• Keine Ergebnisse gefunden

Abstract Melt ‐ InducedFractionationofMajorIonsandTraceElementsinanAlpineSnowpack

N/A
N/A
Protected

Academic year: 2022

Aktie "Abstract Melt ‐ InducedFractionationofMajorIonsandTraceElementsinanAlpineSnowpack"

Copied!
11
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Sven E. Avak1,2,3 , Jürg C. Trachsel4,5, Jacinta Edebeli1,5, Sabina Brütsch1,

Thorsten Bartels‐Rausch1 , Martin Schneebeli4 , Margit Schwikowski1,2,3 , and Anja Eichler1,3

1Laboratory of Environmental Chemistry, Paul Scherrer Institute, Villigen, Switzerland,2Department of Chemistry and Biochemistry, University of Bern, Bern, Switzerland,3Oeschger Centre for Climate Change Research, University of Bern, Bern, Switzerland,4WSLInstitute for Snow and Avalanche Research SLF, Davos Dorf, Switzerland,5Department of Environmental Systems Science, ETH Zurich, Zürich, Switzerland

Abstract

Understanding the impact of melting on the preservation of atmospheric compounds in high‐Alpine snow and glacier ice is crucial for future reconstruction of past atmospheric conditions.

However, detailed studies investigating melt‐related changes of such proxy information are rare. Here we present a series offive snow pit profiles of 6 major ions and 34 trace elements at Weissfluhjoch,

Switzerland, collected between January and June 2017. Atmospheric composition was preserved during the cold season, while melting toward the summer resulted in preferential loss of certain species from the snowpack or enrichment at the base of the snowpack. Increasing mobilization of major ions with meltwater (NH4+< Cl~ Na+< NO3~ Ca2+~ SO42−) can be related to their stronger enrichment at ice crystal surfaces during snow metamorphism. Results for trace elements show that less abundant elements such as Ce, Eu, La, Mo, Nd, Pb, Pr, Sb, Sc, Sm, U, and W were best preserved and may still serve as tracers to reconstruct past natural and anthropogenic atmospheric emissions from melt‐affected snow pit and ice core records. The obtained elution behavior matches thefindings from another high‐Alpine site (upper Grenzgletscher) for major ions and the large majority of investigated trace elements. Both studies indicate that water solubility and location at the microscopic scale are likely to determine the relocation behavior with meltwater and also suggest that the observed species‐dependent preservation from melting snow and ice is representative for the Alpine region, reflecting Central European atmospheric aerosol composition.

1. Introduction

Major ions (MIs) and trace elements (TEs) serve as important proxies for reconstructing past environmental conditions from high‐Alpine snow pits (e.g., Gabrieli et al., 2011; Greilinger et al., 2016; Hiltbrunner et al., 2005; Kuhn et al., 1998; Kutuzov et al., 2013; Nickus et al., 1997) and ice cores (e.g., Döscher et al., 1996;

Eichler et al., 2000; Preunkert et al., 2000; Schwikowski et al., 1999, 2004). For instance, concentration records of ammonium, mainly released from livestock breeding and agriculture (Döscher et al., 1996;

Schwikowski et al., 1999); nitrate, primarily emitted by traffic (Döscher et al., 1995; Preunkert et al., 2003;

Wagenbach et al., 1988); sulfate, typically from fossil fuel burning (Döscher et al., 1995; Preunkert et al., 2001; Schwikowski et al., 1999); and lead, a heavy metal mainly emitted by mining activities, metal production, coal combustion, or the use of leaded gasoline (Schwikowski et al., 2004), revealed the strong impact of Western European industry and society on the atmosphere over the last decades. Concentration records of Ca, Mg, or the rare‐earth elements can be used to reconstruct historic variations of mineral dust emissions to the atmosphere (Gabrieli et al., 2011; Gabrielli et al., 2008). However, postdepositional melting induced by climate warming can significantly alter concentration records of atmospheric trace species from high‐altitude glaciers and snowpack as shown for MIs, organic pollutants, or water stable isotopes (WSIs;

Eichler et al., 2001; Herreros et al., 2009; Kang et al., 2008; Müller‐Tautges et al., 2016; Pavlova et al., 2015; Sinclair & MacDonell, 2016; You et al., 2015). As glaciers, which have served as environmental archives to assess the natural and anthropogenic impact on the atmosphere, are progressively in danger of being affected by melting (Zhang et al., 2015), there is an increasing need to understand the impact of melting on the preservation of various environmental proxies in these archives. Investigation of the fate of MIs during melting of snowpack and glacier ice has been the subject of several studies (Eichler et al., 2001; Ginot et al., 2010; Grannas et al., 2013; Kang et al., 2008; Lee et al., 2008; Z. Li et al., 2006;

©2019. The Authors.

This is an open access article under the terms of the Creative Commons Attribution‐NonCommercial‐NoDerivs License, which permits use and distri- bution in any medium, provided the original work is properly cited, the use is non‐commercial and no modifica- tions or adaptations are made.

Key Points:

Impurity profiles in the seasonal snowpack at Weissuhjoch, Swiss Alps, suggest a well‐preserved atmospheric composition in winter

Melting during warm periods causes preferential elution of selected major ions and trace elements from the snowpack

Ammonium, the rare‐earth elements, and low concentrated trace elements tend to be most resistant to meltwaterinduced relocation

Correspondence to:

A. Eichler, anja.eichler@psi.ch

Citation:

Avak, S. E., Trachsel, J. C., Edebeli, J., Brütsch, S., BartelsRausch, T., Schneebeli, M., et al. (2019).

Meltinduced fractionation of major ions and trace elements in an Alpine snowpack.Journal of Geophysical Research: Earth Surface,124, 16471657. https://doi.org/10.1029/

2019JF005026

Received 12 FEB 2019 Accepted 9 MAY 2019

Accepted article online 31 MAY 2019 Published online 1 JUL 2019

(2)

Virkkunen et al., 2007; Wang et al., 2018; Zong‐Xing et al., 2015). Preferential elution of certain ions relative to others has been observed, being strongly dependent on the respective geographical location of the snow- pack. These studies further indicate that SO42−is generally significantly depleted with melting, while NH4+

appears to be preserved despite meltwater percolation.

In the majority of these studies available in the literature, either melting had strongly affected the MI records making the initial conditions before melting unknown or a direct comparison of concentration records before and after melting to quantify the degree of depletion has not been performed. In contrast to MIs, the effect of melting on the fate of TEs in snow has still not been well characterized. To the best of our knowl- edge, only a few studies have addressed the influence of melting on TE concentration records in snow pits.

Wong et al. (2013) artificially infiltrated meltwater into Greenland snow pits to investigate melt effects on 12 different TE records. They observed that mineral dust particle‐bound TEs remained immobile during melt- water percolation resulting in the preservation of their seasonal chemical signal. On the contrary, Zhongqin et al. (2007) reported from the analyses of snow‐firn pits taken in the Central Asian Tien Shan Mountains that meltwater percolation during the summer may have eluted thefive investigated TEs (Al, Cd, Fe, Pb, and Zn) from the snow‐firn pack. The TE records from the analysis of a meltwater‐affectedfirn part of an Alpine ice core indicated that meltwater percolation led to preferential loss of certain TEs (Avak et al., 2018). Water‐insoluble TEs and low‐abundant water‐soluble TEs remained largely immobile with melt- water. Since this study proposed a geographical site specificity for the preferential elution of TEs (Avak et al., 2018), it remains unclear how representative thesefindings are for the Alpine region.

