<<

1 Late winter oceanography beneath East sea

2 ice during SIPEX

a,b,∗ d,b d,c e a,b 3 G.D. Williams , A.J.S. Meijers , A. Poole , P. Mathiot , T. Tamura , b 4 A. Klocker

a 5 Institute for Low Temperature Science, Hokkaido University, Sapporo, Japan. b 6 Antarctic Climate and Ecosystem Cooperative Research Centre, Sandy Bay, Australia. c 7 Department of the Environment, Water, Heritage and the Arts, Australian Antarctic 8 Division, Kingston, Australia. d 9 Centre for Australian Weather and Climate Research, CSIRO, Hobart, Australia. e 10 Laboratory of Geophysical and Industrial Flow, Grenoble, France

11 Abstract

We report on the late winter oceanography observed beneath Antarctic sea ice offshore from the Sabrina and BANZARE coast of , East (115–125◦E) in September–October 2007 during the Sea Ice Physics and Ecosys- tem eXperiment (SIPEX) research voyage. A pilot program using specifically designed ’through-ice’ Conductivity-Temperature-Depth (CTD) and acoustic Doppler current profiling (ADCP) systems was conducted to opportunistically measure water mass properties and ocean currents during major ice stations. Additional water mass properties across the survey region were collected from Ice-Argo floats deployed during the voyage north of the 3000m isobath. The mean drift of the floats between deployment and the first surface download in summertime was along slope to the west with the Antarctic Slope Current. Vertical profiles of potential temperature reveal the deepest winter mixed layer (WML) in the western sector of the survey northwest of the Dalton Iceberg Tongue polynya. The meridional structure of the Antarctic Slope Front, i.e. the offshore shoaling of the WML across the upper continental slope, is found to be the simillar to previous summer observations. South of the continen- tal shelf break a strong bottom intrusion of modified Circumpolar Deep Water (mCDW) as warm as 0◦C was detected beneath the fast ice near 118◦E. This is in the vicinity of a summertime observation of a shoreward intrusion of mCDW

∗Corresponding author: [email protected] Preprint submitted to Elsevier March 10, 2010 of similar magnitude over ten years earlier. We hypothesise that this strengthens the liklihood that there is a persistent supply of ocean heat to the shelf region and that this is the primary cause of recently reported enhanced melting of the nearby Totten and Moscow University . Interestingly there was no detection of locally formed dense shelf water capable of forming Antarctic Bottom Water at the shelf break locations sampled despite the number of minor polynyas across this region. Ocean current measurements, limited to a maxi- mum period of 24 hours and 50–100m depth by the relative scarcity of backscat- ter, found increased mean vertical speeds at the offshore stations (6–17 cm s−1) relative to the shelf break (2.3–6.4 cm s−1cm s( − 1)). The diurnal variation in the ADCP range suggests that the diel migration of zooplankton was occur- ring beneath the sea ice in late winter, with greater range/abundance offshore. Analysis of concurrent time series of wind, ocean current and their influence on sea ice drift from Global Positioning System (GPS) compass measurements were computed but the length of data acquisitions limited the applicability of this analysis. Overall the pilot ’through-ice’ program was a success and with improvements in the accuracy of instrumentation, and a strategic survey plan, promises to provide invaluable observations for model comparison outside of the current datasets available.

1 Keywords: SIPEX, Under-ice, Oceanography, East Antarctica

2 1 1. Introduction

2 The seasonal formation of sea ice around Antarctic each year is a primary

3 mechanism for water mass transformations in the polar Southern Ocean. Brine-

4 rejection from sea ice production promotes convection and the formation of a

5 deep winter mixed layer in the Antarctic Surface Water (AASW). In discrete lo-

6 cations on the Antarctic continental shelf, enhanced sea ice formation and brine

7 rejection increases the density of shelf waters, such that if transported north-

8 wards with sufficient negative buoyancy to mix down the continental slope,

9 Antarctic Bottom Water (AABW) forms. Oceanographic observations around

10 Antarctica are sparse due to its remote location and this is accentuated dur-

11 ing winter months when sea ice presents additional logistic challenges. Aus-

12 tralian Antarctic research expeditions conducted in winter are predominantly

13 field-based sea-ice studies, i.e. a series of ’ice-stations’ ex-ship, that do not allow

14 for standard ship-based oceanography due to time and safety contraints. For

15 this reason a pilot field based ’through-ice’ oceanography program was under-

16 taken during the Sea Ice Physics and Ecosystem eXperiment (SIPEX) voyage

17 (Worby et al., 2010, this volume) with the goal of sampling the water mass prop-

18 erties and currents beneath the late winter sea ice around the East Antarctic

19 margin offshore from the Sabrina and BANZARE coast of Wilkes Land between ◦ 20 115–130 E (Figure 1).

21 [Figure 1 about here]

22 1.1. The Antarctic Margin of the Sabrina and BANZARE Coast, Wilkes Land

23 The continental shelf region of the Sabrina and BANZARE coasts of Wilkes

24 Land has a combination of floating , iceshelves and grounded iceberg

25 tongues, most notably the Totten Glacier and Moscow University Ice Shelf,

26 together with a series of persistent polynya regions. These range from Cape

27 Poinsett in the west to the in the east and are labelled

28 in Figure 1 following Massom (2003) with estimates of total sea ice produc- −2 29 tion in 2007 (m m ). These sea ice production estimates come from SSM/I

3 1 satellite data with the ERA-Interim reanalysis (ECMWF, 2009) and the heat

2 flux algorithm of Tamura et al. (2007, 2008). These are all examples of recur-

3 ring latent-heat polynyas which form due to ice divergence in the lee of topo-

4 graphic/ice barriers over the continental shelf. The exception is the Voyeykov

5 polynya, which is a ’deep water’ polynya (?) that forms along the fast ice edge ◦ 6 near the shelf break between 122–125 E.

7 The previous observations in this region are limited to the BROKE (Baseline

8 Research on Oceanography Krill and the Enviroment) survey in the austral

9 summer of 1996 (Bindoff et al., 2000), as shown in Figure 1. Vertical sections of

10 potential temperature for the upper 1000m of BROKE meridional transects at ◦ 11 112, 120 and 128 E are shown in Figure 2. These sections demonstrate the key

12 water masses, boundaries and fronts in summertime and their zonal variability

13 across the SIPEX survey region. In this paper we follow definitions used by

14 Orsi et al. (1995); Whitworth et al. (1998) and recently updated by ?. Offshore n 15 of the shelf break, the two major water masses are cold, fresh AASW (γ ¡ n 16 28.00 kg m−3) above warm, saline Circumpolar Deep Water (CDW, 28.00 ¡ γ

17 ¡ 28.27 kg m−3). The CDW properties weaken polewards through mixing with ◦ 18 cold Antarctic water masses and when cooler than 1.5 C (also known as the

