Hostname: page-component-5c6d5d7d68-ckgrl Total loading time: 0 Render date: 2024-08-06T22:36:31.912Z Has data issue: false hasContentIssue false

Na2SO4 and MgSO4 salts during the Holocene period derived by high-resolution depth analysis of a Dome Fuji ice core

Published online by Cambridge University Press:  08 September 2017

Yoshinori Iizuka
Affiliation:
Institute of Low Temperature Science, Hokkaido University, Sapporo 060–0819, Japan. E-mail: iizuka@lowtem.hokudai.ac.jp
Takeo Hondoh
Affiliation:
Institute of Low Temperature Science, Hokkaido University, Sapporo 060–0819, Japan. E-mail: iizuka@lowtem.hokudai.ac.jp
Yoshiyuki Fujii
Affiliation:
National Institute of Polar Research, Kaga, Itabashi-ku, Tokyo 173–8515, Japan
Rights & Permissions [Opens in a new window]

Abstract

We analyzed the profiles of ionic chemical species in three 500 mm sections of an ice core from Dome Fuji, Antarctica, dated 3.0, 8.9 and 13.3 kyr BP (before present), and compared the profiles to those in the surface snow. The 3.0 and 8.9 kyr sections are from the Holocene and the 13.3 kyr section slightly predates the Holocene. The analyses were done on 2 mm thick slices within each section. At each depth, the primary ionic species were Na+, H+, Cl and SO42 A The SO42, Na+ and Mg2+ levels varied with depth in each section over distances ranging from several millimeters to several centimeters. Also, the correlation coefficients between Na+ and SO4 and between Mg2+ and SO42 for each depth were 0.90 or greater, in contrast to the value of 0.59 or less in the surface snow (defined here as 0–3.4 m from the surface). These results suggest that almost all Na+ and Mg2+ in the Holocene ice exists as Na2SO4 and MgSO4 salts, and the formation of these salts occurs not only in the atmosphere during transport, but also in the firn layer.

Type
Research Article
Copyright
Copyright © International Glaciological Society 2006

1. Introduction

Ice-core research is widely conducted to reconstruct regional and global paleoclimates. Ice cores drilled from the inland regions of polar ice sheets are particularly important for reconstructing climates for the past several hundred thousand years, a time period that includes several glacial cycles. In the interior of Antarctica, long-term paleoclimate studies have been carried out for ice cores from Byrd (e.g. Reference Johnsen, Dansgaard, Clausen and LangwayJohnsen and others, 1972), Vostok (e.g. Reference Legrand and DelmasLegrand and Delmas, 1988; Reference PetitPetit and others, 1999), Dome Fuji (e.g. Reference WatanabeWatanabe and others, 2003a, Reference Watanabe, Jouzel, Johnsen, Parrenin and Shojiandb), EPICA DML (e.g. Reference Oerter, Graf, Meyer and WilhelmsOerter and others, 2004) and EPICA Dome C (e.g. EPICA community, 2004).

However, analyses of short-term climatic fluctuations in the ice cores from inland Antarctica have not been done. This is mainly because, owing to the low annual snowfall rates, the annual layers are too thin to provide sufficient material for chemical analyses. Reworking by wind mixes the surface snow, and occasionally a region in inland Antarctica has no net accumulation throughout a year. Furthermore, in this region post-depositional processes are expected to have a larger influence on the variations of chemical concentrations and on the oxygen/hydrogen isotope ratio than is observed in other regions of Antarctica. The post-depositional processes include:

  1. 1. sublimation, condensation and volatilization in surface snow (Reference Sturm and BensonSturm and Benson, 1997; Reference Kreutz, Mayewski, Whitlow and TwicklerKreutz and others, 1998) and

  2. 2. molecular diffusion in firn or ice (Reference Johnsen, Clausen, Cuffey, Hoffmann, Schwander, Creyts and HondohJohnsen and others, 2000; Reference Livingston and GeorgeLivingston and George, 2002; Reference Barnes, Wolff, Mader, Udisti, Castellano and Röthlisberger.Barnes and others, 2003).

