Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-04-30T15:32:49.079Z Has data issue: false hasContentIssue false

Spatial and temporal variations of fractionation of stable isotopes in East-Antarctic snow

Published online by Cambridge University Press:  01 March 2021

Chuanjin Li*
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Jiawen Ren
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Guitao Shi
Affiliation:
Key Laboratory of Geographic Information Science (Ministry of Education), School of Geographic Sciences and Institute of Eco-Chongming, East China Normal University, Shanghai200241, China Key Laboratory for Polar Science of State Oceanic Administration, Polar Research Institute of China, Shanghai200062, China
Hongxi Pang
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing210023, China
Yetang Wang
Affiliation:
College of Geography and Environment, Shandong Normal University, Ji'nan 250358, China
Shugui Hou
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing210023, China
Zhongqin Li
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Zhiheng Du
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Minghu Ding
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China Institute of Tibetan Plateau and Polar Regions Meteorology, Chinese Academy of Meteorological Sciences, Beijing100081, China
Xiangyu Ma
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Jiao Yang
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Aihong Xie
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Puyu Wang
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Xiaoming Wang
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China
Bo Sun
Affiliation:
Key Laboratory for Polar Science of State Oceanic Administration, Polar Research Institute of China, Shanghai200062, China
Cunde Xiao
Affiliation:
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou730000, China State Key Laboratory of Earth Surface Processes and Resource Ecology, Beijing Normal University, Beijing100875, China
*
Author for correspondence: Chuanjin Li, E-mail: lichuanjin@lzb.ac.cn
Rights & Permissions [Opens in a new window]

Abstract

Stable isotope ratios (δ18O and δD) in Antarctic snow and ice are basic proxy indices of climate in ice core studies. The relation between the ratios has important indicative significance for moisture sources. In general, the fractionation characteristics of the two isotopes vary with different meteorological and topographical conditions. This paper presents the spatial and temporal distribution of meteoric water line (MWL) slopes along a traverse from the Zhongshan Station (ZSS) to Dome A in East Antarctica. It is found that the slopes decrease with the increasing distance inland from the coast and the lowest slope occurred at Dome A, where the long-range transported moisture dominates and clear sky snowing have an influence. The slopes in different layers of the snowpack showed a decreasing trend with depth and this is attributed to the fractionation during the interstitial sublimation and re-condensation processes of the water vapor. Frost flower development on the interior plateau surface can greatly alter the depth evolution of the MWL slope. The coastal snow pits also go through the post-depositional smoothing effect, but their influences are not so significant as the inland regions.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press

Introduction

Stable isotope records of snow and ice cores in Antarctica were widely used to infer past local temperature variations and moisture origin (Dansgaard, Reference Dansgaard1964; Landais and others, Reference Landais, Barkan and Luz2008; Steen-Larsen and others, Reference Steen-Larsen2011; Xiao and others, Reference Xiao2013). The isotopic composition in snow, namely the ratio between its heavy and light water molecules, is expressed as δD or δ18O relative to the Vienna Standard Mean Ocean Water (V-SMOW) reference (δD and δ18O, defined as δD = (([HD16O]/[H216O]) sample / ([HD16O]/[H216O]) V−SMOW −1) × 1000; δ18O = (([H218O]/[H216O]) sample/ ([H218O]/[H216O]) V−SMOW-1) × 1000) (Craig, Reference Craig1961; Dansgaard, Reference Dansgaard1964; Masson-Delmotte and others, Reference Masson-Delmotte2008; Steen-Larsen and others, Reference Steen-Larsen2011). More recent ice core studies in Antarctica have been dedicated to obtaining the longest possible climate records in the Polar Regions (e.g., Jouzel and others, Reference Jouzel2007), or characterizing at high-resolution structure of climate variability during glacial (Masson-Delmotte and others, Reference Masson-Delmotte2003; EPICA Community Members, 2006; Stenni and others, Reference Stenni2010) or interglacial periods (Pol and others, Reference Pol2011). Ongoing efforts are directed at documenting the regional variability of climate in various sectors of Antarctica and the related trajectories of air mass transportation and moisture origin. Triple oxygen isotopes (16O, 17O and 18O) and double hydrogen isotopes (1H and 2H) and the second-order parameters (d-excess and 17O-excess, calculations follow the equations d-excess = δD-8 × δ18O and 17O-excess = 106 (ln(δ17O/1000 + 1)-0.528 × ln(δ17O/1000 + 1)), respectively) were widely involved to study the temporal and spatial variations of water isotopes (Dansgaard, Reference Dansgaard1964; Masson-Delmotte and others, Reference Masson-Delmotte2003; Barkan and Luz, Reference Barkan and Luz2007; Hou and others, Reference Hou, Li, Xiao and Ren2007, Reference Hou, Wang and Pang2013; Landais and others, Reference Landais, Barkan and Luz2008; Xiao and others, Reference Xiao2013; Benetti and others, Reference Benetti2014; Pang and others, Reference Pang2015, Reference Pang2019). The isotopic content is strongly influenced by fractionation processes encountered by the air mass throughout its history, such as evaporation conditions, mixture of moisture sources and distillation of the air mass (Dansgaard, Reference Dansgaard1964; Masson-Delmotte and others, Reference Masson-Delmotte2003, Reference Masson-Delmotte2008; Steen-Larsen and others, Reference Steen-Larsen2015). The variations of the relationship between stable isotopes and the temperature arise from changes in ocean evaporation conditions such as sea surface temperatures (SST), relative humidity and ocean surface water isotopic composition (Bonne and others, Reference Bonne2019). When only the equilibrium isotopic fractionation occurs, a mean slope of 8.0 for δD/δ18O of the precipitation will be presented. However, during non-equilibrium phase occurs, such as ocean surface evaporation processes, a kinetic fractionation effect is added to the equilibrium fractionation (Craig and Gordon, Reference Craig, Gordon and Tongiogi1965), which causes a larger kinetic fractionation for H218O than for HDO due to their different molecular diffusivities. As a result, d-excess is strongly imprinted by the kinetic effect. Therefore, the calculation of the d-excess will provide independent isotopic information related to the initial air mass evaporation conditions at the ocean surface (mainly SST and relative humidity) (Merlivat and Jouzel, Reference Merlivat and Jouzel1979; Bonne and others, Reference Bonne2019).