19 Southern Boundary) it is referred to as ’modified’ CDW (mCDW).

20 Over the upper continental slope and continental shelf break this boundary

21 between AASW and mCDW is termed the Antarctic Slope Front (ASF, Jacobs

22 (1991); Whitworth et al. (1998)). There is a broad westward flow, termed the

23 Antarctic Slope Current (ASC), associated with the horizontal desnity gradi-

24 ent across the ASF, that often includes a narrow, fast-flowing jet (up to 50 cm −1 25 s ) topographically pinned between the 750–1250 m isobaths. Though some-

26 times co-located, the ASC is not to be confused with Antarcitc Coastal Current

27 (ACoC), which is associated with the East Wind Drift, shallower and has a more

28 coastal domain. In this paper we will concentrate on the ASC as this is typi- ◦ ◦ 29 cal of the circulation around the East Antarctica margin from 30 E to 150 E

30 (Bindoff et al., 2000; Aoki, 2003; Williams et al., 2008b; Meijers et al., 2010;

31 Williams et al., 2010a) and indeed much of the Antarctic margin away from the

4 1 Weddell Sea and Antarctic peninsula (Fahrbach et al., 1992; Whitworth et al.,

2 1998; Heywood et al., 1998; ?).

3 The vertical structure of the AASW varies seasonally and is shown in Figure

4 2 with a warm, fresh seasonal summer mixed layer (SML) at the surface that

5 forms from the fresh water lens post sea ice melt. The SML deepens through the

6 summer by wind mixing into the temperature minimum layer that is the remnant

7 from the previous winter’s permanent mixed layer (referred to hereafter as the

8 winter mixed layer, WML). The WML forms by convection, driven initially by

9 atmospheric cooling of the surface SML and then by the brine rejection that

10 occurs with sea ice formation. Inshore of the shelf break, in regions of sufficient

11 sea ice formation/brine rejection, the WML extends to the bottom, mixing away

12 all remnants of the summer stratification and reconnects with the base of the n 13 previous year’s WML. If denser than γ = 28.27 kg m−3 and colder than θ ◦ 14 = -1.7–-1.85 C, this new winter water mass is referred to as Shelf Water(SW).

15 Offshore of the continental shelf break, brine rejection is relatively weak and the

16 WML/AASW remains fresher and lighter that the mCDW. In specific locations

17 mCDW can upwell and migrate/intrude onto the continental shelf, transporting

18 heat to the Antarctic coastal domain which has major implications for ocean/ice

19 shelf interactions (Smethie and Jacobs, 2005). The mCDW intrusions are also

20 important for the marine ecosystem if they upwell into the surface layer as they

21 are rich in nutrients. These intrusion are also reportedly an important part

22 of the lifecycle of Antarctic krill, transporting eggs from offshore (Sala et al.,

23 2002). However the influence of mCDW on the continental shelf is ultimately

24 dependent on the local stratification. For example, if SW is present, then by

25 definition mCDW can only occupy the layer above.

26 [Figure 2 about here]

27 The striking feature of the BROKE transects in this region is the south- ◦ ◦ 28 ward migration of the warmer mCDW at 120 E, bringing the sub-surface 0 C

29 isotherm onto the shelf break at station 83 . This was reported as the strongest ◦ 30 mCDW signal of the entire BROKE survey between 80–150 E. This is in broad

5 1 agreement with the southward shift in the Southern Boundary of the Antarctic ◦ 2 Circumpolar Current described by Orsi et al. (1995) east of 115 E (see dashed

3 line in Figure 1). However the BROKE results showed there is more zonal

4 variability in the properties of the mCDW layer, as indicated by the southern ◦ ◦ 5 extent of the 1 C isotherm at 120 E in Figure 2. The other major zonal vari-

6 ation across these transects was in the surface layer over the continental shelf ◦ 7 and the weakly developed seasonal mixed layer at 112 E (i.e. it was observed

8 north of station 72). This was not the result of any temporal lag in the sampling

9 as the transects were conducted from east to west and instead implies greater

10 production and/or persistence of sea ice, which has been found to delay sum-

11 mertime mixed layer development in other regions of East Antarctica (Williams

12 et al., 2008b).

13 In this paper we will present unique late winter observations of the water

14 mass properties beneath sea ice on the continental shelf break (0–550m) and up- ◦ 15 per layer (0–1000m) offshore of the East Antarctic margin between 117–128 E.

16 In addition concurrent measurements of near-surface ocean currents, sea ice

17 drift and surface wind speeds collected at SIPEX ice stations are presented.

18 In section 2 we describe the oceanographic survey and introduce the two main

19 observation platforms, namely a ’through-ice’ Conductivity-Temperature-Depth

20 (CTD) and acoustic Doppler current profiling (ADCP) system, together with

21 Ice-Argo floats deployed on behalf of the University of Washington. In Section 3

22 we present our observations, in particular vertical profiles of potential tempera-

23 ture from the CTD system and Ice-Argo in a description of the ’re-sampling’ of ◦ 24 the mCDW intrusion on the shelf break near 120 E and the spatial variation in

25 the winter mixed layer depth across the survey. This is followed by the vertical

26 profiles of current speed and direction immediately beneath the sea ice at both

27 offshore and fast-ice stations. We analyse the concurrent time series of ocean

28 current and surface wind speed and direction to investigate how each component

29 influences sea ice drift. In Section 4 we discuss the impact of local polynyas and

30 ice shelves on the water mass properties observed in this study and the overall

6 1 success and future of the ’through-ice’ oceanography program.

2 2. Data and Methods

3 This project involved two independent sub-ice observation platforms: A

4 winch-driven Conductivity-Temperature-Depth system for measuring basic wa-

5 ter mass properties and an acoustic Doppler current profiling (ADCP)/GPS

6 system for measuring ocean currents and ice drift. Hereafter these are referred

7 to as the CTD and ADCP systems respectively (Figure 3). Additional water

8 mass data were collected by Ice Argo floats deployed during SIPEX for the

9 University of Washington.

10 [Figure 3 about here]

11 2.1. Ice Argo floats from the University of Washington

12 The Ice-Argo floats deployed during SIPEX are the new generation of Argo

13 float that are capable of autonomously collecting CTD and dissolved oxygen

14 profiles beneath sea-ice. These floats are designed to abort their ascent 10-30m

15 from the surface when sea ice cover is detected by a cold near-surface layer (Klatt

16 et al., 2007). Deployed within the sea-ice field north of the 3000-m isobath, the

17 floats profile the water column (every 2m to 2000m depth) every 7 days without

18 resurfacing, using on-board data storage until the data can be uploaded after

19 the sea-ice melts. Depending on voyage logistics, the floats were either deployed

20 from the back of the ship, through a hole in the ice during ice stations, or from

21 a helicopter into a lead when the ship was south of the 3000-m isobath. The

22 Ice-Argo data were extracted from the publicly-available data portal operated

23 by the National Oceanographic Data Centre (NODC, 2009). We use an abbre-

24 viated World Meteorological Organisation (WMO) naming format, i.e. WMO

25 float number 2900130 is referred to in this paper as a130. Data is expected ◦ 26 to have accuracy of 0.001 C, 0.01 and 5m for temperature, salinity and depth,

27 respectively (Office, 2010). Further information about the Ice-Argo instruments

7 1 and data processing can be accessed through the University of Washington (UW,

2 2009).