The sublimation, condensation and volatilization processes have been studied in ice cores from Greenland and Antarctica (e.g. Reference Dibb, Talbot and BerginDibb and others, 1994; Reference Bergin, Davidson, Dibb, Jaffrezo, Kuhns and Pandis.Bergin and others, 1995; Reference Dibb and JaffrezoDibb and Jaffrezo, 1997; Reference RöthlisbergerRöthlisberger and others, 2002; Reference IizukaIizuka and others, 2004). These processes have been studied in detail for volatile species including:

  1. 1. NO3 (Reference Legrand and DelmasLegrand and Delmas, 1986; Reference Dibb, Talbot, Munger, D.J. Jacob and FanDibb and others, 1998; Reference Honrath, Peterson, S. Guo, Dibb, P.B. Shepson and CampbellHonrath and others, 1999; Reference Munger, Jacob, Fan, Colman and DibbMunger and others, 1999; Reference Wagnon, Delmas and LegrandWagnon and others, 1999),

  2. 2. H2O2 (Reference Neftel, Jacob and KlockowNeftel and others, 1984; Reference Sigg and NeftelSigg and Neftel, 1991; Reference Hutterli, McConnell, Stewart, Jacobi and BalesHutterli and others, 2001, Reference Hutterli, McConnell, Bales and Stewart2003; Reference RöthlisbergerRöthlisberger and others, 2003),

  3. 3. HCHO (Reference Hutterli, Röthlisberger and BalesHutterli and others, 1999, Reference Hutterli, Bales, McConnell and Stewart2002, Reference Hutterli, McConnell, Bales and Stewart2003; Reference RöthlisbergerRöthlisberger and others, 2003) and

  4. 4. Cl (Reference RöthlisbergerRöthlisberger and others, 2003).

In general, these studies found that concentrations of volatile substances in snow layers become lower than those in precipitation owing to their escape from the snow to the atmosphere, particularly in regions with low snow accumulation rates such as inland Antarctica.

The molecular diffusion process of ion species in deep ice has also been studied (e.g. Reference Barnes, Wolff, Mader, Udisti, Castellano and Röthlisberger.Barnes and others, 2003). Diffusion smooths out the initial variations of chemical concentrations, particularly in regions with low snow accumulation rates such as inland Antarctica. Regarding the molecular diffusion process, Reference Barnes, Wolff, Mader, Udisti, Castellano and Röthlisberger.Barnes and others (2003) showed that only events of magnitudes that significantly exceed the background fluctuation in the SO4 2- and Cl levels are detectable after 11 000 years in the Dome C ice core, and post-depositional movement is negligible for Na+.

From 1995 to 1996, the Japanese Antarctic Research Expedition (JARE) conducted a deep ice-coring project at Dome Fuji (77º19′ S, 39º40′ E;3810ma.s.l.). The project team recovered ice down to 2503 m, which covers the past 340 kyr (Reference FujiiFujii and others, 2002; Reference Watanabe, Jouzel, Johnsen, Parrenin and ShojiandWatanabe and others, 2003b). The present annual accumulation rate at Dome Fuji is about 30mmw.e. (Reference Kameda, Azuma, Furukawa, Ageta and TakahashiKameda and others, 1997) and the annual layer thickness of the Dome Fuji ice core is estimated to be 25–35 mm for the Holocene period (Reference WatanabeWatanabe and others, 2003a). In the Dome Fuji region, the snow layers down to 3.4 m depth have summer minima in both the non-sea-salt SO4 2- and Na+ concentrations (Reference IizukaIizuka and others, 2004). If the post-depositional processes have little effect, short-term fluctuations should occur in the Dome Fuji ice core. In this paper, we present high-resolution depth profiles of soluble ions from the Holocene period and discuss chemical reactions that probably occurred during the post-deposition processes.

2. Analytical Procedure

We selected 500 mm long ice sections at three well-separated depths: 119m (119.291–119.770 m), 298 m (298.403–298.878m) and 415 m (415.352–415.818 m). The ages of these ice sections are 3.0, 8.9 and 13.3kyr, respectively, and their annual accumulation rates are estimated to be 27, 29 and 27 mm of ice, respectively, based on δ18O values that are considered to show good correlation with annual accumulation rates (Reference WatanabeWatanabe and others, 2003a). The section from 13.3 kyrBP corresponds to the beginning of the Antarctic Cold Reversal and is thus not strictly from the Holocene. Nevertheless, all sections are more recent than the Last Glacial Maximum and are in or near the Holocene. For convenience, we refer to all three sections as ‘Holocene ice’ even though the oldest slightly predates the Holocene.

