Hostname: page-component-848d4c4894-wg55d Total loading time: 0 Render date: 2024-06-09T18:20:38.505Z Has data issue: false hasContentIssue false

Effects of snow, snowmelting and refreezing processes on air–sea-ice CO2 flux

Published online by Cambridge University Press:  08 September 2017

Daiki Nomura
Affiliation:
Institute of Low Temperature Science, Hokkaido University, Sapporo 060-0819, Japan E-mail: nomura.daiki@nipr.ac.jp Graduate School of Environmental Science and Faculty of Environmental Earth Science, Hokkaido University, Sapporo 060-0810, Japan
Hisayuki Yoshikawa-Inoue
Affiliation:
Graduate School of Environmental Science and Faculty of Environmental Earth Science, Hokkaido University, Sapporo 060-0810, Japan
Takenobu Toyota
Affiliation:
Institute of Low Temperature Science, Hokkaido University, Sapporo 060-0819, Japan E-mail: nomura.daiki@nipr.ac.jp
Kunio Shirasawa
Affiliation:
Institute of Low Temperature Science, Hokkaido University, Sapporo 060-0819, Japan E-mail: nomura.daiki@nipr.ac.jp
Rights & Permissions [Opens in a new window]

Abstract

The air–sea-ice CO2 flux was measured in the ice-covered Saroma-ko, a lagoon on the northeastern coast of Hokkaido, Japan, using a chamber technique. The air–sea-ice CO2 flux ranged from −1.8 to +0.5 mg C m−2 h−1 (where negative values indicate a sink for atmospheric CO2). The partial pressure of CO2 (pCO2) in the brine of sea ice was substantially lower than that of the atmosphere, primarily because of the influence of the under-ice plume from the Saromabetsu river located in the southeastern part of the lagoon. This suggests that the brine had the ability to take up atmospheric CO2 into the sea ice. However, the snow deposited over the sea ice and the superimposed ice that formed from snowmelting and refreezing partially blocked CO2 diffusion, acting as an impermeable medium for CO2 transfer. Our results suggest that the air–sea-ice CO2 flux was dependent not only on the difference in pCO2 between the brine and the overlying air, but also on the status of the ice surface. These results provide the necessary evidence for evaluation of the gas exchange processes in ice-covered seas.

Type
Research Article
Copyright
Copyright © International Glaciological Society 2010

Introduction

Carbon dioxide (CO2) is the primary gas involved in the exchange of carbon between ocean and atmosphere. Because sea ice has been assumed to impede the gas exchange between the ocean and atmosphere, the ice-covered areas of the ocean have not been considered important to the air–sea CO2 flux and the global carbon budget (Reference Tison, Haas, Gowing, Sleewaegen and BernardTison and others, 2002 and references therein). However, research has shown that when the brine volume fraction is higher than 5% (Reference Golden, Ackley and LytleGolden and others, 1998), sea ice is a permeable medium for CO2 (Reference Gosink, Pearson and KelleyGosink and others, 1976; Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). In addition, the partial pressure of CO2 (pCO2) in the brine of sea ice becomes supersaturated or undersaturated with respect to the overlying air (e.g. Reference DelilleDelille, 2006). It is evident that sea ice can influence the exchange of CO2 with the atmosphere (Reference Semiletov, Makshtas, Akasofu and AndreasSemiletov and others, 2004; Reference DelilleDelille, 2006; Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006).

Sea ice covers ∼3–6% of the world’s oceans and persists for up to 6 months (Reference Comiso, Thomas and DieckmannComiso, 2003). During the ice-melting period, uptake of CO2 by sea ice, which contains brine with a low pCO2 compared with that of the atmosphere, has been reported in the Arctic and Antarctic Oceans (Reference Semiletov, Makshtas, Akasofu and AndreasSemiletov and others, 2004; Reference DelilleDelille, 2006; Reference Zemmelink, Delille, Tison, Hintsa, Houghton and DaceyZemmelink and others, 2006). In contrast, CO2 is emitted from sea ice to the overlying air when the brine pCO2 increases with increasing brine salinity (Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). These results suggest that the sea-ice formation/melting processes influence the annual global carbon budget (Reference Rysgaard, Glud, Sejr, Bendtsen and ChristensenRysgaard and others, 2007, Reference Rysgaard, Bendtsen, Pedersen, Ramløv and Glud2009).

During the entire period of sea-ice cover, various conditions occur on the sea-ice surface: rafting and ridging by dynamic ice-growth processes under turbulent conditions, accumulation of snow deposited over sea ice, snow-melting and refreezing (superimposed ice) and formation of melt ponds in the warmer months of spring and summer (e.g. Reference Perovich and LeppärantaPerovich, 1998). Measurements of the air–sea-ice CO2 flux under different sea-ice surface conditions will enable a more complete understanding of the air–sea-ice CO2 flux. However, there have been few such measurements (Reference DelilleDelille, 2006; Reference Zemmelink, Delille, Tison, Hintsa, Houghton and DaceyZemmelink and others, 2006). In particular, there are no detailed descriptions of the relationships between sea-ice surface conditions and air–sea-ice CO2 flux.

We examine how snow deposited over sea ice and the processes of snowmelting and refreezing at the ice surface affect the air–sea-ice CO2 flux. These are important processes to consider when estimating the air–sea-ice CO2 flux.

