ArticlePDF Available

Abstract and Figures

The marginal ice zone is the dynamic interface between the open ocean and consolidated inner pack ice. Surface gravity waves regulate marginal ice zone extent and properties, and, hence, atmosphere-ocean fluxes and ice advance/retreat. Over the past decade, seminal experimental campaigns have generated much needed measurements of wave evolution in the marginal ice zone, which, notwithstanding the prominent knowledge gaps that remain, are underpinning major advances in understanding the region’s role in the climate system. Here, we report three-dimensional imaging of waves from a moving vessel and simultaneous imaging of floe sizes, with the potential to enhance the marginal ice zone database substantially. The images give the direction–frequency wave spectrum, which we combine with concurrent measurements of wind speeds and reanalysis products to reveal the complex multi-component wind-plus-swell nature of a cyclone-driven wave field, and quantify evolution of large-amplitude waves in sea ice. Unprecedented 3D imaging of waves and sea ice floes from a moving icebreaker in the Antarctic marginal ice zone during a polar cyclone reveals a complex wind-plus-swell sea state, where contrasting ice-driven attenuation and wind forcing coexist.
This content is subject to copyright. Terms and conditions apply.
nature communications
Article https://doi.org/10.1038/s41467-022-32036-2
Three-dimensional imaging of waves and
oes in the marginal ice zone during
acyclone
Alberto Alberello
1,2,14
,LukeG.Bennetts
1
, Miguel Onorato
3,4
,
Marcello Vichi
5,6
, Keith MacHutchon
7
,ClareEayrs
8
,
Butteur Ntamba Ntamba
9
, Alvise Benetazzo
10
, Filippo Bergamasco
11
,
Filippo Nelli
12
, Rohinee Pattani
13
,HansClarke
2
,IppolitaTersigni
2
&
Alessandro Toffoli
2
The marginal ice zone is the dynamic interface between the open ocean and
consolidated inner pack ice. Surface gravity waves regulate marginal ice zone
extent and properties, and, hence, atmosphere-ocean uxes and ice advance/
retreat. Over the past decade, seminalexperimentalcampaigns have gener-
ated much needed measurements of wave evolution in the marginal ice zone,
which, notwithstanding the prominent knowledge gaps that remain, are
underpinning major advances in understanding the regions role in the climate
system. Here, we report three-dimensional imaging of waves from a moving
vessel and simultaneous imaging of oe sizes, with the potential to enhance
the marginal ice zone database substantially. The images give the
directionfrequency wave spectrum, which we combine with concurrent
measurements of wind speeds and reanalysis products to reveal the complex
multi-component wind-plus-swell nature of a cyclone-driven wave eld, and
quantify evolution of large-amplitude waves in sea ice.
Improved observational capabilities are needed to understand the
often paradoxical and bafing regional and inter-annual variabilities of
Antarctic sea ice1,2. Autonomous platforms that operate in harsh polar
environments, such as autonomous underwater vehicles3and drones4,
are pushing the boundaries for in-situ observations, generating data
for essential calibration and validation of satellite remote sensing, and
measuring properties beyond the capabilities of contemporary satel-
lites. The marginal ice zone (MIZ), which is characterised by dynamic
interactions between large-amplitude surface waves and relatively
small and thin ice oes, is difcult for satellites to capture5,6and a
major target for improved observations7,8. Wave evolution and ice
properties in the MIZ are intimately coupled911,and,hence,thereis
demand for a technology capable of simultaneously monitoring both
wave activity and ice cover properties, which can capture data during
storms when waveice interactions are most intense.
Historical in-situ measurements of waves in the MIZ show the ice
cover attenuates wave energy exponentially over distance12 at a
frequency-dependent rate that induces a downshift of the peak
frequency13, as well as modifying the directional wave spectrum14.
The attenuation rate has become a research focus, as it informs
Received: 28 September 2021
Accepted: 12 July 2022
Check for updates
1
University of Adelaide, 5005 Adelaide, SA, Australia.
2
The University of Melbourne, 3010 Parkville, VIC, Australia.
3
Università di Torino, 10125 Torino, Italy.
4
INFN, 10125 Torino, Italy.
5
Department of Oceanography, Universityof Cape Town, 7701Rondebosch,South Africa.
6
Marine and Antarctic Research centre for
Innovation and Sustainability (MARIS), University of Cape Town, 7701 Rondebosch, South Africa.
7
Department of Civil Engineering, University of Cape Town,
7701 Rondebosch, South Africa.
8
New York University Abu Dhabi, Abu Dhabi, United Arab Emirates.
9
Cape Peninsula University of Technology, 7535 Cape
Town, South Africa.
10
Istituto di Scienze Marine, Consiglio Nazionale delle Ricerche, 30122 Venice, Italy.
11
Università CaFoscari, 30123 Venice, Italy.
12
Swinburne University of Technology, 3022 Hawthorn, Australia.
13
Atkins, SW1E 5BY London, United Kingdom.
14
Present address: University of East Anglia,
NR4 7TJ Norwich, UK. e-mail: alberto.alberello@outlook.com
Nature Communications | (2022) 13:4590 1
1234567890():,;
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
predictions of the width of the ice-covered region impacted by waves
and, hence, the MIZ extent15,16. Major advances in measuring wave
attenuation in the MIZ have been made over the past decade,
including dedicated campaigns in the Arctic17 and Antarctic11,18.State-
of-the-art in-situ measurements mostly come from arrays of wave
buoys, where the buoys can be traditional open water buoys19,20
deployed between oes in regions of low ice concentration (usually
close to the ice edge) or bespoke buoys deployed on ice oes11,18,20
large enough to support the buoys (usually away from the ice edge)
but small enough that the oes followthe waves. Attenuation rates are
generally calculated by applying an exponential decay ansatz to
measurements provided by neighbouring buoys, in terms of the sig-
nicant wave height11,18,20 or a more detailed analysis in which
the ansatz is applied to each component of the one-dimensional
(frequency) wave spectrum, under the assumption of a stationary
wave eld2023.
The recent surge in measurements (including remote sensing24,25)
has generated a new understanding of wave attenuation in the
MIZ, particularly on how the wave attenuation rate depends on
frequency21,26. Certain theoretical models reproduce the observed
frequency dependence, but the dominant sources of attenuation are
still hotly debated27 and empirical models often rely upon15,28.Further,
the measurements have revealed a large range of attenuation rates23,25,
even at comparable frequencies, which is attributed to dependence on
ice cover properties, such as ice thickness, areal ice concentration
and oe sizes, as well as momentum transfer from winds over the ice
cover. Satellite and model-hindcast data have been used to derive
empirical relationships between measured attenuation rates and ice
concentration18,22,23, ice thickness29 and winds22,23.Incontrast,oe sizes
in the MIZ are below satellite resolutions and have only recently been
integrated into large-scale models16,30, so that coincident oe size data
have been limited to visual observations during deployment. Overall,
data on ice properties are too sparse or unreliable to validate theore-
tical models.
Stereo-imaging techniques are emerging as a tool for in-situ
monitoring of waves and ice properties in the MIZ. In principle, the
images can be used from a moving vessel, as in open waters, to
reconstructthe sea surface elevation in timeand space, thus enabling
analysis of wave dynamics in two-dimensional physical space, the
frequencydirection spectral domain, and wave statistics31.Airborne
synthetic aperture radar (SAR) is an alternative method to measure
frequencydirection wave spectra and has been applied over
6080km long transects of the MIZ32,33. However, stereo-imaging,
being an in-situ technique, can be used to measure sea-ice geome-
trical properties simultaneously34 and can be combined with co-
located meteorological measurements, e.g., wind velocities. Further,
in contrast to SAR35,stereo-imaging resolves wind sea components of
the wave spectrum (short wavelength systems under the inuence of
local winds), as well as swell (long wavelength systems no longer
under the effect of winds).
To date, the use of stereo-imaging techniques in the MIZ has been
limited in scope. Campbell et al.36 use a camera system on a xed
platform on the edge of a lake to quantify incoming and reected
energy uxes of relatively small waves (<0.3m) in pancake and brash
ice. Smith & Thomson37 use camera images froma moored vesselin the
Arctic MIZ during calm conditions (signicant wave heights typically
around 1m) to calculate bulk wave properties and pancake oe velo-
cities. Alberello et al.34 use an autonomous stereo-camera system on a
vessel moving through the winter Antarctic MIZ during a cyclone to
measure pancake oe shapes and sizes.
In this article, we demonstrate the potential to monitor the evo-
lution of the frequencydirection wave spectrum from the images
captured by Alberello et al.34 combined with automated image recon-
struction software. We report the extreme sea state created by the
cyclone over a >40km transect into the Antarctic MIZ, and validate a
subset of the results with co-located buoy measurements. The sea state
deep into the MIZ during the cyclone is shown to be more complex
than previously thought, and partitioning of the two-dimensional
spectra is required to analyse wave evolution of the cyclone-driven
wind sea.Further, evidence is given of momentum transfer from winds
through 100% ice concentration, based on comparing attenuation of
the signicant wave height over distance with an empirical model18,
which is considered to be a benchmark due to the large size of the
underlying data set and that the measurements were made in the
Antarctic MIZ during the sea-ice growth period.