For each ice section, we sliced 3 mm of ice from the outside of the section to decontaminate the surface. This was done on a clean bench in a cold room using a clean ceramic knife. The cleaned ice section was sliced every 2 mm of depth, and each sliced fraction was sealed in a clean polyethylene bag (Whirl Pak®). The estimated thickness error of each sample was <0.5 mm. Thus we obtained 230–240 samples from each 500 mm section.

After melting an ice sample in the pack, it was filtered through a pore size of 0.45 μm. We used an ion chromatograph (Dionex 500) to analyze the liquid for the eight major soluble ions: Cl, NO3 , SO4 2–, CH3SO3 (MSA), Na+, K+, Mg2+ and Ca2+. The concentration of soluble ions was measured with an estimated error of <5% (Reference Igarashi, Kanamori and WatanabeIgarashi and others, 1998). As a reference, we analyzed an ice section frozen from ultra-pure water using the same method, including the same type of bag, and found no peaks on the chromatograph. This result indicated that any contamination from our method would probably not affect our results.

3. Results

The data, listed in Table 1, show that the primary anions are Cl, with average concentrations ranging from 0.92 to 1.90 μmol L–1, and SO4 2–, ranging from 0.95 to 1.19 μmol L–1. The primary cation is Na+ with average concentrations ranging from 1.07 to 2.08 μmol L–1, and the calculated concentrations of H+ range from 0.38 to 3.22 μmol L–1. The concentrations of CH3SO3 , NO3 , K+, Mg2+ and Ca2+ are significantly lower than those of the primary ions.

Table 1. Average concentrations, standard deviations and median absolute deviations of eight ion species (μmol L–1) at depths of 0–3.4 m (surface snow), 119, 298 and 415 m (Holocene ice). The acidity, H+ in the final column, was estimated using the following equation: [H+] = [Cl] + [NO3I + [SO421 + [CH3SO3I – ([Na+] + [K+] + [Mg2+] + [Ca2+]) (μeqL–1)

The depth profiles of Cl are similar in all three Holocene ice sections, as are the profiles for NO3 (Fig. 1). Except for a few spikes in the Cl profiles, the Cl and NO3 profiles are relatively uniform and smooth. In contrast, the SO4 2– and Na+ profiles have fluctuation periods ranging from several millimeters to several centimeters (Fig. 2). Also, the 298 m section has slightly smoother SO4 2– and Na+ profiles than those of the 119 and 415 m ice sections.

Fig. 1. High-resolution depth profiles of Cl and NO3 concentrations at depths of 0–3.4 m (surface snow), 119, 298 and 415 m. The 119 and 298 m sections are from the Holocene, the 415 m section slightly predates the Holocene but, for simplicity, all three will be called ‘Holocene ice’. Analyses were made every 20 mm of depth for surface snow and every 2 mm of depth for the Holocene ice.

Fig. 2. High-resolution depth profiles of SO4 2– and Na+ concentrations in the surface snow and Holocene ice. Analyses were made in the same manner as described for Figure 1.

The Cl and Na+ profiles are consistent with the recent study by Reference Barnes, Wolff, Mader, Udisti, Castellano and Röthlisberger.Barnes and others (2003), who found that events in the Cl profile disappear by diffusion but are retained in the Na+ profile. However, in contrast to the Barnes and others study, our SO4 2– profiles retain millimeter-order fluctuations in deep ice. This contrast is discussed in section 5.

4. Smoothing of Cl and NO3 Profiles by Post-Depositional Processes

We now compare the high-resolution depth profiles of soluble ions in the Holocene ice with those in the surface snow. The surface snow down to 3.4 m was analyzed at a similar resolution to the Holocene ice (Reference IizukaIizuka and others, 2004). The SO4 2– and Na+ concentrations of the surface snow have roughly the same average as those of the Holocene ice (Table 1) with similar fluctuation periods (Fig. 2). This suggests very little smoothing of SO4 2– or Na+ by post-depositional processes.