Materials and Methods

Observation sites

Field observations were carried out in the semi-closed system of Saroma-ko, a lagoon on the Okhotsk Sea coast of Hokkaido, Japan, from 18 February to 9 March 2006 and from 22 February to 4 March 2008 (Fig. 1; Table 1). The surface area of the lagoon is 149.2 km2 and the mean depth is 14.5 m. Sea ice usually covers almost the entire surface of the lagoon from early January through to early April, with large year-to-year variability (Reference Shirasawa and LeppärantaShirasawa and Leppäranta, 2003). A river-water plume develops under the sea ice in the lagoon as a result of increased discharge of river water during the ice-melt period (Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others, 2009). For this study, we chose six ice stations: one (station main) located ∼1.5 km offshore and five others (stations 1–5) forming a transect between the mouth of the Saromabetsu river and the second (east) inlet (Fig. 1; Table 1). For each station, we carried out air–sea-ice CO2 flux measurements, sampling of sea-ice cores, brine and under-ice water, and ice surface observations. Details of the sampling methods for sea-ice cores, brine and under-ice water, and analytical procedures for physical and chemical properties are given in the Appendix.

Fig. 1. Locations of sampling stations in Saroma-ko, Hokkaido, Japan: stations 1–5 (44°05.30′ −44°07.49′ N, 143°56.59′ −143°56.61′ E) and station main (44°07.18′ N, 143°57.34′ E).

Table 1. Sampling dates, locations, properties of snow cover and slush layer over sea ice, ice surface conditions, and sea-ice properties at Saroma-Ko

Air–sea-ice CO2 flux measurements

Air–sea-ice CO2 flux was measured by a closed dynamic chamber method (Fig. 2). This method was used to directly determine CO2 flux in situ to allow for high spatial resolution. We used a stainless-steel chamber 50 cm in diameter and 30 cm high with a serrated bottom edge to prevent gas from leaking through the chamber–ice interface. The chamber was pushed down firmly into the sea-ice surface to prevent lateral diffusion of CO2. Sample air from the chamber passed through Teflon tubes connected to a non-dispersive infrared (NDIR) analyzer (GMP343 carbon dioxide probe; Vaisala Ojy, Vantaa, Finland) that was connected to a data logger (RVR 52 data logger; T&D Corp., Nagano, Japan). Temperatures in the chamber were measured during each flux measurement with a temperature sensor (RVR 52 temperature recorder; T&D Corp., Nagano, Japan). The air in the chamber was circulated with a pump at a flow rate of 0.7 L min−1 through a mass flow controller and chemical desiccant column containing Mg(ClO4)2 (Fig. 2). These devices were installed in a portable adiabatic box (40 cm × 25 cm × 30 cm). Electricity was provided by an electric generator. The CO2 concentration in the chamber was measured every 10 s during experiments lasting 40 min. Four standards (typically 324, 341, 363 and 406 ppm CO2 in natural air) traceable to the World Meteorological Organization (WMO) mole fraction scale (Reference Yoshikawa-Inoue and IshiiYoshikawa-Inoue and Ishii, 2005) were used to calibrate the CO2-measuring system in the laboratory before the experiment in 2006 and both before and after the experiment in 2008. The instrument drift (output voltage) before and after the experiment for each calibration gas agreed to well within 1%. The CO2 flux was calculated from the changes in CO2 concentration in the chamber and the elapsed time (Fig. 3). The precision of the air–sea-ice CO2 flux from duplicate determination was within ±0.13 mg C m−2 h−1.

Fig. 2. Schematic diagram of the CO2-measuring system and stainless-steel chamber. Solid thick lines indicate the closed loop of flowing air for measuring the concentration of CO2 in the air in the chamber. CDC: chemical desiccant column; MFC: mass flow controller; NDIR: non-dispersive infrared analyser; DL: data logger; EG: electric.

Fig. 3. Example of the temporal variation in CO2 concentration in the chamber at station main on 29 February 2008 with the condition depicted in Figure 4b (neither snow nor slush). Open circles and crosses reflect the values obtained before and after the measurement, respectively. The regression line is based on the data obtained after the measurement. The vertical dashed line indicates an elapsed time of 0 min.

During the observation period in 2008, the CO2 flux was measured under two different conditions: (1) a chamber was installed above the sea ice through the snow cover and slush layer deposited over the sea ice (Fig. 4a); and (2) a chamber was installed directly above the sea ice after first removing the snow cover and slush layer with a scoop (Fig. 4b).

Fig. 4. Schematic diagrams showing two situations for measuring the air–sea-ice CO2 flux during the 2008 study period in Saromako. (a) The chamber was installed above the sea ice through the snow cover and slush layer deposited over the sea ice. (b) As in (a) but with snow and slush layer removed artificially.

Observations of snow cover, slush layer and sea-ice surface conditions

We examined the properties of the snow cover and slush layer deposited over sea ice and determined the condition of the ice surface by measuring the snow thickness and the thickness of the slush layer. We then classified the state of the sea-ice surface into three categories: (1) snow+slush; (2) superimposed ice; and (3) snow + superimposed ice. The temperatures of the air, the air–snow interface, the snow–slush-layer interface and the slush– or snow–sea-ice interface were measured using needle-type temperature sensors (SK-250WP; Sato Keiryoki Mfg. Co., Ltd, Tokyo, Japan).

Results and Discussion

Properties of snow cover, slush layer, sea ice and sea-ice surface

The properties of the snow cover, slush layer, sea ice and sea-ice surface conditions during our experiments are shown in Table 1. During the observation period in 2006, there was considerable variation in snow thickness and slush-layer thickness. On 18 and 19 February 2006, there were both snow and slush layers on top of the sea-ice surface (Table 1). However, as air temperature increased, the state of the sea-ice surface changed suddenly. From 23 February to 1 March, because of the melting of the snow and the refreezing of the meltwater, the surface of the sea ice was frequently covered by an icy glaze (herein referred to as ‘superimposed ice’) with neither a snow nor a slush layer (Fig. 5). On 3 March, there was >16 cm of snow on top of the superimposed ice. This loading of snow resulted in an upward flushing of water through brine channels in the ice (Reference Hudier, Ingram and ShirasawaHudier and others, 1995), and a slush layer had formed again at station 3 and station main on 3 March. From 4 to 9 March, the snow melted and the snow thickness decreased with time. During this period, there was no slush layer over the sea ice.