Results
Experimental conditions
On the 4th July 2017, the South African icebreaker S.A. Agulhas II
entered the Antarctic MIZ at 62South and 30East during an explosive
polar cyclone. It encountered the ice edge (the northernmost location
where ice concentration exceeds 10% in a 1km radius around the
vessel38) at 08:00 UTC, and reached 100% ice concentration at 09:00
UTC, approximately 10km from the iceedge. It continued South, whilst
remaining in 100% ice concentration34. Over this time, strong winds
(1819ms1from North-East according to ERA5 reanalysis, which
underestimates in-situ measurements39; Fig. 1a) generated extreme sea
states in the surrounding area, with signicant wave heights (H
S
=4σ
η
,
where σ
η
is the surface elevation standard deviation) up to 10m North-
East of the icebreaker and >6m at the ice edge (i.e., in the 90th
Fig. 1 | Daily averaged environmental conditions on the 4th July 2017 from
ERA5. a Wind speed; btotal wave height and cswellwave height. Vectorsshow the
direction, where in b,cthe length is proportional to the wave period. The hor-
izontal and vertical axis denotes Easting (longitude) and Northing (latitude) in
degrees. b,cThe white area indicates sea-ice concentration 15% (ERA5 wave data
are only provided for ice concentrationup to 15%). The shipposition at the iceedge
is denoted by the red dot.
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 2
Content courtesy of Springer Nature, terms of use apply. Rights reserved
percentile40) when the icebreaker entered the MIZ (Fig. 1b). The mean
wave direction at the ice edge was aligned with the wind throughout
the day41 (from North-East). The mean directional spread, a measure of
the breadth of the wave spectrum in direction42,was60and, like the
mean wave direction, was steady during the day41.Waveheightand
period at the ice edge increased throughout the day, due to the
intensication of the cyclone, thus creating a non-stationary incident
wave eld. A detailed analysisof the sea state indicates that 70% of the
total signicant wave height is due to waves generated locally (wind
sea). Swell contributes the remaining 30%, which also comes from a
north-easterly direction but with a slight offset of less than 20from the
wind sea.
As the icebreaker travelled South into the MIZ, six sequences of
three-dimensional images of the ocean surface were captured by a pair
of synchronised cameras installed on the icebreaker. The measure-
ment locations are dened by average distances from the ice edge,
d=544km, where the distanceis taken along themean wave direction
(Fig. 2a). Each sequence was taken over a 20-minute interval, during
which the ship heading and forward speed was almost constant, with
the rst sequence starting at 08:05 UTC and the last at 11:50 UTC. The
20-minute time interval is a World Meteorological Organisation stan-
dard for analysis of wave measurements43, which balances stationarity
of wave conditions with collecting a large enough number of waves for
a statistically robust analysis.
An automatic algorithm for oe size reconstruction was applied to
the digital images collected along the transect (see Methods)34,and
used to calculate oe size distributions, mean oe diameters and the
areal concentration of oes. At the rst measurement loc ation, close to
the ice edge (d= 5km), the ice concentration was i
c
50%44 and con-
sisted of pancake ice oes (small, approximately circular oes that
form in wavy conditions45,46) and the remaining 50% was water between
the oes (Fig. 2b). At all ve subsequent locations, which were deeper
into the MIZ (d15km; Fig. 2c, d), ice covered 100% of the ocean
surface, in the form of 60% pancake oes and 40% interstitial frazil
ice34, which increased in density with distance from the ice edge.
Pancake oe diameters generally increased with distance from the ice
edge, with the median diameter increasing from 3.0m at the ice edge
to 3.5m at the deepest measurement locations (Fig. 2e), and the oe
size distribution skewing towards larger diameters (Fig. 2f). At all
measurement locations, over 50% of the pancake-covered area was
comprised of oes with diameters in the interval 24m (vertical error
bars in Fig. 2e). Therefore, the ice conditions during the experiment
are considered relatively insensitive to distance from the ice edge, in
comparison to the changes in the incident wave eld.
Fig. 2 | Sea-ice properties on the4th July 2017. a Overview of study area with sea-
ice concentration from AMSR2 on the 4th July 2017 (longitude and latitude are
horizontal and vertical axes, respectively), with bullets indicating six mean mea-
surement locations during 20min acquisitions. bdExample of image acquisitions
(axes in pixels) at d= 5km, 24km and 43km from the ice edge, respectively.
eMedian pancake oe diameter (D, vertical axis) versus distancefrom the edge (d,
horizontal axis) shown as bullets, plus 25th and 75th diameter percentiles (vertical
error bars) and uncertainty in distance from ice edge (horizontal error bars). Sha-
ded background denote the measurement location in intermediate ice con-
centration. fArea weighted oe size distribution (a, vertical axis) as a function of
oe diameter (D, horizontal axis) at each measurement location. Colour coding is
used in all panels to denote the distance from ice edge.
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Three-dimensional imaging of ocean surface and comparison
with buoy data
The three-dimensional images (Fig. 3a) are used to extract ocean sur-
face elevation timeseries, η(t), at each measurement location, from
which the one-dimensional frequency spectra, E(f), and two-
dimensional frequencydirection spectra, E(f,θ), are derived (see
Methods). At the two deepest measurement locations (d=4344km),
wave buoys (waves-in-ice observation systems47 ofthetypeusedinthe
previous studies11,18,20,21), were deployed on ice oes48 and the fre-
quency spectra they provide are used to validate the analysis of the
stereo-camera images (Fig. 3b, c). The overall shape of the corre-
sponding spectra is consistent, noting that the buoys do not show the
multiple peaks clearly due to a low resolution. Discrepancies occur for
the lower (f< 0.05Hz) and upper tails (f> 0.10Hz), which show over
and under estimation of energy, respectively, relative to the buoys.
These modes carry a small amount of energy and make only minor
contributions to integrated parameters, such as the signicant wave
height and mean period, for which values derived from the images and
buoys differ by 5% and 2.5%, respectively.
The uncertainty associated with spectra derived from the stereo-
cameraimages is primarily white noise (i.e., equal across any frequency
band; see Methods). The noise level is negligibl efor the most energetic
modes (0.04 Hz < f<0.16Hz, or 1.4 < log
10
(f)<0.8, corresponding
to periods of 725s; see error bands in Fig. 3b, c insets) and produces
an integrated error of 0.02m, which corresponds to ~0.5% of the
measured signicant wave heights. The spectral density estimated
from buoys is subjected to red noise (i.e., it decreases with increasing
frequency), which also produces a negligible effect on the signicant
wave height49.
Surface elevation and wave spectra as a function of distance
from the ice edge
Surface elevation timeseries around the largest individual waves
recorded (Hmax; maximum crest to trough distance) at the six mea-
surement locations are shown in Fig. 4a. The signicant wave height is
also reported for each timeseries (Fig. 4a; horizontal dashed lines). The
signicant wave height decreases along the transect, from 6.6m close
to the ice edge (d=5km; toppanel)to H
S
=4.6mat d= 44km (bottom
panel), and, hence, wave energy attenuates over distance. Statistical
analysis of the individual waves indicates consistency with Gaussian
(linear) theory. For instance,the kurtosis (fourth order moment of the
probability density function of the surface elevation and a measure of
wave nonlinearity50) is close to three, similar to Gaussian sea states50,51.
The maximum individual wave heights at each location, which are part
Fig. 3 | Three-dimensional surface elevation retrieval and validation against
buoys. a Reconstructed surface elevation, η(axes in pixels). b,cFrequency spectra
were obtained from the surface elevations measured at the two deepest mea-
surement locations (d=4344km; solid curves) and co-located buoy
measurements (broken curves). Insets show thespectra in logarithmicscale and the
shaded area denotes noise level associated to wave spectra derived from stereo
images. b,cThe horizontal axes denote frequency (f) and the vertical ones
energy (E).
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 4
Content courtesy of Springer Nature, terms of use apply. Rights reserved
of energetic wave groups,tend to diminish with distance into the MIZ.
Nevertheless, large waves are recorded tens of kilometres into 100%
ice concentration, e.g., H
max
8m at d= 43km (second to bottom
panel), which corresponds to H
max
/H
S
1.6, close to the maximum
height expected in a linear sea state43. The steepness associated with
the largest individual waves, ε
max
=πH
max
/(2λ), where λis the wave-
length, is a measure of the strength of the wave42. It decreases from
ε
max
0.19 at the ice edge (d=5km) to 0.10 at the deepest mea-
surements locations (d=4344km). Despite the attenuation of the
steepness over distance, these maximum values are expected to be
large enough to have an impact on the ice cover, for example, by
keeping the ice cover unconsolidated46,52.
Attenuation of wave energy with distance into the MIZ is also
evident in the one-dimensional spectra (Fig. 4b). As expected, higher
frequencies (shorter periods) experience greater attenuation than
lower frequencies (longer periods), causing narrowing of the spectral
bandwidth (the breadth of the spectrum in frequency42), with the
bandwidth at the two deepest measurement locations 80% of the
bandwidth close to the ice edge. The frequency spectrum is unimodal
at the rst four measurement locations (d36km) but becomes
bimodal at the deepest two measurement locations (d=4344km)
where a low-frequency peak appears (peak period 15s). The low-
frequency peak indicates the swell initially north of the ice edge
(Fig. 1c) catches up with the higher frequency wind sea generated
close to the ice edge (peak period 12.7s at d= 36km). The swell
overlaps the wind sea system in frequency space, but the two systems
are clearly separated in two-dimensional frequencydirection
space (Fig. 4c) as they are travelling in different directions. The offset
is 20, which is consistent with the difference in direction between
wind sea and swell reported at the ice edge. For further analysis, the
frequencydirection spectra are used to partition the total sea into
swell and wind sea (see Methods). Note that the buoys do not return
the directional spectrum18 and, hence, their measurements cannot
be used to separate wind sea and swell systems, as they overlap in
frequency space.
