Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-19T12:46:19.827Z Has data issue: false hasContentIssue false

Impact of increasing Antarctic ice-shelf melting on Southern Ocean hydrography

Published online by Cambridge University Press:  08 September 2017

Caixin Wang
Affiliation:
Department of Physics, University of Helsinki, Helsinki, Finland. E-mail: caixin.wang@npolar.no Norwegian Polar Institute, Tromsø, Norway
Keguang Wang
Affiliation:
Norwegian Meteorological Institute, Tromsø, Norway
Rights & Permissions [Opens in a new window]

Abstract

Southern Ocean hydrography has undergone substantial changes in recent decades, concurrent with an increase in the rate of Antarctic ice-shelf melting (AISM). We investigate the impact of increasing AISM on hydrography through a twin numerical experiment, with and without AISM, using a global coupled sea-ice/ocean climate model. The difference between these simulations gives a qualitative understanding of the impact of increasing AISM on hydrography. It is found that increasing AISM tends to freshen the surface water, warm the intermediate and deep waters, and freshen and warm the bottom water in the Southern Ocean. Such effects are consistent with the recent observed trends, suggesting that increasing AISM is likely a significant contributor to the changes in the Southern Ocean. Our analyses indicate potential positive feedback between hydrography and AISM that would amplify the effect on both Southern Ocean hydrography and Antarctic ice-shelf loss caused by external factors such as changing Southern Hemisphere winds.

Type
Research Article
Copyright
Copyright © International Glaciological Society 2012

1. Introduction

The Southern Ocean has undergone significant changes in recent decades (see review by Reference JacobsJacobs, 2006): for example, rising temperature in the upper 3000 m (Reference Levitus, Antonov, Boyer and StephensLevitus and others, 2000, Reference Levitus, Antonov and Boyer2005; Reference GilleGille, 2002, 2003), and decreasing salinity in high-latitude waters (Reference Jacobs, Giulivi and MeleJacobs and others, 2002; Reference WhitworthWhitworth, 2002; Reference Boyer, Levitus, Antonov, Locarnini and GarciaBoyer and others, 2005). These changes have been interpreted as being a result of natural variability, anthropogenic effects or both (Reference Levitus, Antonov, Boyer and StephensLevitus and others, 2000). The surface freshening in the Ross Sea has also been linked to the increased precipitation, reduced sea-ice production and increased melting of the West Antarctic ice sheet (Reference Jacobs, Giulivi and MeleJacobs and others, 2002).

In recent decades, Antarctic ice-shelf melting (AISM) has also been observed to be increasing, especially in the Amundsen and Bellingshausen Seas (Reference Shepherd, Wingham and RignotShepherd and others, 2004; Reference Holland, Jenkins and HollandHolland and others, 2010; Reference Jacobs, Jenkins, Giulivi and DutrieuxJacobs and others, 2011; Reference Pritchard, Ligtenberg, Fricker, Vaughan, Van den Broeke and PadmanPritchard and others, 2012). Such increases have been attributed mainly to oceanic forcing (Reference Shepherd, Wingham and RignotShepherd and others, 2004; Reference Walker, Brandon, Jenkins, Allen, Dowdeswell and EvansWalker and others, 2007; Reference Thoma, Jenkins, Holland and JacobsThoma and others, 2008; Reference Wåhlin, Yuan, Björk and NohrWåhlin and others, 2010).

While the interactions between the Southern Ocean and the Antarctic ice shelves are complex, our focus in this study is to investigate the impact of increasing AISM on Southern Ocean hydrography.

2. Method

The ORCA2-LIM model used in this study is a global coupled sea-ice/ocean model (Reference Timmermann, Goosse, Madec, Fichefet, Ethe and DulièreTimmermann and others, 2005). It consists of an ocean model, ORCA, based on a global configuration of OPA (Ocean PArallelise), which is a free sea surface, primitive-equations ocean model (Reference Madec, Delecluse, Imbard and LévyMadec and others, 1998; http://www.lodyc.jussieu.fr/opa/), and a dynamic-thermodynamic sea-ice model, LIM (Louvain-la-Neuve sea-Ice Model; Reference Fichefet and MaquedaFichefet and Morales Maqueda, 1997), which employs a viscous-plastic rheology for sea-ice dynamics following Reference HiblerHibler (1979) and a three-layer snow/ice thermodynamic model following Reference SemtnerSemtner (1976). The model has a mean resolution of 2° in the horizontal, and 31 vertical layers in the ocean model with grid spacing from 10m in the top 100m to 500m at the bottom, and 1 (for dynamics) or 3 (for thermodynamics) vertical layers in the sea-ice model. Its model domain extends to about 78° S in the south, not including any of the Antarctic ice shelves (Reference WangWang, 2011).

Antarctic ice shelves fringe nearly half of the Antarctic coastline (Reference DrewryDrewry, 1983). Ice-shelf melting (ISM) takes place at the base of these shelves (Reference Jacobs, Hellmer, Doake, Jenkins and FrolichJacobs and others, 1992). The thermodynamic interaction between ice shelf and ocean has been parameterized by Reference Beckmann and GoosseBeckmann and Goosse (2003). They assumed that the net heat, Qnet, and freshwater flux, mnet, are proportional to the difference between the oceanic temperature outside the cavity, Tocean, and in situ freezing-point temperature at the base of the ice shelf, Tf, and an effective area of interaction, Aeff:

(1)

(2)

where ρ w and ρ is are the densities of sea water and ice shelf, cpw is the specific heat, γT (= 10-4ms-1) is the transfer coefficient and Lis is the latent heat of fusion of the ice shelf. The ISM parameterization scheme of Reference Beckmann and GoosseBeckmann and Goosse (2003) has been employed in the ORCA2-LIM model by Reference Wang and BeckmannWang and Beckmann (2007) through modification of lateral boundary conditions where the Antarctic ice shelves are located (Fig. 1). Reference Wang and BeckmannWang and Beckmann (2007) showed that the results with the parameterization are comparable to those from regional models with sub-ice-shelf cavities explicitly included.

