ArticlePDF Available

Frost flowers on sea ice as a source of sea salt and their influence on tropospheric halogen chemistry

Wiley
Geophysical Research Letters
Authors:
  • Alfred Wegener Institute for Polar and Marine Research (AWI)

Abstract and Figures

Frost flowers grow on newly-formed sea ice from a saturated water vapour layer. They provide a large effective surface area and a reservoir of sea salt ions in the liquid phase with triple the ion concentration of sea water. Recently, frost flowers have been recognised as the dominant source of sea salt aerosol in the Antarctic, and it has been speculated that they could be involved in processes causing severe tropospheric ozone depletion events during the polar sunrise. These events can be explained by heterogeneous autocatalytic reactions taking place on salt-laden ice surfaces which exponentially increase the reactive gas phase bromine (``bromine explosion''). We analyzed tropospheric bromine monoxide (BrO) and the sea ice coverage both measured from satellite sensors. Our model based interpretation shows that young ice regions potentially covered with frost flowers seem to be the source of bromine found in bromine explosion events.
Content may be subject to copyright.
Frost flowers on sea ice as a source of sea salt and their influence on
tropospheric halogen chemistry
L. Kaleschke, A. Richter, J. Burrows, O. Afe, G. Heygster, and J. Notholt
Institute of Environmental Physics, University of Bremen, Germany
A. M. Rankin and H. K. Roscoe
British Antarctic Survey, Natural Environment Research Council, Cambridge, UK
J. Hollwedel and T. Wagner
Institute of Environmental Physics, University of Heidelberg, Germany
H.-W. Jacobi
Alfred Wegener Institute for Polar and Marine Research, Bremerhaven, Germany
Received 2 June 2004; revised 21 July 2004; accepted 2 August 2004; published 25 August 2004.
[1] Frost flowers grow on newly-formed sea ice from a
saturated water vapour layer. They provide a large effective
surface area and a reservoir of sea salt ions in the liquid phase
with triple the ion concentration of sea water. Recently, frost
flowers have been recognised as the dominant source of sea
salt aerosol in the Antarctic, and it has been speculated that
they could be involved in processes causing severe
tropospheric ozone depletion e vents during the polar
sunrise. These events can be explained by heterogeneous
autocatalytic reactions taking place on salt-laden ice surfaces
which exponentially increase the reactive gas phase bromine
(‘‘bromine explosion’’). We analyzed tropospheric bromine
monoxide (BrO) and the sea ice coverage both measured
from satellite sensors. Our model based interpretation shows
that young ice regions potentially covered with frost flowers
seem to be the source of bromine found in bromine explosion
events.
INDEX TERMS: 0322 Atmospheric Composition and
Structure: Constituent sources and sinks; 1640 Global Change:
Remote sensing; 3309 Meteorology and Atmospheric Dynamics:
Climatology (1620); 3339 Meteorology and Atmospheric
Dynamics: Ocean/atmosphere interactions (0312, 4504); 3360
Meteorology and Atmospheric Dynamics: Remote sensing.
Citation: Kaleschke, L., et al. (2004), Frost flowers on sea ice
as a source of sea salt and their influence on tropospheric halogen
chemistry, Geophys. Res. Lett. , 31, L16114, doi:10.102 9/
2004GL020655.