Wave evolution
The wave age c
p
/u
10
42,wherec
p
is the phase velocity (ratio of wave-
length to wave period) and u
10
is the wind speed, is computed at each
measurement location using wind speeds from the onboard met-
station (Fig. 5a). The values obtained indicate waves are young
(growing in size and length under the action of wind) at the rst four
measurement locations (c
p
/u
10
< 1.2553). The waves switch sharply to
old (c
p
/u
10
> 1.25) at the deepest two measurement locations. The
sharp transition is partially due to the arrival of the swell system, but, as
implied by the similar sharp transition in wave age for the wind
sea (denoted total sea without swell), a sudden drop in wind speed
from 25ms1to 17ms1is the primary cause.
Lengthening of the peak period with distance is evident in the
frequency spectra (Fig. 4b). The peak period of the incident eld also
increases over the duration of the experiment, from 12.4s to 13.5s, and
the measured peak period normalised by the corresponding ERA5
peak period at the ice edge to account for the changing conditions in
the open ocean (based on the mean wave direction and group velocity
of the mean period; see Methods), is relatively insensitive to distance
until the swell system is detected at the deepest two measurement
locations (Fig. 5b), indicating the peak period elongation due to pre-
ferential attenuationof shorter periodcomponents of thespectrum21,54
is negligible. The normalised peak period sharply increases when the
swell system appears at the deepest two measurement locations
(d=4344km), but the increase is relatively small for the wind sea. The
normalised peak period values for the wind sea are consistent with
the MBK spectral attenuation model15,21, where the ERA5 spectrum at
the ice edge is used as the incident eld for the model.
The frequency averaged directional spread (σ
0
) is calculated from
the two-dimensional spectra20 and normalised using the ERA5 spectra
Fig. 4 | Reconstructed surface elevation timeseries and wave spectra. a Surface
elevation timeseries at progressive distances from the ice edge (toptobottom;
time on horizontal axes and surface elevation on vertical axes; colour-coding cor-
responds to locations shown in Fig. 3a) around the largest wave in each record.
Dashed lines indicate signicant wave heights, H
S
.bFrequency and
cfrequencydirection wave spectra at progressive distances from the ice edge.
b,cThe horizontal axis denotes frequency, in bthe vertical axiswave energy and in
cdirection. Two-dimensional spectra are shown in Cartesian coordinates as the
spectracover only a narrow directional range, 60<θ<60
, and normalised by the
peak energy to highlight directional properties (colorbar shown next to the title).
acShaded backgrounds denote the measurement location in intermediate ice
concentration.
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
at the ice edge (Fig. 5c). For the total sea, the normalised directional
spread is relatively insensitive to distance over the rst three mea-
surement locations (0.60 < σ
0
/σ
0,ERA5
< 0.77) but increases sharply at
the nal three locations (σ
0
/σ
0,ERA5
> 1), with the maximum spread (σ
0
/
σ
0,ERA5
1.27) at d= 36km. The wind sea shows little variation in nor-
malised directional spread over all locations (0.44 < σ
0
/σ
0,ERA5
<0.88).
Thus, the peak in directional spread of the total sea at d= 36km is likely
due to a combination of attenuation of the wind sea peak and emer-
genceoftheswellsystem.Thedirectionalspreadofthetotalsea
decreases at the deepest two locations as the swell dominates. The
MBK attenuation model applied to the ERA5 two-dimensional incident
spectra, with propagation distances of the spectral components
dependent on their direction, predicts the normalised directional
spread slightly decreases over distance from 50(σ
0
/σ
0,ERA5
=0.82)at
5km to 43(σ
0
/σ
0,ERA5
=0.52)at4344km, as the components travelling
at oblique directions travel farther and, hence, experience greater
attenuation leading to collimation14. This trend contrasts with the weak
increasing trend in the observed directional spread of the wind sea
by 1.22% per kilometre.
The signicant wave height of the total sea normalised by the
ERA5 ice edge counterpart (to account for the non-stationary wave
conditions at the ice edge, where the signicant wave height
grows from 6.7m to 7.4m during the measurements; see Methods)
attenuates with distance (Fig. 5d). The best-texponentialcurvetothe
normalised measurements, expðαdÞ, gives the attenuation rate
α=8.3×10
6m1. The wind sea attenuates at a greater rate, with the
best-t exponential curve giving the attenuation rate α= 23.5 × 106m1.
Subtracting the contribution of wind input to the wind sea (based on
theory for open water and using winds recorded by the onboard met-
station; see Methods) further increases the attenuation rate to
α=29.4×10
6m1. For comparison, the benchmark empirical model18
gives the rate 32.7 × 106m1(for ice concentrations greater than 80%,
peak periods less than 14s and signicant wave heights up to 6m).
The uncertainty in the calculation of the distance from the ice edge
(see error bars in Fig. 5), which is due to ambiguity introduced by the
coexistence of multiple wave systems, does not affect the reported
trends of wave parameters.
Discussion
The benchmark signicant wave height attenuation rate is greater than
that derived for the total sea from the stereo-camera images by
approximately a factor four. The benchmark rate is based primarily on
measurements of low-energy sea states (H
s
< 1m), where wind speeds
were generally < 10ms1(from ERA5 reanalysis; see Fig. 9 by Montiel
et al.23) and the wave spectra were most likely unimodal (as indicated in
Fig.10cbyKohoutetal.
18). Therefore, we argue that the fairest com-
parison is with the wind sea without wind input (i.e., attenuation of the
large waves generated by the cyclone in the open ocean), although
noting the subtracted wind input represents an upper bound as it does
not consider ice cover. With these modications to the wave eld,
which rely on analysis of the frequencydirection spectrum, the
attenuation rate is less than the benchmark by 10% only. The agree-
ment is remarkable considering the potential differences in conditions
that may affect attenuation rates, such as the extreme wave heights
and strong winds associated with the cyclone, and the ice properties.
Therefore, the results provide support for the benchmark attenuation
rate and evidence it holds for considerably larger waves than pre-
viously recorded in similar ice conditions. However, the results show
the benchmark attenuation rate only applies to single-component
seas, and does not describe the evolution of the complex sea observed
deep into the MIZ during the cyclone.
Theresultsprovideevidencethatstrongwindsfeedwavegrowth
in 100% ice concentration for tens of kilometres over pancake/frazil ice
cover. This contradicts the assumption made in most contemporary
models that wind input scales according to the open water fraction55,56,
i.e., no wind input in 100% ice concentration. The assumption is
already being debated, particularly for frazil, brash and/or pancake
ice conditions22, and theories for wind-to-wave momentum transfer
through ice covers are being proposed57.
The benchmark attenuation rate is greater than the attenuation
rate of the wind sea without wind input possibly due to the properties
of the ice cover. Based on observations during deployment of the
buoys for the benchmark measurements, the ice cover consisted of
unconsolidated pancake/frazil ice, similar to the stereo-camera mea-
surements, and also consolidated, larger, thicker oes and continuous
Fig. 5 | Wave properties at progressive distances from the ice edge. a Wave age;
bpeak period;cdirectional spreading; and dsignicant wave height (vertical axes)
versus distance (horizontal axes). Dots are used to denote total sea, and triangles
the wind sea (total sea without swell). The horizontal error bars denote uncer-
tainties in distancefrom the ice edge due to variability in wind and wave directions.
The shaded area denotes intermediate ice concentration. b,cThedashedlineis
derived by applying the MBK model21 to corresponding ERA5 spectra at the ice
edge. dThe solid grey line denotes the best (exponential) t for the total sea, the
dashed line for the wind sea (total sea without swell) and the dash-dotted line the
wind sea without wind input over ice. The thin dashed line is the benchmark
attenuation derived for i
c
> 0.8 and T
P
<14s
18.
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 6
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ice18, which is likely to cause stronger attenuation58,59.Buildingalarger
database of stereo-camera images in the MIZ will drive improved
understanding of how ice type inuences the attenuation rate. More-
over, collecting stereo-camera images during conditions when the
incoming wave eld is steady will allow the spectral attenuation rate to
be calculated and compared with benchmarks21,23, thus giving more
detailed understanding of the attenuation process.
In conclusion, measurements have been reported of extreme
wave conditions in the winter Antarctic MIZ during an explosive
polar cyclone, which were captured by a stereo-camera system on
a moving vessel over a 44km transect and validated by co-located
buoy measurements. The images empowered analysis of the
frequencydirection wave spectrum evolution, and revealed the
complex multi-component nature of the sea state in the MIZ where
wind sea and swell co-exist and attenuate at different rates. Con-
comitant measurements of winds gave evidence of wind input through
100% ice cover. The success of stereo-imaging system shown in this
study is likely to inuence the design of future eld campaigns in the
MIZ. Moreover, it has the potential to be installed on vessels that
routinely traverse the Antarctic MIZ and autonomously monitor the
sea state, vastly increasing the data available as well as providing
concomitant information on the ice cover. In turn, this will open new
frontiers to advance current knowledge of MIZ dynamics and underpin
the development of the next generation of climate models.
Methods
Image acquisition
The acquisition device consists of two GigE monochrome industrial
CMOS cameras with a 2/3 inch sensor, placed side-by-side at a distance
(i.e., baseline) of 4m. The stereo rig was installed on the monkey bridge
of the icebreaker, 34m from the waterline and tilted 20below the
horizon. The cameras were equipped with 5mm lenses to provide a
eld of view of the ocean surface 90around the port side of the ship
(Fig. 3a). Additionally, an inertial measurement unit (IMU) was rmly
attached close to the two cameras (between the cameras at the same
height and 1m behind them) to capture the stereo-rig movement with
respect to the sea surface during the acquisition. Images were recor-
ded with a resolution of 2448× 2048 pixels and a sampling rate of 2Hz
during daylight on the 4 July 2017 (from 07:00 to 14:00 UTC).