Fig. 5. Photograph of the superimposed ice at station 3 in Saromako on 28 February 2006. The photograph was taken looking west. The length of the standing cores is ∼40 cm.

During 2008 at station main, snow and slush layers were observed throughout the observation period (Table 1). Snow thickness and slush-layer thickness varied according to the amount of snow deposited over the sea ice and the rate of melting.

The sea-ice thickness was almost constant during each experimental period (Table 1). The ice thickness was 42 ± 5 cm (mean ± standard deviation) in 2006 and 45 ± 5 cm in 2008. The sea-ice temperature ranged from −3.1°C to −0.1°C in 2006 and from −2.8°C to −0.9°C in 2008 (Table 1). The temperature of sea ice, snow and slush layer increased overall through the experimental period each year (Table 1). These results suggest a transition from the cold period to the warm or melting period throughout the observation period. The brine volume fraction of sea ice was >5% at all stations (Table 1), indicating that the CO2 in brine can diffuse through the brine channel (Reference Gosink, Pearson and KelleyGosink and others, 1976) due to the high permeability of sea ice (Reference Golden, Ackley and LytleGolden and others, 1998).

Brine pCO2

The brine pCO2 is one of the important factors determining the magnitude of the air–sea-ice CO2 flux (Reference DelilleDelille, 2006; Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). Here we focus on the distributions of the brine pCO2 and discuss the mechanisms controlling it.

The brine pCO2 increased with increasing brine salinity, with a high degree of correlation (r 2 = 0.86; P < 0.05) (Fig. 6). During the observation period in 2006, both brine salinity and brine pCO2 spanned wide ranges: 2.3–29.3 for brine salinity and 2.7–195.1 μatm for brine pCO2. Plots of low brine salinity and low brine pCO2 corresponded to values observed near the river mouth (stations 1–3) during the entire period and through the transect during the warm period 1–9 March 2006. At station main during the 2006 observations, both brine pCO2 and brine salinity decreased drastically (Fig. 7). Brine salinity decreased from 38.2 on 19 February to 8.5 on 7 March. Brine pCO2 decreased from 195.1 μatm on 23 February to 8.6 μatm on 7 March. Near the river mouth (stations 1 and 3) during the 2006 observations, salinity and pCO2 of under-ice water at 1 m depth below the surface of the sea ice ranged from 3.3 to 18.8 and from 26.1 to 187.2 μatm, respectively. These values are low compared with those of under-ice water at 1 m depth below the surface of the sea ice at station 5 on 5 March 2006 (20.2 for salinity and 320.5 μatm for pCO2).

Fig. 6. Relationship between brine salinity and brine pCO2 during observations in 2006 and 2008. The dashed curve indicates the best fit to describe the relationship (y = 0.22x 1.88; r 2 = 0.86; P < 0.05).

Fig. 7. Changes in brine pCO2 and brine salinity at station main during observations in 2006.

During observations in 2008, brine salinity and brine pCO2 at station main stayed within relatively narrow ranges: 23.3–33.2 for brine salinity and 54.9–175.2 μatm for brine pCO2 (Fig. 6). In all experiments, brine pCO2 was under-saturated with respect to the overlying air, with an average value of 388.5 ± 6.1 μatm (mean ± standard deviation) for all observations.

In order to elucidate the process controlling the changes of brine pCO2, we examine the relationships between the salinity-normalized brine dissolved inorganic carbon (n-DIC) and salinity-normalized total alkalinity (n-TA) (Fig. 8). The samples with low salinity and low pCO2 for brine and under-ice water exhibit extremely high n-DIC and n-TA when dissolved inorganic carbon (DIC) and total alkalinity (TA) are normalized to a salinity of 35 (Fig. 8). The relationships between n-DIC and n-TA for brine and under-ice water show good agreement (r 2 = 0.94). The slope of the regression line for brine and under-ice water is 1.3, and plots of brine and under-ice water are slightly deviated above and below the regression line, respectively.

Fig. 8. Relationships between the n-DIC and n-TA (normalized to a constant salinity of 35) for brine and under-ice water during observations in 2006 and 2008. The regression line is based on the brine and under-ice water with slope of 1.3 (r 2 = 0.94). The inset indicates the theoretical slope of the different processes affecting n-DIC and n-TA.

In Saroma-ko, a salinity and chemical substances (nutrients) gradient has been observed moving offshore from the river mouth (Reference Shirasawa and LeppärantaShirasawa and Leppäranta, 2003; Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others, 2009). Based on the salinity–oxygen-isotopic ratio, the fraction of river water just below the sea ice was estimated to be >80% at stations 2–4 during the warm period of the experiment from 1 to 8 March 2006, whereas the fraction of ice meltwater just below the sea ice was almost zero through the transect during the entire period (Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others, 2009). These results suggest that the changes of n-DIC and n-TA for under-ice water shown in Figure 8 were mainly driven by the effect of the river-water plume developed under the sea ice.