The high‐Alpine snowfield site at Weissfluhjoch (WFJ), Switzerland, is well suited for snowpack studies.

Numerous studies have been conducted at this site focusing on snow characterization, snow mechanics, snow metamorphism, and the development of measurement methods since 1936 (Marty & Meister, 2012).

So far, studies investigating the chemical impurities in the snowpack at WFJ focused on SO42−, NO3, Cl, K+, Na+, NH4+

, Ca2+, and Mg2+. Baltensperger et al. (1993) compared consecutive measurements of surface snow sampled from January to March 1988 to that from a snow pit taken at the end of March at WFJ. The snow pit samples were found to be representative of precipitation deposition during the winter. Snow pit sampling, sample handling procedures, and chemical analysis of MIs were further refined by Schwikowski et al. (1997). In addition, the seasonal variability in deposition and the different emission sources of impurities were reflected in the vertical distribution of MIs in the snowpack (Schwikowski et al., 1997).

Here we present thefirst study monitoring the postdepositional fate of an extensive set of environmentally relevant atmospheric impurities in the high‐Alpine snowpack at WFJ. We monitored 6 MIs and 34 TEs dur- ing an entire winter/spring season and investigated their behavior during melting of the snowpack in early summer. Five snow pits sampled from January to June 2017 allowed us to capture both the initial, undis- turbed records of atmospheric compounds in the snowpack and the records after melting had occurred dur- ing the warmer season. Impurity profiles reflecting dry conditions (without significant melting) were compared with profiles reflecting wet conditions to systematically investigate the impact of melting on the preservation of atmospheric tracers in high‐Alpine snow.

2. Materials and Methods

2.1. Study Site and Meteorological Setting

Five snow pit samplings were conducted at the high‐Alpine snowfield site WFJ of the WSL‐Institute for Snow and Avalanche Research, Eastern Switzerland (2,536 m above sea level, 46°49′47″N 9°48′33″E) at reg- ular time intervals in the winter, spring, and early summer seasons in 2017. Sampling dates were 25 January, 22 February, 21 March, 17 April, and 1 June. Earlier snow sampling studies suggest uniform spatial snow deposition and insignificant perturbations of the snow stratigraphy due to strong winds at this site (Baltensperger et al., 1993; Schwikowski et al., 1997). Samples were obtained from the snow pits either 1 day before or after the weekly density measurements with 3‐cm resolution, which were part of an extensive snowpack monitoring program of the WSL‐Institute for Snow and Avalanche Research during the winter season 2016/2017 (Calonne et al., 2016). The snow heights during the sampling were 87 cm (25 January), 126 cm (22 February), 185 cm (21 March), 166 cm (17 April), and 83 cm (1 June; Figure 1). Snow surface (via an infrared radiometer) and 2‐m air temperatures (by an automated weather station) indicated that

(3)

dry conditions without significant melting prevailed for the first three sampling dates on 25 January, 22 February, and 21 March (Figure 1).

Only 6 days with a mean 2‐m air temperature above 0 °C occurred during the period 1 January until 21 March (1.3 ± 0.8 °C on average). Partial melting was identifiable in the snow pit on the fourth sampling date on 17 April due to 13 days between 21 March and 17 April with mean 2‐m air temperatures above 0 °C (1.8 ± 1.3 °C on average). The snowpack was entirely wet on thefifth sampling date on 1 June due to another 25 days between 17 April and 1 June with mean 2‐m air temperatures above 0 °C (4.8 ± 3.4 °C on average).

2.2. Snow Pit Sampling

As snow is particularly sensitive to trace levels of contamination due to its low impurity content, precautions were taken during sampling. Sterile clean room overalls (Tyvek® IsoClean®, DuPont, Wilmington, DE, USA), particulate respirator face masks (3M, Maplewood, MN, USA), and ultra- clean plastic gloves (Semadeni, Ostermundigen, Switzerland) were worn during the sampling. All tools were carefully rinsed with ultrapure water (18 MΩcm quality, arium® pro, Sartorius, Göttingen, Germany) prior to use. Snow pits were sampled with a vertical resolution of 6 cm down to the bottom of the snowpack by pushing a custom‐built rectangular (15 × 24 cm) sampler made from polycarbonate into the pit wall. To allow for sufficient sample volume, the snow was filled into 50‐ml polypropylene vials (Sarstedt, Nümbrecht, Germany) by pushing them twice with the opening facing downward into the snow. Separate vials were used for MI, WSI, and TE analyses. Polypropylene vials were precleanedfive times with ultrapure water for MI and WSI samples. For TE samples, the tubes were precleanedfive times with ultrapure water (18 MΩcm quality, Milli‐Q® Element, Merck Millipore, Burlington, MA, USA) plus once with 0.2 M HNO3prepared from ultrapure HNO3(OptimaTM, Fisher Chemical, Loughborough, UK). Samples for TE analysis were taken from the snow at the front part of the sampler to prevent possible cross contamination of the vials used for MI sampling with HNO3.

2.3. Major ion, water stable isotope, and trace element analysis

A total of 324 samples for MI, WSI, and TE analyses was kept frozen at−20 °C until analysis at the Paul Scherrer Institute (Villigen, Switzerland).

MIs (Na+, NH4+, Ca2+, Cl, NO3, and SO42−) present in the snow pit samples were analyzed after melting at room temperature using ion chromatography (IC; 850 Professional IC equipped with a 872 Extension Module Liquid Handling and a 858 Professional Sample Processor auto sampler, Metrohm, Herisau, Switzerland). Cations were separated using a Metrosep C4 column (Metrohm) and 2.8 mM HNO3as eluent at aflow rate of 1 ml/min. Anions were separated using a Metrosep A Supp 10 column (Metrohm) and were eluted stepwise usingfirst, a 1.5 mM Na2CO3/0.3 mM NaHCO3(1:1 mixture) eluent, and then an 8 mM Na2CO3/1.7 mM NaHCO3(1:1 mixture) eluent at aflow rate of 0.9 ml/min. Possible instrumental drifts were monitored by measuring an in‐house standard after every twentieth sample. The precision of the method was ~5%.

WSI samples were melted at room temperature, and 1 ml aliquots were analyzed forδD andδ18O using a wavelength‐scanned cavity ring down spectrometer (WS‐CRDS, L2130‐iAnalyzer, Picarro, Santa Clara, CA, USA). Samples were injected into the vaporizer (A0211, Picarro, Santa Clara, CA, USA) using a PAL HTC‐xt autosampler (LEAP Technologies, Carrboro, NC, USA). Three in‐house standards were measured after every tenth sample for calibration and to monitor instrumental drifts. The measurement uncertainty was <0.1‰forδ18O and <0.5‰forδD.

Snow pit samples were melted at room temperature, acidified with concentrated ultrapure HNO3to 0.2 M (1% v/v), and analyzed 3–4 hr after acidification following the same procedure as described in Avak et al.