Floe size
An automatic algorithm, developed using the MatLab Image Proces-
sing Toolbox, was applied to extract sea-ice metrics from the recorded
images. To avoid sampling the same oe twice, images every 10s were
analysed. Images were ortho-rectied and projected on a horizontal
plane. A camera-dependent calibration was applied to convert pixel to
metres. The images were processed to eliminate the vessel from the
eld of view, adjust the image contrast andconvertthe grey scales into
a binary map, which isolates the solid ice shapes from background
water or frazil ice. A morphological image processing was applied to
improve the delity of the shape of identied pancakes. Floe area (S)
was calculated based on pixels and approximated by a disk, from
which a characteristic diameter D=ffiffiffiffiffiffiffiffiffiffiffi
4S=π
pwas dened. Further
details on the algorithm and oe size distribution can be found in
Alberello et al.34.
Estimation of the surface elevation
The Wave Acquisition Stereo System (WASS60)wasusedtoestimatea
set of data points in space (a dense 3D point cloud) representing the 3D
ocean surface. WASS analyses left and right stereo images to nd
photometrically distinctive corresponding points (i.e., projections of
the same 3D point in space) that can be triangulated to recover their
original 3D position in space.The operation is performed for each pixel
of the stereo frames to produce a temporal sequence of 3D point
clouds composed of millions of samples each. The data is also
automatically ltered to remove possible outliers and the mean sea-
plane is estimated independently for each frame.
To be effective, the technique requires the geometrical cong-
uration of the two cameras to be known a priori. WASS can estimate
that property as part of the process, with the added advantage of
correcting slight variations in the cameras reciprocal orientation due
to vibrations. The motion of the vessel under the effect of waves,
however, is more signicant than vibrations induced, for example, by
wind. It follows that points clouds from different pair of images lie in a
different reference frame. Measurements of ship motion from the IMU
are therefore used to align and geo-localise each cloud to a common
horizontal plane dening the mean sea level. As the IMU does not
estimate the absolute elevation accurately, an approach combining
surface orientations estimated from the stereo data with the altitude
computed by the IMU was developed to recover the camera motion
throughout the sequence. An unscented Kalman lter is applied to
model the six degrees of freedom position and orientation of the
cameras (i.e., the system state) as a discrete-time random variable. At
each frame, the system state is updated with both the absolute IMU
data (yaw-pitch-roll) and the mean sea-plane distance vector estimated
from the point clouds. Both are modelled as Gaussian distributions in
which the measurement covariance is given by the sensor manu-
facturer specications, in the case of the IMU, or the empirical stereo
estimation error for the cameras31. With the estimated camera motion,
each scattered point cloud is transformed on a common reference
frame with the xyaxes aligned with the mean sea-plane, and inter-
polated on a regular grid to reconstruct a timeseries of 3D surface
elevations. The nal dimension of the reconstructed surfaces is about
150m × 200m, which is 5 times larger than Smith & Thompson37.
A systematic source of uncertainty is the resolution error61, also
referred to as quantisation noise, which depends on the object dis-
tance, the focal length and the camerasresolution, mutual position
and declination. An estimate of this error was calculated as the dif-
ference between a known synthetic surface and its back and forth
transformation, which consists of projecting the known surface onto
the camera coordinate system and re-projecting it onto the original
coordinate system by using the specic geometry of the stereo-camera
setup61. This difference provides the spatial distribution of the error,
the amplitude of which is uniformly distributed across wavenumbers/
frequencies (i.e., white noise).
Estimation of the attenuation rate
The sea state conditions were non-stationary during the observation
period, with wave height increasing 10%from6.7mto7.4m.There-
fore, the attenuation rate αis estimated with respect to the ice edge
and the dimensionless signicant wave height according to the fol-
lowing equation:
log HS
HS,ERA5

=αd,ð1Þ
where dis the distance of each measurement location from the ice
edge calculated along the mean wave direction in the open ocean as
provided by the ERA5 reanalysis, and H
S
/H
S,ERA5
is the signicant wave
height normalised by the incident (open ocean) counterpart from the
ERA5 reanalysis. To estimate the incident wave conditions (H
S,ERA5
), the
delay between the time at which waves enter into the MIZ and the time
at which waves are measured at a specic location was estimated
through the wave group velocity. A standard least square tting is
applied to extrapolate an overall attenuation rate. Owing to the non-
stationarity of the sea state conditions, each frequency components of
the wave eld is subjected to different forcing, besides ice-induced
attenuation. Thereby, the estimate of a frequency-dependent attenua-
tion rate18 is impractical herein as it would be subjected to signicant
uncertainty.
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Note thatthe coexistence of wind sea and swell systems generated
an ambiguity for the selection of the relevant wave direction, which
translates into uncertainties in the distance from the ice edge. By
assuming an overall mean wave direction, the error associated to d
is ± 5% (this is reported as an horizontal error bar in Fig. 5).
Frequencydirection spectrum
The directionalwave spectrum can be computed from three measured
quantities of the ocean surface (for example, a buoy uses either heave,
pitch and roll or three accelerations) with a Fourier expansion
method62. This approach produces accurate mean wave direction, but
the directional spreading is typically too broad and with spurious
bimodal properties. An improvement in the directional resolving
power can be achieved by increasing the number of measured ele-
ments with a spatial array of sensors, which are cross-correlated using
the maximum likelihood method or a wavelet directional method62,63.
With the stereo images, the strategy is to apply an array approach
to extract timeseries ofthe surface elevation and a wavelet directional
method to approximate the directional spectrum, noting that the
estimate of the spectrum from timeseries allows resolving compo-
nents with periods > 11.5s and, hence, wavelength > 200m (i.e., longer
than the eld of view). A virtual array with geometry comprising of a
triangle inscribed in a circle of radius 1m inside a pentagon inscribed
in a circle of radius 3m and with an additional probe in the central
position (a co-array conguration) was used to allow a sufcient
number of non-redundant spatial lags between elements and asym-
metry, which ensure an accurate estimation of the directional
spreading function62. The wavelet directional method resolve modes
around the spectral peak (0.5 < f/f
p
<3,wheref
p
is the peak frequency)
more accuratelythan the maximum likelihood method63. Thereby, it is
the preferred approach to identify the complex multiple peak feature
of the spectrum.
Correction of Doppler shift and true wave spectrum
The sea surface elevation extracted from the images is in a frame of
reference that moves with the forward speed and heading of the ship.
The directional spectrum is ther efore distorted due to a Doppler shift64
and it is typically referred to as the encountered spectrum E
S
(f
S
,θ),
where f
S
is the frequency detected by a moving object. To restore the
original spectral shape, the encounter frequency is converted into the
true frequency fthrough the linear wave dispersion relation as
follows64:
fS=f+4π2f2
gUScos β,ð2Þ
where βis the angle between the ship heading and the (open water)
mean wave direction from the ERA5 reanalysis. The true wave spec-
trum, E(f,θ), is computed by performing a change of variable:
Eðf,θÞ=ESðfS,θÞdfS
df =ESðfS,θÞ1+8π2f
gUScos β

:ð3Þ
where df
S
/df is the Jacobian of the transformation.
Partitioning of spectrum
The partitioning of the wave spectrum is performed using the path of
steepest ascent technique65, which is a specic implementation of the
inverse catchment scheme introduced by Hasselmann et al.66.The
spectral peak that satises the condition
1:2u10
cp
cosðθψÞ>1, ð4Þ
where u
10
is the wind speed, c
p
is the phase velocity, θis the wave
direction and ψis the wind direction, is assumed to be associated with
the wind sea. All other systems are swell and are ranked based on their
energy contents as primary, secondary and tertiary swell. Only primary
swell was considered.
ERA5 incident sea state
The incident sea state at the ice edge is obtained from ERA5 reanalysis,
which provides hourly wave data at a resolution of 40km. The model
performs well in the Southern Ocean but can miss the swell arrival
time67 and the shape of the spectrum is limited by directional resolu-
tion (15), which makes it difcult to discriminate wind seas and swell
that propagate in approximately the same direction.
TheERA5wavemodelusesasimplisticwaveattenuationschemes
in sea ice, which acts only at the outskirts of the MIZ (up to 15% ice
concentration), and assumes waves are completely dissipated for ice
concentration > 15%. To overcome uncertainties related to the model
assumptions, we adopt, as a reference, the wave data in open ocean
(0% ice concentration), in the proximity of the ice edge along the mean
wave direction (the distances from the ice edge specied in Fig. 2aare
the one projected in the direction of the mean wave direction). We also
account for the time delay due to wave energy advection (based on the
wave group velocity from the incident ERA5 mean period) to associate
sea states in the MIZ with the open ocean counterpart.
Comparison against ERA5 reanalysis
To provides evidence of consistency between the ERA5 reanalysis and
the observations presented herein, and thus to allow using the former
to produce incident sea states, a comparison against open ocean
conditions recorded on July 2nd (between 12:00 and 14:00 UTC at 52
South and 26East) and July 3rd (between 12:00 and 13:00 UTC at 56
South and 28East) is shown in Table 1. For this comparison, the ERA5
reanalysis data refers to the value at the closest grid point, while
observations are an average over the time period. Overall the agree-
mentisgood,withdifferenceinwaveheight<0.2m(<4%),waveper-
iod < 0.2s (< 2%) and directional spread <0. 3.A more thorough
evidence of the accuracy of the ERA5 reanalysis in the Southern Ocean
through comparison against a large data set acquired during the
Antarctic Circumnavigation Expedition is detailed by Derkani68.