3 [Figure 4 about here]

4 Figure 4 shows the deployment locations of Ice-Argo floats during SIPEX,

5 and their first ’download’ position, with the mean speed necessary to cover this

6 distance. The broad pattern for the time mean drift of the floats is one of along

7 slope to the west with the Antarctic Slope Current. The floats deployed in the

8 western sector of the survey were transported west, but with a greater northward

9 component, again following the bathymetry. The one exception to this was a

10 float that surfaced near Cape Poinsett. Of course, the exact trajectories and

11 drift speeds of these floats whilst under sea-ice is unknown.

12 2.2. Conductivity-Temperature-Depth System

13 The CTD system comprised of an Falmouth Scientific Institute (FSI) CTD

14 instrument, a tripod and over 1000m of polyethylene rope on a winch/drum

15 attached to a metal sled (see Figure 3). The winch system consisted of a control

16 box and was powered by a portable generator. The FSI was itself housed in

17 a protective fibreglass case to shield the sensors from rough treatment in the

18 sub-ice environment. Initial deployments demonstrated the sensitivity of the

19 instrument to the cold water and so the deployment procedure thereafter began

20 by taking the instrument down to approximately 50 m and leaving it there for

21 3–5 minutes before returning to near the surface and beginning the cast. The −1 22 rate of descent was kept below 1 m s on the way down. Post-deployment the

23 FSI was returned to the ship and the data downloaded and assessed to ensure

24 the data were copacetic. An interesting side-effect of the CTD system was that

25 the ’noise’ from the speed controller box effectively blocked out the ships HF

26 radio communications. As a result, the CTD system could not be operated

27 when helicopter operations were being conducted within 20 km of the ship.

28 Overall, the conductivity/salinity data collected by the FSI were poor (for

29 further details, please consult Rosenberg (????)). In the later stages of SIPEX,

8 1 a shipboard CTD was deployed (Seabird SBE911+ CTD) with the FSI attached

2 for calibration. The vertical profiles from this cast are shown in the top row

3 of Figure 5 and clearly show a large offset in salinity (> 0.1). The first profile

4 from the nearest Ice-Argo float (a120) show that the float performed better

5 than the FSI, closely matching the Seabird profile. An attempt to calibrate

6 all FSI profiles using the combined FSI/Seabird result was conducted with the

7 results again compared to nearby Ice-Argo profiles. This appeared successful

8 for Station 2 (float a130), in conjunction with a nearby profile from BROKE

9 (Stn 109), but not thereafter, with Stations 10 and 11 showing the ’calibrated’

10 salinity data as still much fresher than at the nearest Ice-Argo data (floats a114

11 and a125 respectively). Similar residual offsets in salinity post-calibration were

12 found when the calibrated salinity profiles at Station 8 were compared at with

13 nearby salinity data from BROKE (Stn 82). These offsets in salinity are simply

14 too large, in particular at depth below the seasonally variable surface layer, for

15 this to be accounted for by spatial or temporal variability.

16 [Figure 5 about here]

17 Upon completion of the SIPEX voyage, further testing of the FSI instrument

18 revealed that the conductivity sensor was highly sensitive to the protective fi-

19 breglass casing it was deployed in, and that minor movements (1–2mm) in the

20 orientation of the instrument relative to the casing during battery changes were

21 likely responsible for the offset and variable behavior during the voyage. For

22 this reason we dismiss the absolute value of salinity from this instrument, but

23 present it here nonetheless for its vertical structure, in particular for the stations

24 on the shelf break (lower panel, Figure 5) that were not sampled by Ice-Argo.

25 2.3. Acoustic Doppler Current Profiling System

26 The acoustic Doppler current profiling (ADCP) system was designed and

27 constructed by Marine Science Support at the Australian Antarctic Division

28 with the goal of observing the ocean current regime beneath the sea ice. To de-

29 termine the absolute ocean current, a GPS compass was included in the system

9 1 to account for sea ice drift. The system consisted of a three-legged support-

2 ing platform that included a central, downward facing retractable pylon for the

3 RDI Workhorse ADCP operating at 307.2 kHz, a flat tray for storing the power

4 supply/control system, mounting points for the radio modem antenna and an

5 upward facing extendable pylon for the Furuno SC50 GPS compass (see Figure

6 3). The observation site was preferentially in the zone perpendicular to the port

7 bow of the ship, between 50–100 m away to minimise any contamination of the

8 flow field from the ships bow thrusters.

9 At each station a hole through the sea ice was prepared and the ADCP

10 lowered up to 1.5 m, being fully submerged and clear of the base of the ice.

11 The ADCP, GPS and radio modem were connected to the 24 volt recharge-

12 able portable battery through a control box. The system was designed so the

13 Furuno GPS compass and ADCP remained in fixed position relative to each

14 other, and the entire system was positioned parallel to the ships heading. The

15 basic procedure began with the initialisation of the GPS compass and establish-

16 ing of communication with the ADCP computer on board the ship via the radio

17 modem. Thereafter the ADCP was initialised from the ship using the RDI Win-

18 River software. After initial testing determined a lack of available backscatter in

19 the water column, attributed to the scarcity of biological matter in late winter,

20 the sampling bin size was set to 8 m to achieve a maximum range of 50–100 m

21 with a 1.76 m blanking window below the transducer head. This subsequently

22 ruled out any small-scale sampling of the immediate boundary layer below the

23 ice. The ADCP collected single ping ensembles every 0.5 seconds, operating

24 using the RDI default settings.

25 The ADCP data were post-processed during the voyage using the RDI Win-

26 River software. This synchronised the GPS heading data with the ADCP ve-

27 locity data streams, producing velocity headings relative to true north and the

28 ADCP position, and removed bad data points where data were below the beam

29 correlation threshold or the water velocity exceeded the threshold of 107 cm −1 30 s . The ice mounted ADCP frame ensured that ADCP tilt was fixed near 0

31 degrees and all four ADCP beams were required to produce solutions. Ensemble

10 1 averaging over three minutes (360 pings) reduced the rms error to below 0.5 cm −1 −1 2 s for all stations except station 3, where it is below 1 cm s . The water

3 velocities relative to the ADCP were converted to absolute u and v velocities

4 by adding the ADCP (ice) drift velocity components calculated based on the 90

5 second boxcar filter smoothed GPS compass position. Due to the high clarity

6 of the water column from the relative absence of biological backscatter in late

7 winter, the ADCP maximum range was limited to between 50-100 dbar, with

8 a strong diurnal variation. All stations exhibited strong vertical coherence in −1 9 velocity with standard deviations of the shear of less than 1–5 cm s absolute

10 velocity (Table 1). Because of the relatively narrow range of depths sampled

11 and the strong vertical coherence observed, we use the calculated vertical mean

12 when referring to horizontal components of velocity.