Fig. 1. The name, extent and position of the Antarctic ice shelves included in the ORCA2-LIM-ISP model. EW: Riiser-Larsen and Brunt Ice Shelves.

Two experiments, one with and the other without AISM, were run for two passes forced with the same CORE (Coordinated Ocean-ice Reference Experiments) forcing data for 1958–2000. Using the same 43 year forcing data, the first pass was set to spin up the model state, and the second was for analysis. Reference Wang and BeckmannWang and Beckmann (2007) had reported changes in temperature and salinity at 216 m depth after inclusion of AISM. In this study, we further analyze these twin experiments by looking at hydrographic variability at different depths.

The simulated total annual mean circumpolar freshwater flux from AISM (Fig. 2) is ~106 mSv (1 mSv= 103 m3 s-1) in this study, with a range from ~102 mSv in the mid-1960s to ~112 mSv in the 1980s and 1990s (a total range of ~10% of the mean). These values are a factor of 3–4 higher than previous estimates (26–32 mSv; see Section 3.1). Our two simulations, with zero AISM and mean AISM of ~106mSv, therefore represent at least an order of magnitude more variability than in the actual record of AISM. Since we do not expect Southern Ocean hydrography to respond linearly to changes in AISM over such a broad range, the sensitivity to AISM estimated from these runs is only a qualitative guide to true sensitivity over realistic ranges of variability. Nevertheless, the comparison provides valuable qualitative information on how the Southern Ocean responds to changing AISM, and the model trends can be compared with observations during the last four decades.

Fig. 2. Time series of the simulated monthly (black) and annual (red) mean of the total circumpolar freshwater flux from AISM.

3. Results

3.1. Freshwater flux from AISM

It is difficult to estimate the circumpolar freshwater flux from AISM. Our knowledge of the melt rate of ice shelves is mostly from estimations based on glaciological and oceanographic observations or numerical simulations, since direct measurements are still very limited due to difficulties in accessing the sub-ice-shelf cavity. However, measurements or numerical simulations have not been done for all Antarctic ice shelves. Different approaches may yield large differences in estimated ISM (e.g. due to the assumption of a steady-state ice sheet (Reference Shepherd, Wingham and RignotShepherd and others, 2004) and the selection of outflow depth (Reference Payne, Holland, Shepherd, Rutt, Jenkins and JoughinPayne and others, 2007)).

The simulated total annual mean circumpolar freshwater flux from AISM, 106 mSv, is considerably larger than previous estimates (e.g. 26.14 mSv (Reference Jacobs, Hellmer and JenkinsJacobs and others, 1996), 31.6 mSv (Reference Beckmann, Hellmer and TimmermannBeckmann and others, 1999) and 28.42 mSv (Reference HellmerHellmer, 2004)). However, closer inspection of the individual ISM values reveals that many of these estimates are comparable to other studies, particularly for the major ice shelves (Table 1). For example, the freshwater flux from the Filchner-Ronne Ice Shelf (FRIS) is 6.52 mSv in this study, in good agreement with an earlier glaciological shelf-wide estimate (Reference Jacobs, Hellmer and JenkinsJacobs and others, 1996) and the numerical results from cavity-resolving models (e.g. Reference Timmermann, Hellmer and BeckmannTimmermann and others, 2002). The freshwater fluxes from the other large ice shelves (e.g. Ross (8.36mSv) and Amery (5.44mSv), as well as eastern Weddell (4.89 mSv)) in this study are also comparable to the other studies from regional cavity-resolving models (e.g. Reference Beckmann, Hellmer and TimmermannBeckmann and others, 1999; Reference Timmermann, Hellmer and BeckmannTimmermann and others, 2002; Reference HellmerHellmer, 2004). The freshwater flux from the Fimbul Ice Shelf is ~19.80 mSv in this study, overestimated by about a factor of two compared with other studies (e.g. Reference Beckmann, Hellmer and TimmermannBeckmann and others, 1999; Reference Beckmann and TimmermannBeckmann and Timmermann, 2001; Reference Timmermann, Hellmer and BeckmannTimmermann and others, 2002; Reference HellmerHellmer, 2004; Reference Walkden, Heywood, Nicholls and AbrahamsenWalkden and others, 2009), due to the warm bias in our simulation because of the coarse model resolution (Reference WangWang, 2011). Recent in situ observations indicate that the water temperature under the Fimbul Ice Shelf is generally <0.1 °C above the surface freezing point (Reference Hattermann, NØst, Lilly and SmedsrudHattermann and others, 2012), suggesting that basal melting here cannot simply be parameterized by coastal deep ocean temperatures. The modeled freshwater flux from the Getz Ice Shelf (16.4 mSv) is more than twice that estimated from glaciological measurements (Reference Shepherd, Wingham and RignotShepherd and others, 2004). The reason for this discrepancy requires further investigation. The high melt rate in the Amundsen and Bellingshausen Seas is generally consistent with available estimates. For example, the freshwater flux from the George VI Ice Shelf is 5.35 mSv in this study, corresponding to a melt rate of 3 m a-1, which is close to recent measurements of 3–5 m a-1 (Reference Jenkins and JacobsJenkins and Jacobs, 2008) and 2.49 m a-1 (Reference Holland, Jenkins and HollandHolland and others, 2010), although higher than previous estimates of 1.9 m a-1 (Reference Jacobs, Hellmer and JenkinsJacobs and others, 1996) and 0.43 m a-1 (Reference HellmerHellmer, 2004). Pine Island Glacier Ice Shelf (PIG) is experiencing extremely high melt rate due to the warm circumpolar deep water (CDW) flowing onto the continental shelf. Estimates of melt rates for PIG are in a wide range, from 10 m a-1 (Reference Jacobs, Hellmer and JenkinsJacobs and others, 1996) to 24 ± 4ma-1 (Reference RignotRignot, 1998). The freshwater flux from our simulation is 1.62 mSv, corresponding to a melt rate of 21 m a-1, close to the simulation of 20.6 m a-1 by Reference Payne, Holland, Shepherd, Rutt, Jenkins and JoughinPayne and others (2007) with a plume model, and approximating the simulation of 28 m a-1 using NASA IceBridge data by Reference Schodlok, Menemenlis, Rignot and StudingerSchodlok and others (2012). The freshwater fluxes from the minor ice shelves (Lazarev, Voyeykov, Cook and Dotson) need further verification, since there have been very few studies of these ice shelves.