MBK attenuation model
Based on eld measurements in the Antarctic MIZ, MBK (Meylan
et al.21) proposed an(exponential) attenuation rate for each frequency
component in the spectrum as
βðfÞ=af2+bf4where a=2:12 × 103and b=4:59 × 102:ð5Þ
The MBK attenuation rate is used to propagate the incident
directional wave spectrum into sea ice up to the measurements loca-
tions using the expression
Eðf,θÞ=EERA5ðf,θÞexpðicβðfÞdðθÞÞ:ð6Þ
Table 1 | Integrated spectral wave parameters from our measurements and ERA5
H
S,WASS
[m] H
S,ERA5
[m] T
P,WASS
[s] T
P,ERA5
[s] σ
0,WASS
[]σ
0,ERA5
[]
2nd July 2017 4.93 4.94 10.18 10.37 35.6 35.9
3rd July 2017 4.16 4.35 8.01 7.84 36.3 36.2
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 8
Content courtesy of Springer Nature, terms of use apply. Rights reserved
The attenuation rate depends on the ice concentration (i
c
), the fre-
quency of the wave component and the distance from the ice edge on
the direction of each component.
Wind input
In the open ocean, wind transfers momentum to the sea surface, for-
cing ocean waves to grow over distance (fetch). Wave growth is esti-
mated with empirical models, which are expressed as42
~
HS=a1
~
Fb1ð7Þ
where
~
F=gF=u2
10 and
~
HS=gHS=u2
10 are the dimensionless fetch and
wave height, respectively, and a
1
=2.88×10
3and b
1
= 0.4569.Eqn.(7)is
used to estimate the fetch at the ice edge (F
ERA5
) from the known ERA5
wave height (H
S,ERA5
). The distance from the edge (d) is added to the
fetch
F=FERA5+d:ð8Þ
Eqn. (7) is then applied again to give an updated estimate of the wave
height (H
S,wind
) over theentire distance (F). The difference between the
updatedwaveheightandtheoneattheedgeprovidesanestimateof
the wave height attributed to wind input, i.e.
ΔHS=HS,wind HS,ERA5:ð9Þ
Data availability
Data sets for this research (reconstructed surface elevations) are
available through the Australian Antarctic Data Centre (AADC):
Alberello et al. (2021) Wave acquisition stereo-camera system mea-
surements (WASS) from a voyage of the S.A. Agulhas II, July2017, Ver. 1,
Australian Antarctic Data Centrehttps://doi.org/10.26179/q9bd-5f74.
Code availability
MATLAB was used for the analysis. Data processing is described within
the manuscript and code is available from the rst author upon
request.
References
1. Turner,J.&Comiso,J.SolveAntarcticas sea-ice puzzle. Nature 547,
275277 (2017).
2. Eayrs,C.,Li,X.,Raphael,M.N.&Holland,D.M.Rapiddeclinein
Antarctic sea ice in recent years hints at future change. Nat. Geosci.
14,460464 (2021).
3. Williams, G. et al. Thick and deformed Antarctic sea ice
mapped with autonomous underwater vehicles. Nat. Geosci. 8,
6167 (2015).
4. Li,T.etal.Resolvingne-scale surface features on polar sea ice: a
rst assessment of UAS photogrammetry without ground control.
Remote Sens. 11, 784 (2019).
5. Hobbs, W. R. et al. A review of recent changes in Southern Ocean
sea ice, their drivers and forcings. Glob. Planet. Change 143,
228250 (2016).
6. Brouwer, J. et al. Altimetric observation of wave attenuation
through the antarctic marginal ice zone using icesat-2. Cryosphere
16,23252353 (2022).
7. Pope, A. et al. Community review of Southern Ocean satellite data
needs. Antarct. Sci. 29,97138 (2017).
8. Newman,L.etal.Delivering sustained, coordinated, and integrated
observations of the Southern Ocean for global impact. Front. Mar.
Sci. 6, 433 (2019).
9. Williams,T.D.,Bennetts,L.G.,Squire,V.A.,Dumont,D.&Bertino,L.
Wave-ice interactions in the marginal ice zone. Part 1: theoretical
foundations. Ocean Model. 71,8191 (2013).
10. Williams,T.D.,Bennetts,L.G.,Squire,V.A.,Dumont,D.&Bertino,L.
Wave-ice interactions in the marginal ice zone. Part 2: Numerical
implementation and sensitivity studies along 1D transects of the
ocean surface. Ocean Model. 71,92101 (2013).
11. Kohout, A., Williams, M., Dean, S. & Meylan, M. Storm-induced sea-
ice breakup and the implications for ice extent. Nature 509,
604 (2014).
12. Robin, G. D. Q. Wave propagation through elds of pack ice. Philos.
Trans. Roy. Soc. A 255,313339 (1963).
13. Squire, V. A. & Moore, S. M. Direct measurement of the attenuation
of ocean waves by pack ice. Nature 283,365368 (1980).
14. Wadhams, P., Squire, V. A., Ewing, J. A. & Pascal, R. W. The effect of
the marginal ice zone on the directional wave spectrum of the
ocean. J. Phys. Oceanogr. 16,358376 (1986).
15. Bennetts, L. G., OFarrell, S. & Uotila, P. Impacts of ocean-wave-
induced breakup of Antarctic sea ice via thermodynamics in a
stand-alone version of the CICE sea-ice model. Cryosphere 11,
10351040 (2017).
16. Roach,L.A.,Horvat,C.,Dean,S.M.&Bitz,C.M.Anemergentseaice
oe size distribution in a global coupled ocean-sea ice model. JGR
Oceans 123,43224337 (2018).
17. Thomson, J. et al. Overview of the Arctic sea state and boundary
layer physics program. J. Geophys. Res. Oceans 123,
86748687 (2018).
18. Kohout, A. et al. Observations of exponential wave attenuation in
Antarctic sea ice during the pipers campaign. Ann. Glaciol. 61,
114 (2020).
19. Doble, M. J. & Bidlot, J. R. Wave buoy measurements at the Antarctic
sea ice edge compared with an enhanced ECMWF WAM: Progress
towards global waves-in-ice modelling. Ocean Model. 70,
166173 (2013).
20. Montiel, F., Squire, V. A., Doble, M., Thomson, J. & Wadhams, P.
Attenuation and directional spreading of ocean waves during a
storm event in the autumn Beaufort Sea marginal ice zone. J.
Geophys. Res. Oceans 123,59125932 (2018).
21. Meylan, M. H., Bennetts, L. G. & Kohout, A. L. In situ measurements
and analysis of ocean waves in the Antarctic marginal ice zone.
Geophys. Res. Lett. 41,50465051 (2014).
22. Rogers, W. E., Meylan, M. H. & Kohout, A. L. Estimates of spectral
wave attenuation in Antarctic sea ice, using model/data inversion.
Cold Reg. Sci. Technol. 182, 103198 (2020).
23. Montiel, F., Kohout, A. L. & Roach, L. A. Physical drivers of ocean
wave attenuation in the marginal ice zone. J. Phys. Oceanog.
https://doi.org/10.1175/JPO-D-21-0240.1 (2022).
24. Ardhuin, F. et al. Estimates of ocean wave heights and attenuation in
sea ice using the sar wave mode on sentinel-1a. Geophys. Res. Lett.
42,23172325 (2015).
25. Stopa, J. E., Sutherland, P. & Ardhuin, F. Strong and highly variable
push of ocean waves on Southern Ocean sea ice. Proc. Natl. Acad.
Sci. 115, 58615865 (2018).
26. Meylan, M. H. et al. Dispersion relations, power laws, and energy
loss for waves in the marginal ice zone. J. Geophys. Res. Oceans
123, 33223335 (2018).
27. Squire, V. A. Ocean wave interactions with sea ice: a reappraisal.
Annu.Rev.FluidMech.52,3760 (2020).
28. Meylan, M. H., Horvat, C., Bitz, C. M. & Bennetts, L. G. A oe size
dependent scattering model in two- and three-dimensions for wave
attenuation by ice oes. Ocean Model. 161, 101779 (2021).
29. Doble, M. J., De Carolis, G., Meylan, M. H., Bidlot, J.-R. & Wadhams,
P. Relating wave attenuation to pancake ice thickness, using eld
measurements and model results. Geophys. Res. Lett. 42,
44734481 (2015).
30. Boutin, G. et al. Towards a coupled model to investigate wave-sea
ice interactions in the Arctic marginal ice zone. Cryosphere 14,
709735 (2020).
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
31. Benetazzo, A. et al. Stereo wave imaging from moving vessels:
practical use and applications. Coast. Eng. 109,114127 (2016).
32. Sutherland, P. & Gascard, J.-C. Airborne remote sensing of ocean
wave directional wavenumber spectra in the marginal ice zone.
Geophys. Res. Lett. 43,51515159 (2016).
33. Sutherland, P., Brozena, J., Rogers, W. E., Doble, M. & Wadhams, P.
Airborne remote sensing of wave propagation in the marginal ice
zone. J. Geophys. Res. Oceans 123,41324152 (2018).
34. Alberello, A. et al. Pancake ice oe size distribution during the
winter expansion of the Antarctic marginal ice zone. Cryosphere 13,
4148 (2019).
35. Collard, F., Ardhuin, F. & Chapron, B. Monitoring and analysis of
ocean swell elds from space: new methods for routine observa-
tions. J. Geophys. Res.114, C07023 (2009).
36. Campbell, A. J., Bechle, A. J. & Wu, C. H. Observations of surface
waves interacting with ice using stereo imaging. J. Geophys. Res.
Oceans 119,32663284 (2014).