13 [Table 1 about here]

14 2.3.1. Ocean Tides

15 As the ADCP time series were too short (maximum of 24 hours) to meaning-

16 fully estimate the tidal contribution to currents through frequency analysis, tidal

17 velocities were estimated using the Circum-Antarctic Tidal Simulation (CATS -

18 vers. 2008b) (Padman et al., 2002). Subtracting the tidal predictions from the

19 ADCP data only successfully reduced the variance of the u and v components

20 of velocity at stations 2, 8, 13 and 14, suggesting that the phase of the tide

21 model is inaccurate in the SIPEX region. This inaccuracy may also be caused

22 by non-tidal currents being in phase with the tidal signal, but without a longer

23 time series of observations this ambiguity cannot be resolved. We therefore only

24 use the tide model as a gauge of the tidal velocity magnitudes in this analysis.

25 Table 1 shows that north of the shelf break tidal velocities are approximately an

26 order of magnitude smaller than the ADCP mean current absolute velocities,

27 while over the shelf at the shallower stations, the tidal currents are of a similar

28 magnitude to the mean ADCP velocity and may significantly influence it. The

29 meridional component of station 8 in particular exhibits a in particular exhibits

30 a fluctuation over its duration (11.52 hours) that would be consistent with tidal

11 1 influence.. However, without significantly longer observations or an accurate

2 tide model we are unable to unambiguously separate out the tidal component

3 from the ADCP velocities and the velocities are left unmodified in this study.

4 3. Results

5 3.1. Water Mass Properties from ’through-ice’ CTD and Ice-Argo

6 We begin with the water mass properties observed during SIPEX and focus

7 on two main results: the detection of Modified Circumpolar Deep Water on the

8 continental shelf across fast-ice stations 5, 6 and 8, and the large-scale vertical

9 stratification, i.e. the winter mixed layer depth, across the offshore stations and

10 Ice-Argo profiles.

11 3.1.1. Modified Circumpolar Deep Water on the continental shelf north of the

12 Totten Glacier

13 The most significant oceanographic result from SIPEX is the detection of

14 mCDW on the shelf break at station 8. Figures 5 and 6 show that the bottom ◦ 15 layer of Station 8 had a 50-70m layer of mCDW that is warmer (up to 2 C)

16 and more saline (at least 0.2) relative to the water column above, and that this

17 is consistent with the mCDW layer previously recorded in summertime during

18 BROKE at station 82. The strength of the warming signal implies that the

19 Antarctic Slope Front is on or south of the shelf break, as in the western Ross

20 Sea (Whitworth III and Orsi, 2006; Muench et al., 2009). That this feature was

21 observed during SIPEX in late winter, more than ten years after the summertime

22 observation during BROKE, implies this intrusion is likely to be a persistent

23 feature.

24 [Figure 6 about here]

25 In other regions of East Antarcrtica, most notably the Ad´elieand George V

26 Land continental shelf break, mCDW intrusions are much weaker in terms of ◦ 27 their warming signal relative to the 0 C observed here (?), in particular during

12 1 winter when the mCDW is almost unidentifiable due to strong convective mixing

2 during dense shelf water formation (Williams and Bindoff, 2003). Additionally

3 the mCDW intrusions for the AGV region are lighter than the local shelf water

4 and penetrate the continental shelf water column at mid-depths (300–400m).

5 This is important because this means the heat flux associated with the mCDW

6 cannot directly influence ocean/ice shelf interactions at the grounding lines (>

7 1000m, Rignot and Jacobs (2002)) of the Mertz and in that

8 region. We note that for the the mCDW intrusions observed

9 during both BROKE and SIPEX occupy the bottom layer of the observed water

10 mass profiles. This is significant for two reasons. It suggests firstly that there

11 was no dense shelf water present in this location at the time of sampling and

12 secondly that when mCDW is the densest water mass, it can influence a greater

13 depth range of the nearby Totten Glacier which has an even deeper grounding

14 line than the Mertz and Ninnis glaciers Rignot and Jacobs (2002).

15 3.1.2. Winter Mixed Layer Depth over the Continental Slope and Rise

16 As introduced earlier, the winter mixed layer depth (WML) defining the

17 base of the AASW layer is the product of the current season’s sea ice produc-

18 tion and the local mesoscale circulation on the continental shelf. It is observed

19 to exhibit a meridional structure from summertime surveys, shoaling north of

20 upper continental slope to form the horizontal density gradient of the Antarctic

21 Slope Front and generate the baroclinic Antarctic Slope Current. One of the

22 goals of SIPEX was to firstly determine if this meridional structure exists im-

23 mediately after the sea ice production season, and secondly to examine whether

24 any large-scale variability in the winter mixed layer depth could be attributed

25 to the spatial variability of sea-ice production across the survey region.

26 [Figure 7 about here]

27 We separate the sampling stations into eastern and western sectors of the

28 SIPEX survey and start by presenting vertical profiles of potential temperature

29 for the eastern sector in Figure 7. The WML is relatively shallow (∼100m)

13 1 at the northern stations, i.e. a127, 3 and 2 and then there is a transition to

2 the deeper WML (from ∼100 to sim200m) across the Antarctic Slope Front

3 between the initial profiles for Ice-Argo float a130 and a115 to a117, northeast

4 of the Voyeykov polynya. The profile from Station 2 is unique in that there is a

5 double-step across the permanent thermocline separating the base of the WML

6 and the mCDW below, indicative of advection.

7 [Figure 8 about here]

8 [Figure 9 about here]

9 The vertical profiles from the western sector of the survey (see Figure 8)

10 show a deepening of the WML depth, relative to the eastern sector, to more than

11 300m. The deepest WML are at FSI-CTD stations 11, 13 and 14. The spatial

12 variability in both WML depth and sub-surface Tmax below the WML during

13 late winter and summer from SIPEX CTD and Ice-Argo respectively is shown

14 in Figure 9. There is a shift to shallower WML depths from south to north,

15 similar to that observed in summer (Figure 2), in relation to the meridional

16 barotropic fields associated with the Antarctic Slope Current. The region of

17 deepest WML, extending furthest offshore from the shelf break is found between ◦ 18 115–120 E. This is northwest of the Dalton Iceberg Tongue polynya region and

19 suggests a link to the ice production therein. However the northern extent of

20 the ’deep’ WML is most likely due to the northward shift in the bathymetry

21 in that region deflecting the Antarctic Slope Current to the north-west. The

22 spatial distribution of sub-surface Tmax below the WML also clearly indicates

23 the southern migration of the warmest mCDW/CDW east of this region at ◦ 24 120 E. The lack of true repeat sampling between BROKE and SIPEX disallows

25 definitive comparison of seasonal and decadal variation.