Table 1. Freshwater flux (mSv) from ISM in different studies

The modeled total circumpolar freshwater flux from the AISM during the integration period 1958–2000 (Fig. 2) shows significant interannual variability, with a clear increase since the mid-1960s, consistent with the observed increase in basal melting during recent decades (Reference Rignot and JacobsRignot and Jacobs, 2002; Reference Shepherd, Wingham and RignotShepherd and others, 2004). An increase in basal melting has been suggested to be a major cause of the thinning of ice shelves in the Amundsen Sea (Reference Shepherd, Wingham and RignotShepherd and others, 2004), owing to the ocean warming (Reference Shepherd, Wingham and RignotShepherd and others, 2004) and changes in wind patterns in the Southern Hemisphere (Reference Pritchard, Ligtenberg, Fricker, Vaughan, Van den Broeke and PadmanPritchard and others, 2012; Reference Steig, Ding, Battisti and JenkinsSteig and others, 2012).

3.2. Impact of AISM on temperature

Figure 3 shows climatological mean temperature differences ΔT (with minus without AISM) at different depths. The impact of including AISM in temperature is very complex. Water in the vicinity of the Antarctic ice shelves generally becomes colder above 216 m (Fig. 3a-c), due to the cold fresh water from ISM (Reference HellmerHellmer, 2004; Reference Wang and BeckmannWang and Beckmann, 2007). The largest cooling is found at 216m (Fig. 3c), the base of most ice shelves’ edges in the model (except the FRIS and the Ross Ice Shelf (RIS)), where the heat and freshwater fluxes due to AISM are applied. Substantial cooling at 216 m is found in the coastal areas of the Amundsen, Weddell and Ross seas, with the most extreme values of ΔT in each region being about –2.2°C, –1.5°C and –1.25°C, respectively. The ISM of the smaller ice shelves in East Antarctica tends to cause much smaller ΔT in the vicinity of the continent, generally <0.2°C; the exception is the region around Amery Ice Shelf.

Fig. 3. Climatology mean temperature differences ΔT (°C; with minus without AISM) at depths of (a) 5 m, (b) 55 m, (c) 216 m, (d) 512 m, (e) 1033 m and (f) 3752 m. Positive temperature differences represent temperature increase (shaded red); negative temperature differences indicate temperature decrease (shaded blue).

The coastal cooling at 512 m (Fig. 3d) is weak, only occurring near the much thicker ice shelves FRIS and RIS. Close to the surface, cooling affects a much larger area (Fig. 3a and b). At 55 m (Fig. 3b), the cooling becomes weaker but broader next to the continent, as a result of the rising cold buoyancy fresh water from the ice shelves and offshore advection by mean wind stress. At the surface (Fig. 3a), ΔT is smaller than at mid-depths, implying a more extensive mixing between the colder ice-shelf water and the warmer CDW, and a fast exchange with the atmosphere. The most pronounced cooling at the surface is not next to the continent, but in the Antarctic Circumpolar Current (ACC) region north of the Weddell Sea (down to –1.0°C). The surface cooling next to the continent is mostly <0.4°C except in two regions: one is outside the Amery Ice Shelf, and the other is close to Cape Adare in the western Ross Sea, both being about –0.8°C.

Most of the Southern Ocean is undergoing significant warming when AISM is included (Fig. 3). Above 1000 m, ΔT is generally not >1.0°C, except in the Amundsen and Bellingshausen Seas and the eastern Weddell Sea. The largest warming occurs at 216 m (Fig. 3c), up to 1.2°C. Large warming is also seen in the Atlantic and Pacific sectors of the ACC regions. Further down, the magnitude of warming tends to decrease but the warming area tends to increase. Below ~1000m, almost all the Southern Ocean water south of 50° S is warmer with AISM (Fig. 3e and f). However, the magnitude of warming is spatially inhomogeneous. For example, at 1033 m (Fig. 3e), the largest ΔT is found in the Weddell Sea, up to 1.2°C, and near the continent, generally >0.6°C. In other areas, warming is much smaller, generally <0.4°C. Down to the bottom, ΔT decreases. At the same time, the warming pattern also changes. At 3752 m (Fig. 3f), the largest warming (>0.6°C) takes place in the Australia- Antarctic Basin in the Indian Ocean, not next to the continent as in the shallower layers.