Unlike the SO4 2– and Na+ trends, the Cl and NO3 profiles in the Holocene ice are much smoother than those in the surface snow. The average concentrations of NO3 in the Holocene ice sections are much lower than those in surface snow. This decrease in concentration indicates that the smoothing below 0.6 m depth is due to volatilization of NO3 from the snow to the atmosphere (Reference Legrand and DelmasLegrand and Delmas, 1986; Reference Dibb, Talbot, Munger, D.J. Jacob and FanDibb and others, 1998; Reference Honrath, Peterson, S. Guo, Dibb, P.B. Shepson and CampbellHonrath and others, 1999; Reference Munger, Jacob, Fan, Colman and DibbMunger and others, 1999; Reference Wagnon, Delmas and LegrandWagnon and others, 1999; Reference RöthlisbergerRöthlisberger and others, 2002). Also, the snow accumulation rate of the Dome Fuji region is lower than 40mmw.e. a–1, which means that Cl loss by volatilization can occur (Rothlisber-ger and others, 2003). If we assume that the average Cl/Na+ ratio in the surface snow (1.57 mol L–1 (mol L–1)–1) is the initially deposited ratio at Dome Fuji, and that this is constant throughout the Holocene, then the Cl loss in the Holocene ice sections is estimated to be 27.9%, 23.0% and 71.8% of its initial concentration for the 119, 298 and 415 m deep sections, respectively.

The residual Cl in the Holocene ice (i.e. 72.1%, 77.0% and 28.2% of the initial Cl in the 119, 298 and 415 m ice) shows smoother profiles than those observed in the surface snow (Fig. 1). The median absolute deviations of the Cl profiles are 0.16, 0.21 and 0.21 μmol L–1 for the 119, 298 and 415 m ice sections, respectively (Table 1), values that are significantly lower than the surface snow value of 0. 31 μmol L–1. We argue that this smoothing is a result of significant diffusion of Cl. (This is in addition to the postdepositional loss by volatilization.) Moreover, the Cl profiles are relatively smooth by the time the ice reaches 119m depth, and apparently little additional smoothing has occurred below 119m (Fig. 1). As 119m is close to where the firn transforms to ice, the smoothing process of Cl is probably due to its diffusion in firn.

The Cl profiles also contain a few spikes. Most Cl spikes occur at the same depth as Na+ spikes (Fig. 2), which suggests that some of the Na+ and Cl did not diffuse in the Holocene ice but instead formed NaCl salt by anion–cation trapping (Reference Livingston and GeorgeLivingston and George, 2002). Alternatively, sea-salt NaCl may not have transformed to HCl and Na2SO4. Although these spikes are dramatic changes to the profile, their occurrence is rare. Overall, we think that the binding of Na+ with Cl is insignificant compared to the binding of Na+ with SO4 2–, as discussed in detail below.

5. Post-Depositional NA2SO4·10H2O and MGSO4·12H2O Salt Formation

In the surface snow the correlation coefficients between anions and cations are low, except for the high value (r 2 = 0. 89. between Na+ and Mg2+ (Table 2). Moreover, the ratio of Mg2+ to Na+ in the surface snow is equal to that of sea salt, suggesting that most Na+ and Mg2+ comes from sea salt. The Cl is also derived from NaCl and MgCl2 in sea salts; however, this is not reflected in the correlations in Table 1. The reason for this lack of correlation is that the depth profiles of Na+ and Cl in the surface snow may be disturbed by various processes including:

  1. 1. chemical reactions of NaCl and H2SO4 during transport through the atmosphere (Reference Legrand and DelmasLegrand and Delmas, 1988),

  2. 2. ion fractionation on the snow surface by sublimation and condensation (lizuka and others, 2004) and

  3. 3. volatilization of HCl from the surface snow (Reference RöthlisbergerRöthlisberger and others, 2003).

Table 2. Correlation coefficient matrix for the eight measured ion species in the surface snow and in the Holocene ice. Correlation coefficients exceeding 0.9 are in bold

We suggest that the same processes apply to the depth profile of Mg2+ (with MgCl2 taking the place of NaCl). Due to these processes, the surface snow probably has various combinations of salts (NaCl, MgCl2, Na2SO4 and MgSO4) and acids (HCl and H2SO4), with the result that it has low correlation coefficients between cations (Na+ and Mg2+) and anions (Cl and SO4 2–).