The mean slope between δD and δ18O during the equilibrium fractionation (8.0) also known as the mean meteoric water line (MWL) was widely used to study the fractionation of the water isotopes in the worldwide precipitation (Dansgaard and others, Reference Dansgaard1964). However, a lower mean MWL slope (7.75) was detected on Antarctica ice sheet (Masson-Delmotte and others, Reference Masson-Delmotte2008) and showed great spatial variations (Xiao and others, Reference Xiao2013; Li and others, Reference Li2016). The evaporation, transportation, condensation and the post-depositional processes on water vapor have influences on both the evolution of the water stable isotopes and MWL slopes (both for the equilibrium fractionation and kinetic fractionation) in Antarctic snowpack (Masson-Delmotte and others, Reference Masson-Delmotte2008). The influencing factors at the deposition site mainly include the meteorological and topographical conditions (Xiao and others, Reference Xiao2013), precipitation type and intermittency (Cuffey and Steig, Reference Cuffey and Steig1998; Helsen and others, Reference Helsen2006; Laepple and others, Reference Laepple2018), redistribution of snow (Ekaykin and others, Reference Ekaykin, Lipenkov, Barkov, Petit and Masson-Delmotte2002; Ekaykin and Lipenkov, Reference Ekaykin and Lipenkov2009; Laepple and others, Reference Laepple2016; Münch and others, Reference Münch, Kipfstuhl, Freitag, Meyer and Laepple2016), cyclones and cyclone paths (Qin and others, Reference Qin, Petit, Jouzel and Stievenard1994), surface and interstitial sublimation and re-condensation processes (Ding and others, Reference Ding2010, Reference Ding2011; Steen-Larsen and others, Reference Steen-Larsen2013; Xiao and others, Reference Xiao2013; Madsen and others, Reference Madsen2019), etc. Calculations of the d-excess based on the routine MWL slope would result in great discrepancies in different sites. By now, few systematic studies on the spatial variations on MWL slopes were executed in Antarctica (Masson-Delmotte and others, Reference Masson-Delmotte2008). In order to accurately interpret the stable water isotope records from deep ice cores, it is critical to evaluate the spatial and temporal distribution of MWL slopes, which needs to be documented from modern measurements and processes studies (Xiao and others, Reference Xiao2013).

In addition, the influences of the fractionation on the initial deposition signals of the water isotopes and the alteration of climatic information were another research topic for inland Antarctic regions (Hoshina and others, Reference Hoshina2014, Reference Hoshina, Fujita, Iizuka and Motoyama2016; Casado and others, Reference Casado2016, Casado, Reference Casado2018; Ritter and others, Reference Ritter2016). During the past decades, there existed some opposing views on what drives that deposition of water isotopes were due to climate variations or post-deposition noise. Some sites seemed to be greatly influenced by the signal (e.g., Dome F, Dome C and Vostok) while other sites were not (e.g. DML and South Pole) (Steen-Larsen and others, Reference Steen-Larsen2014; Münch and others, Reference Münch, Kipfstuhl, Freitag, Meyer and Laepple2016). To make these controversial debates clear, more inland and coastal sites should be included and detailed studies on fractionation effects on temporal isotopic evolution process should be executed.

Here we present a modern process study on spatial and temporal variations of water stable isotopes (δ18O and δD) in the snow along a traverse from coast to interior plateau in eastern Antarctica, and two topics are mainly involved: (1) spatial distribution of the MWL slopes at different sites with varied environmental and geographical conditions along the traverse and (2) the temporal variations of the MWL slopes in different depth of the snow pits and ice core and the possible causes for their spatial and temporal variations.

Samples and data

Samples and analysis

Dome A is located at the highest point of Eastern Antarctica ice sheet (4093 m) and 1248 km from the nearest coast (Fig. 1) (Hou and others, Reference Hou, Li, Xiao and Ren2007; Reference Hou, Wang and Pang2013; Xiao and others, Reference Xiao2008; Reference Xiao2013). According to in-situ measurement at the Dome A station (Hou and others, Reference Hou, Li, Xiao and Ren2007; Xiao and others, Reference Xiao2008; Ma and others, Reference Ma, Bian, Xiao, Allison and Zhou2010), the mean annual temperature at Dome A is −58.5°C, the accumulation rate is 0.023 m a−1 w.e. and the mean wind velocity is 2.2 m s−1. These values are the lowest among the stations in Antarctica. During the 29th Chinese National Antarctica Research Expedition (CHINARE) inland expedition from coastal Zhongshan Station (ZSS, 69°37′31″S, 76°37′22″E) to inland Dome A (80°25′01″S, 77°06′58″E) in the austral summer in 2012/13, 13 snow pits and 115 surface samples were approximately evenly captured along the transect (Fig. 1 and Table 1) (Li and others, Reference Li2016). The depth of the snow pits varied from 2 to 3 m and the period covered changed from ~4 years at the coastal region to ~36 years at the interior regions. For the inland two snow pits (29-M and 29-L, Table 1), the mean sampling resolution was 3 cm, amounted to 0.3–0.4 year per sample depending on the density. The sampling depth for the surface samples along the traverse was 3–5 cm and the time covered varied from ~1 year in the interior regions to only the summer precipitation episodes at the coast. During the 32nd CHINARE inland expedition in 2015/16, another coastal snow pit (32-A) with larger depth (3 m) and higher sampling resolution (~4 cm per sample) was dug. During the same expedition, an ice core (CA2016-75) with the length 33.24 m were drilled (75 km from the coast) in a dry hole with a mechanical drill. The length, weight and diameter of each piece of ice core were measured just after its extraction.

Fig. 1. Surface sampling sites along the CHINARE inland traverse between Zhongshan Station and Dome A. Red dots show the snow pit locations, the blue dotting line shows the surface sample sampling sites and the green column shows the location of the coastal ice core (CA2016-75).

Table 1. Statistics of the slopes of δD and δ18O in snow samples along the ZSS-Dome A traversea

a For the detailed data and information, please refer to Fig. S1 and Table S1 to S16 in the supplementary materials.