1. Introduction
[2] The discovery of events of low ozone concentration
in the atmospheric boundary layer at measurement stations
located at high latitudes in the Northern Hemisphere has
prompted much research into their origin. These events
were found to be associated with enhanced amounts of
inorganic bromine compounds [Barrie et al., 1988;
McConnell et al., 1992; Fan and Jacob, 1992; Foster et
al., 2001] . Similar epi sodes have been observed in the
Southern Hemisphere at high latitudes [Wessel et al.,
1998; Frieß et al., 2004]. The advent of the measureme nt
of tropospheric trace gases from space by the Global Ozone
Monitoring Experiment, GOME, led to the discovery of
enhanced amounts of BrO close to regions of sea ice in the
Northern and the Southern Hemisphere [Richter et al.,
1998; Wagner and Platt, 1998]. GOME measures the light
scattered from the atmosphere and reflected by the ground
between 240 and 790 nm wavelength with a horizontal
ground resolution of about 320 40 km
2
[Burrows et al. ,
1999]. A retrieval algorithm, based on differential optical
absorption spectroscopy (DOAS) and stratospheric BrO
modeled by a three-dimensional radiative- dynamical-
chemical model, yields the tropospheric f raction of the
column density of BrO [Richter et al., 1998; Wagner and
Platt, 1998; Chipperfield, 1999]. Bromine destroys ozone
very efficiently in two interlinked catalytic cycles which
produce BrO and HOBr in the gas-phase [Foster et al.,
2001]. Gaseous HOBr reacts with Br ions in a slightly acidic
sea salt solution and releases Br
2
and BrCl into the gas phase
[Fickert et al., 1999; Adams et al., 2002]. The photolabile
Br
2
molecule is subsequently photo-dissociated into atomic
Br [Foster et al., 2001]. Therefore, every Br atom of the
HOBr molecule entering the liquid phase has the potential to
release two Br atoms to the gas phase. The above gives a
simplified description of the heterogeneous autocatal ytic
reaction that causes an exponential increase of gaseous Br
radicals, the so-called bromine explosion. The main source
of bromine over the open oceans in the marine boundary
layer outside the polar regions was identified to be sea salt
aerosol generated by breaking waves on the ocean surface
[Tang and McConnell, 1996; Vogt et al., 1996; Sander et al.,
2003]. The processes and sources unique to the polar ocean
surfaces still remained unidentified, though the highest BrO
amounts have been observed over the sea ice during the
polar sunrise [Ridley et al., 2003 ; Zeng et al., 2003; Frieß et
al., 2004]. Recently, the potential role of frost flowers
(Figure 1) in this processes has been raised [Rankin et al.,
2002]. Frost flo wers are ice crystals which grow on frozen
leads (linear breaks in the sea ice cover) an d polynyas
(openings between drift ice and fast ice or the coast). Frost
flowers exhibit enhanced salinities and bromide ion concen-
trations of about three times of that of bulk seawater
[Perovich and Richter-Menge, 1994; Rankin et al., 2002].
GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L16114, doi:10.1029/2004GL020655, 2004
Copyright 2004 by the American Geophysical Union.
0094-8276/04/2004GL020655$05.00
L16114 1of4
Frost flowers only last for a few days until they are blown
away by strong winds or covered by drifting snow [Perovich
and Richter-Menge, 1994]. The aerosol produced by frost
flowers was identified in Antarctic ice cores and at coastal
stations due to its depleted sulfate to sodium ratio compared
to the aerosol originating from the open ocean [Wagenbach
et al., 1998; Rankin et al., 2002; Rankin and Wolff, 2003].
2. Methods and Data Sets
[3] It is assumed, that the open water area (which is given
by one minus the sea ice concentration) will be covered
soon with thin new ice on which frost flowers can grow.
The sea ice concentration is determined from analyses of the
thermal microwave emission me asured by the Special
Sensor Microwave Imager (SSM/I) aboard the Defense
Meteorological Satellite Program (DMSP) platform using
the ASI algorit hm [Kaleschke et al., 2001; Maslanik and
Stroeve, 2003]. Briefly this algorithm takes advantage of the
higher polarisability of a specularly reflecting water surface
compared to the more diffusely reflecting sea ice surface
[Kaleschke et al., 2001; Kern et al., 2003].
[
4] A one dimensional thermodynamic model has been
developed to calculate the frost flower coverage (Figure 2).
The model combines a frost flower growth parameterization
obtained from laboratory experiments with the equations of
sea ice heat balance which are described in more detail by
Martin et al. [1996] and Maykut [1986]. We assume that
the growth of frost flowers depends only on two basic
prerequisites, the existence of new ice which is formed in
leads or polynyas and of a strong negative temperature
gradient above the ice surface. The two model input
parameters, the surface air temperature T
a
and the open
water (OW) fraction, are taken from numerical weather
prediction reanalysis data (NCEP/NCAR) and from satellite
passive microwave measurements, respectively. The ice
thickness H = 1.33Q
0.53
[cm] is calculated from the cumu-
lative freezing days Q =
R
(T
f
T
a
)dt with air temperature
T
a
and the freezing point of sea water T
f
= 1.9C. The sea
ice surface temperature T
0
=
15:96 þ HT
a
8:4 þ H
is approximated for
thin ice using an averaged heat transfer coefficient which
describes both sensible and latent heat exch ange [Maykut,
1986]. Assuming that the influence of a varying wind field
and the insulating effect of frost flowers on the heat flux are
negligible, then the frost flower growth rate g =
0.000785e
0.478(T
0
T
a
)
is readily calculated using coefficients
from laboratory experiments [Martin et al., 1996; Maykut,
1986]. The area coverage F
t
is calculated using the recursive
expression F
t
= F
tdt
+ g(1 F
tdt
)dt for a time step dt.
