J. Geophys. Res., 104(D24), 31,05331,065.Google Scholar
Johnsen, S.J., Dansgaard, W. and White, J.W.C.. 1989. The origin of Arctic precipitation under present and glacial conditions. Tellus, 41B(4), 452468.CrossRefGoogle Scholar
Jones, P.D. and Moberg, A.. 2003. Hemispheric and large-scale surface air temperature variations: an extensive revision and an update to 2001. J. Climate, 16(2), 206223.2.0.CO;2>CrossRefGoogle Scholar
Jouzel, J. 1999. Calibrating the isotopic paleothermometer. Science, 286(5441), 910911.CrossRefGoogle Scholar
Jouzel, J. and 12 others. 1997. Validity of the temperature reconstruction from water isotopes in ice cores. J. Geophys. Res., 102(C12), 26,47126,487.CrossRefGoogle Scholar
King, J.C., Turner, J., Marshall, G.J., Connolley, W.M. and Lachlan-Cope, T.A.. 2003. Antarctic Peninsula climate variability and its causes as revealed by analysis of instrumental records. In Domack, E.W., Burnett, A., Leventer, A., Conley, P., Kirby, M. and Bindschadler, R., eds. Antarctic Peninsula climate variability: a historical and paleoenvironmental perspective. Washington, DC, American Geophysical Union, 1730. (Antarctic Research Series 79.)Google Scholar
Levitus, S., Antonov, J.I., Boyer, T.P. and Stephens, C.. 2000. Warming of the world ocean. Science, 287(5461), 22252229.CrossRefGoogle Scholar
Mann, M.E. and Jones, P.D.. 2003. Global surface temperatures over the past two millennia. Geophys. Res. Lett., 30(15), 1820. (10.1029/2003GL017814.)CrossRefGoogle Scholar
Mann, M.E., Bradley, R.S. and Hughes, M.K.. 1999. Northern Hemisphere temperatures during the past millennium: inferences, uncertainties and limitations. Geophys. Res. Lett., 26(6), 759762.CrossRefGoogle Scholar
Marshall, G.J., Lagun, V. and Lachlan-Cope, T.A.. 2002. Changes in Antarctic Peninsula tropospheric temperatures from 1956 to 1999: a synthesis of observations and reanalysis data. Int. J. Climatol., 22(3), 291310.CrossRefGoogle Scholar
McConnell, J.R. and 6 others. 2001. Annual net snow accumulation over southern Greenland from 1975 to 1998. J. Geophys. Res., 106(D24), 33,82733,838.CrossRefGoogle Scholar
Moberg, A., Sonechkin, D.M., Holmgren, K., Datsenko, N.M. and Karlén, W.. 2005. Highly variable Northern Hemisphere temperatures reconstructed from low- and high-resolution proxy data. Nature, 433(7026), 613617.CrossRefGoogle ScholarPubMed
Mosley-Thompson, E. 1992. Paleoenvironmental conditions in Antarctica since A.D. 1500: ice core evidence. In Bradley, R.S. and Jones, P.D., eds. Climate since A.D. 1500. London and New York, Routledge, 572591.Google Scholar
Mosley-Thompson, E. 1996. Holocene climate changes recorded in an East Antarctica ice core. In Jones, P.D., Bradley, R.S. and Jouzel, J., eds. Climatic variations and forcing mechanisms of the last 2000 years. Berlin, Springer-Verlag, 263279. (NATO Advanced Research Series 41.)CrossRefGoogle Scholar
Mosley-Thompson, E. and Thompson, L.G.. 2003. Ice core paleoclimate histories from the Antarctic Peninsula: Where do we go from here? In Domack, E., Burnett, A. and Leventer, A., eds. Antarctic Peninsula climate variability: a historical and paleoenvironmental perspective. Washington, DC, American Geophysical Union, 115127. (Antarctic Research Series 79.)Google Scholar
Mosley-Thompson, E., Thompson, L.G., Dai, J., Davis, M. and Lin, P.N.. 1993. Climate of the last 500 years: high resolution ice core records. Quat. Sci. Rev., 12(6), 419430.CrossRefGoogle Scholar
Mosley-Thompson, E. and 8 others. 2001. Local to regional-scale variability of annual net accumulation on the Greenland ice sheet from PARCA cores. J. Geophys. Res., 106(D24), 33,83933,851.CrossRefGoogle Scholar
Mosley-Thompson, E., Mashiotta, T.A. and Thompson, L.G.. 2003. Ice core records of late Holocene volcanism: current and future contributions from the Greenland PARCA cores. In Robock, A. and Oppenheimer, C., eds. Volcanism and the Earth’s atmosphere. Washington, DC, American Geophysical Union, 153164. (AGU Monograph 139.)CrossRefGoogle Scholar
Oerlemans, J. 2005. Extracting a climate signal from 169 glacier records. Science, 308(5722), 675677.CrossRefGoogle ScholarPubMed
Overpeck, J. and 17 others. 1997. Arctic environmental change of the last four centuries. Science, 278(5341), 12511256.CrossRefGoogle Scholar
Peel, D.A. 1992. Ice core evidence from the Antarctic Peninsula region. In Bradley, R.S. and Jones, P.D., eds. Climate since A.D. 1500. London and New York, Routledge, 549571.Google Scholar