All the samples were transported frozen to Lanzhou and all the analyzing procedures were executed at the State Key Laboratory of the Cryospheric Science (SKLCS) in Lanzhou China under 100–1000 class clean environment. Cations were analyzed on a Dionex ISC 3000 ion chromatograph (Dionex ISC 3000, Thermo Scientific, USA) using an Ion Pac CS12A column, 20 m mol L−1 MSA eluent and a cation electrolytically regenerated suppressor (CERS). The anions were analyzed on a Dionex ISC3000 ion chromatograph (Dionex ISC3000, Thermo Scientific, USA) using an Ion Pac AS11-HC column, 25 m mol L−1 KOH eluent, and an anion electrolytically regenerated suppressor (ASRS). The detection limit, defined as 3 times the standard deviation of the baseline noise, is ~1 ng g−1 for all major ions. The stable isotope (D and 18O) compositions of all samples were analyzed by a liquid-water isotope analyzer (DLT 100, Los Gatos, USA) based on off-axis integrated cavity output spectroscopy (OA-ICOS) at the SKLCS. The isotopic ratios are expressed in per mil (‰) units relative to Vienna Standard Mean Ocean Water (V-SMOW) (Xiao and others, Reference Xiao2013; Li and others, Reference Li2016). The accuracies of δD and δ18O measurements are ±0.6 and ±0.2‰, respectively.

The linear second-order parameters on water isotopes of d-excess and the logarithmic d-excess (d-ln) are both adopted in this study to show their spatial and temporal variations and the related influencing factors and the calculation of d-ln is based on the equation: d-ln = ln(1 + δD) + 0.0285 (ln(1 + δ18O))2–8.47 ln(1 + δ18O) (Uemura and others, Reference Uemura2012; Markle and others, Reference Markle2017).

Ancillary data

The coastal ice core (CA2016-75) was dated by the annual layer counting from the seasonal variation profiles of marine ions (Na+, SO42−) and black carbon (BC). The separated dating results on Na+ and BC are well consistent with the uncertainty of ±2 years for 102 years (1915–2016 A.D.). The calculated annual mean accumulation rates on the ice core is ~0.180 m a−1 w.e., met well with the previous results (Ding and others, Reference Ding2011). The two inland snow pits (29-M and 29-L) were dated according to the volcanic deposit signals (Li and others, Reference Li2014) and the accumulation rate data. Because of the relative calm wind condition, the influence of the drifting snow was negligible at Dome A. The calculated accumulation rates for the whole duration (37 years for 29-M and 36 years for 29-L) are consistent with the measurements by the automatic weather station (AWS) (Ma and others, Reference Ma, Bian, Xiao, Allison and Zhou2010).

Results and Discussions

Spatial distribution of slopes along the transect

δD and δ18O both show decreasing trends while d-excess increases with the distance from the coast along the ZSS to Dome A traverse (Xiao and others, Reference Xiao2013; Pang and others, Reference Pang2015, Reference Pang2019; Li and others, Reference Li2016). The mean slope of δD with the surface temperature was 6.4 ± 0.2‰ per °C, like the average for East Antarctica (Masson-Delmotte and others, Reference Masson-Delmotte2008; Xiao and others, Reference Xiao2013). The mean MWL slope in surface snow samples in this study (7.78 ± 0.04, R 2 = 0.99) is little larger compared with the values measured before (7.5 ± 0.1, R 2 = 0.99) along the same traverse (Xiao and others, Reference Xiao2013). We speculate the different sampling depth is responsible for the discrepancy. Our sampling depth (3–5 cm) is less compared with Xiao's study (5 cm) (Xiao and others, Reference Xiao2013).

According to the spatial variations of the accumulation rates and the impurities in snow (Ma and others, Reference Ma, Bian, Xiao, Allison and Zhou2010; Ding and others, Reference Ding2011; Xiao and others, Reference Xiao2013; Li and others, Reference Li2014, Reference Li2016; Shi and others, Reference Shi2019; Ma and others, Reference Ma2020), the ZSS-Dome A traverse can be divided the into three different sections, the coastal section (0–400 km), the intermediate section (400–900 km) and the interior section (900–1248 km). The mean slopes between δD and δ18O in the coastal, intermediate and interior sections are 8.18 ± 0.15 (R 2 = 0.99, n = 30), 8.03 ± 0.09 (R 2 = 0.99, n = 50) and 7.79 ± 0.10 (R 2 = 0.99, n = 35), respectively. The general decreasing trend of the MWL slopes in three sections may be caused by the increasing distance from the source regions and intensified fractionation effects (Masson-Delmotte and others, Reference Masson-Delmotte2008; Xiao and others, Reference Xiao2013; Pang and others, Reference Pang2015, Reference Pang2019).

To get a detailed spatial distribution pattern of the slopes along the traverse, the MWL slopes in the 13 snow pits were calculated (Fig. 2). In general, the slopes have a decreasing trend with the increasing distance inland. The snow pits located near the coast always show relatively higher slopes (Table 1) and the highest slope occurs in snow pit 29-C (8.61 ± 0.26, R 2 = 0.98, n = 20) and this value was significantly larger than the global mean MWL (8.0) (Dansgaard and others, Reference Dansgaard1964) and also larger than the mean slope of the Antarctic snow (7.75 ± 0.02, R 2 = 0.998) (Masson-Delmotte and others, Reference Masson-Delmotte2008) (Fig. 2). Coastal regions of Antarctic ice sheet usually receive moisture from adjacent offshore water while the interior regions get sources from the distant open sea (Reijmer and others, Reference Reijmer, Van den Broeke and Scheele2002; Masson-Delmotte and others, Reference Masson-Delmotte2008; Sodemann and Stohl, Reference Sodemann and Stohl2009). The higher MWL slopes at the coastal section may indicate less fractionation during the evaporation and the transportation processes (Hou and others, Reference Hou, Wang and Pang2013; Xiao and others, Reference Xiao2013). In addition, the high accumulation at coastal section can cause a quick bury of the deposited snow, which is of benefit to stop (or reduce) the surface sublimation and condensation processes (Ding and others, Reference Ding2010). The slopes in the intermediate section (400–900 km) generally show a ‘V’ spatial pattern and the slopes decrease at the initial part potentially implies a decrease of water vapor sources from the adjacent coastal ocean. However, after ~700 km, the slope shows an increasing trend till the ~900 km inland. We speculate that the intensified wind speed, special surface topography and the post-depositional effects may have influences to some extent (Xiao and others, Reference Xiao2013; Münch and others, Reference Münch, Kipfstuhl, Freitag, Meyer and Laepple2016). The most interior snow pit at Dome A (29-M) has the lowest slope (7.24 ± 0.11, R 2 = 0.98, n = 100), much lower than the global mean MWL and the Antarctic snow samples. Another interior snow pit 29-L also shows a low slope value (7.84 ± 0.10, R 2 = 0.98, n = 100). The distance between the two sites is 148 km and they share similar environmental conditions (Ren and others, Reference Ren and Qin1995, Reference Ren2010).