Because the growth rate decreases rapidly as the ice
thickness increases, the model yields a maximum percent-
age area F
max
(T
a
) covered by frost flowers for a given
surface air temperature. This is defined as the re lative
potential frost flower (PFF) area. The total PFF area is
obtained by weighting the total area with the new ice
fraction. The predicted relative PFF areas for different
integration times and air temperatures are presented in
Figure 2. The height and hence the volume of the frost
flowers cannot be calculated as the parameterization of the
growth rate was derived from video ima ges which measured
the area coverage. As there is currently insufficient micro-
physical understanding about the initial nucleation and
growth of frost flowers as well as about the decay processes,
a minimum frost flower area for a given set of conditions
cannot be estimated [Perovich and Richter-Menge, 1994;
Martin et al., 1995, 1996]. The original aim of the above
described method was to identify dates and regions worth to
be analysed using costly very high resolution satellite
images for the future development of a more direct frost
flower retrieval algorithm [Kaleschke and Heygster, 2004].
3. Comparison of Model Results and BrO Data
[5] One typical example of the resulting PFF and the BrO
data for the Antarctic is shown in Figure 3. The overall
mechanism requires the release of bromine atom precursors,
either directly on the frost flower surface or within its
aerosol. We calculated forward air trajectories starting at
regions with a high probability of frost flowers from the
NCEP/NCAR Reanalysis surface wind field in order to
account for the atmospheric transport of the aerosol or
BrO. Some regions with a high probability of frost flower
occurrence for example at the Ronne-Filchner Ice Shelf in
the Weddell Sea show no corresponding BrO plumes
Figure 1. Frost flowers on sea ice covering a lead: Stellar
dendrites of about 1 to 2 cm height on young sea ice
(Courtesy of Stefan Kern, University of Hamburg). The
photograph was taken at 7558
0
N2534
0
E, 24 March 2003.
The air temperature was about 18C.
Figure 2. Frost flower model and frost flower coverage as
a function of air temperature for different integration times.
The three curves represent the integration times of one, five
and ten days, respectively. The theoretical upper limit of
frost flower coverage F
max
(T
a
) is approximately given by
the upper curve.
L16114 KALESCHKE ET AL.: FROST FLOWERS AND TROPOSPHERIC CHEMISTRY L16114
2of4
because sunlight is needed for the photochemical reactions.
The polynya at the Ross Ice Shelf that occurs frequently due
to strong katabatic winds is shown to be a strong source of
frost flower aerosol responsible for the enhanced BrO
concentration farther north in the illum inated area. The
trajectories are only a rough approximation as errors could
occur in convective regions which are not well represented
in the atmospheric boundary layer of the NCEP/NCAR
model over sea ice [Kaleschke et al. , 2001]. Nevertheless,
more than ninety percent of the trajectories hit the enhanced
BrO areas in Figure 3. One key area of frequently occurring
ozone depletion events in the Arctic is shown in Figure 4.
This typical example shows the huge (almost 500 50 km
2
)
recurring shore polynya in the northwest of the Hudson Bay
which is potentially covered with frost flowers. This polynya
frequently appears under offshore wind conditions.
[
6] We investigated the entire 1996 to 2002 dataset and
found commonly more than two third of the PFF trajectories
hitting the enhanced BrO areas for the Arctic and Antarctic
during polar sunrise. Almost all cases of enhanced BrO
amounts were associated with a high probability of frost
flowers on the previous days. Occasionally enhanced PFF
values appear without enhanced BrO amounts. However,
this does not reject the hypothesis of frost flowers causing
BrO production as the BrO retrieval could be hampered
by clouds [Richter et al., 1998; Wagner and Platt, 1998;
Frieß et al., 2004]. Furthermore, the PFF is a potential
theoretical upper limit. Specific meteorological conditions
could have prevented the actual growth of frost flowers. The
influence of the wind is ambiguous: the dynamical opening
of leads and polynyas is a wind driven effect and a
prerequisite for thin ice production, but persisting strong
winds could prevent the growth of frost flowers [Perovich
and Richter-Menge, 1994]. Changing wind fields s uch as
passing cyclones probably support the growth of frost
flowers.