Our study area was located in the southernmost part of the seasonal sea-ice zone in the Northern Hemisphere, where the sea ice is relatively thinner, ice temperature is higher and sea-ice brine volume fraction is higher than in polar seas (Table 1). Water exchange between the well-developed brine channels and under-ice water readily occurred through upward flushing caused by loading of snow deposited over the sea ice (Reference Hudier, Ingram and ShirasawaHudier and others, 1995; Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others, 2009). Therefore, it can be considered that the changes of n-DIC and n-TA for brine were caused by the exchange with under-ice water characterized by the river-water plume under the ice (Fig. 8). At first glance, the slope of the regression line for brine and under-ice water is also close to the theoretical slope for CaCO3 dissolution (slope = 2.0) (Fig. 8). However, because the high concentration of dissolved inorganic phosphorus in brine (Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others, 2009) and the high ice temperature (Table 1) (compared with the polar seas) impedes the formation of CaCO3 (Reference Killawee, Fairchild, Tison, Janssens and LorrainKillawee and others, 1998; Reference PapadimitriouPapadimitriou and others, 2005), we suggest that CaCO3 formation is minimal in sea ice. In addition, in order to demonstrate the wide distributions of n-DIC and n-TA for brine (>10 000 μmol kg−1) (Fig. 8), large quantities of CaCO3 are needed (Reference DelilleDelille, 2006). The slight difference between the plots for brine (above the regression line in Fig. 8) and under-ice water (below the regression line) might be attributed to other processes such as photosynthesis. Chlorophyll-a concentrations of brine were higher than in under-ice water during our observation period (Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others, 2009). This can lead to the photosynthetic activity of ice algae in brine and depletion of the brine pCO2 relative to the under-ice water.

Reference DelilleDelille (2006) reported that brine pCO2 changes dramatically from the season of sea-ice formation to the season of melting conditions in the Antarctic pack ice. During sea-ice formation, brine pCO2 increases with increasing DIC and with changes in CO2 solubility and the dissociation constants of carbonic acid (Reference Papadimitriou, Kennedy, Dieckmann and ThomasPapadimitriou and others, 2003; Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). Therefore, brine pCO2 is supersaturated with respect to the air. On the Antarctic pack ice, the brine pCO2 increases up to 900 μatm in late winter (Reference DelilleDelille, 2006). In contrast, during the ice-melting season, brine pCO2 becomes undersaturated with respect to the air as a result of the increased net primary production, dilution by melting ice and dissolution of carbonate minerals (Reference Gleitz, van der Loeff, Thomas, Dieckmann and MilleroGleitz and others, 1995; Reference DelilleDelille, 2006). In Arctic fast ice in spring, brine pCO2 is 220–280 μatm (Reference Semiletov, Makshtas, Akasofu and AndreasSemiletov and others, 2004), and in Antarctic pack ice in summer, brine pCO2 is about 50 μatm (Reference DelilleDelille, 2006).

Air–sea-ice CO2 flux

The values of the air–sea-ice CO2 flux obtained in this study ranged from −1.8 to +0.5 mg C m−2 h−1 (where negative values indicate a sink for atmospheric CO2) (Fig. 9). The values agree well with the results obtained over bare Antarctic pack ice (−2.6 to +1.0 mg C m−2 h−1; Reference DelilleDelille, 2006). However, more negative flux values (i.e. increased CO2 flux from air to sea or water) were observed over melt ponds and open brine channels in the Arctic fast ice (−19.3 to −9.8 mg C m−2 h−1; Reference Semiletov, Makshtas, Akasofu and AndreasSemiletov and others, 2004) and over snow and slush layers on Antarctic pack ice (−9.1 to −3.3 mg C m−2 h−1; Reference Zemmelink, Delille, Tison, Hintsa, Houghton and DaceyZemmelink and others, 2006), where there is a highly productive ice algae community.

Fig. 9. The air–sea-ice CO2 flux vs brine pCO2 for all experiments. The horizontal dotted line indicates an air–sea-ice CO2 flux of 0 mg C m−2 h−1.

Previous studies report that the air–sea-ice CO2 flux is mainly influenced by the difference between pCO2 of the brine in sea ice and that of the overlying air (Reference DelilleDelille, 2006; Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). We saw during our period of observations that the pCO2 of the overlying air was almost constant compared with that of the brine. Therefore, the air–sea-ice CO2 flux will depend primarily on the brine pCO2. However, in our data there was no clear relationship between the air–sea-ice CO2 flux and brine pCO2 when we plotted all the data corrected during the observation period (Fig. 9). These results suggest that there are other processes, in addition to the difference in pCO2 between the brine and overlying air, that influence the air–sea-ice CO2 flux. We now examine these other processes.

Relationships between air–sea-ice CO2 flux and ice surface properties

In order to clarify the process controlling the air–sea-ice CO2 flux, in addition to the difference of pCO2 between the brine and overlying air, we examine the relationships between the air–sea-ice CO2 flux and ice surface properties.

The air–sea-ice CO2 flux for each ice surface type as a function of the snow thickness is shown in Figure 10 and Table 2. The air–sea-ice CO2 flux approaches zero with increasing snow thickness for sea-ice surface conditions of snow + slush (Fig. 10). When the snow is deeper than about 9 cm, most of the air–sea-ice CO2 flux values are near zero (Fig. 10; Table 2). In contrast, when a chamber was installed above the sea ice after first removing the snow and slush layer (Fig. 4b), the negative air–sea-ice CO2 flux ranged from −1.8 to −0.6 mg C m−2 h−1 (Fig. 10; Table 2). Even when the snow thickness was less than about 9 cm, the air–sea-ice CO2 flux values were near zero under conditions where the sea-ice surface was classified as snow + superimposed ice or superimposed ice (Fig. 10; Table 2). These results suggest that a superimposed ice surface or a melted snow deposit deeper than 9 cm acts as a physical barrier to CO2 diffusion from air to sea ice in this area during the ice-melting season, even though the brine in the sea ice has the potential (i.e. relatively low pCO2) to take up CO2 from the atmosphere under all ice surface conditions (Fig. 6). These results were coincident with the limited measurement of air–sea-ice CO2 flux over the superimposed ice in the Antarctic pack ice (Reference DelilleDelille, 2006).