Fig. 2. Spatial distribution of the MWL (meteoric water line) between δD and δ18O in 13 snow pits (29-M to 29-A, black square) and surface snow (red square) along the traverse from Zhongshan Station to Dome A together with the d-excess (bottom) and d-ln (middle). Different sections of the traverse are divided by the vertical dashed lines and the mean values of the parameters in different sections are shown in squares. The horizontal dotted lines show the global mean meteoritic water line (8.0, Dansgaard, Reference Dansgaard1964) and the mean slope in Antarctic snow (7.75, Masson-Delmotte and others, Reference Masson-Delmotte2008).

In the previous studies, most interior snow samples usually show the lowest water isotope ratios (Masson-Delmotte and others, Reference Masson-Delmotte2008; Xiao and others, Reference Xiao2013; Li and others, Reference Li2016). The water isotope ratios at the snow pit 29-M vary between −394.80 and −485.96‰ with the mean −450.93 ± 16.25‰ for δD and between −63.25 and −50.67 ‰ with the mean −58.46 ± 2.22‰ for δ18O. These values are the lowest in Antarctica up to date. Two potential reasons can be put forward to be mainly accountable for the lowest slope at the Dome A region. The first is the moisture source. The moisture arriving in the interior region is mainly from the low to mid-latitudes open oceans and has experienced long-distance transportation in the upper troposphere or lower stratosphere, and thus more intensive fractionation effects may occur during the long pathway (Xiao and others, Reference Xiao2013). The second is that at Dome A, low temperature together with the special precipitation form and accumulation have an intensifying effect on the surface evaporation and re-condensation of the water vapor at this highest location of Eastern Antarctica (Fujita and others, Reference Fujita and Abe2006; Ding and others, Reference Ding2015). Clear sky precipitation (diamond dust) is particular in interior regions on the Antarctic ice sheet and can occupy a large fraction of the total accumulation (Hou and others, Reference Hou, Li, Xiao and Ren2007). During the days with weak wind and higher temperature at the snow surface compared to the air, the vapor pressure of the air mass close to the snow surface can reach or even exceed saturation, resulting in frost flower growth at the snow surface (Hou and others, Reference Hou, Li, Xiao and Ren2007; Xiao and others, Reference Xiao2013). According to Hou and others (Reference Hou, Li, Xiao and Ren2007), the frost flower can comprise ~7% of the whole accumulation at Dome A. The diamond dust and frost flowers both present a large specific surface area (300–590 cm2 g−1) and hence are favorable to intensify the surface sublimation and re-condensation processes of water vapor. During the sublimation of ice grains and re-condensation of water vapor onto ice grains under extremely low temperatures, the isotopic fractionations will occur in the porous snow layers (Town and others, Reference Town, Warren, Walden and Waddington2008; Wang and others, Reference Wang2012; Hoshina and others, Reference Hoshina, Fujita, Iizuka and Motoyama2016; Casado and others, Reference Casado2016, Casado, Reference Casado2018; Madsen and others, Reference Madsen2019).

Comparison between the slopes and the mean d-excess values in the 13 snow pits shows that they have a generally opposite spatial variability (R = −0.71, P < 0.01, N = 13), the slopes decrease with the distance inland while the d-excess increase simultaneously (Fig. 2). They also show some different variation patterns at the coastal and the intermediate sections. The d-excess value in Antarctic snow is mainly controlled by the relative humidity, sea surface temperature and wind velocities at the moisture source regions (Hou and others, Reference Hou, Wang and Pang2013; Bonne and others, Reference Bonne2019) and also influenced by the condensation conditions at the deposition site (Markle and others, Reference Markle2017). The slope between δD and δ18O is mainly controlled by the climatic and meteorological conditions of snowfall (air temperature, relative humidity, wind velocity, etc.) (Masson-Delmotte and others, Reference Masson-Delmotte2008). Therefore, the difference between the slope and the d-excess may indicate that the local climatic and meteorological conditions had significant influences on the fractionation processes of the water isotopes. One point should be pointed out that the relation between the slopes and the source region conditions is mainly based on distillation modeling and future observations are in great demand to testify that (Masson-Delmotte and others, Reference Masson-Delmotte2008).

The lognormal statistical on deuterium excess (d-ln) can effectively reduce the influences from the non-source factors and is regarded as an optimal index to study the climatic variations at the moisture source regions (Uemura and others, Reference Uemura2012; Markle and others, Reference Markle2017). We made some comparisons among the slopes, d-excess and the d-ln values along the traverse (Fig. 2). The d-ln clearly indicates a consistently increasing trend inland, implying a constant shift of different source regions and this is consistent with previous results (Masson-Delmotte and others, Reference Masson-Delmotte2008; Xiao and others, Reference Xiao2013). The difference between the d-ln and the slopes may be mainly caused by the influences from the local environment, especially on the intermediate section.

Temporal variations of water isotopes

The mean water isotopes (δD and δ18O) values in the 13 snow pits are systematically lower than the surface snow in the same areas, implying depleted isotope signals exist in the snow pits (Xiao and others, Reference Xiao2013; Li and others, Reference Li2016). We further investigate the temporal variations of the fractionation between δD and δ18O in the snow pits and ice core. The temporal variations of the slopes in the two inland snow pits (29-M and 29-L) are firstly investigated. The two pits are vertically divided into three parts (1 m resolution for 29-M and 0.83 m for 29-L). It shows a distinct increase (K = 0.26, P = 0.15) of the slope with depth in the 29-M (6.99 ± 0.30, 7.14 ± 0.14 and 7.51 ± 0.11 for the top, middle and bottom sections, respectively), but 29-L shows an opposite variation pattern (K = −0.29, P = 0.23) (8.01 ± 0.18, 7.91 ± 0.15 and 7.44 ± 0.18 for the top, middle and bottom sections, respectively). This difference between the two snow pits should reflect the different fractionation processes at the two sites. During expeditions inland by the CHINARE, surface frost flower has been only seen at the 1150–1248 km on the traverse and no frost flower layer had been detected at 29-L. The fractionations of the water isotopes will sustain in the snow layer during the snow metamorphism and cause an increase of the isotopic composition in the upper layer. Isotope exchange and diffusion within the porous matrix of the snowpack affect snow isotopic composition (Langway, Reference Langway1970; Johnsen, Reference Johnsen1977; Ebner and others, Reference Ebner, Steen-Larsen, Stenni, Schneebeli and Steinfeld2017; Casado and others, Reference Casado2018) and this procedure will be sustained until the firn transformation into ice. The different fractionation procedures between HDO and H218O in the snowpack may be the causes for the increase of the slopes at the upper layers at 29-L (Masson-Delmotte and others, Reference Masson-Delmotte2003). The inverse variations of the slopes at 29-M should be caused by the intensified fractionation effects on the frost flower with much lower slopes. The mixing effects within the surface snow will decrease the mean slope in the upper snow layers and the mixing effect decrease with depth can cause the slope increase. Because no surface frost flower is developed, no specific fractionation mechanism exists at the surface of 29-L, so the slope vertical variation shows an opposite pattern to 29-M,i.e. the slope decreases with depth at 29-L.