(2018) using discrete inductively coupled plasma sector field mass spectrometry (Element 2, Thermo Fisher Scientific, Bremen, Germany). Low (LR,R= 300) or medium resolution (MR,R= 4,000) data were Figure 1.The 2m air and snow surface temperatures and daily precipita-

tion rate and snowpack heights at the Weissuhjocheld site, Swiss Alps, during the winter season of 2016/2017. Snow pit samplings were conducted both in the cold season, where dry conditions without signicant melting prevailed, and in the warm season, where severe melting of the snowpack occurred.

(4)

acquired for Ag (LR), Al (MR), Ba (LR), Bi (LR), Ca (MR), Cd (LR), Ce (LR), Co (MR), Cs (LR), Cu (LR and MR), Eu (LR), Fe (MR), La (LR), Li (LR and MR), Mg (MR), Mn (MR), Mo (LR), Na (MR), Nd (LR), Ni (LR and MR), Pb (LR), Pr (LR), Rb (LR), Sb (LR), Sc (MR), Sm (LR), Sr (LR), Tl (LR), Th (LR), U (LR), V (MR), W (LR), Yb (LR), Zn (LR and MR), and Zr (LR). If both LR and MR resolution data were available, MR data were used for further data evaluation. Quantification of intensity values was performed by internal and external calibrations using a Rh‐standard and seven liquid standards covering the typical concentration range of TEs in natural snow and ice samples, respectively. Linear regressions of the calibration curves consistently revealed correlation coefficients >0.999.

2.4. Data Evaluation

IC raw data were processed using the MagIC Net 3.2 software (Metrohm, Herisau, Switzerland). Sample concentrations were not blank‐corrected as the concentrations of the blank, determined by analyzing ultrapure water, were below the detection limit (DL). Sample concentrations below the DL were substituted with half the value of the DL. On average, 10% of the measurement values were below the DL. Inductively coupled plasma mass spectrometry raw data were evaluated following the method described by Knüsel et al. (2003). Concentrations were blank‐corrected by subtracting a measurement blank consisting of four measurements of ultrapure 0.2 M HNO3. The instrumental DL was defined as 3σof the mea- surement blank, and concentrations below the DL were substituted with half of the value of the DL. 109Ag was excluded from the data set for further evaluation and discussion as concentrations of all samples were below the DL.

Total depths of the snow profiles were converted to water equivalents (w.

eq.) by multiplying with the respective density. As 17 April had the largest snow depth, the profiles of 25 January, 22 February, 21 March, and 1 June were aligned relative to the profile of 17 April, with 0‐cm w.eq. being the surface on this day and assuming the same bottom depth for all profiles.

Allfive profiles cover the depth interval 45‐to 65‐cm w.eq. The profiles of 22 February, 21 March, 17 April, and 1 June cover 36‐to 65‐cm w.eq., the profiles of 22 February, 21 March, and 17 April cover 30‐to 65‐cm w.eq., and the profiles of 21 March and 17 April cover 5‐to 65‐cm w.eq.

The sample at the base of each snow pit was omitted from all chemical profiles to exclude a possible influence of the soil at WFJ.

3. Results and Discussion

3.1. Major ions

3.1.1. Comparison of the Five MI Concentration Profiles

The chemical profiles of selected MIs (NH4+, Cl, and Ca2+) and theδ18O records in thefive snow pits are shown in Figure 2. The general MI patterns of the four profiles, reflecting the winter and spring periods (January–April), show a strong correspondence in the respective overlapping parts. Apart from possible changes caused by melting (17 April), slight concentration differences in the profiles can be attributed to the spatial variability of impurities within the snowpack as the locations of the individual snow pit sam- plings were several meters (up to 20 m) apart. The resemblance between the four winter and spring snow pits is also visible for the correspondingδ18O profiles which support the depth assignment of the different snow pits.

The chemical profiles of NH4+

, Cl, and Ca2+are exemplary for the different emission sources and transport characteristics of MIs in Alpine snow and ice. The depths 0‐to 30‐cm w.eq. (21 March and 17 April) indicate Figure 2.Concentration proles of NH4+, Cl, and Ca2+of theve snow

pits taken at the Weissuhjocheld site during winter/spring (black/blue curves) and early summer (red curve). For comparison, the corresponding δ18O records are shown. The gray dashed line indicates the bottom of the snowpack. Prole depths were aligned to the one of 17 April, and 0cm w.eq.

reects the surface of the snowpack on this day. The surface sample of the snow pit from 1 June is shown with a dashed line (see text for details). The shaded area indicates the depth interval (45to 65cm w.eq.) used for cal- culation of the concentration ratio and determining an elution sequence. w.

eq. = water equivalents.

(5)

precipitation occurring in spring after 22 February, while 30‐ to 65‐cm w.eq. represent winter precipitation (Figure 1). The NH4+

springtime concentrations are roughly three times higher than win- tertime concentrations. NH4+detected in snow pits and ice cores from European high‐Alpine sites is nowadays generally of anthropo- genic origin (Döscher et al., 1996; Gabrieli et al., 2011). NH4+also shows a pronounced maximum toward summer due to enhanced emissions from agricultural activities and stronger convection (Baltensperger et al., 1997). Higher concentrations of NH4+

in the springtime snowfall compared to wintertime snowfall at WFJ were also observed by Schwikowski et al. (1997). Ca2+and Cl, unlike NH4+

, have higher concentrations in the wintertime snow than in spring. At Alpine sites, Ca2+and Clare indicative of the input from mineral dust (Bohleber et al., 2018; Schwikowski et al., 1995;

Wagenbach et al., 1996) and sea salt aerosols (Eichler et al., 2004), respectively. A very pronounced peak in the Ca2+and Na+profiles between 50‐and 60‐cm w.eq. (Figure 2) can be most likely attributed to a Saharan dust event. This is corroborated by back trajectory ana- lysis for the beginning of January 2017 and measurements by the Swiss Federal Office of Meteorology and Climatology at the Jungfraujoch high‐Alpine research station on 5 January 2017 (MeteoSwiss, 2014).

After several weeks with temperatures above 0 °C, the snowpack was very wet from meltwater (and rainwater) at the beginning of the sum- mer (1 June; Figure 1). The uppermost (surface) sample of the snow pit from 1 June (red dotted line in Figure 2) is most likely influenced by residual snow impurities after surface melting and wet or dry deposition between 17 April and 1 June (Figure 1) and was thus not taken into consideration. Theδ18O profile of 1 June reveals a strong smoothing compared to the previous profiles, indicative of strong melting (Thompson et al., 1993). However, a signature between 45‐ and 65‐cm w.eq. resembling that from the previous profiles is still visible. This indicates that at this depth, meltwater percolation was concentrated to the porous space in the snowpack and primarily per- colated along ice surfaces, keeping the ice matrix in principle pre- served. The MI profiles of 1 June compared to those of the first three sampling dates are almost unaltered (NH4+), slightly depleted (Na+and Cl), or strongly depleted (Ca2+, NO3, and SO42−), most likely due to different elution behavior of the investigated ions with meltwater (Eichler et al., 2001).