4. Conclusion
[7] It can be summarized that the sea salt and associated
halogen flux from the ocean to the atmosphere is governed
by different processes inside and outside the polar regions:
Outside the polar regions , the sea salt is injected into
the atmosphere by breaking waves on the ocean surface
dominated by the wind [Sander et al., 2003]. Whereas the
process inside the sea ice covered re gions is mainly
modulated by the air temperature [Zeng et al., 2003; Frieß
et al., 2004]. Previous work has been insufficient to localize
the potential bromine sources [Richter et al., 1998; Wagner
Figure 3. Comparison: Example total PFF coverage
(green stars) and enhanced BrO amounts (red isolines) over
the Antarctic ocean. The maximum values of two
consecutive days are shown: from 28 to 29 and 29 to
30 August 1997 for PFF and BrO, respectively. The green
stars mark the endpoints of 24h air trajectories starting at a
frost flower area coverage of greater than 0.4%, sampled
every 187 km. The PFF field was smoothed to 300 km
spatial re solution to approximately match the GOME
resolution. The total PFF coverage is proportional to the
size of the stars. Two stars of 1% and 10% area coverage are
shown in the lower right corner for comparison. The
red isoline corresponds to enhanced BrO amounts of 3.6
10
13
molec/cm
2
which is the mean (1.4) plus one standard
deviation (2.2). The black circle indicates the almost dark
latitudes (solar zenith angle > 80). Sea ice covered regions
are presented in white and the open ocean is coloured in
blue.
Figure 4. Example total PFF coverage (green stars) and
enhanced BrO amounts (red isolines) over the Hudson
Bay for 11 March 2001. Symbols and lines are similar to
Figure 3, but the PFF coverage is displayed color coded.
The 24h air trajectories start at PFF greater than 1% sampled
every 62.5 km. The red isoline corresponds to enhanced
BrO amounts of 8.8 10
13
molec/cm
2
which is the mean
(6.2) plus two standard deviations (1.3). The isoline is
broken due to the data gaps of the GOME coverage.
L16114 KALESCHKE ET AL.: FROST FLOWERS AND TROPOSPHERIC CHEMISTRY L16114
3of4
and Platt, 1998; Zeng et al., 2003; Ridley et al., 2003; Frieß
et al., 2004]. We provided a method to localize the young
ice regions potentially covered with frost flowers that seem
to be a prerequisite for the bromine explosion. This provides
a crucial step for a better understanding of the exchange
processes between ocean, sea ice and atmosphere that are of
great importance for the Earth’s climate system [Shepson et
al., 2003].
[
8] Acknowledgments. L. K. and HW. J. gratefully acknowledge the
German Research Foundation (DFG) for funding. L. K. thanks Gunnar
Spreen, Stefan Kern, Roland von Glasow, Eric W. Wolff, Mark Drinkwater,
Robert Ezraty, Wolfgang Dierking, Christof Lu¨pkes, Thomas Busche and
Christian Haas for discussions.
References
Adams, J. W., N. S. Holmes, and J. N. Crowley (2002), Uptake and reaction
of HOBr on frozen and dry NaCl/NaBr surfaces between 253 and 233 K,
Atmos. Chem. Phys., 2, 79 91.
Barrie, L. A., J. W. Bottenheim, R. C. Schnell, P. J. Crutzen, and R. A.
Rasmussen (1988), Ozone destruction and photochemi cal reactions at
polar sunrise in the lower Arctic atmosphere, Nature, 334, 138 141.
Burrows, J. P., et al. (1999), The global ozone monitoring experiment
(GOM E): Mission concept and first scientific resul ts, J. Atmos. Sci.,
56(2), 151 175.
Chipperfield, M. P. (1999), Multiannual simulations with a three-
dimensional chemical transport model, J. Geophys. Res., 104(D1),
1781 1805.
Fan, S. M., and D. J. Jacob (1992), Surface ozone depletion in Arctic spring
sustained by bromine reactions on aerosols, Nature, 359, 524 552.
Fickert, S., J. W. Adams, and J. N. Crowley (1999), Activation of Br-2 and
BrCl via uptake of HOBr onto aqueous salt solutions, J. Geophys. Res.,
104(D19), 23,719 23,727.
Foster, K. L., R. A. Plastridge, J. W. Bottenheim, P. B. Shepson, B. J.