The vertical temperature gradients in different layers of the snow pits may be another important factor for the fractionation. The temperature gradient can cause the convection of the water vapor in the firn layers. Especially during the autumn and winter seasons, when the surface is significantly cooled down while the deeper layer still kept a relatively higher temperature, the water vapor sublimates at the lower depth can move upward and re-condenses at the shallower layers accompanied by the kinetic fractionation (Masson-Delmotte and others, Reference Masson-Delmotte2008). The formation of depth hoar in the snow layers also has a great influence on the fractionation of water isotopes, especially at the sites in the inland with low accumulations (Satake and Kawada, Reference Satake and Kawada1997; Neumann and Waddington, Reference Neumann and Waddington2004; Neumann and others, Reference Neumann, Waddington, Steig and Grootes2005). The widespread depth hoar in East Antarctic snow layers during the winter is a good proof for the interstitial sublimation and re-condensation processes (Qin and others, Reference Qin and Ren1992, Reference Qin, Ren, Wang, Petit and Jouzel1993; Ekaykin and others, Reference Ekaykin and Lipenkov2009). However, no depth hoar layers were observed in our snow pits since all the snow pits were dug in the summer season.

Previous studies on the variations of water isotopes with depth in the inland Antarctic snow pits (e.g. Vostok, Dome F, Dome C) showed a mean 20 cm cycle, and these intervals were ascribed to accumulation adjusting, wind erosion and post-depositional redistribution, which can significantly alter the initial seasonal depositional signals and climatic information (Hoshina and others, Reference Hoshina2014, Reference Hoshina, Fujita, Iizuka and Motoyama2016; Casado and others, Reference Casado2016; Laepple and others, Reference Laepple2016; Casado, Reference Casado2018). We check the deposition signals of δ18O in the snow pits along the traverse and find a similar variation pattern (Fig. 3). The two interior snow pits display the cycle periodicity varied between 16 and 51 cm with the mean value 32.38 ± 11.36 cm at 29-M and between 10 and 34 cm with the mean value of 23.89 ± 8.95 cm at 29-L. Moreover, both the snow pits show enlarged cycle periodicity with the depth accompanied by the decreasing variation amplitudes, implying that the smoothing effects of the isotopes is intensified with depth. The smoothing effects on the water stable isotopes with the depth can also be seen at the coastal and intermediate sections (Fig. 3). 29-G shows an extreme smoothing trend in the bottom part, suggesting that besides the interstitial sublimation and re-condensation effects, the intensified wind erosion and mixing effects on the snow have great influences on the variations of the isotopes at this site. Sodium (Na+) is usually regarded as a good indicator of the seasonal variations of the marine incursion, but its seasonal signals seem to be smoothed at the inland snow layers with similar depth intervals to the water isotopes (Fig. 3). Opposite variation patterns between the Na+ and the water isotopes always exist generally in our inland snow pits and this phenomenon seems widespread among the eastern Antarctica (Hoshina and others, Reference Hoshina, Fujita, Iizuka and Motoyama2016). However, at the coastal snow pits, such as 32-A, Na+ has clear seasonal variations but the water isotopes do not.

Fig. 3. Temporal variations of the δ18O (black) and Na+ (red) in different snow pits along the traverse between ZZS and Dome A.

Because of the smoothing influences on the water isotopes in the inland snow pits, we select the ice core (CA2016-75) at the coastal region to study the annual variations of the water isotopes and the potential influencing factors (Fig. 4). The CA2016-75 ice core drilling site has a relatively high accumulation rate (~0.18 m a−1 w.e.) and thickness of annual layers vary between 30 and 40 cm, so the annual mean depth usually has seven–ten samples and is sufficient to study the seasonal variations. Based on the dating results, the annual mean water isotopes and the second-order proxies (d-excess and d-ln) are calculated (Fig. 4). The seasonal variations of the δ18O and δD can be detected for the top 10 m depth (since ~1990s). The smoothing effect is significantly below 10 m depth, eliminating the annual and seasonal signals. From this phenomenon, we can find the post-depositional sublimation and re-condensation of the water vapor is also significant in the coastal region, even for the sites with the high accumulation rates.

Fig. 4. Temporal variations of the δD (red), δ18O (blue), d-excess (pink), d-ln (green) in the CA2016-75 ice core. The vertical dotted lines show the dating results of the ice core and the thick solid lines show the 10-point smoothing results (red for δD, blue for δ18O, pink for d-excess and green for d-ln, respectively).

The MWL slope for the whole ice core (8.54 ± 0.11) is similar with the coastal 29-A and 32-A snow pits and a decrease of the slopes with depth is also detected (8.56 ± 0.08 for the upper 10 m and 8.37 ± 0.09 for the lower 23.24 m, Table 1). The temporal variability of water isotopes at CA2016-75 ice core is different from the Law Dome ice core (DE08-2) which shows significant seasonal cycles till 150 m depth (Masson-Delmotte and others, Reference Masson-Delmotte2003). Higher accumulation for DE08-2 ice core site (between 1.00 and 1.50 m a−1 w.e.) and lower wind speed may be the dominant factors for the good preservation of the seasonal cycles of water isotopes in the ice core (Masson-Delmotte and others, Reference Masson-Delmotte2003).