37. Smith, M. & Thomson, J. Pancake sea ice kinematics and dynamics
using shipboard stereo video. Ann. Glaciol. 111, https://doi.org/10.
1017/aog.2019.35 (2019).
38. Worby, A. P. A technique for making ship-based observations of
Antarctic sea ice thickness and characteristics. 14 (Antarctic
CRC, 1999).
39. Vichi, M. et al. Effects of an explosive polar cyclone crossing the
Antarctic marginal ice zone. Geophys. Res. Lett. https://doi.org/10.
1029/2019GL082457 (2019).
40. Young, I. R., Fontaine, E., Liu, Q. & Babanin, A. V. The wave climate of
the southern ocean. J. Phys. Oceanogr. 50,14171433 (2020).
41. Copernicus Climate Change Service. ERA5: Fifth generation of
ECMWF atmospheric reanalyses of the global climate (Copernicus
Climate Change Service Climate Data Store (CDS), 2017).
42. Holthuijsen, L. H. Waves in Oceanic and Coastal Waters (Cambridge
University Press, 2007).
43. World Meteorological Organization. Guidetowaveanalysisand
forecasting, WMO-No 702 (World Meteorological Organization
Geneva, Switzerland, 1998).
44. de Jong, E. et al. Sea iceconditions within theAntarctic marginal ice
zone in winter 2017, onboard the SA Agulhas II, https://doi.org/10.
1594/PANGAEA.885211 (2018).
45. Wadhams, P., Aulicino, G., Parmiggiani, F., Persson, P. O. G. & Holt,
B. Pancake ice thickness mapping in the Beaufort sea from wave
dispersion observed in SAR imagery. J. Geophys. Res. Oceans
https://doi.org/10.1002/2017JC013003 (2018).
46. Shen, H. H., Ackley, S. F. & Hopkins, M. A. A conceptual model for
pancake-ice formation in a wave eld. Ann. Glaciol. 33,
361367 (2001).
47. Kohout, A. L., Penrose, B., Penrose, S. & Williams, M. J. A device for
measuring wave-induced motion of ice oes in the Antarctic mar-
ginal ice zone. Ann. Glaciol. 56,415424 (2015).
48. Alberello, A. et al. Drift of pancake ice oes in the winter Antarctic
marginal ice zone during polar cyclones. J. Geophys. Res. Oceans
125, e2019JC015418 (2020).
49. Thomson, J., Hošeková, L., Meylan, M. H., Kohout, A. L. & Kumar, N.
Spurious rollover of wave attenuation rates in sea ice caused by
noise in eld measurements. J. Geophys. Res. Oceans 126,
e2020JC016606 (2021).
50. Onorato, M. et al. Statistical properties of directional ocean waves:
the role of the modulational instability in the formation of extreme
events. Phys. Rev. Lett. 102,114502(2009).
51. Toffoli, A. et al. Wind generated rogue waves in an annular wave
ume. Phys.Rev.Lett.118,144503(2017).
52. Womack, A., Vichi, M., Alberello, A.&Toffoli,A.Atmosphericdrivers
of a winter-to-spring lagrangian sea-ice drift in the eastern antarctic
marginal ice zone. J. Glaciol. 115, https://doi.org/10.1017/jog.2022.
14 (2022).
53. Aouf, L. et al. New directional wave satellite observations: Towards
improved wave forecasts and climate description in southern
ocean. Geophys. Res. Lett. 48, e2020GL091187 (2021).
54. Bennetts, L. G. & Squire, V. A. On the calculation of an attenuation
coefcient for transects of ice-covered ocean. Proc.R.Soc.Math.
Phys.Eng.Sci.468,136162 (2012).
55. Li,J.,Kohout,A.L.&Shen,H.H.Comparisonofwavepropagation
through ice covers in calm and storm conditions. Geophys. Res.
Lett. 42,59355941 (2015).
56. Liu, Q., Rogers, W. E., Babanin, A., Li, J. & Guan, C. Spectral mod-
eling of ice-induced wave decay. J. Phys. Oceanogr. 50,
15831604 (2020).
57. Zhao, X. & Zhang, C. A theoretical model of wind-wave
growth over an ice-covered sea. Bound.-Layer Meteorol. 178,
119 (2020).
58. Dolatshah,A.etal.Hydroelasticinteractions between water waves
and oating freshwater ice. Phys. Fluids 30,091702(2018).
59. Ardhuin, F., Otero, M., Merrield, S., Grouazel, A. & Terrill, E. Ice
breakup controls dissipation of wind waves across
southern ocean sea ice. Geophys. Res. Lett. 47,
e2020GL087699 (2020).
60. Bergamasco, F., Torsello, A., Sclavo, M., Barbariol, F. & Benetazzo,
A. WASS: an open-source pipeline for 3D stereo reconstruction of
ocean waves. Comput. Geosci. 107,2836 (2017).
61. Benetazzo, A. Measurements of short water waves using
stereo matched image sequences. Coast. Eng. 53,
10131032 (2006).
62. Young, I. R. On the measurement of directional wave spectra. Appl.
Ocean Res. 16,283294 (1994).
63. Donelan, M. A., Drennan, W. M. & Magnusson, A. K. Nonstationary
analysis of the directional properties of propagating waves. J. Phys.
Oceanogr. 26,19011914 (1996).
64. Faltinsen, O. Sea loads on ships and offshore structures, vol. 1
(Cambridge university press, 1993).
65. Hanson, J. L. & Phillips, O. M. Automated analysis of ocean surface
directional wave spectra. J. Atmos. Ocean. Technol. 18,
277293 (2001).
66. Hasselmann, S., Brüning, C., Hasselmann, K. & Heimbach, P. An
improved algorithm for the retrieval of ocean wave spectra from
synthetic aperture radar image spectra. J. Geophys. Res. 101,
1661516629 (1996).
67. Jiang, H., Babanin, A. V. & Chen, G. Event-based validation of swell
arrival time. J. Phys. Oceanogr. 46,35633569 (2016).
68. Derkani,M.H.Waves in the Southern Ocean and Antarctic Marginal
Ice Zone: Observations and modelling. Ph.D. thesis (2021).
69. Kahma, K. K. & Calkoen, C. J. Reconciling discrepancies in the
observed growth of wind-generated waves. J. Phys. Oceanogr. 22,
13891405 (1992).
Acknowledgements
The expedition was funded by the South African National Antarctic
Programme through the National Research Foundation. This work was
motivated by the Antarctic Circumnavigation Expedition (ACE) and
partially funded by the ACE Foundation and Ferring Pharmaceuticals.
A.A., L.B. and A.T. were supported by the Australian Antarctic Science
Programme (project 4434). A.A. acknowledges support from the Japa-
nese Society for the Promotion of Science (PE19055). L.G.B. is supported
by the Australian Research Council (FT190100404). L.G.B. and A.T. are
supported by the Australian Research Council (DP200102828). M.O. was
supported by the Simons Collaboration on Wave Turbulence, Award No.
617006, and from the Departments of Excellence 2018-2022Grant
awarded by the Italian Ministry of Education, University and Research
(MIUR, L.232/2016). M.O. acknowledges the EU H2020 FET Open
BOHEME, Grant No. 863179. M.V. and K.M. were supported by the NRF
SANAP contract UID118745. C.E. was supported under NYUAD Center for
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Global Sea Level Change project G1204. We are indebted to Captain
Knowledge Bengu and the crew of the S.A. Agulhas II for their invaluable
contribution to data collection. ERA5 reanalysis was obtained using
Copernicus Climate Change Service Information. M.O. acknowledges
B. GiuliNico for interesting discussions. A.A., A.T. and M.O. thank
L. Fascette for technical support during the cruise.
Author contributions
Conceptualisation: A.A., L.G.B., M.O., A.T. Methodology: A.A., L.G.B.,
M.O., M.V., K.M.H., C.E., B.N.N., A.B., F.B., F.N., R.P., H.C., A.T. Investi-
gation: A.A., L.G.B., M.O., M.V., A.T. Visualisation: A.A., L.G.B., I.T., A.T.
Supervision: A.A., L.G.B., A.T. Writing-original draft: A.A., L.G.B., M.O.,
M.V., A.B., F.B., A.T. Writing-review & editing: A.A., L.G.B., M.O., M.V.,
K.M.H., C.E., B.N.N., A.B., F.B., F.N., R.P., H.C., I.T., A.T.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information The online version contains
supplementary material available at
https://doi.org/10.1038/s41467-022-32036-2.
Correspondence and requests for materials should be addressed to
Alberto Alberello.
Peer review information Nature Communications thanks Clarence
Collins and Dany Dumont for their contribution to the peer review of this
work. Peer reviewer reports are available.
Reprints and permission information is available at
http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jur-
isdictional claims in published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as
long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license, and indicate if
changes were made. The images or other third party material in this
article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not
included in the articles Creative Commons license and your intended
use is not permitted by statutory regulation or exceeds the permitted
use, you will need to obtain permission directly from the copyright
holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2022
Article https://doi.org/10.1038/s41467-022-32036-2
Nature Communications | (2022) 13:4590 11
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... This remote region comprises an uninterrupted band of water around Antarctica [34]. It is dominated by strong westerly winds, the notorious roaring forties, furious fifties and screaming sixties, which give rise to some of the fiercest waves on the planet over almost infinite fetches all the year round [18,[34][35][36]. As a reference, the significant wave height-an average measure of the highest third of waves in a given sea state and a proxy for wave height statistics in the ocean-has a 50-th percentile of ≈ 5 m and a 90-th percentile of ≈ 7 m in the austral winter [34]. ...