However, in the Holocene ice the correlation coefficients between Na+, Mg2+ and SO4 2_ are greater than 0.9 (bold entries in Table 2). Hence, some redistribution processes must have caused these ion pairs to become closely correlated in the deep ice. Moreover, the profile for SO4 2_ is nearly equal to that for the sum of Na+ and Mg2+ in molar equivalents in the 119 and 415 m ice sections (Fig. 3). In the 298m section these two profiles have the same peak and trough positions, but the amount of SO4 2_ generally exceeds the sum of Na+ and Mg2+. These results suggest that almost all Na+ and Mg2+ ions in the Holocene ice coexist with SO4 2¯ In the 298 m ice, 73.2% of the SO4 2_ is estimated to have existed as a compound with Na+ or Mg2+.

Fig. 3. High-resolution profiles of SO4 2–, Mg2+ and the sum of Mg2+ and Na+ in the surface snow and Holocene ice. The SO4 2¯ and Mg2+ profiles are solid black lines and the sum of the Mg2+ and Na+ profiles is a dotted line. These concentrations are in equivalent (μeq L–1).

According to the thermodynamic phase diagram for Na2SO4 and MgSO4 in Reference Usdowski and DietzelUsdowski and Dietzel (1998), the salts Na2SO4·10H2O and MgSO4·12H2O should exist at the temperature and average Na+, Mg2+ and SO4 2_ concentrations found in the Holocene ice. The existence of these salts in our samples is further supported by the findings of Reference Ohno, Igarashi and HondohOhno and others (2005), in which most of the soluble impurities in the Dome Fuji ice core consisted of salt inclusions within ice grains. Moreover, they found that the inclusions were primarily Na2SO4·10H2O and MgSO4·12H2O in the Holocene ice. Taken together, these results suggest that almost all the Na+, Mg2+ and SO4 2_ existed as salts of Na2SO4·10H2O and MgSO4·12H2O. The formation of the post-depositional salts is mostly completed within the firn because the high correlation coefficients between Na+, Mg2+ and SO4 2_ occurred in the 119 m ice. These sulfate salts have eutectic points of–1.56 and –3.67°C for Na2SO4·10H2O and MgSO4·12H2O, respectively and thus would exist as a solid in firn and ice. This solidity means that Na+ and Mg2+ would have low mobilities in the Holocene ice.

We argue that the excess amount of SO4 in the 298 m section is due to the relatively low cation concentrations in that section (Table 1) and this excess SO4 exists with H+ in the form of liquid H2SO4. The shape of the SO4 profile is nearly the same as that of the sum of the Na+ and Mg2+ profiles. When we subtract the two curves, the excess SO4 is found to have a smooth profile, much like that of Cl. The smooth profile suggests that the excess SO4 was fully diffused, as would occur if this SO4 combined with H+ to form liquid H2SO4, which readily diffuses in firn or ice. Following this smoothing process our findings for the SO42– agree with those of Reference Barnes, Wolff, Mader, Udisti, Castellano and Röthlisberger.Barnes and others (2003).

We interpret our findings to mean that the salts of Na2SO4·10H2O and MgSO4·12H2O can form within the firn when liquid-phase SO4 diffuses to relatively immobile Na+ and Mg2+, reacts and then becomes relatively immobile. It is also known that these salts form in the atmosphere (Reference Legrand and DelmasLegrand and Delmas, 1988), and thus some salts probably originated in the firn and some in the atmosphere; but whatever the source, the salts should be relatively immobile and have a more rapidly fluctuating profile. The key points here are that the SO4 ions that form salts of Na2SO4·10H2O and MgSO4·12H2O become essentially immobile, whereas the SO4 2– ions that form an acid remain mobile and thus should have a smooth profile.

To summarize, our findings can be explained by the following processes:

  1. 1. liquid-phase SO4 diffuses within the firn (Reference Barnes, Wolff, Mader, Udisti, Castellano and Röthlisberger.Barnes and others, 2003),

  2. 2. SO42– becomes immobile after reacting with Na+ or Mg2+,

  3. 3. Na+, Mg2+ and SO4 remain as Na2SO4·10H2O and MgSO4·12H2O salts in ice (Reference Ohno, Igarashi and HondohOhno and others, 2005) and

  4. 4. liquid-phase SO42– is present in the ice if the Na+ or Mg2+ concentrations are insufficient to react with all of the SO4 2–.