Variations of water isotope ratios (δD, δ18O, Fig. 4) show a general flat pattern between 1910s and mid-1980s, after that significant increases during 1980s–90s (K = 2.89, P < 0.0001, N = 92 for δD and K = 0.36, P < 0.0001, N = 92 for δ18O, respectively) and the ratios stay at a relatively higher level till the sampling time. The d-excess and d-ln show similar temporal variations for the whole duration (mid-1910s–2015A.D.) (Fig. 4). They both show relatively flat variations between 1910s and 1990s, and significant decreasing trends are followed till the mid-2000s and then a quick increasing trend till the sampling time. The reasons for the variations and variability of the isotopes and the second-order parameters (d-excess and d-ln) are beyond the scope of this paper and more detailed discussions will be included in the following studies.

Conclusions

Spatial and temporal distributions of fractionation slopes between δD and δ18O in the surface snow, snow pits and ice core samples along the traverse from ZSS to Dome A were calculated. Three spatial sections were divided along the traverse, the low slope existed at the interior section and high values at the coastal region, the intermediate section showed large variations implying that complex mechanisms were involved in the fractionation processes and more efforts are needed in the future. The lowest slope existed at Dome A and was much lower than the mean MWL, it was speculated that the long-range transportation of moisture to this interior plateau and the particular precipitation and accumulation styles at Dome A are the dominant influencing factors. Inland snow pits showed significant smoothing effects on the water isotopes as ever been found in other inland sites (Vostok, Dome C, Dome F, etc.) in eastern Antarctica and the original deposition signals of the water isotopes has been significantly altered. Moreover, not only the inland snow pits, the coastal snow pits in our study were also found for the post-depositional smoothing effects. The persistent sublimation, migration, and re-condensation of the water vapor in the snow (firn) layers caused by the deviations of the temperature in different depths and the wind sweeping effects may be the main influencing factors. The slopes of the isotopes in different depth changed significantly. Variations of the slopes with depth at Dome A snow pit (29-M) showed opposite distribution pattern to 29-L and the existence of the frost flower or not may be mainly accountable for the discrepancy. Coastal snow pits and ice cores showed different smoothing effects on the snow impurities which presented significant seasonal variations and multi-year oscillation. More efforts on other water isotopes tracers (e.g. δ17O and tritium) in the near future are in demand to study the sources of the moisture along the traverse between ZSS and Dome A and their fractionation procedures under different environmental conditions and also for their relationships with the climatic variations.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/jog.2021.5

Acknowledgement

We thank all the members who participated in the 2012/13 and 2016/17 CHINARE field campaigns for sample collection. Special thanks should also be given to analyzing personnel on the water isotopes in the State Key Laboratory of the Cryospheric Sciences. Dr Xiaoqing Cui, Ms. Xiaoxiang Wang, Yan Liu and Tingting Su and Yuman Zhu from the cold and arid region environmental and engineering research institute, Chinese Academy of Sciences who completed parts of the analysis. This work is financially supported by the Innovative Research Group, the National Natural Science Foundation of China (No. 41671063, 41830647, 41625012, and 41961144028), the State Key Laboratory of Cryospheric Sciences (supporting fund No. SKLCS-ZZ-2018-01) and the Youth Innovation Promotion Association, Chinese Academy of Sciences.