... Complementing atmospheric data from the ship's meteorological station as well as navigation records from the ship's voyage data recorder were acquired throughout the journey. Two wave buoys were also deployed during the expeditions to provided supplementary records of the surface elevation [33,35]. ...
... An example of stereo pair and reconstructed surface is shown in Fig. 1b,c. From each sequence, the directional wave spectrum was derived (see details in [35]), and spectral parameters were computed. These are reported in Fig. 2 as a function of the wave age relative to the spectral peak. ...
Article
Full-text available
We report direct observations of surface waves from a stereo camera system along with concurrent measurements of wind speed during an expedition across the Southern Ocean in the austral winter aboard the South African icebreaker S.A. Agulhas II. Records include water surface elevation across a range of wave conditions spanning from early stages of wave growth to full development. We give experimental evidence of rogue seas, i.e., sea states characterized by heavy tails of the probability density function well beyond the expectation based on bound mode theory. These conditions emerge during wave growth, where strong wind forcing and high nonlinearity drive wave dynamics. Quasiresonance wave-wave interactions, which are known to sustain the generation of large amplitude rogue waves, capture this behavior. Wave statistics return to normality as the wind forcing ceases and waves switch to a full developed condition.
... Waves in sea ice are measured using several different techniques, including, but not limited to, in situ observations by means of buoys or accelerometers placed on the ice (e.g., Squire and Moore, 1980;Cheng et al., 2017;Voermans et al., 2019;Kohout et al., 2020;Wahlgren et al., 2023), analysis of ship motion (e.g., Collins et al., 2015), remote observations from ships by stereo imaging (e.g., Smith and Thomson, 2019;Alberello et al., 2022) and from aircrafts by laser scanning (Sutherland and Gascard, 2016;Sutherland et al., 2018), satellite-based methods (e.g., Stopa et al., 2018a, b;Horvat et al., 2020;Brouwer et al., 2022;Huang and Li, 2023), or, recently, distributed acoustic sensing of seafloor cables (Smith et al., 2023). Taken together, the result of this observational effort is a large, very valuable body of data encompassing thousands of measured wave energy spectra. ...
... As a result, analyses based on that data reduce to comparisons of spectra from two data points (notably, even if more than two buoys are used, they are often analyzed on a pairwise basis, as in, e.g., Cheng et al., 2017;Kohout et al., 2020;Montiel et al., 2022). If we denote those two locations by A and B, then the so-called apparent attenuation coefficient a a is computed as (e.g., Stopa et al., 2018b;Kohout et al., 2020;Alberello et al., 2022;Montiel et al., 2022): ...
Article
Full-text available
Numerical modeling of waves in sea ice covered regions of the oceans is important for many applications, from short-term forecasting and ship route planning up to climate modeling. In spite of a substantial progress in wave-in-ice research that took place in recent years, spectral wave models – the main tool for wave modeling at regional and larger scales – still don’t capture the underlying physics and have rather poor predictive skills. This article discusses recent developments in wave observations and spectral wave modeling in sea ice, identifies problems and shortcomings of the approaches used so far, and sketches future directions that, in the opinion of the author, have the potential to improve the performance of wave-in-ice models.
... In the MIZ, modeling ice drift, breakup, and formation is particularly challenging. Many physical processes in the MIZ occur at scales smaller than the model's grid resolution, which brings challenges in parameterization as well as in characterizing the floe sizes (see floe size distributions, for example, in [2,71,63]). Understanding the interaction between ocean waves and sea ice in MIZs is crucial for accurately modeling and predicting sea ice dynamics, especially in the context of climate change and its impact on polar regions; we refer the reader to [74,75,28,4,42] and the references therein for details. ...
... last access: 5 October 2022). The wave model data are available through the Integrated Ocean Waves for Geophysical and Other Applications (IOWAGA) project: https://tds3.ifremer.fr/thredds/catalog.html (Ardhuin et al., 2011). The analysis uses and modifies icesat2-toolkit (https://read-icesat-2.readthedocs.io/, ...
Article
Full-text available
Sea ice is important for Earth's energy budget as it influences surface albedo and air–sea fluxes in polar regions. On its margins, waves heavily impact sea ice. Routine and repeat observations of waves in sea ice are currently lacking, and therefore a comprehensive understanding of how waves interact with sea ice and are attenuated by it is elusive. In this paper, we develop methods to separate the two-dimensional (2D) surface wave spectra from sea-ice height observations made by the ICESat-2 (IS2) laser altimeter, a polar-orbiting satellite. A combination of a linear inverse method, called generalized Fourier transform (GFT), to estimate the wave spectra along each beam and a Metropolis–Hastings (MH) algorithm to estimate the dominant wave's incident angle was developed. It allows us to estimate the 2D wave signal and its uncertainty from the high-density, unstructured ATL03 ICESat-2 photon retrievals. The GFT is applied to re-binned photon retrievals on 25 km segments for all six beams and outperforms a discrete Fourier transform (DFT) in accuracy while having fewer constraints on the data structure. The MH algorithm infers wave direction from beam pairs every 25 km using coherent crests of the most energetic waves. Assuming a dominant incident angle, both methods together allow a decomposition into 2D surface wave spectra with the advantage that the residual surface heights can potentially be attributed to other sea-ice properties. The combined GFT–MH method shows promise in routinely isolating waves propagating through sea ice in ICESat-2 data. We demonstrate its ability on a set of example ICESat-2 tracks, suggesting a detailed comparison against in situ data is necessary to understand the quality of retrieved spectra.
Article
The seasonal sea ice zone encompasses the region between the winter maximum and summer minimum sea ice extent. In both the Arctic and Antarctic, the majority of the ice cover can now be classified as seasonal. Here, we review the sea ice physics that governs the evolution of seasonal sea ice in the Arctic and Antarctic, spanning sea ice growth, melt, and dynamics and including interactions with ocean surface waves as well as other coupled processes. The advent of coupled wave–ice modeling and discrete-element modeling, together with improved and expanded satellite observations and field campaigns, has yielded advances in process understanding. Many topics remain in need of further investigation, including rheologies appropriate for seasonal sea ice, wave-induced sea ice fracture, welding for sea ice freeze-up, and the distribution of snow on seasonal sea ice. Future research should aim to redress biases (such as disparities in focus between the Arctic and Antarctic and between summer and winter processes) and connect observations to modeling across spatial scales.
Article
Full-text available
Sea ice–wave interactions have been widely studied in the marginal ice zone, at relatively low wind speeds and wave frequencies. Here, we focus on very different conditions typical of coastal polynyas: extremely high wind speeds and locally generated, short, steep waves. We overview available parameterizations of relevant physical processes (nonlinear wave–wave interactions, energy input by wind, whitecapping and ice‐related dissipation) and discuss modifications necessary to adjust them to polynya conditions. We use satellite‐derived data and spectral modeling to analyze waves in 10 polynya events in the Terra Nova Bay, Antarctica. We estimate the wind‐input reduction factor over frazil/grease ice in the wave‐energy balance equation at 0.56. By calibrating the model to satellite observations we show that exact treatment of quadruplet wave–wave interactions (as opposed to the default Discrete Interaction Approximation) is necessary to fit the model to data, and that the power n > 4 in the sea‐ice source term Sice ∼ fⁿ (where f denotes wave frequency) is required to reproduce the strong attenuation of high‐frequency waves observed for frazil streaks. We use a very‐high resolution satellite image of a fragment of one of the polynyas to determine whitecap fraction. We show that there are more than twofold differences in whitecap fraction over ice‐free and ice‐covered regions, and that the model produces realistic whitecap fractions without any tuning of the whitecapping source term. Finally, we estimate the polynya‐area‐integrated wind input, energy dissipation due to whitecapping, and whitecap fraction to be on average below 25%, 10% and 30%, respectively, of the corresponding open‐water values.
Preprint
Full-text available
A holistic review is given of the Southern Ocean dynamic system, in the context of the crucial role it plays in the global climate and the profound changes it is experiencing. The review focuses on connections between different components of the Southern Ocean dynamic system, drawing together contemporary perspectives from different research communities, with the objective of “closing loops” in our understanding of the complex network of feedbacks in the overall system. The review is targeted at researchers in Southern Ocean physical science with the ambition of broadening their knowledge beyond their specific field and facilitating better-informed interdisciplinary collaborations. For the purposes of this review, the Southern Ocean dynamic system is divided into four main components: large-scale circulation; cryosphere; turbulence; and gravity waves. Overviews are given of the key dynamical phenomena for each component, before describing the linkages between the components. The reviews are complemented by an overview of observed Southern Ocean trends and future climate projections. Priority research areas are identified to close remaining loops in our understanding of the Southern Ocean system.
Article
Full-text available
Two ensembles of buoys, deployed in the marginal ice zone (MIZ) of the northeastern Weddell Sea region of the Southern Ocean, are analysed to characterise the dynamics driving sea ice drift and deformation during the winter-growth and the spring-retreat seasons of 2019. The results show that although the two buoy arrays were deployed within the same region of ice-covered ocean, their trajectory patterns were vastly different. This indicates a varied response of sea ice in each season to the local winds and currents. Analyses of the winter data showed that the Antarctic Cir-cumpolar Current modulated the drift near the sea ice edge. This led to a highly energetic and mobile ice cover, charac-terised by free-drift conditions. The resulting drift and deformation were primarily driven by large-scale atmospheric forcing, with negligible contributions due to the wind-forced inertial response. For this highly advective coupled ice-ocean system, ice drift and deformation linearly depended on atmospheric forcing. We also highlight the limits of commercial floating ice velocity profilers in this regime since they may bias the estimates of sea ice drift and the ice type detection. On the other hand, the spring drift was governed by the inertial response as increased air temperatures caused the ice cover to melt and break up, promoting a counterintuitively less wind-driven ice-ocean system that was more dominated by inertial oscillations. In fact, the deformation spectra indicate a strong decoupling to large-scale atmospheric forcing. Further analyses, extended to include the deformation datasets from different regions around Antarctica, indicate that, for similar spatial scales, the magnitude of deformation varies between seasons, regions, and the proximity to the sea ice edge and the coastline. This implies the need to develop rheology descriptions that are aware of the ice types in the different regions and seasons to better represent sea ice dynamics in the MIZ.