Our findings also suggest that the low mobility of Na+ and Mg2+ in deep ice should be useful for reconstructing the past Holocene environment with millimeter resolution, which corresponds to several months in the Holocene (Reference Watanabe, Jouzel, Johnsen, Parrenin and ShojiandWatanabe and others, 2003b). There will be some inherent uncertainties in the procedure; for example, the levels of Na2SO4·10H2O in the ice include sources from chemical reactions in the atmosphere and thus do not solely reflect the past climate (Reference Legrand and DelmasLegrand and Delmas, 1988). In addition, some years have no net snow accumulation. Nevertheless, there is a need for additional climate analysis methods for the Holocene, and the levels of Na+ and Mg2+ in deep ice show promise as one such method.

6. Conclusions

We have analyzed the ion profiles in ice-core sections from 3.0, 8.9 and 13.3 kyr BP, and compared the profiles to the ion profiles in the top 3.4m of the core. We found that the correlation coefficients among Na+, Mg2+ and SO4 2– in the Holocene ice exceed 0.9 (Table 2), whereas the surface snow has low correlation coefficients between Na+ and SO4 2– and also between Mg2+ and SO4 2–. The analyses indicates the presence of the salts of Na2SO4·10H2O and MgSO4·12H2O, and the amount of these salts relative to the other ions suggests some of the salt forms within the firn through a process in which liquid-phase SO4 2– diffuses towards relatively immobile Na+ and Mg2+ and then reacts. Of the total SO4 2– concentration, the component that is probably bound in the salts of Na2SO4·10H2O and MgSO4·12H2O has millimeter-order fluctuations in the depth profiles, whereas the remaining SO4 2–, which we argue was part of the acid of H2SO4, has smooth profiles. Na+ and Mg2+ are expected to have a low mobility in the Holocene ice since they form the solid salts Na2SO4·10H2O and MgSO4·12H2O. The low mobility of Na+ and Mg2+ should be useful for reconstructing the past Holocene environment with millimeter resolution.

Acknowledgements

We thank the Dome Fuji drilling team and all the participants in the JARE Dome Fuji traverse. The paper was significantly improved as a result of comments by R. Röthlisberger, an anonymous reviewer and the Scientific Editor, D. Peel, to whom we are greatly indebted. This study was supported by Grant-in-Aids for Creative Scientific Research (grant No. 14GS0202), for Scientific Research (B) (grant No. 16710012) and for Scientific Research (S) (grant No. 15101001).