References

Barkan, E and Luz, B (2007) Diffusivity fractionations of H216O/H217O and H216O/H218O in air and their implications for isotope hydrology. Rapid Communication Mass Spectrom 21, 29993005. doi: 10.1002/rcm.3180CrossRefGoogle Scholar
Benetti, M and 6 others (2014) Deuterium excess in marine water vapor: dependency on relative humidity and surface wind speed during evaporation. Journal of Geophysical Research: Atmospheres 119(2), 584593.Google Scholar
Bonne, JL and 9 others (2019) Resolving the controls of water vapour isotopes in the Atlantic sector. Nature Communications 10, 1632.CrossRefGoogle ScholarPubMed
Casado, M (2018) Antarctic Stable Isotopes, Reference Module in Earth Systems and Environmental Sciences. Amsterdam: Elsevier. doi: 10.1016/B978-0-12-409548-9.11655-0.Google Scholar
Casado, M and 12 others (2016) Continuous measurements of isotopic composition of water vapour on the East Antarctic Plateau. Atmospheric Chemistry and Physics 16, 85218538.CrossRefGoogle Scholar
Craig, H (1961) Isotopic variation in meteoric water. Science 133(3465), 17021703.CrossRefGoogle Scholar
Craig, H and Gordon, LI (1965) Deuterium and oxygen 18 variations in the ocean and the marine atmosphere. In Tongiogi, E (ed.), Stable Isotopes in Oceanographic Studies and Paleotemperatures. Pisa, Spoleto, Italy: V. Lishi e F., pp. 9130.Google Scholar
Cuffey, KM and Steig, EJ (1998) Isotopic diffusion in polar firn: implications for interpretation of seasonal climate parameters in ice core records, with emphasis on central Greenland. Journal of Glaciology 44, 273284.CrossRefGoogle Scholar
Dansgaard, W (1964) Stable isotopes in precipitation. Tellus 16, 436468.CrossRefGoogle Scholar
Ding, M and 5 others (2010) Distribution of δ18O in surface snow along a transect from Zhongshan Station to Dome A, East Antarctica. Chinese Science Bulletin 55(24), 27092714.CrossRefGoogle Scholar
Ding, M and 6 others (2011) Spatial variability of surface mass balance along a traverse route from Zhongshan station to Dome A, Antarctica. Journal of Glaciology 57(204), 658666.Google Scholar
Ding, M and 8 others (2015) Surface mass balance and its climate significance from the coast to Dome A, East Antarctica. Science in China: Earth Science 58, 17871797. doi: 10.1007/s11430-015-5083-9CrossRefGoogle Scholar
Ebner, PP, Steen-Larsen, HC, Stenni, B, Schneebeli, M and Steinfeld, A (2017) Experimental observation of transient δ18O interaction between snow and advective airflow under various temperature gradient conditions. The Cryosphere 11, 17331743.CrossRefGoogle Scholar
Ekaykin, AA and Lipenkov, VY (2009) Formation of the ice core isotopic composition. Physics of ice core records, Low Temperature Science 68(Supplement), 299314.Google Scholar
Ekaykin, AA, Lipenkov, VY, Barkov, NI, Petit, JR and Masson-Delmotte, V (2002) Spatial and temporal variability in isotope composition of recent snow in the vicinity of Vostok Station, Antarctica: implications for ice-core record interpretation. Annals of Glaciology 35(1), 181186.CrossRefGoogle Scholar
EPICA Community Members (2006) One-to-one coupling of glacial climate variability in Greenland and Antarctica. Nature 444(7116), 195198.CrossRefGoogle Scholar
Fujita, K and Abe, O (2006) Stable isotopes in daily precipitation at Dome Fuji, East Antarctica. Geophysical Research Letters 33, L18503. doi: 10.1029/2006GL026936CrossRefGoogle Scholar
Helsen, MM and 6 others (2006) Modeling the isotopic composition of Antarctic snow using backward trajectories: simulation of snow pit records. Journal of Geophysical Research 111, D15109. doi: 10.1029/2005JD006524CrossRefGoogle Scholar
Hoshina, Y and 8 others (2014) Effect of accumulation rate on water stable isotopes of near-surface snow in inland Antarctica, Journal of Geophysical Research - Atmosphere 119, 274283.CrossRefGoogle Scholar
Hoshina, Y, Fujita, K, Iizuka, Y and Motoyama, H (2016) Inconsistent relationships between major ions and water stable isotopes in Antarctic snow under different accumulation environments. Polar Sciences 10, 110.CrossRefGoogle Scholar
Hou, S, Li, Y, Xiao, C and Ren, J (2007) Recent accumulation rate at Dome A, Antarctica. Chinese Science Bulletin 52, 428431.CrossRefGoogle Scholar
Hou, S, Wang, Y and Pang, H (2013) Climatology of stable isotopes in Antarctic snow and ice: Current status and prospects. Chinese Science Bulletin 58, 10951106.CrossRefGoogle Scholar
Johnsen, S (1977) Stable isotope homogenization of polar firn and ice, in Proc. of Symp. on Isotopes and Impurities in Snow and Ice, I.U.G.G. XVI, General Assembly, Grenoble, pp. 210219, IAHS-AISH Publ. 118.Google Scholar
Jouzel, J and 31 others (2007) Orbital and millennial Antarctic climate variability over the past 800000 years. Science 317, 793796.CrossRefGoogle ScholarPubMed
Laepple, T and 5 others (2016) Layering of surface snow and firn at Kohnen Station, Antarctica: noise or seasonal signal? Journal of Geophysical Research-Earth Surface 121, 18491860.CrossRefGoogle Scholar
Laepple, T and 5 others (2018) On the similarity and apparent cycles of isotopic variations in East Antarctic snow pits. The Cryosphere 12, 169187.CrossRefGoogle Scholar
Landais, A, Barkan, E and Luz, B (2008) The record of δ18O and δ17Oexcess in ice from Vostok, Antarctica, during the last 150 000 years. Geophysical Research Letters 35, L02709. doi: 10.1029/ 2007GL032096Google Scholar
Langway, CC (1970) Stratigraphic analysis of a deep ice core from Greenland, Special Paper 125. Boulder, CO: The Geological Society of America.CrossRefGoogle Scholar
Li, C and 8 others (2014) Temporal variations in marine chemical concentrations in coastal areas of eastern Antarctica and associated climatic causes. Quaternary International 352, 1625.CrossRefGoogle Scholar
Li, C and 5 others (2016) Spatial and temporal variability of marine-origin matter along a transect from Zhongshan Station to Dome A, Eastern Antarctica. Journal of Environmental Sciences 46, 190202.CrossRefGoogle Scholar
Ma, X and 11 others (2020) Spatial and temporal variations of refractory black carbon along the transect from Zhongshan Station to Dome A, eastern Antarctica. Atmospheric Environment 242, 117816.CrossRefGoogle Scholar
Ma, Y, Bian, L, Xiao, C, Allison, I and Zhou, X (2010) Near surface climate of the traverse route from Zhongshan Station to Dome A, East Antarctica. Antarctic Science 22(04), 443459.CrossRefGoogle Scholar
Madsen, MV and 9 others (2019) Evidence of isotopic fractionation during vapor exchange between the atmosphere and the snow surface in Greenland. Journal of Geophysical Research-Atmospheres 124, 29322945. doi: 10.1029/2018JD029619CrossRefGoogle ScholarPubMed
Markle, BR and 9 others (2017) Global atmospheric teleconnections during Dansgaard-Oeschger events. Nature Geoscience 10(1), 3640. doi: 10.1038/ngeo2848CrossRefGoogle Scholar
Masson-Delmotte, V and 6 others (2003) Recent southern Indian ocean climate variability inferred from a Law Dome ice core: new insights for the interpretation of coastal Antarctic isotopic records. Climate Dynamics 21, 153166.CrossRefGoogle Scholar
Masson-Delmotte, V and 35 others (2008) A review of Antarctic surface snow isotopic composition: observations, atmospheric circulation, and isotopic modeling, Journal of Climate 21, 33593387.CrossRefGoogle Scholar
Merlivat, L and Jouzel, J (1979) Global climatic interpretation of the deuterium-oxygen 18 relationship for precipitation. Journal of Geophysical Research 84, 50295033. doi: 10.1029/JC084iC08p05029CrossRefGoogle Scholar
Münch, T, Kipfstuhl, S, Freitag, J, Meyer, H and Laepple, T (2016) Regional climate signal vs. Local noise: a two-dimensional view of water isotopes in Antarctic firn at Kohnen Station, Dronning Maud Land. Climate of the Past 12, 15651581.CrossRefGoogle Scholar
Neumann, TA and Waddington, ED (2004) Effects of firn ventilation on isotopic exchange. Journal of Glaciology 50, 183194. doi: 10.3189/172756504781830150CrossRefGoogle Scholar
Neumann, TA, Waddington, ED, Steig, EJ and Grootes, PM (2005) Non-climate influences on stable isotopes at Taylor Mouth, Antarctica. Journal of Glaciology 51, 248258.CrossRefGoogle Scholar
Pang, H and 12 others (2015) Spatial distribution of 17O-excess in surface snow along a traverse from Zhongshan station to Dome A, East Antarctica. Earth and Planetary Science Letters 44, 126133.CrossRefGoogle Scholar
Pang, H and 18 others (2019) Influence of summer sublimation on δD, δ18O, and δ17O in precipitation, East Antarctica, and implications for climate reconstruction from ice cores. Journal of Geophysical Research: Atmospheres 124, 73397358.Google Scholar
Pol, K and 11 others (2011) Links between MIS 11 millennial to sub-millennial climate variability and long-term trends as revealed by new high-resolution EPICA Dome C deuterium data – a comparison with the Holocene. Climate of the Past 7(2), 437450.CrossRefGoogle Scholar
Qin, D, Petit, JR, Jouzel, J and Stievenard, M (1994) Distribution of stable isotopes in surface snow along the route of the 1990 International Trans-Antarctica Expedition. Journal of Glaciology 40(134), 107118.Google Scholar
Qin, D and Ren, J (1992) A study on snow profiles and surface characteristics along 6000km Trans-Antarctic route. Science in China 35(3), 366374.Google Scholar
Qin, D, Ren, J, Wang, W, Petit, JR and Jouzel, J (1993) Distribution of δD in 25-cm surface snow along Trans-Antarctic Route (II) – the “1990 International Trans-Antarctic Expedition” glaciological research. Science in China 36(3), 375384.Google Scholar
Reijmer, CH, Van den Broeke, MR and Scheele, MP (2002) Air parcel trajectories and snowfall related to five deep drilling locations in Antarctica based on the ERA-15 dataset. Journal of Climate 15, 19571968.2.0.CO;2>CrossRefGoogle Scholar
Ren, J and 6 others (2010) A 2680 year volcanic record from the DT-401 East Antarctic ice core. Journal of Geophysical Research 115(D11), D11301. doi: 10.1029/2009JD012892CrossRefGoogle Scholar
Ren, J and Qin, D (1995) Distribution of deuterium excess in surface snow of the Antarctic ice sheet. Chinese Scientific Bulletin 40(19), 16291633.Google Scholar
Ritter, F and 9 others (2016) Isotopic exchange on the diurnal scale between near-surface snow and lower atmospheric water vapor at Kohnen station, East Antarctica. The Cryosphere 10, 16471663. doi: 10.5194/tc-10-1647-2016CrossRefGoogle Scholar
Satake, H and Kawada, K (1997) The quantitative evaluation of sublimation and the estimation of original hydrogen and oxygen isotope ratios of a firn core at East Queen Maud Land, Antarctica. Bulletin of Glacier Research 15, 9397.Google Scholar
Shi, G and 9 others (2019) Organic tracers from biomass burning in snow from the coast to the ice sheet summit of East Antarctica. Atmospheric Environment 201, 231241.CrossRefGoogle Scholar
Sodemann, H and Stohl, A (2009) A symmetries in the moisture origin of Antarctic precipitation. Geophysical Research Letters 36, L22803. doi: 10.1029/2009GL040242CrossRefGoogle Scholar
Steen-Larsen, HC and 23 others (2011) Understanding the climatic signal in the water stable isotope records from the NEEM shallow firn/ice cores in northwest Greenland. Journal of Geophysical Research 116, D06108. doi: 10.1029/ 2010JD014311CrossRefGoogle Scholar
Steen-Larsen, HC and 23 others (2013) Continuous monitoring of summer surface water vapor isotopic composition above the Greenland Ice Sheet. Atmospheric Chemistry and Physics 13, 48154828.CrossRefGoogle Scholar
Steen-Larsen, HC and 18 others (2014) What controls the isotopic composition of Greenland surface snow? Climate of the Past 10, 377392.CrossRefGoogle Scholar
Steen-Larsen, HC and 9 others (2015) Moisture sources and synoptic to seasonal variability of North Atlantic water vapor isotopic composition. Journal of Geophysical Research: Atmosphere 120, 57575774.Google Scholar
Stenni, B and 14 others (2010) The deuterium excess records of EPICA Dome C and Dronning Maud Land ice cores (East Antarctica). Quaternary Science Reviews 29, 146159.CrossRefGoogle Scholar
Town, MS, Warren, SG, Walden, VP and Waddington, ED (2008) Effect of atmospheric water vapor on modification of stable isotopes in near-surface snow on ice sheets. Journal of Geophysical Research 113, D24303. doi: 10.1029/2008JD009852CrossRefGoogle Scholar
Uemura, R and 5 others (2012) Ranges of moisture-source temperature estimated from Antarctic ice cores stable isotope records over glacial–interglacial cycles. Climate of the Past 8, 11091125.CrossRefGoogle Scholar
Wang, Y and 5 others (2012) Snow accumulation and its moisture origin over Dome Argus, Antarctica. Climate Dynamics 40, 731742. doi: 10.1007/s00382-012-1398-9.CrossRefGoogle Scholar
Xiao, C and 5 others (2008) Preliminary evidence indicating Dome A (Antarctica) satisfying pre conditions for drilling the oldest ice core. Chinese Science Bulletin 53, 102106.CrossRefGoogle Scholar
Xiao, C and 10 others (2013) Stable isotopes in surface snow along a traverse route from Zhongshan station to Dome A, East Antarctica. Climate Dynamics 41, 24272438. doi: 10.1007/s00382-012-1580-0CrossRefGoogle Scholar
Figure 0