Figure 2 suggests that total AISM increased ~10% between the mid-1960s (102 mSv) and the mid-1980s, after which it stayed fairly constant between 110 and 112 mSv. If the effect of AISM on Southern Ocean hydrography is correctly captured by the model, we should see similar trends in data although with much smaller total change.

Reference GilleGille (2002) reported a warming of 0.17°C in the Southern Ocean waters at 700–1100 m depth between the 1950s and 1980s. Most observed warming occurred in the ACC region between 45°S and 60°S, especially in the Atlantic and Indian sectors. This is almost identical to the warming pattern shown above as a result of increasing AISM (Fig. 3d and e). Reference Aoki, Yoritaka and MasuyamaAoki and others (2003) showed another example at depths of 200–900 m around 30–150° E in the ACC region, where warming over three decades (1966–98) suggested that the southern ACC front had shifted southward. A similar warming is reproduced in our model (Fig. 3c-e). Reference Levitus, Antonov and BoyerLevitus and others (2005) reported changes in ocean heat content vs time for 0–300, 0–700 and 0–3000 m integrations, finding that about half of the increased heat content was from the Atlantic Ocean. They also noted the cooling above ~600m and between 1100 and 1450 m at high southern latitudes. The cooling above ~600m is reproduced near the coastal regions (Fig. 3), whereas the cooling below ~ 1100 m is not reproduced in our simulation. This is possibly caused by the dynamic responses of ISM to changing wind (Reference Pritchard, Ligtenberg, Fricker, Vaughan, Van den Broeke and PadmanPritchard and others, 2012; Reference Steig, Ding, Battisti and JenkinsSteig and others, 2012), or by the Antarctic Surface Water flushing and melting the ice base (Reference Hattermann, NØst, Lilly and SmedsrudHattermann and others, 2012), which are not captured by the model.

The warming of the deep and bottom waters in the Southern Ocean over the past half-century has been reported by a number of studies. Reference Fahrbach, Meyer, Rohardt, Schröder and WoodgateFahrbach and others (1998) observed that the bottom water in the central Weddell Sea became warmer during the interval 1989–95. This warming trend was shown to affect most of the Weddell Sea when the data were extended to the early 2000s (Reference Fahrbach, Hoppema, Rohardt, Schröder and WisotzkiFahrbach and others, 2004). Observations dating back to 1912 indicated that the warming of the Weddell Sea Warm Deep Water had occurred since the 1970s (Reference Robertson, Visbeck, Gordon and FahrbachRobertson and others, 2002). The warming of the bottom water was also observed in the Indian sector (Reference WhitworthWhitworth, 2002; Reference Johnson, Purkey and BullisterJohnson and others, 2008) and Pacific sector (Reference Ozaki, Obata, Naganobu and GamoOzaki and others, 2009) of the Southern Ocean. In particular, Reference Ozaki, Obata, Naganobu and GamoOzaki and others (2009) revealed a long-term warming of the bottom layer of the Ross Sea from 1969 to 2004. Reference Johnson, Purkey and BullisterJohnson and others (2008) reported warming of the abyssal waters in the eastern Indian Ocean between 1994/95 and 2007. These warming trends are consistent with the effects of including AISM (Fig. 3e and f).

3.3. Impact of AISM on salinity

The impact of AISM on the Southern Ocean salinity at different levels is shown in Figure 4. AISM inputs fresh water into the adjacent ocean. As for temperature, the strongest freshening is seen close to 216 m (Fig. 4c), but with large and extensive freshening at shallower depths as well (Fig. 4a and b). At 216 m, the largest freshening is in the vicinity of the continent (down to –1), where most of the ice-shelf/ocean interaction was applied in the model. In addition, freshening is dominant over the whole Southern Ocean at this depth, except for patches of slightly higher salinity (up to 0.1) in the central Kerguelen, Weddell and Ross gyres.

Fig. 4. Same as Figure 3, but for salinity.

The substantial freshening at the surface (Fig. 4a and b) arises because the fresh subsurface water rises to the surface due to its low density. The freshening at the surface is particularly remarkable in the Ross and Weddell Seas (down to –0.5), the Amundsen and Bellingshausen Seas (down to –0.4) and Prydz Bay (down to –0.3). In the Indian Ocean and western Weddell Sea, the freshening is generally much weaker. In the region east of 45° E, the water even becomes more saline, up to 0.2. At 55 m, almost all of the Southern Ocean becomes fresher (Fig. 4b). Unlike at the surface, the freshening is more pronounced next to the continent at this depth. Aside from the general freshening, small areas of more saline water up to 0.1 are found around the 45° E meridian and in the southern Ross Sea.

At depths of 512 and 1033 m, inclusion of AISM generally causes higher salinity in the Southern Ocean (Fig. 4d and e). Freshening is only found in the Weddell and Ross Seas close to the continent, where the decrease in salinity is up to 0.3. At the bottom (Fig. 4f), most of the Southern Ocean south of ~55° S shows a trend of reduced salinity when AISM is included, although the changes are small (<0.02).