Fig. 10. Relationship between snow thickness and air–sea-ice CO2 flux for various ice surface conditions. Crosses reflect the conditions depicted in Figure 4b. The horizontal dotted line indicates an air–sea-ice CO2 flux of 0 mg C m−2 h−1.

Table 2. Air–sea-ice CO2 flux under different sea-ice surface conditions

We also examined the relationships between the air–seaice CO2 flux and the difference in pCO2 between the air and brine in sea ice (pCO2_air − pCO2 brine = Δ pCO2_air–brine) for the two situations depicted in Figure 4 (Fig. 11). We found a notable difference between the two situations. When the snow and slush layer were removed artificially (Fig. 4b), there was a clear relationship between the air–sea-ice CO2 flux and ΔpCO2_air–brine (r 2 = 0.59; P < 0.05), showing that the air–sea-ice CO2 flux was driven mainly by the difference in pCO2 between the brine and overlying air (Reference DelilleDelille, 2006; Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). However, when the air–sea-ice CO2 flux was measured over the snow cover (>9 cm) and slush layer (Fig. 4a) the flux was near zero.

Fig. 11. Relationships between air–sea-ice CO2 flux and the difference in pCO2 between the air and the brine in sea ice (ΔpCO2_air–brine = pCO2_air − pCO2_brine) for the two situations depicted in Figure 4 during observations in 2008. Open circles reflect the conditions depicted in Figure 4b. The solid line indicates the linear regression of CO2 flux vs ΔpCO2_air–brine for open circles (r 2 = 0.59; P < 0.05). The cross indicates the result obtained on 23 February with a snow thickness of 5.5 cm (Table 1). The horizontal dotted line indicates an air–sea-ice CO2 flux of 0 mg C m−2 h−1.

There are some related studies concerning the air–soil CO2 flux through snow deposited over soils (Reference Sommerfeld, Mosier and MusselmanSommerfeld and others, 1993, Reference Sommerfeld, Massman, Musselman and Mosier1996; Reference TakagiTakagi and others, 2005; Reference Schindlbacher, Zechmeister-Boltenstern, Glatzel and JandlSchindlbacher and others, 2007). These studies found that the snow cover deposited over soils is a porous medium for gases before snowmelting begins. While CO2 concentrations in soils drive the overall CO2 flux through the snow cover, temporal and spatial variability of the CO2 flux may be caused by specific snowpack properties (Reference Schindlbacher, Zechmeister-Boltenstern, Glatzel and JandlSchindlbacher and others, 2007). In particular, during the snowmelting season, because of the meltwater input to the pores in snow (Reference DenothDenoth, 1982), refreezing of meltwater and the formation of a specific ice layer in the snowpack (Reference Schindlbacher, Zechmeister-Boltenstern, Glatzel and JandlSchindlbacher and others, 2007), the air–soil CO2 flux through the snow cover is restricted (Reference Schindlbacher, Zechmeister-Boltenstern, Glatzel and JandlSchindlbacher and others, 2007). These same conditions could have occurred in the snow in our experiments because of the relatively high temperatures during our experimental periods (Table 1). Although we have no data regarding the detailed properties of the snow, our results provide direct evidence for the effect of the snow deposited over sea ice on the air–sea-ice CO2 flux during the ice-melting season. To understand the effects of snow properties on the air–sea-ice CO2 flux, further investigations are highly desirable to state the snow density, grain shape composition and moisture in more detail.

Conclusions

We examined effects of snow on sea ice, snowmelting and refreezing processes on air–sea-ice CO2 flux in the semi-closed lagoon system using a chamber technique. The air–sea-ice CO2 flux ranged from −1.8 to +0.5 mg C m−2 h−1 (where negative values indicate a sink for atmospheric CO2). The partial pressure of CO2 (pCO2) in the brine of sea ice was substantially lower than that of the atmosphere, primarily because of the influence of the under-ice plume from the Saromabetsu river located in the southeastern part of the lagoon. The formation of superimposed ice on sea ice and melted snow deposited over sea ice (>9 cm) acted as physical barriers to CO2 exchange between the air and sea ice even though the brine within the sea ice has the potential to absorb CO2 from the atmosphere. Our results suggest that the air–sea-ice CO2 flux is dependent not only on the difference in pCO2 between the brine and the overlying air but also on the condition of the sea-ice surface, especially during the ice-melting season, indicating that measurements of the air–sea-ice CO2 flux should also include detailed observations of the sea-ice surface conditions.

Clearly, further investigations will be required to understand the effects of snow properties on the air–sea-ice CO2 flux and the applicability of the limited results from this study to the general field (e.g. Arctic and Antarctic seas), where the variability in ice surface snow- and ice-melting processes must be an important factor affected by global warming. However, these data provide clues to understanding the evaluation of the gas exchange processes in the ice-covered seas. Indeed, in the Arctic Ocean, formation of superimposed ice early in the melt season is common and has been linked to retention of meltwater at the ice surface in the early stage of melt (Reference Eicken, Grenfell, Perovich, Richter-Menge and FreyEicken and others, 2004). The overall role of sea ice in the climate system remains to be investigated in other seasonally ice-covered seas (Reference Rysgaard, Glud, Sejr, Bendtsen and ChristensenRysgaard and others, 2007, Reference Rysgaard, Bendtsen, Pedersen, Ramløv and Glud2009).

Acknowledgements

We thank S. Noriki, H. Hattori, S. Taguchi, S. Aoki, S. Nihashi, Y. W. Watanabe, K. Suzuki, T. Kawamura, M. Ishikawa, T. Takatsuka and Y. Kusumi for sampling assistance, for providing measurement devices and for useful comments. We also thank M. Lange, B. Delille and an anonymous reviewer for useful comments and editorial assistance. This work is partly supported by the Research Fellowship of the Japan Society for the Promotion of Science (#17654098 and #195968).