26 3.2. Ocean Currents from the ’Through-Ice’ ADCP System

27 3.2.1. Offshore stations

28 Of the observations made north of the shelf break, over bathymetry of 1500-

29 4000 m, three stations (2, 10 and 14) were conducted for periods close to 24

14 1 hours, while the remaining three (1, 3 and 13) all lasted for less than seven hours.

2 All stations are highly coherent in the vertical with RMS differences between

3 the mean heading and the heading at individual bin depths being between 5–

4 12 degrees for all stations. The influence of tides is small north of the shelf −1 5 break (mean magnitude of order 1 cm s , Table 1) and their removal does not

6 influence the heading or magnitude of the currents observed at these stations

7 significantly. All statistics and figures show the observations with the tidal

8 component included.

9 [Table 1 about here]

10 [Figure 10 about here]

11 The column mean speed and heading of the water beneath the sea ice at

12 the long period stations are shown in Figure 10. The top 100m below the sea −1 13 ice tends to be relatively fast (mean magnitude 6–17 cm s ), with maximum −1 14 speeds of over 30 cm s at stations 2, 3 and 14. At the three stations occupied

15 for longer than 7 hours the currents vary strongly on time scales of a few hours

16 in both magnitude and direction. This is particularly apparent at stations 10

17 and 14 where the flow reverses direction over a period of less than five hours.

18 Similar sharp changes in flow direction occur at station 3, although the change

19 in heading is not as dramatic. In almost all cases these changes occur over

20 relatively short times and are separated by longer periods where the heading and

21 velocity is relatively steady. This non-continuous nature of the ocean transport

22 heading, in combination with the fact that changes are almost universally in a

23 clockwise sense argues against these being the result of inertial oscillations of

24 the water column. −1 −1 25 A mean westward velocity of 3-33 cm s (variance 1.3–9.7 cm s ) is ob-

26 served at all of the offshelf stations, which is consistent with other hydrographic

27 observations (Bindoff et al., 2000) and the presence of the ASC north of the

28 shelf break. Periods of eastward transport do occur at stations 3, 10 and 14,

29 but with the exception of station 14, all are relatively brief. Less consistency is

15 −1 1 apparent in the meridional component (variance 2.6–18.1 cm s ) and the long

2 period stations all exhibit periods of both northward and southward transport −1 3 with magnitudes greater than 10 cm s . However, the northward transport

4 tends to dominate and the net flux over the observed periods is northwards at

5 all stations except station 10, where it is southward (mean meridional velocity −1 6 -4 cm s ).

7 3.2.2. Fast ice stations near the continental shelf break

8 The stations occupied over the shelf (<800 m) were distinctly different from

9 those north of the shelf break. Smaller currents were encountered (2.3–6.4 cm −1 10 s ) and tidal velocities were stronger with mean magnitudes of up to 3 cm −1 −1 11 s and peak velocities of over 6 cm s . The mean current direction was also

12 more variable. While offshelf currents were broadly northwestward, there was

13 no consistent mean heading between the three onshelf stations, although rela- −1 14 tively consistent eastward currents of 1.6–5.8 cm s were observed at stations

15 6 and 8. Additionally, the onshelf current meters are significantly less vertically

16 coherent than those offshelf, with a RMS heading difference between the mean

17 water column heading and individual bin headings of 22–40 degrees. This re-

18 duction of vertical coherence may be a result of the generally weaker currents

19 and reduced signal to noise ratios. The presence of tidal Ekman spirals may

20 also reduce vertical coherence but the short period of observations (< 18 hours)

21 and relatively large ADCP bin sizes (8 m) means there is insufficient data to

22 resolve such vertical structures reliably.

23 [Figure 11 about here]

24 3.2.3. ADCP range as proxy for the diel migration and relative abundance of

25 zooplankton during SIPEX

26 Previous studies have used ADCP backscatter to detect zooplankton(Cresswell

27 et al., 2009; Kaneda et al., 2002). For the SIPEX results, the clear diurnal

28 variation in ADCP range invites a qualitative assessment, albeit extremely sim-

29 plified, of the diel migration of zooplankton beneath sea-ice in this region in

16 1 late-winter. Figure 12 shows the diel variation in ADCP backscatter and rela-

2 tive distribution of maximum ADCP range across the SIPEX survey. Given the

3 basic assumption that ADCP range is directly proportional to the presence of

4 zooplankton, in particular in late-winter before the sea ice melts and contributes

5 detritus to the backscatter of the water columns, then we find that diel migra-

6 tion of zooplankton was occurring beneath sea ice in late winter. The presence

7 of zooplankton was confirmed by direct observations/sampling through the ice

8 holes and camera surveys conducted after each ’through-ice’ CTD. If we make

9 the further assumption that the species of zooplankton and associated acoustic

10 backscatter was constant, then we also find that there was greater abundance of

11 zooplankton at the offshore stations relative to the stations on the shelf break

12 beneath the fast-ice.

13 [Figure 12 about here]

14 3.3. Wind, Ocean Current and Sea Ice Drift

15 Sea ice drift is the result of the summation of the forces acting upon it,

16 most notably the wind, ocean currents and coriolis force, and to a less extent

17 internal ice dynamics (particularly in regions of weak sea ice concentration) and

18 sea surface tilt. The GPS compass used in the ADCP system to determine the

19 absolute ocean current provides highly accurate measurements of sea ice drift.

20 Simultaneous measurements of wind from the underway system on board the

21 RV Aurora Australis present an opportunity to assess the influence of wind

22 and ocean currents on the sea ice drift measured during SIPEX ice stations.

23 The wind data were taken as a maximum of ten minute averages collected on

24 both the port and starboard side of the ship, at ∼ 10m height.

25 In Figure 13 we present the speed and adjusted heading for ocean current,

26 wind and ice drift at the longer period stations (2, 10 and 14). The heading ◦ 27 is adjusted to ensure there are no step across the 359⇒0 transition. There

28 is a characteristic increase in speed at the end of each station when the winds

29 reached 30 knots and the ice stations deemed unsafe. There are clearly strong

17 1 correlations between the components of the ice/ocean/wind system, with the

2 speeds of all three varying together closely. Additionally there is also an obvious

3 dependence between the heading of the ice and the headings of the wind and

4 ocean. However, it is not always clear whether the wind or ocean is dominant

5 in driving the ice heading, or that the wind dictates the surface ocean heading,

6 although the correlated increase in speed does suggest this may be the case. For

7 example, the ice and ocean headings are almost exactly matched at station 2,

8 while the wind heading is almost 100 degrees counter clockwise of these, much

9 greater and opposite to what might be expected if the wind was forcing an

10 Ekman rotation. In a counter example, the ice and wind heading are closely

11 matched at the beginning of the station 10 time series, but suddenly diverge

12 and the ocean heading appears to drive the ice between day 273.6–273.7. The

13 ocean and wind rotate in opposite directions at station 10 at day 273.6, as well

14 as at station 14 between days 280.6–280.7. Otherwise, however, the ocean and

15 wind generally have very similar headings and trends.