Finlayson-Pitts, and C. W. Spicer (2001), The role of Br-2 and BrCl in
surface ozone destruction at polar sunrise, Science, 291(5503), 471 474.
Frieß, U., J. Hollwedel, G. Konig-Langlo, T. Wagner, and U. Platt (2004),
Dynamics and chemistry of tropospheric bromine explosion events in the
Antarctic coastal region, J. Ge ophys. Res ., 109(D6) , D06305,
doi:10.1029/2003JD004133.
Kaleschke, L., and G. Heygster (2004), Towards multisensor microwave
remote sensing of frost flowers on sea ice, Ann. Glaciol., in press.
Kaleschke, L., C. Lu¨pkes, T. Vihma, J. Haarpaintner, A. Bochert,
J. Hartmann, and G. Heygster (2001), SSM/I sea ice remote sensing
for mesoscale ocean-atmosphere interaction analysis, Can. J. Remote
Sens., 27(5), 526 537.
Kern, S., L. Kaleschke, a nd D. A. Clausi (2003), A c omparison of
two 85 GHz SSM/I ice concentration algorithms with AVHRR and
ERS-SAR, IEEE Trans. Geosci. Remote Sens., 41(10), 2294 2306.
Martin, S., R. Drucker, and M. Fort (1995), A laboratory study of frost
flower growth on the surface of young sea ice, J. Geophys. Res., 100(C4),
7027 7036.
Martin, S., Y. Yu, and R. Drucker (1996), The temperature dependence of
frost flower growth on laboratory sea ice and the effect of the flowers on
infrared observations of the surface, J. Geophys. Res., 101(C5), 12,111
12,125.
Maslanik, J., and J. Stroeve (2003), DMSP SSM/I daily polar gridded
brightness temperatures [CD-ROM ], Natl. Snow and Ice Data Cent.,
Boulder, Colo.
Maykut, G. A. (1986), The surface heat and mass balance, in The Geophys-
ics of Sea Ice, NATO ASI B146, edited by N. Untersteiner, pp. 395 463,
Martinus Nijhoff, Zoetermeer, Netherlands.
McConnell, J. C., G. S. Henderson, L. A. Barrie, J. Bottenheim, H. Niki,
C. H. Langford, and E. M. J. Templeton (1992), Photochemical bromine
production implicated in Arctic boundary-layer ozone depletion, Nature,
355, 150 152.
Perovich, D., and J. A. Richter-Menge (1994), Surface characteristics of
lead ice, J. Geophys. Res., 99(C8), 16,341 16,350.
Rankin, A., and E. W. Wolff (2003), A year-long record of size segregated
aerosol composition at Halley, Antarctica, J. Geophys. Res., 108(D24),
4775, doi:10.1029/2003JD003993.
Rankin, A. M., E. W. Wolff, and S. Martin (2002), Frost flowers: Implica-
tions for tropospheric chemistry and ice core interpretation, J. Geophys.
Res., 107(D23), 4683, doi:10.1029/2002JD002492.
Richter, A., F. Wittrock, M. Eisinger, and J. P. Burrows (1998), GOME
observations of tropospheric BrO in Northern Hemispheric spring and
summer 1997, Geophys. Res. Lett., 25(4), 2683 2686.
Ridley, B. A., et al. (2003), Ozone depletion events observed in the high
latitude surface layer during the TOPSE aircraft program, J. Geophys.
Res., 108(D4), 8356, doi:10.1029/2001JD001507.
Sander, R., et al. (2003), Inorganic bromine in the marine boundary layer: A
critical review, Atmos. Chem. Phys., 3, 1301 1336.
Shepson, P., P. Matrai, L. Barrie, and J. Bottenheim (2003), Ocean-atmo-
sphere-sea ice snowpack interactions in the Arctic and global change, Eos
Trans. AGU, 84(36), 349 355.
Tang, T., and J. C. McConnell (1996), Autocatalytic release of bromine
from Arctic snow pack during polar sunrise, Geophys. Res. Lett., 23(18),
2633 2636.
Vogt, R., P. J. Crutzen, and R. Sander (1996), A mechanism for halogen
release from sea-salt aerosol in the remote marine boundary layer, Nature,
383, 327 330.
Wagenbach, D., F. Ducroz, R. Mulvaney, L. Keck, A. Minikin, M. Legrand,
J. S. Hall, and E. W. Wolff (1998), Sea-salt aerosol in coastal Antarctic
regions, J. Geophys. Res., 103(D9), 10,961 10,974.