Appendix

Methods for sampling sea-ice cores, brine and under-ice water

Sea-ice cores were collected with a SIPRE-type (US Army Snow, Ice and Permafrost Research Establishment) ice corer with an inner diameter of 7.5 cm. Immediately after ice cores were collected the ice-core length was measured using a ruler, and temperatures were measured by inserting a needle-type temperature sensor (SK-250WP; Sato Keiryoki Mfg. Co., Ltd, Tokyo, Japan) into holes drilled into the core. The ice cores were placed in a polythene bag and stored horizontally in a cooler along with snow to maintain low temperatures and to minimize brine drainage from sea-ice cores. Thereafter, the ice cores were transported to the Saroma Research Center for Aquaculture and stored in a freezer at −30°C. The ice cores were later analysed to determine ice bulk salinity and to estimate the brine volume fraction of the sea ice.

Samples of the sea-ice brine were collected by the gravity drainage method. Full details of this technique are presented by Reference Nomura, Takatsuka, Ishikawa, Kawamura, Shirasawa and Yoshikawa-InoueNomura and others (2009). The brine collected into 200 mL sample bottles was divided into subsamples: one in a 10 mL glass vial for measuring salinity and one in a 120 mL glass vial for measuring brine DIC and brine TA. Mercuric chloride (HgCl2; 50 μL) was added to the DIC sample to stop biological activity.

Under-ice water samples were collected through the ice-core holes using a 500 mL Teflon water sampler (GL Science Inc., Tokyo, Japan) at 1 m depth below the surface of the sea ice. These water samples were collected approximately 15 min after drilling the ice cores, in order to avoid any effects from the disturbance caused by drilling. The under-ice water samples were stored in the same manner as the brine samples for further analysis.

Sample analysis

Salinity of the brine and under-ice water samples was measured using a salt analyser (SAT-210; Toa Electronics Ltd, Tokyo, Japan). The standard deviation for salinity calculated from 15 subsamples taken from a reference water sample (S=10.00) was 0.03 (Reference Nomura, Yoshikawa-Inoue and ToyotaNomura and others, 2006). DIC was determined by coulometry (Reference Johnson, King and SieburthJohnson and others, 1985). DIC measurements were calibrated with working sea-water standards traceable to the Certified Reference Material distributed by A.G. Dickson (Scripps Institution of Oceanography, La Jolla, California, USA). The precision of DIC analysis from duplicate determinations was within ±0.1% (Reference Wakita, Watanabe, Watanabe, Ono, Tsurushima and TsunogaiWakita and others, 2005). TA was analysed by the improved single-point titration method (Reference Culberson, Pytkowicz and HawleyCulberson and others, 1970). We measured TA using a glass electrode calibrated with Tris buffer and 2-aminopyridine buffer (Reference Dickson and GoyetDickson and Goyet, 1994). The precision of the TA analysis from duplicate determinations was within ±0.2% (Reference Wakita, Watanabe, Watanabe, Ono, Tsurushima and TsunogaiWakita and others, 2005). The pCO2 of brine and under-ice water were computed from DIC and TA using the computer program CO2SYS, version 01.05 (Reference Lewis and WallaceLewis and Wallace, 1998). We used the carbonate dissociation constants (K 1 and K 2) of Reference Mehrbach, Culberson, Hawley and PytkowiczMehrbach and others (1973) as refit by Reference Dickson and MilleroDickson and Millero (1987), and the K SO4 determined by Reference DicksonDickson (1990). Air pCO2 was calculated from air CO2 concentration by assuming a barometric pressure of 1 atm and saturated water vapour of brine (Reference Weiss and PriceWeiss and Price, 1980).

For estimation of the brine volume fraction, the ice-core sample was cut to a 4.5 cm × 2.5 cm rectangular cross-section with a bandsaw and then sliced into 3 cm thick sections in the cold room at −15°C. The bulk volume and mass of the 3 cm thick sea-ice sections were measured to calculate the density of the sea ice. The sea-ice sections were then melted to measure salinity, and finally the brine volume fraction of sea ice was calculated with the equations of Reference Cox and WeeksCox and Weeks (1983) and Reference Eicken, Thomas and DieckmannEicken (2003). The brine volume fraction can be derived as:

(A1)

where V b is the brine volume, Vis the sea-ice bulk volume, p is the sea-ice bulk density and S i is the salinity after melting. In a temperature range higher than −23°C, brine salinity, S b, and density, ρ b (g cm−3), as a function of sea-ice temperature and salinity can be approximated by the following:

(A2)

(A3)

where T is the temperature (°C).