3.1.2. Preferential Elution of MIs: Elution Sequence and Discussion

To quantify and compare the apparent preferential elution of MIs, a concentration ratio cwet/cdryfor the overlapping depth (45‐to 65‐cm w.eq.) was calculated for each MI (Table 1). cwetcorresponds to the integrated MI concentration for the wet profile of 1 June, whereas cdryrepresents the mean of the integrated concentration profiles during dry (no significant melting) periods (25 January, 22 February, and 21 March). The profiles on 17 April were not included as the snowpack was neither completely dry nor very wet. Corresponding to the concentration ratio, an elution sequence was established: NH4+< Cl~ Na+< NO3~ Ca2+~ SO42−, where NH4+is the least mobile ion and SO42−the most mobile ion. Similar elution sequences were reported by Eichler et al.

(2001; from another melt‐affected Alpine site), Z. Li et al. (2006), Ginot et al. (2010), and recently by Zong‐Xing et al. (2015), who all observed a particularly pronounced leaching of SO42−and Ca2+, while NH4+

and Clwere retained in meltingfirn and snowpack.

Table 1

Concentration Ratio, Classicationa, Mean Concentration of the Dry Snow Pits (25 January, 22 February, and 21 March), and Detection Limits of Major Ions and Trace Elements Investigated in This Study

Concentration ratio/classication

Mean concentration (ng/L)

Detection limit (ng/L)

Na+ 0.62 ± 0.45 17,000 ± 2,840 500

NH4+ 0.77 ± 0.27 26,300 ± 4,580 500

Ca2+ 0.37 ± 0.18 71,500 ± 11,400 10,000

Cl 0.64 ± 0.43 34,000 ± 4,980 1,000

NO3 0.37 ± 0.11 249,000 ± 21,200 1,000

SO42− 0.36 ± 0.10 57,300 ± 5,330 5,000

Al i.b.s. 4,870 ± 792 115

Ba 0.39 ± 0.26/group 2 262 ± 67 3.5

Bi 0.50 ± 0.28/group 2 4.5 ± 0.91 0.01

Ca 0.46 ± 0.24/group 2 110,000 ± 21,400 498

Cd 0.55 ± 0.23/group 2 1.9 ± 0.26 0.27

Ce 1.5 ± 1.1/group 1 14 ± 2.9 0.03

Co i.b.s. 24 ± 4.5 0.22

Cs i.b.s. 2.1 ± 0.33 0.07

Cu i.b.s. 48 ± 5.6 2.7

Eu 1.1 ± 0.79/group 1 0.51 ± 0.10 0.13

Fe i.b.s. 5,990 ± 1,180 245

La 1.5 ± 1.1/group 1 6.1 ± 1.2 0.03

Li i.b.s. 6.1 ± 0.90 3.8

Mg i.b.s. 37,000 ± 6,600 30

Mn i.b.s. 612 ± 110 0.69

Mo 1.1 ± 0.59/group 1 3.1 ± 0.46 1.8

Na 0.50 ± 0.32/group 2 15,900 ± 2,620 1,280

Nd 1.2 ± 0.92/group 1 7.7 ± 1.7 0.01

Ni i.b.s. 101 ± 22 9

Pb 0.84 ± 0.36/group 1 180 ± 15 2.7

Pr 1.1 ± 0.84/group 1 1.8 ± 0.38 0.01

Rb i.b.s. 15 ± 2.1 0.54

Sb 0.88 ± 0.36/group 1 9.1 ± 1.2 0.13

Sc 1.5 ± 1.4/group 1 1.8 ± 0.41 0.17

Sm 1.2 ± 0.88/group 1 2.0 ± 0.46 0.02

Sr 0.37 ± 0.21/group 2 270 ± 55 3.4

Th 0.45 ± 0.30/group 2 1.4 ± 0.25 0.03

Tl i.b.s. 0.39 ± 0.04 0.01

U 1.5 ± 0.99/group 1 2.3 ± 0.64 0.01

V i.b.s. 19 ± 2.5 0.37

W 1.2 ± 0.69/group 1 0.54 ± 0.07 0.18

Yb i.b.s. 0.63 ± 0.13 0.02

Zn 0.47 ± 0.36/group 2 2,780 ± 463 28

Zr 0.13 ± 0.08/group 2 5.4 ± 0.75 0.26

aTrace elements were classied depending on whether the prole of the snow pit of 1 June showed a strong enrichment from the soil at Weissuhjoch below 55cm water equivalent (inuenced by soil [i.b.s.]), an unbiased prole pattern compared to the dry condition proles (group 1), or a heavily depleted concen- tration prole (group 2).

(6)

Elution sequences of MIs were also determined through laboratory stu- dies (Cragin et al., 1996; Tranter et al., 1992; Tsiouris et al., 1985). We recently performed an elution experiment where, for thefirst time, homo- genous impurity‐doped artificial snow was exposed to well‐defined snow metamorphism conditions prior to leaching with 0 °C ultrapure water.

This experiment was conducted to determine enrichment differences of MIs between ice crystal interiors and surfaces after metamorphism (Trachsel et al., 2017). Snow undergoes drastic structural transformation cycles during metamorphism (Pinzer et al., 2012), which may result in sig- nificant impurity redistribution. Our elution sequence at the WFJ agrees well with the findings of the laboratory‐based elution experiment (Figure 3). MIs having a higher solubility in ice such as NH4+and Cl (Feibelman, 2007; Hobbs, 1974) showed less mobility most likely due to incorporation into the crystal interior during snow metamorphism.

They can be incorporated either by substituting water molecules located on lattice sites of the ice crystal (Zaromb & Brill, 1956) or by occupying interstitial spaces of the crystal structure (Petrenko & Whitworth, 2002).

On the other hand, Ca2+and SO42−were enriched on ice crystal surfaces with progressing snow metamorphism (Figure 3); explaining their avail- ability for mobilization with meltwater. This is supported by previous location studies of salts in ice. Mulvaney et al. (1988) used a combination of scanning electron microscopy and energy‐dispersive X‐ray microanaly- sis to show that H2SO4concentrations at triple junctions are several orders of magnitude higher compared to grain interiors in polar ice from Antarctica. Similar observations for H2SO4and HNO3using Raman spec- troscopy were reported by Fukazawa et al. (1998). Accumulation of MgSO4at grain boundaries in ice from the Greenland Ice Sheet Project (GISP) ice core 2 was revealed by low‐vacuum scanning electron microscopy‐energy‐dispersive X‐ray (Baker et al., 2003). The varying elution behavior of MIs observed in the snow pit of 1 June with NO3, Ca2+, and SO42−being most heavily depleted can therefore be explained by their microscopic locations on ice surfaces, exposing them to relocation during melting.

Ourfindings show that the atmospheric composition of MIs is well preserved in the snowpack at WFJ during the cold season. Melting during the warm season leads to preferential leaching of MIs depending on their microscopic location either on the ice surface or in the ice interior. The loss of MIs is particularly significant for Ca2+and SO42−. In contrast, the strong persistence of NH4+

emphasizes that NH4+

can still serve as environmental tracer for the interpretation of snow pit and ice core records affected by melting.