Rogers, J.C. 1985. Atmospheric circulation changes associated with the warming over the northern north-Atlantic in the 1920s. J. Appl. Meteorol., 24(12), 13031310.2.0.CO;2>CrossRefGoogle Scholar
Rozanski, K., Araguás-Araguás, L. and Gonfiantini, R.. 1992. Relation between long-term trends of oxygen-18 isotope composition of precipitation and climate. Science, 258(5084), 981985.CrossRefGoogle ScholarPubMed
Rozanski, K., Araguás-Araguás, L. and Gonfiantini, R.. 1993. Isotopic patterns in modern global precipitation. In Swart, P.K., Lohmann, K.C., McKenzie, J.A. and Savin, S., eds. Climate change in continental isotopic records. Washington, DC, American Geophysical Union, 136. (Geophysical Monograph 78.)Google Scholar
Rozanski, K., Johnsen, S.J., Schotterer, U. and Thompson, L.G.. 1997. Reconstruction of past climates from stable isotope records of palaeo-precipitation preserved in continental archives. Journal des Sciences Hydrologiques, 42(5), 725745.CrossRefGoogle Scholar
Thomas, R.H. and Investigators, PARCA. 2001. Program for Arctic Regional Climate Assessment (PARCA): goals, key findings, and future directions. J. Geophys. Res., 106(D24), 33,69133,705.CrossRefGoogle Scholar
Thompson, D.W.J. and Solomon, S.. 2002. Interpretation of recent Southern Hemisphere climate change. Science, 296(5569), 895899.CrossRefGoogle ScholarPubMed
Thompson, L.G., Mosley-Thompson, E., Dansgaard, W. and Grootes, P.M.. 1986. The Little Ice Age as recorded in the stratigraphy of the tropical Quelccaya ice cap. Science, 234(4774), 361364.CrossRefGoogle ScholarPubMed
Thompson, L.G. and 8 others. 1989. 100,000 year climate record from Qinghai–Tibetan Plateau ice cores. Science, 296(4929), 474477.CrossRefGoogle Scholar
Thompson, L.G. and 7 others. 1994. Climate since AD 1510 on Dyer Plateau, Antarctic Peninsula: evidence for recent climate change. Ann. Glaciol., 20, 420426.CrossRefGoogle Scholar
Thompson, L.G. and 7 others. 1995. Late glacial stage and Holocene tropical ice core records from Huascarán, Peru. Science, 269(5220), 4650.CrossRefGoogle ScholarPubMed
Thompson, L.G. and 9 others. 1997. Tropical climate instability: the last glacial cycle from a Qinghai–Tibetan ice core. Science, 276(5320), 18211825.CrossRefGoogle Scholar
Thompson, L.G. and 11 others. 1998. A 25,000-year tropical climate history from Bolivian ice cores. Science, 282(5395), 18581864.CrossRefGoogle ScholarPubMed
Thompson, L.G., Mosley-Thompson, E. and Henderson, K.A.. 2000. Ice-core palaeoclimate records in tropical South America since the Last Glacial Maximum. J. Quat. Sci., 15(4), 377394.3.0.CO;2-L>CrossRefGoogle Scholar
Thompson, L.G., Mosley-Thompson, E., Davis, M.E., Lin, P.N., Henderson, K. and Mashiotta, T.A.. 2003. Tropical glacier and ice core evidence of climate change on annual to millennial timescales. Climatic Change, 59(1–2), 137155.CrossRefGoogle Scholar
Turner, J. and 8 others. 2005. Antarctic climate change during the last 50 years. Int. J. Climatol., 25, 279294.CrossRefGoogle Scholar
Vaughan, D.G., Marshall, G.J., Connolley, W.M., King, J.C. and Mulvaney, R.. 2001. Climate change: devil is in the detail. Science, 293(5536), 17771779.CrossRefGoogle ScholarPubMed
Vuille, M. and 6 others. 2003. Modeling δ18O in precipitation over the tropical Americas: 2. Simulation of the stable isotope signal in Andean ice cores. J. Geophys. Res., 108(D6), 4175. (10.1029/ 2001JD002039.)Google Scholar
Yao, T., Thompson, L.G., Mosley-Thompson, E., Zhihong, Y., Xingping, Z. and Lin, P.N.. 1996. Climatological significance of δ18O in north Tibetan ice cores. J. Geophys. Res., 101(D23), 29,53129,537.CrossRefGoogle Scholar
Figure 0

Fig. 1. Decadally averaged δ18O histories are shown from north to south (top to bottom). Horizontal bars in each record show the average δ18O values for the pre- and post-1900 time intervals. The 1000 year Northern Hemisphere reconstruction is from Mann and others (1999) and updated by Mann and Jones (2003). Superimposed is the observed near-surface temperature record (Jones and Moberg, 2003). Maps show locations for the cores.

Figure 1

Table 1. Information for all ice cores discussed in the text Timescale under revision.

Figure 2

Fig. 2. The 5 year running mean of the annual δ18O histories from five new Greenland ice cores are shown along with the 1989 Site T record collected at the Summit (GISP2) site. Vertical shaded bars highlight specific time intervals discussed in the text. Horizontal bars in each record show the average δ18O value before and after the large cooling in 1917–19 that was followed by a rapid, widespread warming over much of the Arctic.