References

Comiso, J.C. 2003. Large-scale characteristics and variability of the global sea ice cover. In Thomas, D.N. and Dieckmann, G.S., eds. Sea ice: an introduction to its physics, chemistry, biology and geology. Oxford, etc., Blackwell, 112142.CrossRefGoogle Scholar
Cox, G.F.N. and Weeks, W.F.. 1983. Equations for determining the gas and brine volumes in sea-ice samples. J. Glaciol., 29(102), 306316.Google Scholar
Culberson, C., Pytkowicz, R.M. and Hawley, J.E.. 1970. Seawater alkalinity determination by the pH method. J. Mar. Res., 28(1), 1521.Google Scholar
Delille, B. 2006. Inorganic carbon dynamics and air–ice–sea CO2 fluxes in the open and coastal waters of the Southern Ocean. (PhD thesis, University of Liège.)Google Scholar
Denoth, A. 1982. The pendular–funicular liquid transition and snow metamorphism. J. Glaciol., 28(99), 357364.Google Scholar
Dickson, A.G. 1990. Thermodynamics of the dissociation of boric acid in synthetic seawater from 273.15 to 318.15 K. Deep-Sea Res. I, 37(5), 755766.CrossRefGoogle Scholar
Dickson, A.G. and Goyet, C., eds. 1994. Handbook of methods for the analysis of the various parameters of the carbon dioxide system in sea water, version 2.0. Oak Ridge, TN, US Department of Energy, Oak Ridge National Laboratory. (ORNL/CDIAC-74.)CrossRefGoogle Scholar
Dickson, A.G. and Millero, F.J.. 1987. A comparison of the equilibrium constants for the dissociation of carbonic acid in seawater media. Deep-Sea Res. I, 34(10), 17331743.Google Scholar
Eicken, H. 2003. From the microscopic, to the macroscopic, to the regional scale: growth, microstructure and properties of sea ice. In Thomas, D.N. and Dieckmann, G.S., eds. Sea ice: an introduction to its physics, chemistry, biology and geology. Oxford, Blackwel, 2281.Google Scholar
Eicken, H., Grenfell, T.C., Perovich, D.K., Richter-Menge, J.A. and Frey, K.. 2004. Hydraulic controls of summer Arctic pack ice albedo. J. Geophys. Res., 109(C8), C08007. (10.1029/2003JC001989.)Google Scholar
Gleitz, M., van der Loeff, M.M.R., Thomas, D.N., Dieckmann, G.S. and Millero, F.J.. 1995. Comparison of summer and winter inorganic carbon, oxygen and nutrient concentrations in Antarctic sea ice brine. Mar. Chem., 51(2), 8191.CrossRefGoogle Scholar
Golden, K.M., Ackley, S.F. and Lytle, V.I.. 1998. The percolation phase transition in sea ice. Science, 282(5397), 22382241.Google Scholar
Gosink, T.A., Pearson, J.G. and Kelley, J.J.. 1976. Gas movement through sea ice. Nature, 263(5572), 4142.Google Scholar
Hudier, E.J.J., Ingram, R.G. and Shirasawa, K.. 1995. Upward flushing of sea water through first year sea ice. Atmos.–Ocean, 33(3), 569580.Google Scholar
Johnson, K.M., King, A.E. and Sieburth, J.M.. 1985. Coulometric TCO2 analyses for marine studies; an introduction. Mar. Chem., 16(1), 6182.Google Scholar
Killawee, J.A., Fairchild, I.J., Tison, J.L., Janssens, L. and Lorrain, R.. 1998. Segregation of solutes and gases in experimental freezing of dilute solutions: implications for natural glacial systems. Geochim. Cosmochim. Acta, 62(23–24), 36373655.Google Scholar
Lewis, E. and Wallace, D.W.R.. 1998. Program developed for CO2 system calculations. Oak Ridge, TN, US Department of Energy. Oak Ridge National Laboratory. (ORNL/CDIAC-105.)CrossRefGoogle Scholar
Mehrbach, C., Culberson, C.H., Hawley, J.E. and Pytkowicz, R.M.. 1973. Measurement of the apparent dissociation constants of carbonic acid in seawater at atmospheric pressure. Limnol. Oceanogr., 18(6), 897907.CrossRefGoogle Scholar
Nomura, D., Yoshikawa-Inoue, H. and Toyota, T.. 2006. The effect of sea-ice growth on air–sea CO2 flux in a tank experiment. Tellus B, 58(5), 418426.CrossRefGoogle Scholar
Nomura, D., Takatsuka, T., Ishikawa, M., Kawamura, T., Shirasawa, K. and Yoshikawa-Inoue, H.. 2009. Transport of chemical components in sea ice and under-ice water during melting in the seasonally ice-covered Saroma-ko Lagoon, Hokkaido, Japan. Estuar. Coast. Shelf Sci., 81(2), 201209.CrossRefGoogle Scholar
Papadimitriou, S., Kennedy, H., Dieckmann, G.S. and Thomas, D.N.. 2003. Experimental evidence for carbonate precipitation and CO2 degassing during sea ice formation. Geochim. Cosmochim. Acta, 68(8), 17491761.Google Scholar
Papadimitriou, S. and 6 others. 2005. Biogeochemical composition of natural sea ice brines from the Weddell Sea during early austral summer. Limnol. Oceanogr., 52(5), 18091823.Google Scholar
Perovich, D.K. 1998. The optical properties of sea ice. In Leppäranta, M., ed. Physics of ice-covered seas. Helsinki, University of Helsinki Press, 195230.Google Scholar
Rysgaard, S., Glud, R.N., Sejr, M.K., Bendtsen, J. and Christensen, P.B.. 2007. Inorganic carbon transport during sea ice growth and decay: a carbon pump in polar seas. J. Geophys. Res., 112(C3), C03016. (10.1029/2006JC003572.)Google Scholar
Rysgaard, S., Bendtsen, J., Pedersen, L.T., Ramløv, H. and Glud, R.N.. 2009. Increased CO uptake due to sea ice growth and decay in the Nordic Seas. J. Geophys. Res., 114(C9), C09011. (10.1029/2008JC005088.)Google Scholar
Schindlbacher, A., Zechmeister-Boltenstern, S., Glatzel, G. and Jandl, R.. 2007. Winter soil respiration from an Austrian mountain forest. Agric. Forest Meteorol., 146(3–4), 205215.Google Scholar
Semiletov, I., Makshtas, A., Akasofu, S.-I. and Andreas, E.L.. 2004. Atmospheric CO balance: the role of Arctic sea ice. Geophys. Res. Lett., 31(5), L05121. (10.1029/2003GL017996.)Google Scholar
Shirasawa, K. and Leppäranta, M.. 2003. Hydrometeorological and sea ice conditions in Saroma-ko lagoon, Hokkaido, Japan. Rep. Ser. Geophys. 46, 161168.Google Scholar
Sommerfeld, R.A., Mosier, A.R. and Musselman, R.C.. 1993. CO2, CH4 and N2O flux through a Wyoming snowpack and implications for global budgets. Nature, 361(6408), 140142.Google Scholar
Sommerfeld, R.A., Massman, W.J., Musselman, R.C. and Mosier, A.R.. 1996. Diffusional flux of CO2 through snow: spatial and temporal variability among Alpine–subalpine sites. Global Biogeochem. Cycles, 10(3), 473482.Google Scholar
Takagi, K. and 8 others. 2005. Dynamic carbon dioxide exchange through snowpack by wind-driven mass transfer in a conifer–broadleaf mixed forest in northernmost Japan. Global Biogeochem. Cycles, 19(GB2), GB2012. (10.1029/2004GB002272.)CrossRefGoogle Scholar
Tison, J.L., Haas, C., Gowing, M.M., Sleewaegen, S. and Bernard, A.. 2002. Tank study of physico-chemical controls on gas content and composition during growth of young sea ice. J. Glaciol., 48(161), 177191.Google Scholar
Wakita, M., Watanabe, S., Watanabe, Y.W., Ono, T., Tsurushima, N. and Tsunogai, S.. 2005. Temporal change of dissolved inorganic carbon in the subsurface water at Station KNOT (44°N, 155°E) in the Western North Pacific subpolar region. J. Oceanogr., 61(1), 129139.Google Scholar
Weiss, R.F. and Price, B.A.. 1980. Nitrous oxide solubility in water and seawater. Mar. Chem., 8(4), 347359.Google Scholar
Yoshikawa-Inoue, H. and Ishii, M.. 2005. Variations and trends of CO2 in the surface seawater in the Southern Ocean south of Australia between 1969 and 2002. Tellus B, 57(1), 5869.Google Scholar
Zemmelink, H.J., Delille, B., Tison, J.L., Hintsa, E.J., Houghton, L. and Dacey, J.W.H.. 2006. CO2 deposition over the multi-year ice of the western Weddell Sea. Geophys. Res. Lett., 33(13), L13606. (10.1029/2006GL026320.)Google Scholar
Figure 0