3.2. Trace elements

3.2.1. TE Concentration Profiles

Concentrations of Co, Fe, Ce, Sb, Ca, and Sr in the snow pits on 25 January, 22 February, 21 March, 17 April, and 1 June are exemplarily shown in Figure 4. The general pattern of the profiles of thefirst four snow pits (January to April) agrees well in the overlapping depths. As for the MI profiles, misalignment in peaks can be attributed to small‐scale spatial variability of impurities in the snowpack. The reproducibility of chemical profiles with ultratrace level concentrations (e.g., Ce and Sb) from different sampling campaigns demon- strates that significant contamination during the snow pit samplings did not occur. Inductively coupled plasma mass spectrometry measurements of Ca and Na concentration profiles are in good agreement with the records of Ca2+and Na+concentrations obtained by IC (Figures 2 and 4). The uppermost (surface) sam- ple of the snow pit from 1 June revealed highly elevated concentrations for many TEs (red dotted line in Figure 4). This is most likely due to residual snow impurities after surface melting and TEs that reached the snowpack by wet or dry deposition between 17 April and 1 June (Figure 1). For this reason, the upper- most sample of the snow pit from 1 June was not taken into consideration for the following discussion.

The 34 TEs can be either of geogenic origin or emitted by anthropogenic sources. At high‐Alpine sites, Al, Ba, Bi, Ca, Cs, Fe, Li, Mg, Mn, Na, Rb, Sr, Th, Tl, U, W, Zr, and the rare‐earth elements (Ce, Eu, La, Nd, Pr, Sc, Figure 3.Time evolution of relative enrichment of Na+, NH4+, Ca2+, Cl,

and SO42−concentrations at ice crystal surfaces determined by leaching articial snow, homogeneously doped with known concentrations of MIs, and exposed for different time periods to a temperature gradient mimicking snow metamorphism, with 0 °C ultrapure water (adapted from Trachsel et al., 2017).

(7)

Sm, and Yb) are mainly deposited with mineral dust (Gabrieli et al., 2011;

Gabrielli et al., 2008), whereas Ag, Cd, Co, Cu, Mo, Ni, Pb, Sb, V, and Zn in Alpine snow and ice are characteristic of anthropogenic atmospheric pol- lution (Barbante et al., 2004; Gabrieli et al., 2011; Gabrielli et al., 2008;

Schwikowski et al., 2004; van de Velde et al., 1999, 2000). The chemical profiles of Ca, Ce, Fe, and Sr (Figure 4) are representative of TEs originat- ing from mineral dust particles and reveal an enrichment in the depths reflecting winter precipitation (30‐to 65‐cm w.eq.), where dry conditions prevailed (25 January, 22 February, and 21 March). These TEs show a major peak between 50‐and 60‐cm w.eq. due to a Saharan dust event beginning of January 2017 (see section 3.1.1). As explained above for NH4+, TEs indicative of anthropogenic influence such as Sb (Figure 4) are enriched in depths reflecting spring precipitation deposited after 22 February (0‐to 30‐cm w.eq.). This is due to the typical seasonality of con- vective air mass transport from the planetary boundary layer to high‐ Alpine sites occurring in spring and summer (Baltensperger et al., 1997).

Based on the concentration profiles on 1 June, the 34 investigated TEs were divided into three groups (Table 1). Group 1 (Ce, Eu, La, Mo, Nd, Pb, Pr, Sb, Sc, Sm, U, and W) is characterized by an almost unchanged profile pattern between 45‐and 65‐cm w.eq. compared to the dry condi- tion profiles (e.g., Ce and Sb, Figure 4). This indicates that group 1 TEs were not affected by melting. Group 2 TEs (Ba, Bi, Ca, Cd, Na, Sr, Th, Zn, and Zr) were strongly depleted on 1 June compared to winter and springtime records (as shown for Ca and Sr in Figure 4).

Group 3 TEs (Al, Co, Cs, Cu, Fe, Li, Mg, Mn, Ni, Rb, Tl, V, and Yb) exhibit a strong enrichment below 55‐cm w.eq. close to the soil (see, e.g., Co and Fe, Figure 4). The bottom of the snowpack on 1 June was very wet, prob- ably enriched with impurities not of atmospheric origin, but present in the soil at WFJ. We therefore attribute the strong increase in concentration below 55‐cm w.eq. to the contact with soil and denote this group as

“influenced by soil.” Group 3 TEs were therefore not included in further evaluation.

3.2.2. Different Preservation of TEs Under Melt Conditions A quantitative classification of TE fractionation during melting of the snowpack (groups 1 and 2) was performed by calculating a concentration ratio cwet/cdryfor each TE in the overlapping depths (see above, Table 1).

The highest concentration ratio was obtained for La, indicative of its well preserved concentration record on 1 June, whereas the concentration pro- file of Zr is most severely affected by melting, having the lowest concentra- tion ratio. Concentration ratios of TEs classified as group 1 are close to 1 and range between 1.5 ± 1.1 (La) and 0.84 ± 0.36 (Pb). Concentration ratios of group 2 TEs are significantly below 1 and range between 0.55 ± 0.23 (Cd) and 0.13 ± 0.08 (Zr). Arranging the concentration ratios according to size and plotting this rank against the ratio indicates that each TE belonging to groups 1 and 2 was differently preserved in the snowpack during melting (Figure 5).

The preservation of group 1 TEs with observed concentration ratios >0.7 (Ce, Eu, La, Mo, Nd, Pb, Pr, Sb, Sc, Sm, U, and W) is in excellent agreement withfindings from another Alpine site (Avak et al., 2018, Figure 5). At this

~180‐km distant site, meltwater percolation led to postdepositional distur- bance of a 16‐mfirn section of a high‐Alpine ice core from the upper Grenzgletscher (Monte Rosa massif, southern Swiss Alps, 4,200 m above sea level, 45°55′28″N 7°52′3″E). Preferential elution of TEs led to significant concentration depletion in the records of Ba, Ca, Cd, Co, Mg, Mn, Na, Ni, Sr, and Zn, whereas Figure 4.Concentration proles of Co, Fe, Ce, Sb, Ca, and Sr of theve

snow pits taken at the Weissuhjocheld site between winter/spring (black/blue curves) and early summer (red curve). The gray dashed line indicates the bottom of the snowpack. Prole depths were aligned to the one of 17 April, and 0cm water equivalent reects the surface of the snowpack on this day. The surface sample of the snow pit from 1 June is shown with a dashed line (see text for details). The shaded area indicates the depth inter- val (45to 65cm water equivalent) used for calculation of the concentration ratio. Trace elements were classied asinuenced by soil,or allocated to group 1 or 2 (retained or depleted concentration prole, according to the pattern of 1 June).