As with temperature (Section 3.2), we can use these model results to predict the change in the Southern Ocean salinity associated with the modeled 10% increase in AISM between the 1960s and 1980s (Fig. 2). Reference Wong, Bindoff and ChurchWong and others (1999) found a large-scale freshening of intermediate water in the Pacific and Indian Oceans between the periods 1930–80 and 1985–94, and partly attributed this to surface water freshening in the source region of the Southern Ocean. They estimated a net increase in freshwater flux of ~31 mm a-1 (~22 mSv) for the Southern Ocean between 55° S and 65° S. The increasing freshwater input is 8–12.5 mm a-1 (~6–9 mSv) from coupled atmosphere-ocean models, with increasing atmospheric greenhouse-gas concentration for the same period (Reference Manabe, Bryan and SpelmanManabe and others, 1990; Reference Gordon and O’FarrellGordon and O’Farrell, 1997), which could explain about one-third of the net increase in freshwater flux. The increase in the simulated freshwater flux from AISM between the mid-1960s and mid-1980s is ~10 mSv (Fig. 2). It is likely a significant contributor to the freshwater flux variability. In addition, the simulated large salinity decrease in the vicinity of ice shelves is not inconsistent with summer observations near the RIS by Reference Jacobs, Giulivi and MeleJacobs and others (2002), who found a rapid freshening in the surface water in the Ross Gyre and much deeper close to the ice-shelf edge. Reference Jacobs, Giulivi and MeleJacobs and others (2002) attributed this freshening to the increasing AISM in the Bellingshausen and Amundsen Seas, which is also found in later in situ observations (e.g. Reference Shepherd, Wingham and RignotShepherd and others, 2004; Reference WåhlinWåhlin and others, in press).

The modeled spatial extent of surface freshening down to 500 m in the Pacific, Atlantic and Indian sectors is generally consistent with the observed freshening (Reference Boyer, Levitus, Antonov, Locarnini and GarciaBoyer and others, 2005). It is seen from the observations that strong freshening occurs in the Pacific and Atlantic sectors of the high-latitude Southern Ocean, with the most significant salinity decrease in the top surface waters. In the Indian sector of the Southern Ocean, however, the observed overall salinity is slightly increasing (Reference Boyer, Levitus, Antonov, Locarnini and GarciaBoyer and others, 2005). This pattern is qualitatively depicted in the present study (Fig. 4), where the freshening in the Indian sector is much weaker than in the Pacific and Atlantic sectors. Consistent with recent observations by Reference RintoulRintoul (2007), Reference Johnson, Purkey and BullisterJohnson and others (2008) and Reference Ozaki, Obata, Naganobu and GamoOzaki and others (2009), increasing AISM would cause freshening of the bottom water in the Southern Ocean (Fig. 4). Such overall consistency presented in this study has not been simulated in any previous study.

3.4. Impact of AISM on density

Potential density is determined by both temperature and salinity. However, in most regions of the cold Southern Ocean, density is mainly controlled by salinity. Therefore, the general patterns of density change (Fig. 5) are similar to those of salinity at most of the vertical levels, where density generally decreases. The largest decrease in density (>0.2 kg m-3) is found at the surface (Fig. 5a), particularly in the Weddell, Bellingshausen, Amundsen and Ross seas. Below the surface layer, a large decrease in density is mainly confined to the continental margins down to 216 m. At 1033 m, the water decreases slightly in density in a fairly large area (Fig. 5e), where both temperature (Fig. 3e) and salinity (Fig. 4e) increase due to the inclusion of AISM. Here the increase in temperature plays a dominant role in the density change. At 3752 m the decrease is generally <0.1 kgm-3.

Fig. 5. Same as Figure 3, but for density (kgm-3).

The simulated density decrease in the bottom water is consistent with observations (Reference WhitworthWhitworth, 2002; Reference Aoki, Yoritaka and MasuyamaAoki and others, 2003; Reference RintoulRintoul, 2007; Reference Johnson, Purkey and BullisterJohnson and others, 2008; Reference Ozaki, Obata, Naganobu and GamoOzaki and others, 2009). Due to the rapid freshening of the bottom water, the decrease in density is also significant, resulting in loss of the densest water classes (e.g. Reference Johnson, Purkey and BullisterJohnson and others, 2008). Additionally, warming of the intermediate water would also cause a decrease in density. Such changes may reduce the stability of the water column, enhancing eddy formation due to baroclinic instability (e.g. Reference Meredith and HoggMeredith and Hogg, 2006; Reference Screen, Gillett, Stevens, Marshall and RoscoeScreen and others, 2009).

4. Conclusions

We have investigated the effects of increasing AISM on Southern Ocean hydrography. Meltwater from AISM is mainly confined to the vicinity of the ice shelves. Due to its lower density, the freshened water rises, causing strong freshening and cooling of the upper-ocean Antarctic coastal waters. The freshened upper-ocean water is advected northward by wind stress, and is replaced by warmer and more saline deep water flowing southward towards the ice fronts.

Our model run with AISM suggests that the total circumpolar meltwater flux increased ~10% from a minimum in the mid-1960s to a fairly stable maximum value from the mid-1980s (Fig. 2). Southern Ocean hydrographic changes reported in many studies (e.g. Reference Wong, Bindoff and ChurchWong and others, 1999; Reference Levitus, Antonov, Boyer and StephensLevitus and others, 2000; Reference GilleGille, 2002; Reference Robertson, Visbeck, Gordon and FahrbachRobertson and others, 2002; Reference Aoki, Yoritaka and MasuyamaAoki and others, 2003; Reference Boyer, Levitus, Antonov, Locarnini and GarciaBoyer and others, 2005; Reference JacobsJacobs, 2006; Reference RintoulRintoul, 2007; Reference Ozaki, Obata, Naganobu and GamoOzaki and others, 2009) are qualitatively consistent with the modeled effect of increasing AISM. Ice-shelf melting tends to freshen the surface water, warm the intermediate and deep waters, and freshen and warm the bottom water in the Southern Ocean.