Fig. 1. Surface sampling sites along the CHINARE inland traverse between Zhongshan Station and Dome A. Red dots show the snow pit locations, the blue dotting line shows the surface sample sampling sites and the green column shows the location of the coastal ice core (CA2016-75).

Figure 1

Table 1. Statistics of the slopes of δD and δ18O in snow samples along the ZSS-Dome A traversea

Figure 2

Fig. 2. Spatial distribution of the MWL (meteoric water line) between δD and δ18O in 13 snow pits (29-M to 29-A, black square) and surface snow (red square) along the traverse from Zhongshan Station to Dome A together with the d-excess (bottom) and d-ln (middle). Different sections of the traverse are divided by the vertical dashed lines and the mean values of the parameters in different sections are shown in squares. The horizontal dotted lines show the global mean meteoritic water line (8.0, Dansgaard, 1964) and the mean slope in Antarctic snow (7.75, Masson-Delmotte and others, 2008).

Figure 3

Fig. 3. Temporal variations of the δ18O (black) and Na+ (red) in different snow pits along the traverse between ZZS and Dome A.

Figure 4

Fig. 4. Temporal variations of the δD (red), δ18O (blue), d-excess (pink), d-ln (green) in the CA2016-75 ice core. The vertical dotted lines show the dating results of the ice core and the thick solid lines show the 10-point smoothing results (red for δD, blue for δ18O, pink for d-excess and green for d-ln, respectively).

Supplementary material: PDF

Li et al. supplementary material

Li et al. supplementary material

Download Li et al. supplementary material(PDF)
PDF 1 MB