(8)

the seasonality in the Al, Bi, Ce, Cs, Cu, Eu, Fe, La, Li, Mo, Nd, Pb, Pr, Rb, Sb, Sc, Sm, Tl, Th, U, V, W, Yb, and Zr records was well preserved (Avak et al., 2018). The different behavior of TEs during meltwater percolation was related to their varying water solubility and location at the micro- scopic scale. TEs mainly present in water‐insoluble mineral dust particles at the upper Grenzgletscher site (Al, Fe, Zr, and the rare‐earth elements) were enriched on firn grain surfaces and mostly preserved, since their insolubility in water resulted in immobility with meltwater. TEs likely pre- sent in water‐soluble compounds (Ba, Bi, Ca, Cd, Co, Cs, Cu, Li, Mg, Mn, Mo, Na, Ni, Pb, Rb, Sb, Sr, Tl, Th, U, V, W, and Zn; Avak et al., 2018;

Birmili et al., 2006; Greaves et al., 1994; Hsu et al., 2010; T. Li et al., 2015) revealed variable mobility with meltwater presumably due to the different microscopic location of the ions in the ice structure. For those rather water‐soluble TEs, the concentration at the upper Grenzgletscher site was found to be primarily decisive determining either incorporation into the ice interior during snow metamorphism or segregation to ice sur- faces because of exceeded solubility limits in ice.

In correspondence with the upper Grenzgletscher study and the work reported by Wong et al. (2013), also at WFJ, records of investigated TEs present as water‐insoluble mineral dust particles (rare‐earth elements:

Ce, Eu, La, Nd, Pr, Sc, and Sm) are well preserved most probably due to their immobility with meltwater. Contrariwise, Zhongqin et al.

(2007) observed a depletion of mineral dust‐bond elements, such as Fe and Al at Urumqi glacier 1 (eastern Tien Shan) in summer and relate this to meltwater relocation. However, it cannot be excluded that the low con- centrations in summer observed for all investigated elements are due to dilution effects during the wet season.

There is also strong agreement in the behavior of rather water‐soluble TEs between upper Grenzgletscher and WFJ, showing a dependence on the concentration level. While TEs present in ultratrace amounts tend to be preserved, more abundant TEs were preferentially eluted (Figure 5). The only three TEs revealing a different behavior between the two sites are Bi, Th, and Zr, which still show a preservation of the seasonality at upper Grenzgletscher, but significant depletion at WFJ. One explanation could be a higher proportion of the water‐soluble fraction of Bi, Th, and Zr deposited during winter at WFJ, favoring higher meltwater mobility. Another possible reason is dif- fering concentrations. However, Th and Zr concentrations are not significantly different between the two sites. Only the higher Bi concentrations at WFJ could explain stronger melt‐induced relocation.

Thesefindings in the preservation of TEs corroborate our previous hypothesis (Avak et al., 2018) that for ice cores and snow pits from Alpine areas, representing central European atmospheric aerosol composi- tion, and affected by partial melting, Ce, Eu, La, Mo, Nd, Pb, Pr, Sb, Sc, Sm, U, and W may still be used as robust environmental proxies. However, chemical profiles of Ba, Ca, Cd, Na, Zn, and Sr are prone to be depleted by melting. Records of Bi, Th, and Zr behaved differently at the two Alpine sites. The potential of TEs classified as influenced by soil here and previously found to be preserved (Al, Cs, Cu, Fe, Li, Rb, Tl, V, and Yb) or depleted (Co, Mg, Mn, and Ni) infirn (Avak et al., 2018) could not be further investigated at the WFJ site because of the influence of the soil in the presence of meltwater.

4. Conclusions

Melting processes induced by climate warming can significantly alter concentration records of atmospheric pollutants deposited in high‐altitude glacier ice and snowpacks. Understanding the postdepositional beha- vior of atmospheric trace species is essential for reconstructing past environmental conditions from these natural archives. The analysis offive snow pits sampled during the winter/spring season of 2017 at the high‐Alpine site WJF, Switzerland, is presented here. Comparison of impurity profiles representing dry and wet snow conditions allowed to systematically investigate the impact of melting on the preservation Figure 5.Rank of preservation plotted against the concentration ratio cwet/

cdryin the overlapping part (45to 65cm water equivalent depth) for each trace element classied into group 1 (retained concentration prole) or 2 (depleted concentration prole). cwetcorresponds to the integrated trace element concentration of the wet prole (1 June), whereas cdryrepresents the mean of the integrated concentration proles during dry conditions (25 January, 22 February, and 21 March). Symbols in bold indicate a similar behavior as observed for highAlpinern affected by meltwater percolation (Avak et al., 2018). Circle sizes represent the mean concentrations in the dry snow pits where no melting occurred.

(9)

of atmospheric tracers in high‐Alpine snow. While the atmospheric composition was found to be well pre- served during the winter, melting in the spring and early summer caused a preferential loss of certain MIs and TEs. We propose that the elution behavior of MIs is depending on their microscopic location, which is most likely defined by redistribution processes occurring during snow metamorphism. Among the six investigated MIs (Ca2+, Cl, Na+, NH4+, NO3, and SO42−), NH4+was retained the strongest, probably due to its preferred location in the ice interior. Ca2+and SO42−concentrations were significantly depleted;

this can be explained by their predominant occurrence on ice crystal surfaces. Variable mobility was also observed for TEs, with Ce, Eu, La, Mo, Nd, Pb, Pr, Sb, Sc, Sm, U, and W being the best‐preserved elements.

In tendency, high concentrations of TEs were found to favor relocation with meltwater, while low abundant TEs were preferably retained.

The MIs and 18 out of 21 investigated TEs revealed a consistent behavior with meltwater percolation at two Alpine sites, Weissfluhjoch and upper Grenzgletscher, which are 180 km apart. This indicates that the observed species dependent preservation from melting snow and ice is particularly valid for the Alpine region, reflecting central European atmospheric aerosol composition. Based on the observations at the two Alpine sites, we propose that NH4+

, TEs deposited as water‐insoluble particles, and TEs of water‐soluble particles occurring in low concentrations may still be applicable as environmental proxies in snow pits and ice cores affected by melting. Corroborating the existence of meltwater‐resistant proxies is particularly rele- vant as many high‐mountain glaciers worldwide, which provide avenues for assessing natural and anthro- pogenic impact on the atmosphere, are increasingly affected by melting.

References

Avak, S. E., Schwikowski, M., & Eichler, A. (2018). Impact and implications of meltwater percolation on trace element records observed in a highAlpine ice core.Journal of Glaciology,64(248), 877886. https://doi.org/10.1017/jog.2018.74

Baker, I., Cullen, D., & Iliescu, D. (2003). The microstructural location of impurities in ice.Canadian Journal of Physics,81(1‐2), 1–2), 1–9.

https://doi.org/10.1139/p03030

Baltensperger, U., Gäggeler, H. W., Jost, D. T., Lugauer, M., Schwikowski, M., Weingartner, E., & Seibert, P. (1997). Aerosol climatology at the highalpine site Jungfraujoch, Switzerland.Journal of Geophysical Research,102(D16), 19,70719,715. https://doi.org/10.1029/

97JD00928

Baltensperger, U., Schwikowski, M., Gäggeler, H. W., Jost, D. T., Beer, J., Siegenthaler, U., et al. (1993). Transfer of atmospheric consti- tuents into an alpine snowfield.Atmospheric Environment. Part A. General Topics,27(12), 1881–1890. https://doi.org/10.1016/0960‐

1686(93)902938

Barbante, C., Schwikowski, M., Döring, T., Gäggeler, H. W., Schotterer, U., Tobler, L., et al. (2004). Historical record of European emissions of heavy metals to the atmosphere since the 1650s from Alpine snow/ice cores drilled near Monte Rosa.Environmental Science &