If we assume that the effect of increasing AISM is linear through the range of our two model runs, we can scale predicted hydrographic changes by 10% to estimate expected changes from the mid-1960s to the mid-1980s and later. For example, in the Weddell Sea the temperature difference at 1033m between the simulations with and without AISM is ~0.6°C. This suggests that the actual warming caused by increasing AISM would be ~0.06°C between 1965 and 2000, about one-third of the observed 0.17°C between 1950 and 1980 in the same area (Reference GilleGille, 2002). This implies that increasing AISM is likely a significant contributor to recent changes in the Southern Ocean. However, due to the difficulty of exactly determining AISM and its circumpolar amount, as well as the sparsity of samples in the high-latitude Southern Ocean, more studies are needed to clarify their exact relations.

We have not investigated the detailed causes of the modeled changes in AISM between 1958 and 2000 (Fig. 2), although a likely factor is changes in hemispheric wind fields (Reference Pritchard, Ligtenberg, Fricker, Vaughan, Van den Broeke and PadmanPritchard and others, 2012; Reference Steig, Ding, Battisti and JenkinsSteig and others, 2012). Our results suggest that, if AISM increases for some reason, the wind-driven offshore transport of near-surface freshened water will be associated with southward transport of warmer intermediate layers, leading to further increases in AISM. That is, there is a positive feedback between AISM and hydrography that can increase the effect of climate perturbations on Antarctic ice-sheet mass balance. However, further studies are required to quantify other causes of hydrographic variability, including changes in winds and sea-ice production in, and export from, Antarctic coastal regions.

Acknowledgements

We thank the DRAKKAR group for providing the ORCA2-LIM configuration, especially Jean-Marc Molines for his assistance in setting up the model in Helsinki. We are grateful to Aike Beckmann and anonymous reviewers for constructive comments which substantially improved the manuscript. This study was supported by the University of Helsinki through the project EPOS. The simulations were carried out at the Finnish Center for Scientific Computing.