16 [Figure 13 about here]

17 Qualitatively this unclear relationship is continued in Figure 14, where we

18 see the progressive vector diagrams for ice, ocean and wind tracers at SIPEX

19 stations 1, 2, 3, 10, 13 and 14. For the wind tracer we use 2% of the wind

20 speed. Of the longer period stations, stations 2 and 14 show the ocean as more

21 dominant, and the wind more dominant on station 10. The resultant of the

22 combined ocean and wind vectors agreed best with ice drift at Station 13. In

23 Figure 15 we present the principal component and correlation analysis for the

24 ice drift, ocean current and wind time series shown in Figure 13. At station 2,

25 the percentage of variation explained by the 1st principal component of speed

26 for ice/ocean and ice/wind pairs was 89 and 81% with similarly large overlaps

27 of correlation at other stations. This demonstrates that there is a significant

28 degree of coupling between the wind-ocean system, suggesting that the wind

29 acts to substantially influence the surface ocean flow on the time and spatial

30 scales observed here.

18 1 [Figure 14 about here]

2 [Figure 15 about here]

3 4. Discussion

4 The most interesting and important water mass result from SIPEX was the ◦ ◦ 5 observation of modified Circumpolar Deep Water at 117 E warmer than 0 C.

6 Bindoff et al. (2000) reported similar intrusions of mCDW near this location

7 in summer (Jan. 1996) and noted that this was the primary location sampled ◦ 8 during the BROKE survey between 80–150 E to have this feature. Williams and

9 Bindoff (2003) showed ’highly’ modified circumpolar deep water in the Ad´elie

10 Depression in winter that crossed the shelf break at mid-depths above the denser

11 shelf water below. Williams et al. (2008a) showed that these intrusions increased

12 dramatically in summer after the sea ice production season. However our result

13 here finds mCDW that is much warmer and bears more similarity to shelf break

14 exchanges in the western Ross Sea and Amundsen sea where the Antarctic Slope

15 Front extends onto the continental shelf. Clearly two observations of mCDW

16 in the vicinity of each other over a 10 year period does not prove that these

17 warm, saline intrusions have been constant over this time period. Nonetheless

18 this additional observation of mCDW in late winter suggests that these mCDW

19 intrusions are not limited to the austral summer as in other locations and we

20 speculate that this strengthens the case for for these mCDW intrusions being

21 persistent intra- and inter annually.

22 This is very relevant to the melting of continental ice in this region as mCDW

23 provides the necessary heat flux. In addition to the strength of the mCDW

24 intrusion in this region, the density of these intrusions relative to the local shelf

25 water is also critical. To truly influence the melting of ice shelves, the mCDW

26 must be dense enough to reach the grounding line. For the in

27 the Ad´elieDepression, the mCDW intrusions occur at mid-depths and occupy

28 the upper layer of the water column, providing some heat flux to the base of

29 the floating glacier (300–400m), but have little influence on the groundling line

19 1 (>1200m). Here, the mCDW is the densest water mass occupying the bottom

2 layer, implying that should it continue its path south, it could reach nearby

3 grounding lines. Though the East Antarctic is the most stable part

4 of Antarctica, the nearby Totten and Moscow University IceShelf were reported

5 by Rignot and Thomas (2002) to be in a state of negative mass balance. Most

6 recently Pritchard et al. (2009) report that the glacial thinning in this region

7 could be three times more than previously estimated. Enhanced ocean heat

8 flux from a persistent, bottom-trapped inflow of mDCW would certainly play a

9 major role in such changes.

10 There are two mechanisms that could contribute to mCDW being the densest

11 water mass on the shelf in this location, despite its warm temperature. The

12 influence of cold, fresh Ice Shelf Water (ISW), from ocean/ice shelf interactions

13 beneath the nearby Totten Glacier and Moscow Univsersity Ice Shelf, is likely

14 to provide a freshening feedback to the local shelf water properties. In addition

15 the sea-ice production in the nearby polynya regions may simply be insufficient

16 to produce truly dense shelf water. To examine this second point further we

17 present the sea ice production for the polynyas identified in this region in Figure 3 18 16. The Cape Poinsett and Dibble Iceberg Tongue polynyas produce ∼60 km

19 of sea ice each year. These values are relatively low in comparison with the

20 total sea ice production in the Mertz Glacier and Cape Darnley polynya/dense 3 21 shelf water formation regions (∼ 180 km ), though the region east of the Mertz 3 22 Glacier has similar production to the Dibble (∼ 60 km ) and this has been

23 shown to be sufficient to produce low salinity shelf water cascades resulting in

24 Antarctic Bottom Water (Williams et al., 2010b). The Dalton Iceberg Tongue 3 25 polynya is the next largest at 40 km , with the Voyeykov and Blodgett averaging 3 26 20 km per year, and although are persistent with relatively small interannual

27 variability, are most likely too small to generate sufficient dense shelf water for

28 Antarctic Bottom Water, let alone in conjunction with negative feedback from

29 the fresh water flux from nearby ice shelves.

30 [Figure 16 about here]

20 1 However we note that the it has been shown in modelling studies (?) that

2 dense shelf water overflows and mCDW intrusions can be intrinsically linked.

3 That is, cascades of dense shelf water over the shelf break can induce a compen-

4 sating upwelling of mCDW. Following this idea, there may yet be dense shelf ◦ 5 water formation, upstream or downstream of the mDCW intrusion at 117 E.

6 The BROKE survey detected evidence of dense shelf water overflows and AABW

7 production downstream from the Dibble Iceberg Tongue polynya in the elevated

8 bottom layer concentrations of offshore CFC-11 (Williams et al., 2010b). We

9 demonstrate this in Figure 17, which shows the bottom layer in meridional

10 vertical sections of potential temperature across the BANZARE and Sabrina ◦ 11 coasts. The Antarctic Bottom Water layer, in particular the -0.3 C isotherm ◦ 12 shoals to 1500m at 128 E to the west of the Dibble and then there is a sub- ◦ 13 sequent decrease in temperature below 3000m at 120 E. Bindoff et al. (2000)

14 reported evidence of modified Shelf Water on the upper continental slope west

15 of the Cape Poinsett polynya region. It is likely that given its modest amount of

16 sea-ice production and the significant amount of melting occurring locally that

17 there is very little dense shelf water being exported from the Dalton Iceberg

18 Tongue Polynya.

19 [Figure 17 about here]

20 As enticing as the concurrent measurements of sea ice drift, ocean currents

21 and wind were for examining the comparative influence of wind and ocean on

22 sea ice drift, the mixed results in Figures 13–15 reflected the fact that our time