Technology,38(15), 4085–4090. https://doi.org/10.1021/es049759r

Birmili, W., Allen, A. G., Bary, F., & Harrison, R. M. (2006). Trace metal concentrations and water solubility in sizefractionated atmospheric particles and influence of road traffic.Environmental Science & Technology,40(4), 1144–1153. https://doi.org/10.1021/

es0486925

Bohleber, P., Erhardt, T., Spaulding, N., Hoffmann, H., Fischer, H., & Mayewski, P. (2018). Temperature and mineral dust variability recorded in two lowaccumulation Alpine ice cores over the last millennium.Climate of the Past,14(1), 2137. https://doi.org/10.5194/

cp‐14‐21‐2018

Calonne, N., Cetti, C., Fierz, C., van Herwijnen, A., Jaggi, M., Löwe, H., et al. (2016). A unique time series of daily and weekly snowpack measurements at Weissfluhjoch, Davos, Switzerland. In International snow science workshop proceedings 2016. International snow science workshop, ISSW 2016, Breckenridge, CO, USA, October 27 (pp. 684689).

Cragin, J. H., Hewitt, A. D., & Colbeck, S. C. (1996). Grain‐scale mechanisms influencing the elution of ions from snow.Atmospheric Environment,30(1), 119127. https://doi.org/10.1016/13522310(95)00232N

Döscher, A., Gäggeler, H. W., Schotterer, U., & Schwikowski, M. (1995). A 130 years deposition record of sulfate, nitrate and chloride from a highalpine glacier.Water, Air, & Soil Pollution,85(2), 603609. https://doi.org/10.1007/BF00476895

Döscher, A., Gäggeler, H. W., Schotterer, U., & Schwikowski, M. (1996). A historical record of ammonium concentrations from a glacier in the Alps.Geophysical Research Letters,23(20), 27412744. https://doi.org/10.1029/96GL02615

Eichler, A., Schwikowski, M., Furger, M., Schotterer, U., & Gäggeler, H. W. (2004). Sources and distribution of trace species in Alpine precipitation inferred from two 60year ice core paleorecords.Atmospheric Chemistry and Physics Discussions,4(1), 71108. https://doi.

org/10.5194/acpd‐4‐71‐2004

Eichler, A., Schwikowski, M., & Gäggeler, H. W. (2000). An Alpine icecore record of anthropogenic HF and HCl emissions.Geophysical Research Letters,27(19), 3225–3228. https://doi.org/10.1029/2000GL012006

Eichler, A., Schwikowski, M., & Gäggeler, H. W. (2001). Meltwaterinduced relocation of chemical species in Alpinern.Tellus Series B:

Chemical and Physical Meteorology,53(2), 192–203. https://doi.org/10.3402/tellusb.v53i2.16575

Feibelman, P. J. (2007). Substitutional NaCl hydration in ice.Physical Review B: Condensed Matter and Materials Physics,75(21), 214113.

https://doi.org/10.1103/PhysRevB.75.214113

Fukazawa, H., Sugiyama, K., Mae, S., Narita, H., & Hondoh, T. (1998). Acid ions at triple junction of Antarctic ice observed by Raman scattering.Geophysical Research Letters,25(15), 2845–2848. https://doi.org/10.1029/98GL02178

Gabrieli, J., Carturan, L., Gabrielli, P., Kehrwald, N., Turetta, C., Cozzi, G., et al. (2011). Impact of Po Valley emissions on the highest glacier of the Eastern European Alps.Atmospheric Chemistry and Physics,11(15), 8087–8102. https://doi.org/10.5194/acp‐11‐

80872011 Acknowledgments

The authors would like to thank Philipp Baumann and Matthias Jaggi (both from SLF) for support during the sampling campaigns of 22 February and 1 June, respectively. The authors also acknowledge Franziska Aemisegger (Institute for Atmospheric and Climate Science, ETH Zurich) for performing back trajectory analyses and Martine Collaud (Swiss Federal Office of Meteorology and Climatology) for providing a list of Saharan dust event detected at the Jungfraujoch high‐

Alpine research station between November 2016 and February 2017.

Funding was provided by the Swiss National Science Foundation (SNSF) under Grant 155999. The corresponding data is available at the NOAA (National Oceanic and Atmospheric

Administration) data center for paleoclimate (ice core sites; https://

www.ncdc.noaa.gov/paleo/study/

26750). Finally, we greatly appreciate the comments and suggestions of the two anonymous referees and Dominic Winski, which helped to improve the clarity of the manuscript.

(10)

Gabrielli, P., Cozzi, G., Torcini, S., Cescon, P., & Barbante, C. (2008). Trace elements in winter snow of the Dolomites (Italy): A statistical study of natural and anthropogenic contributions.Chemosphere,72(10), 1504–1509. https://doi.org/10.1016/j.

chemosphere.2008.04.076

Ginot, P., Schotterer, U., Stichler, W., Godoi, M. A., Francou, B., & Schwikowski, M. (2010). Influence of the Tungurahua eruption on the ice core records of Chimborazo, Ecuador.The Cryosphere,4(4), 561568. https://doi.org/10.5194/tc45612010

Grannas, A. M., Bogdal, C., Hageman, K. J., Halsall, C., Harner, T., Hung, H., et al. (2013). The role of the global cryosphere in the fate of organic contaminants.Atmospheric Chemistry and Physics,13(6), 32713305. https://doi.org/10.5194/acp1332712013

Greaves, M. J., Statham, P. J., & Elderfield, H. (1994). Rare earth element mobilization from marine atmospheric dust into seawater.Marine Chemistry,46(3), 255260. https://doi.org/10.1016/03044203(94)900817

Greilinger, M., Schöner, W., Winiwarter, W., & Kasper‐Giebl, A. (2016). Temporal changes of inorganic ion deposition in the seasonal snow cover for the Austrian Alps (19832014).Atmospheric Environment,132, 141152. https://doi.org/10.1016/j.atmosenv.2016.02.040 Herreros, J., Moreno, I., Taupin, J. D., Ginot, P., Patris, N., de Angelis, M., et al. (2009). Environmental records from temperate glacier ice

on Nevado Coropuna saddle, southern Peru.Advances in Geosciences,22, 2734. https://doi.org/10.5194/adgeo22272009

Hiltbrunner, E., Schwikowski, M., & Körner, C. (2005). Inorganic nitrogen storage in alpine snow pack in the Central Alps (Switzerland).

Atmospheric Environment,39(12), 22492259. https://doi.org/10.1016/j.atmosenv.2004.12.037 Hobbs, P. V. (1974).Ice Physics. Oxford: Oxford University Press.

Hsu, S.C., Wong, G. T. F., Gong, G. C., Shiah, F. K., Huang, Y. T., Kao, S. J., et al. (2010). Sources, solubility, and dry deposition of aerosol trace elements over the East China Sea.Marine Chemistry,120(1‐4), 116–127. https://doi.org/10.1016/j.marchem.2008.10.003 Kang, S., Huang, J., & Xu, Y. (2008). Changes in ionic concentrations andδ18O in the snowpack of Zhadang glacier, Nyainqentanglha

mountain, southern Tibetan Plateau.Annals of Glaciology,49, 127–134. https://doi.org/10.3189/172756408787814708

Knüsel, S., Piguet, D. E., Schwikowski, M., & Gäggeler, H. W. (2003). Accuracy of continuous icecore traceelement analysis by inductively coupled plasma sectorfield mass spectrometry.Environmental Science & Technology,37(10), 2267–2273. https://doi.org/10.1021/

es026452o

Kuhn, M., Haslhofer, J., Nickus, U., & Schellander, H. (1998). Seasonal development of ion concentration in a high alpine snow pack.