References

Aoki, S, Yoritaka, M and Masuyama, A (2003) Multidecadal warming of subsurface temperature in the Indian sector of the Southern Ocean. J. Geophys. Res., 108(C4), 8081 (doi: 10.1029/2000JC000307)Google Scholar
Beckmann, A and Goosse, H (2003) A parameterization of ice shelf-ocean interaction for climate models. Ocean Model., 5(2), 157170 (doi: 10.1007/s10712–011–9123–5)Google Scholar
Beckmann, A and Timmermann, R (2001) Circumpolar influences on the Weddell Sea: indication of an Antarctic circumpolar coastal wave. J. Climate, 14(17), 37853792 (doi: 10.1175/1520- 0442(2001)014<3785:CI0TWS>2.0.C0;2)2.0.CO;2>CrossRefGoogle Scholar
Beckmann, A, Hellmer, HH and Timmermann, R (1999) A numerical model of the Weddell Sea: large-scale circulation and water mass distribution. J. Geophys. Res., 104(C10), 23 37523 392 (doi: 10.1029/1999JC900194)Google Scholar
Boyer, TP, Levitus, S, Antonov, JI, Locarnini, RA and Garcia, HE (2005) Linear trends in salinity for the World Ocean, 1955–1998. Geophys. Res. Lett., 32(1), L01604 (doi: 10.1029/2004GL021791)Google Scholar
Drewry, DJ (1983) Antarctica: glaciological and geophysical folio. Scott Polar Research Institute, Cambridge Google Scholar
Fahrbach, E, Meyer, R, Rohardt, G, Schröder, M and Woodgate, RA (1998) Gradual warming of the Weddell Sea deep and bottom water. FRISP Rep. 12, 2434 (hdl:10013/epic.18699)Google Scholar
Fahrbach, E, Hoppema, M, Rohardt, G, Schröder, M and Wisotzki, A (2004) Decadal-scale variations of water mass properties in the deep Weddell Sea. Ocean Dyn., 54(1), 7791 (doi: 10.1007/s10236–003–0082–3)Google Scholar
Fichefet, T and Morales Maqueda, MA (1997) Sensitivity of a global sea ice model to the treatment of ice thermodynamics and dynamics. J. Geophys. Res., 102(C6), 12 60912 646 (doi: 10.1029/97JC00480)CrossRefGoogle Scholar
Gille, ST (2002) Warming of the Southern Ocean since the 1950s. Science, 295(5558), 12751277 (doi: 10.1126/science.1065863)Google Scholar
Gille, ST (2003) Float observations of the Southern Ocean. Part I: estimating mean fields, bottom velocities, and topographic steering. J. Phys. Oceanogr., 33(6), 11671181 (doi: 10.1175/1520–0485(2003)033<1167:FOOTSO>2.0.CO;2)Google Scholar
Gordon, HB and O’Farrell, SP (1997) Transient climate change in the CSIRO coupled model with dynamic sea ice. Mon. Weather Rev., 125(5), 875907 (doi: 10.1175/1520–0493(1997)125 <0875:TCCITC>2.0.CO;2)Google Scholar
Hattermann, T, NØst, OA, Lilly, JM and Smedsrud, LH (2012) Two years of oceanic observations below the Fimbul Ice Shelf, Antarctica. Geophys. Res. Lett., 39(12), L12605 (doi: 10.1029/2012GL051012)Google Scholar
Hellmer, HH (2004) Impact of Antarctic ice shelf basal melting on sea ice and deep ocean properties. Geophys. Res. Lett., 31(10), L10307 (doi: 10.1029/2004GL019506)Google Scholar
Hibler, WD III (1979) A dynamic thermodynamic sea ice model. J. Phys. Oceanogr., 9(7), 815846 Google Scholar
Holland, PR, Jenkins, A and Holland, DM (2010) Ice and ocean processes in the Bellingshausen Sea, Antarctica. J. Geophys. Res., 115(C5), C05020 (doi: 10.1029/2008JC005219)Google Scholar
Jacobs, S (2006) Observations of change in the Southern Ocean. Philos. Trans. R. Soc. London, Ser. A, 364(1844), 16571681 (doi: 10.1098/rsta.2006.1794)Google Scholar
Jacobs, SS, Hellmer, HH, Doake, CSM, Jenkins, A and Frolich, RM (1992) Melting of ice shelves and the mass balance of Antarctica. J. Glaciol., 38(130), 375387 Google Scholar
Jacobs, SS, Hellmer, HH and Jenkins, A (1996) Antarctic ice sheet melting in the southeast Pacific. Geophys. Res. Lett., 23(9), 957960 (doi: 10.1029/96GL00723)Google Scholar
Jacobs, SS, Giulivi, CF and Mele, PA (2002) Freshening of the Ross Sea during the late 20th century. Science, 297(5580), 386389 (doi: 10.1126/science.1069574)Google Scholar
Jacobs, SS, Jenkins, A, Giulivi, CF and Dutrieux, P (2011) Stronger ocean circulation and increased melting under Pine Island Glacier ice shelf. Nature Geosci., 4(8), 519523 (doi: 10.1038/ngeo1188)Google Scholar
Jenkins, A and Jacobs, S (2008) Circulation and melting beneath George VI Ice Shelf, Antarctica. J. Geophys. Res., 113(C4), C04013 (doi: 10.1029/2007JC004449)Google Scholar
Johnson, GC, Purkey, SG and Bullister, JL (2008) Warming and freshening in the abyssal southeastern Indian Ocean. J. Climate, 21(20), 53515363 (doi: 10.1175/2008JCLI2384.1)Google Scholar
Levitus, S, Antonov, JI, Boyer, TP and Stephens, C (2000) Warming of the world ocean. Science, 287(5461), 22252229 (doi: 10.1126/science.287.5461.2225)Google Scholar
Levitus, S, Antonov, JI and Boyer, TP (2005) Warming of the world ocean, 1955–2003. Geophys. Res. Lett., 32(2), L02604 (doi: 10.1029/2004GL021592)Google Scholar
Madec, G, Delecluse, P, Imbard, M and Lévy, C (1998) OPA 8.1 Ocean General Circulation Model reference manual. Institut PierreSimon Laplace (IPSL), Paris (Note du Pole de modélisation 11)Google Scholar
Manabe, S, Bryan, K and Spelman, MJ (1990) Transient response of a global ocean-atmosphere model to a doubling of atmospheric carbon dioxide. J. Phys. Oceanogr., 20(5), 722749 (doi: 10.1175/1520–0485(1990)020<0722:TROAGO>2.0.CO;2)2.0.CO;2>CrossRefGoogle Scholar
Meredith, MP and Hogg, AM (2006) Circumpolar response of Southern Ocean eddy activity to a change in the Southern Annular Mode. Geophys. Res. Lett., 33(16), L16608 (doi: 10.1029/2006GL026499)Google Scholar
Ozaki, H, Obata, H, Naganobu, M and Gamo, T (2009) Long-term bottom water warming in the north Ross Sea. J. Oceanogr., 65(2), 235244 (doi: 10.1007/s10872–009–0022-z)Google Scholar