Article
Full-text available
The Antarctic marginal ice zone (MIZ) is a highly dynamic region where sea ice interacts with ocean surface waves generated in ice-free areas of the Southern Ocean. Improved large-scale (satellite-based) estimates of MIZ extent and variability are crucial for understanding atmosphere–ice–ocean interactions and biological processes and detection of change therein. Legacy methods for defining the MIZ are typically based on sea ice concentration thresholds and do not directly relate to the fundamental physical processes driving MIZ variability. To address this, new techniques have been developed to measure the spatial extent of significant wave height attenuation in sea ice from variations in Ice, Cloud and land Elevation Satellite-2 (ICESat-2) surface heights. The poleward wave penetration limit (boundary) is defined as the location where significant wave height attenuation equals the estimated error in significant wave height. Extensive automated and manual acceptance/rejection criteria are employed to ensure confidence in along-track wave penetration width estimates due to significant cloud contamination of ICESat-2 data or where wave attenuation is not observed. Analysis of 304 ICESat-2 tracks retrieved from four months of 2019 (February, May, September and December) reveals that sea-ice-concentration-derived MIZ width estimates are far narrower (by a factor of ∼ 7 on average) than those from the new technique presented here. These results suggest that indirect methods of MIZ estimation based on sea ice concentration are insufficient for representing physical processes that define the MIZ. Improved large-scale measurements of wave attenuation in the MIZ will play an important role in increasing our understanding of this complex sea ice zone.
Chapter
Full-text available
In this Summary, the techniques needed to describe the chaos of ocean waves are introduced. The basic concept for this is the random-phase/amplitude model that leads to the definition of the variance density spectrum. Interpreted as the energy density spectrum, this provides the basis for modeling the physical aspects of the waves. . This document summarizes just a chapter of the above-mentioned book published by Cambridge University Press, 2010. The summary is not intended to include all information in the chapter; thus, the readers are strongly encouraged to read the original text to gain a broad perspective on the topic. Notwithstanding some minor changes and adjustments in the text, most sentences, figures, tables, and all equations are directly gotten from the book to keep the originality.
Article
Full-text available
Sea-ice drift in the Antarctic marginal ice zone (MIZ) is discussed using data from a 4-month- long drift of a buoy deployed on a pancake ice floe during the winter sea-ice expansion. We demonstrate increased meandering and drift speeds, and changes in the dynamical regimes of the absolute dispersion during cyclone activity, together with high correlations between drift velocities and wind from atmospheric reanalyses. This indicates a dominant physical control of wind forcing on ice drift and the persistence of free-drift conditions. These conditions occurred despite the buoy remaining largely in >80% ice concentrations and at distances >200 km from the estimated ice edge. The drift is additionally characterised by a strong inertial signature at 13.47 h, which appears initiated by passing cyclones. A wavelet analysis of the buoy’s velocity confirms that the momentum transfer from winds at the multi-day frequencies is due to atmospheric forcing, while the initiation of inertial oscillations of sea ice has been identified as the secondary effect. Propagating storm-generated waves may initiate inertial oscillations by increasing the mobility of floes and enhance the drag of the inertial current. This analysis indicates that the Antarctic MIZ in the Indian Ocean sector remains much wider and mobile, during austral winter-to-spring, than defined by sea-ice concentration. (Open access)
Article
Full-text available
The effects of instrument noise on estimating the spectral attenuation rates of ocean waves in sea ice are explored using synthetic observations in which the true attenuation rates are known explicitly. The spectral shape of the energy added by noise, relative to the spectral shape of the true wave energy, is the critical aspect of the investigation. A negative bias in attenuation that grows in frequency is found across a range of realistic parameters. This negative bias decreases the observed attenuation rates at high frequencies, such that it can explain the rollover effect commonly reported in field studies of wave attenuation in sea ice. The published results from five field experiments are evaluated in terms of the noise bias, and a spurious rollover (or flattening) of attenuation is found in all cases. Remarkably, the wave heights are unaffected by the noise bias, because the noise bias occurs at frequencies that contain only a small fraction of the total energy.
Article
Full-text available
A model-data inversion is applied to an extensive observational dataset collected in the Southern Ocean north of the Ross Sea during late autumn to early winter, producing estimates of the frequency-dependent rate of dissipation by sea ice. The modeling platform is WAVEWATCH III® which accounts for non-stationarity, advection, wave generation, and other relevant processes. The resulting 9477 dissipation profiles are co-located with other variables such as ice thickness to quantify correlations which might be exploited in later studies to improve predictions. An average of dissipation profiles from cases of thinner ice near the ice edge is fitted to a simple binomial. The binomial shows remarkable qualitative similarity to prior observation-based estimates of dissipation, and the power dependence is consistent with at least three theoretical models, one of which assumes that dissipation is dominated by turbulence generated by shear at the ice-water interface. Estimated dissipation is lower closer to the ice edge, where ice is thinner, and waveheight is larger. The quantified correlation with ice thickness may be exploited to develop new parametric predictions of dissipation.
Article
Full-text available
In spite of continuous improvements of ocean wave models in the last decades, large errors still remain in particular under strongly forced conditions, often encountered in the Southern Ocean, where strong westerly winds generate some of the fiercest waves on Earth in almost unlimited fetch conditions. The newly launched China‐France Oceanography SATellite (CFOSAT) provides directional spectra of ocean waves for both wind seas and swells. Compared to Synthetic Aperture Radar (SAR), it can resolve shorter wavelengths in all directions, which dominate in non‐fully developed wind waves. Here, the assimilation of these CFOSAT wavenumber components is proved to bring more accurate predictions of wave growth compared to the assimilation of significant wave height alone. A notable reduction of model bias is found in the Southern Ocean, especially in the Pacific Ocean sector. Results further exhibit a downward shift of the wave age, consistent with theoretical wave growth curves.
Article
Despite a recent resurgence of observational studies attempting to quantify the ice-induced attenuation of ocean waves in polar oceans, the physical processes governing this phenomenon are still poorly understood. Most analyses have attempted to relate the spatial rate of wave attenuation to wave frequency, but have not considered how this relationship depends on ice, wave, and atmospheric conditions. An in-depth analysis of the wave-buoy data collected during the 2017 Polynyas, Ice Production, and Seasonal Evolution in the Ross Sea (PIPERS) program in the Ross Sea is conducted. Standard techniques are used to estimate the spatial rate of wave attenuation α , and the influence of a number of potential physical drivers on its dependence on wave period T is investigated. A power law is shown to consistently describe the α ( T ) relationship, in line with other recent analyses. The two parameters describing this relationship are found to depend significantly on sea ice concentration, mean wave period, and wind direction, however. Looking at cross correlations between these physical drivers, three regimes of ice-induced wave attenuation are identified, which characterize different ice, wave, and wind conditions, and very possibly different processes causing this observed attenuation. This analysis suggests that parameterizations of ice-induced wave decay in spectral wave models should be piecewise, so as to include their dependence on local ice, wave, and wind conditions. Significance Statement This work attempts to quantify how ice, wave, and wind conditions in polar oceans affect the way that ocean waves decay as a result of their interactions with sea ice. In situ wave data collected in the Ross Sea are analyzed along with several freely available ice, wave, and wind datasets. A simple relationship is shown to describe how wave attenuation due to sea ice depends on the wave period consistently across all data analyzed. However, the parameters of this relationship are significantly affected by sea ice concentration, mean wave period, and wind direction. This finding suggests that large-scale wave models need to account for this dependence on ice, wave, and wind conditions to improve wave forecast in ice-covered oceans.
Article
Following years of record highs, an unexpected and precipitous reduction in Antarctic sea-ice extent started in 2016. This decline, lasting three years, was the most pronounced of the satellite era, equivalent to 30 years of sea-ice loss in the Arctic. Here, we synthesize recent work showing this sea-ice reduction probably resulted from the interaction of a decades-long ocean warming trend and an early spring southward advection of atmospheric heat, with an exceptional weakening of the Southern Hemisphere mid-latitude westerlies in late spring. We discuss what this event reveals about the underlying atmospheric and oceanic dynamical processes that control sea ice in the region and the ways in which shifting climate variability and remote forcings, especially from the tropics, influence these processes. Knowledge gaps show that further work is needed to improve future projections of changes in one of the largest seasonal phenomena on the planet. The combined effects of decades-long warming and particularly vigorous injections of atmospheric heat from lower latitudes were the likely culprits for sharp declines in sea-ice extent around Antarctica starting in 2016.
Article
Two- and three-dimensional models are proposed for ocean-wave attenuation due to scattering by ice floes in the marginal ice zone, in which the attenuation rate depends on the horizontal size of the individual floes. The scattering models are shown to reproduce the behaviour of wave attenuation over short wave periods. However, it is shown that scattering alone cannot explain the observed asymptotic dependence of attenuation at long wave periods. Based on these findings, it is proposed that attenuation models consist of a scattering component supplemented by an empirical damping term based on measurements, so that attenuation over all periods is correctly modelled. Computer code to calculate wave attenuation through a field of ice floes is provided in the supplementary material.