Atmospheric Environment,32(23), 40414051. https://doi.org/10.1016/S13522310(97)002161

Kutuzov, S., Shahgedanova, M., Mikhalenko, V., Ginot, P., Lavrentiev, I., & Kemp, S. (2013). High‐resolution provenance of desert dust deposited on Mt. Elbrus, Caucasus in 20092012 using snow pit andrn core records.The Cryosphere,7(5), 14811498. https://doi.org/

10.5194/tc‐7‐1481‐2013

Lee, J., Nez, V. E., Feng, X., Kirchner, J. W., Osterhuber, R., & Renshaw, C. E. (2008). A study of solute redistribution and transport in seasonal snowpack using natural and artificial tracers.Journal of Hydrology, 243–254. https://doi.org/10.1016/j.

jhydrol.2008.05.004

Li, T., Wang, Y., Li, W. J., Chen, J. M., Wang, T., & Wang, W. X. (2015). Concentrations and solubility of trace elements infine particles at a mountain site, southern China: regional sources and cloud processing.Atmospheric Chemistry and Physics,15(15), 89879002. https://

doi.org/10.5194/acp‐15‐8987‐2015

Li, Z., Edwards, R., MosleyThompson, E., Wang, F., Dong, Z., You, X., et al. (2006). Seasonal variability of ionic concentrations in surface snow and elution processes in snow–firn packs at the PGPI site on Ürümqi glacier No. 1, eastern Tien Shan, China.Annals of Glaciology, 43, 250256. https://doi.org/10.3189/172756406781812069

Marty, C., & Meister, R. (2012). Long‐term snow and weather observations at Weissfluhjoch and its relation to other high‐altitude obser- vatories in the Alps.Theoretical and Applied Climatology,110(4), 573583. https://doi.org/10.1007/s0070401205843

MeteoSwiss. (2014). Saharan dust events. https://www.meteoswiss.admin.ch/home/climate/the‐climate‐of‐switzerland/specialties‐of‐the‐

swissclimate/saharandustevents.html (date of access: 07/05/2019)

Müller‐Tautges, C., Eichler, A., Schwikowski, M., Pezzatti, G. B., Conedera, M., & Hoffmann, T. (2016). Historic records of organic com- pounds from a high Alpine glacier: Inuences of biomass burning, anthropogenic emissions, and dust transport.Atmospheric Chemistry and Physics,16(2), 1029–1043. https://doi.org/10.5194/acp‐16‐1029‐2016

Mulvaney, R., Wolff, E. W., & Oates, K. (1988). Sulphuric acid at grain boundaries in Antarctic ice.Nature,331(6153), 247249. https://doi.

org/10.1038/331247a0

Nickus, U., Kuhn, M., Baltensperger, U., Delmas, R., Gäggeler, H. W., Kasper, A., et al. (1997). SNOSP: Ion deposition and concentration in high Alpine snow packs.Tellus Series B: Chemical and Physical Meteorology,49(1), 56–71. https://doi.org/10.1034/j.1600‐0889.49.

issue1.4.x

Pavlova, P. A., Jenk, T. M., Schmid, P., Bogdal, C., Steinlin, C., & Schwikowski, M. (2015). Polychlorinated biphenyls in a temperate Alpine glacier: 1. Effect of percolating meltwater on their distribution in glacier ice.Environmental Science & Technology,49(24), 1408514091.

https://doi.org/10.1021/acs.est.5b03303

Petrenko, V. F., & Whitworth, R. W. (2002).Physics of Ice. Oxford: Oxford University Press. https://doi.org/10.1093/acprof:oso/

9780198518945.001.0001

Pinzer, B. R., Schneebeli, M., & Kaempfer, T. U. (2012). Vaporux and recrystallization during dry snow metamorphism under a steady temperature gradient as observed by time‐lapse micro‐tomography.The Cryosphere,6(5), 1141–1155. https://doi.org/10.5194/tc‐6‐1141‐

2012

Preunkert, S., Legrand, M., & Wagenbach, D. (2001). Sulfate trends in a Col du Dôme (French Alps) ice core: A record of anthropogenic sulfate levels in the European midtroposphere over the twentieth century.Journal of Geophysical Research,106(D23), 31,99132,004.

https://doi.org/10.1029/2001JD000792

Preunkert, S., Wagenbach, D., & Legrand, M. (2003). A seasonally resolved alpine ice core record of nitrate: Comparison with anthropo- genic inventories and estimation of preindustrial emissions of NO in Europe.Journal of Geophysical Research,108(D21), 4681. https://

doi.org/10.1029/2003JD003475

Preunkert, S., Wagenbach, D., Legrand, M., & Vincent, C. (2000). Col du Dôme (Mt Blanc Massif, French Alps) suitability for ice‐core studies in relation with past atmospheric chemistry over Europe.Tellus Series B: Chemical and Physical Meteorology,52(3), 9931012.

https://doi.org/10.3402/tellusb.v52i3.17081

Schwikowski, M., Barbante, C., Doering, T., Gaeggeler, H. W., Boutron, C., Schotterer, U., et al. (2004). Post17thcentury changes of European lead emissions recorded in high‐altitude Alpine snow and ice.Environmental Science & Technology,38(4), 957–964. https://

doi.org/10.1021/es034715o

Schwikowski, M., Brütsch, S., Gäggeler, H. W., & Schotterer, U. (1999). A high‐resolution air chemistry record from an Alpine ice core: Fiescherhorn glacier, Swiss Alps.Journal of Geophysical Research,104(D11), 13,70913,719. https://doi.org/10.1029/

1998JD100112

Referenzen

ÄHNLICHE DOKUMENTE

How does the scale factor evolve in time, if the universe is dominated by dust, photons or vacuum

How does the scale factor evolve in time, if the universe is dominated by dust, photons or vacuum

61 The proposal was rejected by most of ASEAN member states for three main reasons. First, the multilateral defense cooperation would send a wrong signal to major powers. It

A charged region of the membrane insertion complex, present at the cytoplasmic face, could retain a charged region of the nascent chain adjacent to the

The aim of the EGYPT (Eddies and GYres Paths Tracking) programme is to study the circulation of the water masses in the eastern basin of the Mediterranean.. The surface circulation

GENETIC VARIABILITY, SHELL AND SPERM MORPHOLOGY SUGGEST THAT THE SURF CLAMS DONAX MARINCOVICHI ANDD. OBESULUS ARE

A RAFOS sound source array and acoustically-tracked floats in the Weddell Sea are already yielding valuable information on ocean circulation and structure beneath the sea ice

The need for business intelligence systems can be easily explained if we study the current economic context: to survive in today's competitive agricultural and economic conditions a