Payne, AJ, Holland, PR, Shepherd, AP, Rutt, IC, Jenkins, A and Joughin, I (2007) Numerical modeling of ocean-ice interactions under Pine Island Bay’s ice shelf. J. Geophys. Res., 112(C10), C10019 (doi: 10.1029/2006JC003733)Google Scholar
Pritchard, HD, Ligtenberg, SRM, Fricker, HA, Vaughan, DG, Van den Broeke, MR and Padman, L (2012) Antarctic ice-sheet loss driven by basal melting of ice shelves. Nature, 484(7395), 502505 (doi: 10.1038/nature10968)Google Scholar
Rignot, EJ (1998) Fast recession of a West Antarctic glacier. Science, 281(5376), 549551 (doi: 10.1126/science.281.5376.549)Google Scholar
Rignot, E and Jacobs, SS (2002) Rapid bottom melting widespread near Antarctic ice sheet grounding lines. Science, 296(5575), 20202023 (doi: 10.1126/science.1070942)Google Scholar
Rintoul, SR (2007) Rapid freshening of Antarctic Bottom Water formed in the Indian and Pacific oceans. Geophys. Res. Lett., 34(6), L06606 (doi: 10.1029/2006GL028550)CrossRefGoogle Scholar
Robertson, R, Visbeck, M, Gordon, AL and Fahrbach, E (2002) Longterm temperature trends in the deep waters of the Weddell Sea. Deep-Sea Res. II, 49(21), 47914806 (doi: 10.1016/S0967- 0645(02)00159–5)Google Scholar
Schodlok, M, Menemenlis, D, Rignot, E and Studinger, M (2012) Sensitivity of the ice-shelf/ocean system to the sub-ice-shelf cavity shape measured by NASA IceBridge in Pine Island Glacier, West Antarctica. Ann. Glaciol., 53(60 Pt 2), 156162 (doi: 10.3189/2012AoG60A073)CrossRefGoogle Scholar
Screen, JA, Gillett, NP, Stevens, DP, Marshall, GJ and Roscoe, HK (2009) The role of eddies in the Southern Ocean temperature response to the southern annular mode. J. Climate, 22(3), 806818 (doi: 10.1175/2008JCLI2416.1)Google Scholar
Semtner, AJ Jr (1976) A model for the thermodynamic growth of sea ice in numerical investigations of climate. J. Phys. Oceanogr., 6(5), 379389 (doi: 10.1175/1520–0485(1976)006<0379: AMFTTG>2.0.CO;2)Google Scholar
Shepherd, A, Wingham, D and Rignot, E (2004) Warm ocean is eroding West Antarctic Ice Sheet. Geophys. Res. Lett., 31(23), L23404 (doi: 10.1029/2004GL021106)Google Scholar
Steig, EJ, Ding, Q, Battisti, DS and Jenkins, A (2012) Tropical forcing of Circumpolar Deep Water Inflow and outlet glacier thinning in the Amundsen Sea Embayment, West Antarctica. Ann. Glaciol., 53(60 Pt 1), 1928 (doi: 10.3189/2012AoG60A110)Google Scholar
Thoma, M, Grosfeld, K and Lange, MA (2006) Impact of the Eastern Weddell Ice Shelves on water masses in the eastern Weddell Sea. J. Geophys. Res., 111(C12), C12010 (doi: 10.1029/2005JC003212)Google Scholar
Thoma, M, Jenkins, A, Holland, D and Jacobs, S (2008) Modelling circumpolar deep water intrusions on the Amundsen Sea continental shelf, Antarctica. Geophys. Res. Lett., 35(18), L18602 (doi: 10.1029/2008GL034939)Google Scholar
Timmermann, R, Hellmer, HH and Beckmann, A (2002) Simulations of ice-ocean dynamics in the Weddell Sea 2. Interannual variability 1985–1993. J. Geophys. Res., 107(C3), 3025 (doi: 10.1029/2000JC000742)Google Scholar
Timmermann, R, Goosse, H, Madec, G, Fichefet, T, Ethe, C and Dulière, V (2005) On the representation of high latitude processes in the ORCA-LIM global coupled sea ice-ocean model. Ocean Model., 8(1–2), 175201 (doi: 10.1016/j.oce- mod.2003.12.009)Google Scholar
Wåhlin, AK, Yuan, X, Björk, G and Nohr, C (2010) Inflow of warm circumpolar deep water in the Central Amundsen Shelf. J. Phys. Oceanogr., 40(6), 14271434 (doi: 10.1175/2010JPO4431.1)Google Scholar
Wåhlin, A and 6 others (in press) Some implications of Ekman layer dynamics for cross shelf exchange in the Amundsen Sea. J. Phys. Oceanogr. (doi: 10.1175/JPO-D-11–041.1)Google Scholar
Walkden, GJ, Heywood, KJ, Nicholls, KW and Abrahamsen, P (2009) Freshwater transport at Fimbulisen, Antarctica. J. Geophys. Res., 114(C8), C08014 (doi: 10.1029/2008JC005028)Google Scholar
Walker, DP, Brandon, MA, Jenkins, A, Allen, JT, Dowdeswell, JA and Evans, J (2007) Oceanic heat transport onto the Amundsen Sea shelf through a submarine glacial trough. Geophys. Res. Lett., 34(2), L02602 (doi: 10.1029/2006GL028154)Google Scholar
Wang, C (2011) Antarctic ice shelf melting and its impact on the global sea ice-ocean system. Rep. Ser. Geophys. 67 Google Scholar
Wang, C and Beckmann, A (2007) Investigation of the impact of Antarctic ice-shelf melting in a global ice-ocean model (ORCA2-LIM). Ann. Glaciol., 46, 7882 (doi: 10.3189/172756407782871602)Google Scholar
Whitworth, TI (2002) Two modes of bottom water in the Australian-Antarctic Basin. Geophys. Res. Lett., 29(5), 1073 (doi: 10.1029/2001GL014282)Google Scholar
Wong, APS, Bindoff, NL and Church, JA (1999) Large-scale freshening of intermediate waters in the Pacific and Indian oceans. Nature, 400(6743), 440443 (doi: 10.1038/22733)Google Scholar
Figure 0

Fig. 1. The name, extent and position of the Antarctic ice shelves included in the ORCA2-LIM-ISP model. EW: Riiser-Larsen and Brunt Ice Shelves.

Figure 1

Fig. 2. Time series of the simulated monthly (black) and annual (red) mean of the total circumpolar freshwater flux from AISM.

Figure 2

Table 1. Freshwater flux (mSv) from ISM in different studies

Figure 3

Fig. 3. Climatology mean temperature differences ΔT (°C; with minus without AISM) at depths of (a) 5 m, (b) 55 m, (c) 216 m, (d) 512 m, (e) 1033 m and (f) 3752 m. Positive temperature differences represent temperature increase (shaded red); negative temperature differences indicate temperature decrease (shaded blue).

Figure 4

Fig. 4. Same as Figure 3, but for salinity.

Figure 5

Fig. 5. Same as Figure 3, but for density (kgm-3).