23 series were simply too short to address this conclusively. At times the ocean

24 sometimes appears to drive the ice, and sometimes the wind (and sometimes

25 both). Similarities in the ocean and wind headings/velocities and overlapping

26 correlation coefficients strongly indicate that the wind is important in driving

27 the ocean, and hence directly or indirectly the sea ice too. However as men-

28 tioned, this is only very short space/time scales, i.e. 24 hours. Over longer

29 timescales and in the presence of stronger ocean currents such as the Antarctic

30 Slope Current, the ocean is clearly dominating, as indicated by the Ice Argo

21 1 drift pattern in Figure 4 and reported by Heil et al. (2010, this volume) from

2 drifting buoys. South of the shelf break, in the absence of strong currents, the

3 synoptic wind patterns, tidal signals and clear inertial oscillations shown by Heil

4 et al. (2010, this volume) show that sea ice ice drift is much more of a random

5 walk than in locations where there are strong ocean currents. Ideally future

6 deployments of this ’through-ice’ ADCP system can be run for much longer

7 periods and perhaps even modified to run autonomously over several months.

8 4.1. Future work and challenges for observing winter oceanography in the SIZ

9 The pilot ’through-ice’ oceanography program initiated on SIPEX was suc-

10 cessful but can be improved in two key areas. Firstly the accuracy of the salinity

11 measurements needs to be improved to facilitate greater comparison of results

12 with more precise ship-based CTD observations. Secondly, the length of ADCP

13 sampling must be increased to better understand the processes driving the intra-

14 diurnal variability found during SIPEX. Perhaps the best result during SIPEX

15 came from the Ice Argo floats deployed on behalf of the University of Wash-

16 ington. The majority of these floats have continued to work over two years

17 capturing two seasonal cycles. This dataset will greatly improve the under-

18 standing of the seasonal formation and partial decay of the winter mixed layer

19 each year and will provide a comprehensive benchmark against which the cur-

20 rent and future generations of climate models can compare with. However the

21 Ice-Argo floats remain limited to sampling the region north of the continental

22 slope, that is north of the 3000-m isobath, and for this reason the ’through-ice’

23 CTD system used during SIPEX is a relatively inexpensive and easy way to

24 sample the regions on the upper slope and on the continental shelf break during

25 sea-ice research voyages.

26 The biggest challenge that remains after not being met on SIPEX is the

27 effective wintertime sampling of the Antarctic Slope Front and Antarctic Slope

28 Current region beneath the sea ice between the shelf break and the 1500-m iso-

29 bath. We can report that the Antarctic Slope Current was ’experienced’ by the

30 RV Aurora Australis when heading south towards the continental shelf break

22 1 when it began drifting within a ’river of ice’ that was several tens of metres wide.

2 The ship was in fact powerless to manuever efficiently due to the loose nature

3 of the crushed sea ice within it and for several minutes found itself drifting

4 helplessly west, less than a kilometre from numerous large icebergs grounded on

5 the shelf break, until it eventually regained thrust and escaped, with all further

6 requests to conduct CTDs in this region declined. Some modelling studies have

7 found that the Antarctic Slope Current increases in winter beneath the sea-ice

8 (Mathiot et al., 2009), perhaps in response to the increased winds and the in-

9 creased air/ice drag coefficient. More work is needed to understand the physical

10 mechanisms behind this phenomena and its seasonal variability. However in

11 situ observations of the ASC in winter remain a challenge.

12 Acknowledgements

13 We would like to thank the officers and crew of the RV Aurora Australis

14 and the Marine Science Support staff from the Australian Antarctic Division for

15 their professionalism and support during the SIPEX voyage. The majority of

16 this equipment used in the CTD system was previously used during the Amery

17 Ice Shelf drilling project and Russell Brand constructed the overall system. The

18 ADCP system was designed and built by the Marine Science Support group at

19 the Australian Antarctic Division, with the loan of the ADCP kindly arranged

20 by Mike Craven. Mark Rosenberg completed the post-processing of the FSI

21 CTD data. Bathymetry data provided by ETOPO1 (2009). Ice Argo floats

22 deployed during SIPEX and presented in this paper were from the University

23 of Washington and we would like to thank Steve Riser, Annie Wong, Dana

24 Swift and Rick Rupin for their efforts in providing these data. These data were

25 collected and made freely available by the International Argo Project and the na-

26 tional programs that contribute to it.(http://www.argo.ucsd.edu,http://argo.jcommops.org).

27 Argo is a pilot program of the Global Ocean Observing System.

28

29 References

30 Aoki, S., 2003. Seasonal and Spatial Variations of Iceberg Drift off Dronning

23 1 Maud Land, Antarctica, Detected by Satellite Scatterometers. Journal of

2 Oceanography 59, 629–635.

3 Bindoff, N. L., Rosenberg, M. A., Warner, M. J., 2000. On the circulation and

4 water-masses over the Antarctic continental slope and rise between 80 and ◦ 5 150 East. Deep-Sea Research II 47, 2299–2326.

6 Cresswell, K. A., Tarling, G. A., Thorpe, S. E., Burrows, M. T., Wiendenmann,

7 J., 2009. Diel vertical migration of Antarctic krill (Euphasia superba) is flex-

8 ible during advection across the Scotia Sea. Journal of Plankton Research

9 doi:10.1093/plankt/ecss.fbp062, –.

10 ECMWF, 2009. ERA-Interim Reanalysis.

11 URL http://www.ecmwf.int/research/era/do/get/era-interim

12 ETOPO1, 2009. Global Predicted Bathymetry v11. Smith, W. H. F. and D.

13 Sandwell, Global seafloor topography from satellite altimetry and ship depth

14 soundings, science, 277, p.1956-1962, 1997.

15 URL http://topex.ucsd.edu/marine topo/

16 Fahrbach, E., Rohardt, G., Krause, G., 1992. The Antarctic Coastal Current in

17 the southeastern Weddell Sea. Polar Biology 12, 171–182.

18 Heil, P., Allison, I., Massom, R., Worby, A., 2010, this volume. Effects of in-

19 tensified atmospheric forcing and near-coastal currents on meso-scale sea-ice

20 dynamics off East Antarctica. Deep-Sea Research II.

21 Heywood, K. R., Locarnini, R. A., Frew, R. D., Dennis, P. F., King, B. A.,

22 1998. Transport and water masses of the antarctic slope front system in the

23 eastern Weddell Sea. Antarctic Research Series 75. S. Jacobs and R. Weiss.

24 Washington, American Geophysical Union 75, 203–214.

25 Jacobs, S. S., 1991. On the Nature and Significance of the Antarctic Slope Front.

26 Marine Chemistry 35(1-4), 9–24.

24 1 Kaneda, A., Takeoka, H., Koizumi, Y., 2002. Periodic Occurence of Diurnal Sig-

2 nal of ADCP backscatter strength in Uchiumi Bay, Japan. Estaurine, Coastal

3 and Shelf Science 55, doi:10.1006/ecss.2001.0908, 323–330.