References

Barnes, P.R.F., Wolff, E.W., Mader, Udisti, R. E. Castellano, H.M. and Röthlisberger., R. 2003. Evolution of chemical peak shapes in the Dome C, Antarctica, ice core. J. Geophys. Res., 108(D3), 4126. (10.1029/2002JD002538.) Google Scholar
Bergin, M.H., Davidson, C.I., Dibb, Jaffrezo, J.L. H.D. Kuhns, J.E. and Pandis., S. N. 1995. A simple model to estimate atmospheric concentrations of aerosol chemical species based on snow core chemistry at Summit, Greenland. Geophys. Res. Lett., 22(24), 35173520. Google Scholar
Dibb, J.E. and Jaffrezo, J.L.. 1997. Air–snow exchange investigations at Summit, Greenland: an overview. J. Geophys. Res., 102(C12), 26,795-26,807. Google Scholar
Dibb, J.E., Talbot, R.W. and Bergin, M.H.. 1994. Soluble acid species in air and snow at Summit, Greenland. Geophys. Res. Lett., 21(15), 16271630. Google Scholar
Dibb, J.E., Talbot, R.W., Munger, D.J. Jacob, J.W. and Fan, S.M.. 1998. Air–snow exchange of HNO3 and NOy at Summit, Greenland. J. Geophys. Res., 103(D3), 34753486. Google Scholar
EPICA community. 2004. Eight glacial cycles from an Antarctic ice core. Nature, 429(6992), 623628. Google Scholar
Fujii, Y. and 25 others. 2002. Deep ice core drilling to 2503 m depth at Dome Fuji, Antarctica. Nat. Inst. Polar Res. Mem. (Spec. Issue) 56, 103116. Google Scholar
Honrath, R.E., Peterson, S. Guo, M.C., Dibb, P.B. Shepson, J.E. and Campbell, B.. 1999. Evidence of NOx production within or upon ice particles in the Greenland snowpack. Geophys. Res. Lett., 26(6), 695698. CrossRefGoogle Scholar
Hutterli, M.A., Röthlisberger, R. and Bales, R.C.. 1999. Atmosphere-to-snow-to-firn transfer studies of HCHO at Summit, Greenland. Geophys. Res. Lett., 26(12), 16911694. Google Scholar
Hutterli, M.A., McConnell, J.R.R.W. Stewart, , Jacobi, H.W. and Bales, R.C.. 2001. Impact of temperature-driven cycling of hydrogen peroxide (H2O2) between air and snow on the planetary boundary layer. J. Geophys. Res., 106(D14), 15,39515,404. Google Scholar
Hutterli, M.A., Bales, R.C., McConnell, J.R. and Stewart, R.W.. 2002. HCHO in Antarctic snow: preservation in ice cores and air–snow exchange. Geophys. Res. Lett., 29(8), 1235. (10.1029/2001GL014256.) Google Scholar
Hutterli, M.A., McConnell, J.R., Bales, R.C. and Stewart, R.W.. 2003. Sensitivity of hydrogen peroxide (H2O2) and formaldehyde (HCHO) preservation in snow to changing environmental conditions: implications for ice core records. J. Geophys. Res., 108(D1), 4023. (10.1029/2002JD002528.) Google Scholar
Igarashi, M., Kanamori, N. and Watanabe, O.. 1998. [Analytical method for small amount of polar snow and ice samples by ion chromatography.]. Antarct. Rec., 42(1), 6480. [In Japanese with English summary.] Google Scholar
Iizuka, Y. and 6 others. 2004. SO4 2– minimum in summer snow layer at Dome Fuji, Antarctica, and the probable mechanism. J. Geophys. Res., 109(D4), D04307. (10.1029/2003JD004138.) Google Scholar
Johnsen, S.J., Dansgaard, W., Clausen, H.B. and Langway, C.C. Jr. 1972. Oxygen isotope profiles through the Antarctic and Greenland ice sheets. Nature, 235(5339), 429434. CrossRefGoogle Scholar
Johnsen, S.J., Clausen, H.B., Cuffey, K.M., Hoffmann, G., Schwander, J. and Creyts, T.. 2000. Diffusion of stable isotopes in polar firn and ice: the isotope effect in firn diffusion. In Hondoh, T., ed. Physics of ice core records. Sapporo, Hokkaido University Press, 121140. Google Scholar
Kameda, T., Azuma, N., Furukawa, T., Ageta, Y. and Takahashi, S.. 1997. Surface mass balance, sublimation and snow temperatures at Dome Fuji Station, Antarctica, in 1995. In Proceedings of the NIPR Symposium on Polar Meteorology and Glaciology. Tokyo, Japan, National Institute of Polar Research, 2434. Google Scholar
Kreutz, K.J., Mayewski, P.A., Whitlow, S.I. and Twickler, M.S.. 1998. Limited migration of soluble ionic species in a Siple Dome, Antarctica, ice core. Ann. Glaciol., 27, 371377. Google Scholar
Legrand, M. and Delmas, R.J.. 1986. Relative contributions of tropospheric and stratospheric sources to nitrate in Antarctic snow. Tellus, 38B(3-4), 236249. Google Scholar
Legrand, M.R. and Delmas, R.J.. 1988. Formation of HCl in the Antarctic atmosphere. J. Geophys. Res., 93(D6), 71537168. Google Scholar
Livingston, F.E. and George, S.M.. 2002. Effect of sodium on HCl hydrate diffusion in ice: evidence for anion–cation trapping. J. Phys. Chem., 106A, 51145119. Google Scholar
Munger, J.W., Jacob, D.J., Fan, S.M., Colman, A.S. and Dibb, J.E.. 1999. Concentrations and snow–atmosphere fluxes of reactive nitrogen at Summit, Greenland. J. Geophys. Res., 104(D11), 13,72113,734. Google Scholar
Neftel, A., Jacob, P. and Klockow, D.. 1984. Measurements of hydrogen peroxide in polar ice samples. Nature, 311, 4345. CrossRefGoogle Scholar
Oerter, H., Graf, W., Meyer, H. and Wilhelms, F.. 2004. The EPICA ice core Dronning Maud Land: first results from stable isotope measurements. Ann. Glaciol., 39, 307312. Google Scholar
Ohno, H., Igarashi, A. and Hondoh, T.. 2005. Salt inclusions in polar ice core, location and chemical form of water-soluble impurities. Earth Planet Sci. Lett., 232(1-2), 171178. Google Scholar
Petit, J.R. and 18 others. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399(6735), 429436. Google Scholar
Röthlisberger, R. and 10 others. 2002. Nitrate in Greenland and Antarctic ice cores: a detailed description of post-depositional processes. Ann. Glaciol., 35, 209216. Google Scholar
Röthlisberger, R. and 8 others . 2003. Limited dechlorination of sea salt aerosols during the last glacial period: evidence from the EPICA Dome C ice core. J. Geophys. Res., 108(D16), 4256. (10.1029/2003JD003604.) Google Scholar
Sigg, A. and Neftel, A.. 1991. Evidence for a 50% increase in H2O2 over the past 200 years from a Greenland ice core. Nature, 351(6327), 557559. Google Scholar
Sturm, M. and Benson, C.S.. 1997. Vapor transport, grain growth and depth-hoar development in the subarctic snow. J. Glaciol., 43(143), 4259. Google Scholar
Usdowski, E. and Dietzel, M.. 1998. Atlas and data of solid-solution equilibria of marine evaporites, Berlin, Heidelberg, New York, Springer-Verlag. CrossRefGoogle Scholar
Wagnon, P., Delmas, R.J. and Legrand, M.. 1999. Loss of volatile acid species from upper firn layers at Vostok, Antarctica. J. Geophys. Res., 104(D3), 34233431. Google Scholar
Watanabe, O. and 6 others. 2003a. Dating of the Dome Fuji, Antarctica deep ice core. Nat. Inst. Polar Res. Mem. (Spec. Issue) 57, 2537. Google Scholar
Watanabe, O.,J. Jouzel, S., Johnsen, F. Parrenin, H. Shojiand, N.Yoshida. 2003b. Homogeneous climate variability across East Antarctica over the past three glacial cycles. Nature, 422(6931), 509512. Google Scholar
Figure 0