Fig. 1. Locations of sampling stations in Saroma-ko, Hokkaido, Japan: stations 1–5 (44°05.30′ −44°07.49′ N, 143°56.59′ −143°56.61′ E) and station main (44°07.18′ N, 143°57.34′ E).

Figure 1

Table 1. Sampling dates, locations, properties of snow cover and slush layer over sea ice, ice surface conditions, and sea-ice properties at Saroma-Ko

Figure 2

Fig. 2. Schematic diagram of the CO2-measuring system and stainless-steel chamber. Solid thick lines indicate the closed loop of flowing air for measuring the concentration of CO2 in the air in the chamber. CDC: chemical desiccant column; MFC: mass flow controller; NDIR: non-dispersive infrared analyser; DL: data logger; EG: electric.

Figure 3

Fig. 3. Example of the temporal variation in CO2 concentration in the chamber at station main on 29 February 2008 with the condition depicted in Figure 4b (neither snow nor slush). Open circles and crosses reflect the values obtained before and after the measurement, respectively. The regression line is based on the data obtained after the measurement. The vertical dashed line indicates an elapsed time of 0 min.

Figure 4

Fig. 4. Schematic diagrams showing two situations for measuring the air–sea-ice CO2 flux during the 2008 study period in Saromako. (a) The chamber was installed above the sea ice through the snow cover and slush layer deposited over the sea ice. (b) As in (a) but with snow and slush layer removed artificially.

Figure 5

Fig. 5. Photograph of the superimposed ice at station 3 in Saromako on 28 February 2006. The photograph was taken looking west. The length of the standing cores is ∼40 cm.

Figure 6

Fig. 6. Relationship between brine salinity and brine pCO2 during observations in 2006 and 2008. The dashed curve indicates the best fit to describe the relationship (y = 0.22x1.88; r2 = 0.86; P < 0.05).

Figure 7

Fig. 7. Changes in brine pCO2 and brine salinity at station main during observations in 2006.

Figure 8

Fig. 8. Relationships between the n-DIC and n-TA (normalized to a constant salinity of 35) for brine and under-ice water during observations in 2006 and 2008. The regression line is based on the brine and under-ice water with slope of 1.3 (r2 = 0.94). The inset indicates the theoretical slope of the different processes affecting n-DIC and n-TA.

Figure 9

Fig. 9. The air–sea-ice CO2 flux vs brine pCO2 for all experiments. The horizontal dotted line indicates an air–sea-ice CO2 flux of 0 mg C m−2 h−1.

Figure 10

Fig. 10. Relationship between snow thickness and air–sea-ice CO2 flux for various ice surface conditions. Crosses reflect the conditions depicted in Figure 4b. The horizontal dotted line indicates an air–sea-ice CO2 flux of 0 mg C m−2 h−1.

Figure 11

Table 2. Air–sea-ice CO2 flux under different sea-ice surface conditions

Figure 12

Fig. 11. Relationships between air–sea-ice CO2 flux and the difference in pCO2 between the air and the brine in sea ice (ΔpCO2_air–brine = pCO2_air − pCO2_brine) for the two situations depicted in Figure 4 during observations in 2008. Open circles reflect the conditions depicted in Figure 4b. The solid line indicates the linear regression of CO2 flux vs ΔpCO2_air–brine for open circles (r2 = 0.59; P < 0.05). The cross indicates the result obtained on 23 February with a snow thickness of 5.5 cm (Table 1). The horizontal dotted line indicates an air–sea-ice CO2 flux of 0 mg C m−2 h−1.