4 Klatt, O., Boebel, O., Farhbach, E., 2007. A Profiling Float’s Sense of Ice. Jour-

5 nal of Atmospheric and Oceanic Technology 24, DOI: 10.1175/JTECH2026.1,

6 1301–1308.

7 Massom, R., 2003. Recent iceberg calving events in the Ninnis Glacier Region,

8 East Antarctica. Antarctic Science 15 (2), 303–313.

9 Mathiot, P., Barnier, B., Galle, H., Molines, J., Penduff, T., 2009. Correction of

10 katabatic winds in era40 and its effect on polynya and shelf water in antarc-

11 tica. Ocean Modelling submitted.

12 Meijers, A., Klocker, A., Bindoff, N. L., Williams, G. D., Marsland, S. J., 2010. ◦ 13 The large-scale circulation off the east antarctic coast (30–80 E). Deep-Sea

14 Research II BROKE-West Special Volume, in press, –.

15 Muench, R. D., Padman, L., Gordon, A. L., Orsi, A., 2009. A dense shelf outflow

16 from the ross sea, antarctica: Mixing and the contribution of tides. Journal

17 of Marine Systems in press, doi: 10.1016/j.jmarsys.2008.11.003.

18 NODC, 2009. Operational Oceanography Group: Global Argo Data Reposi-

19 tory. April 2007. U.S. Department of Commerce, National Oceanic and Atmo-

20 spheric Administration, National Oceanographic Data Center, Silver Spring,

21 Maryland, 20910. 2009.

22 URL http://www.nodc.noaa.gov/argo/accessData.htm

23 Office, A. P., 2010. Argo home page.

24 URL http://www.argo.ucsd.edu/

25 Orsi, A. H., Whitworth, T., Nowlin, W., 1995. On the meridional extent and

26 fronts of the Antarctic Circumpolar Current. Deep-Sea Research Part I 42,

27 641–673.

25 1 Padman, L., Fricker, H. A., Coleman, R., Howard, S., Erofeeva, S., 2002. A new

2 tidal model for the Antarctic ice shelves and seas. Annals of 34,

3 247–254.

4 Pritchard, H. D., Arthern, R. J., Vaughan, D. G., Edwards, L. A., 2009. Ex-

5 tensive dynamic thinning on the margins of the Greenland and Antarctic ice

6 sheets. Nature doi:10.1038/nature08471, 1–5.

7 Rignot, E., Jacobs, S. S., 2002. Rapid Bottom Melting Widespread near Antarc-

8 tic Ice Sheet Grounding Lines. Science 296, 2020–2023.

9 Rignot, E., Thomas, R. H., 2002. Mass Balance of Polar Ice Sheets. Science 297

10 (5586), 1502–1506.

11 Rosenberg, M., ???? Sipex, Marine Science Cruise AU0701 - FSI CTD data

12 processing. Tech. rep.

13 Sala, A., Azzali, M., Russo, A., 2002. Krill of the Ross Sea: distribution, abun-

14 dance and demography of Euphausia superba and Euphausia crystallorophias

15 during the Italian Antarctic Expedition (January-February 2000). Scientia

16 Marina 66, 123–133.

17 Smethie, W., Jacobs, S., 2005. Circulation and melting under the Ross Ice Shelf:

18 estimates from evolving CFC, salinity and temperature fields in the Ross Sea.

19 Deep-Sea Research II 52, 959–978.

20 Tamura, T., Ohshima, K. I., Markus, T., Cavalieri, D., Nihashi, S., Hirasawa,

21 N., 2007. Estimation of thin ice thickness and detection of fast ice from SSM/I

22 data in the Antarctic Ocean. Journal of Atmospheric and Oceanic Technology

23 24, 1757–1772.

24 Tamura, T., Ohshima, K. I., Nihashi, S., 2008. Mapping of sea ice pro-

25 duction for Antarctic coastal polynyas. Geophysical Research Letters 35,

26 doi:10.1029/2007GL032903.

26 1 UW, 2009. University of Washington ARGO Profiling Drifters.

2 URL http://flux.ocean.washington.edu/argo/info/general-info.shtml

3 Whitworth, T., Orsi, A., Kim, S.-J., W. D. Nowlin, J., Locarnini, R., 1998.

4 Water masses and mixing near the Antarctic Slope Front. Antarctic Research

5 Series 75. S. Jacobs and R. Weiss. Washington, American Geophysical Union

6 75, 1–27.

7 Whitworth III, T., Orsi, A. H., 2006. Antarctic Bottom Water production

8 and export by tides in the Ross Sea. Geophysical Research Letters 33, doi:

9 10.1029/2006GL026357.

10 Williams, G., Nicol, S., Aoki, S., Meijers, A., Bindoff, N., Iijima, Y., Mars-

11 land, S., Klocker, A., 2010a. Surface oceanography of BROKE-West, along ◦ 12 the Antarctic margin of the south-west Indian Ocean (30–80 E). Deep-Sea

13 Research II BROKE-West Special Volume, in press.

14 Williams, G. D., Aoki, S., Jacobs, S., Rintoul, S., Tamura, T., Bindoff, N.,

15 2010b. Antarctic Bottom Water from the Ad´elieand George V Land coast, ◦ 16 East Antarctica (140–149 e). Journal of Geophysical Research in press, –.

17 Williams, G. D., Bindoff, N. L., 2003. Wintertime oceanography of the Ad´elie

18 Depression. Deep-Sea Research II: Topical studies in Oceanography Recent

19 investigations of the Mertz Polynya and George Vth Land continental margin,

20 East Antarctica, 1373 – 1392.

21 Williams, G. D., Bindoff, N. L., Marsland, S. J., Rintoul, S. R., 2008a. Formation

22 and export of dense shelf water from the Adlie Depression, East Antarctica.

23 Journal of Geophysical Research 113, C04039, doi:10.1029/2007JC004346.

24 Williams, G. D., Nicol, S., Raymond, B., Meiners, K., 2008b. On the sum-

25 mertime mixed layer development in the marginal sea-ice zone off the Maw-

26 son coast, East Antarctica. Deep Sea Research Part II: Topical Studies in

27 Oceanography. Dynamics of Plankton, Krill, and Predators in Relation to En-

28 vironmental Features of the Western Antarctic Peninsula and Related Areas:

27 1 SO GLOBEC Part II. Volume 55, Issues 3-4. DOI:10.1016/j.dsr2.2007.11.007,

2 365–376.

3 Worby, A., Steer, A., Leiser, J., Galin, N., Yi, D., Allison, I., Heil, P., Massom,

4 R., Zwally, H., 2010, this volume. Regional-scale sea ice and snow thickness

5 distributions from in situ and satellite measurements over East Antarctica

6 during the Sea Ice Physics and Ecosystem eXperiment (SIPEX). Deep-Sea

7 Research II.

28