Table 1. Average concentrations, standard deviations and median absolute deviations of eight ion species (μmol L–1) at depths of 0–3.4 m (surface snow), 119, 298 and 415 m (Holocene ice). The acidity, H+ in the final column, was estimated using the following equation: [H+] = [Cl] + [NO3I + [SO421 + [CH3SO3I – ([Na+] + [K+] + [Mg2+] + [Ca2+]) (μeqL–1)

Figure 1

Fig. 1. High-resolution depth profiles of Cl and NO3 concentrations at depths of 0–3.4 m (surface snow), 119, 298 and 415 m. The 119 and 298 m sections are from the Holocene, the 415 m section slightly predates the Holocene but, for simplicity, all three will be called ‘Holocene ice’. Analyses were made every 20 mm of depth for surface snow and every 2 mm of depth for the Holocene ice.

Figure 2

Fig. 2. High-resolution depth profiles of SO42– and Na+ concentrations in the surface snow and Holocene ice. Analyses were made in the same manner as described for Figure 1.

Figure 3

Table 2. Correlation coefficient matrix for the eight measured ion species in the surface snow and in the Holocene ice. Correlation coefficients exceeding 0.9 are in bold

Figure 4

Fig. 3. High-resolution profiles of SO42–, Mg2+ and the sum of Mg2+ and Na+ in the surface snow and Holocene ice. The SO42¯ and Mg2+ profiles are solid black lines and the sum of the Mg2+ and Na+ profiles is a dotted line. These concentrations are in equivalent (μeq L–1).