Wagner, T., and U. Platt (1998), Satellite mapping of enhanced BrO con-
centrations in the troposphere, Nature, 395, 486 490.
Wessel, S., S. Aoki, P. Winkler, R. Weller, A. Herber, H. Gernandt, and
O. Schrems (1998), Tropospheric ozone depletion in polar regions: A
comparison of observations in the Arctic and Antarctic, Tellus, Ser. B,
50(1), 34 50.
Zeng, T., Y. H. Wang, K. Chance, E. V. Browell, B. A. Ridley, and E. L.
Atlas (2003), Widespread persistent near-surface ozone depletion at
northern high latitudes in spring, Geophys. Res. Lett., 30(24), 2298,
doi:10.1029/2003GL018587.
O. Afe, J. Burrows, G. Heygster, L. Kaleschke, J. Notholt, and A. Richter,
Institute of Environmental Physics, P.O. Box 330440, D-28334 Bremen,
Germany. (lkalesch@iup.physik.uni-bremen.de)
J. Hollwedel and T. Wagner, Institute of Environme ntal Physics,
University of Heidelberg, Germany.
H.-W. Jacobi, Alfred Wegener Institute for Polar and Marine Research,
Bremerhaven, Germany.
A. M. Rankin and H. K. Roscoe, British Antarctic Survey, Natural
Environment Research Council, Cambridge, UK.
L16114 KALESCHKE ET AL.: FROST FLOWERS AND TROPOSPHERIC CHEMISTRY L16114
4of4
... However, the dominant source bromine and the underlying processes involved remain unclear, 65 with more than half a dozen different candidates proposed. These include frost flowers (Kaleschke et al., 2004;Piot and von Glasow, 2008), first-year sea ice surface 2007b), open leads/polynyas (e.g., Peterson et al., 2016;Kirpes et al., 2019;Criscitiello et al., 2021), snowpack on tundra (Pratt et al., 2013), snowpack on sea ice (Custard et al., 2017;Peterson et al., 2019), snowpack on ice sheets (Thomas et al., 2011), and sea salt aerosols from blowing snow (Yang et al., 2008;2010;Frey et al., 2020;Huang et al., 2020). Significant progress has been made in recent decades, with data 70 ...
... All the data will be archived in BAS Polar Data Centre. Phys., 4, 2427−2440, 2004 Surface snow salinity probability distribution over sea ice inland sites Table S4. Table S4 gives more statistical details of the linear regressions in each panel. ...
Preprint
Full-text available
This study explores the role of snowpack in polar boundary layer chemistry, especially as a direct source of reactive bromine (BrOX=BrO+Br) and nitrogen (NOX=NO+NO2) in the Arctic springtime. Surface snow samples were collected daily from a Canadian high Arctic location at Eureka, Nunavut (80° N, 86° W) from the end of February to the end of March in 2018 and 2019. The snow was sampled at several sites representing distinct environments: sea ice, inland close to sea level, and a hilltop ~600 m above sea level (asl). At the inland sites, surface snow salinity has a double-peak distribution with the first and lowest peak at 0.001–0.002 practical salinity unit (psu), which corresponds to the precipitation effect, and the second peak at 0.01–0.04 psu, which is likely related to the salt accumulation effect (due to loss of water vapour by sublimation). Snow salinity on sea ice has a triple-peak distribution; its first and second peaks overlap with the inland peaks, and the third peak at 0.2–0.4 psu is likely due to the sea water effect (due to upward migration of brine on sea ice). At all sites, snow sodium and chloride concentrations increase by almost 10-fold from the top 0.2 cm to ~1.5 cm in depth. Surface snow bromide at sea level is significantly enriched, indicating a net sink of atmospheric bromine. Moreover, surface snow bromide at sea level has an increasing trend over the measurement time period, with mean slopes of 0.024 in the 0–0.2 cm layer and 0.016 μM d-1 in the 0.2–0.5 cm layer. Surface snow nitrate at sea level also shows a significant increasing trend, with mean slopes of 0.27, 0.20, and 0.07 μM d-1 in the top 0.2 cm, 0.2–0.5 cm, and 0.5–1.5 cm layers, respectively. Using these trends, an integrated net deposition flux of bromide of 1.01×107 molecules cm-2 s-1 and an integrated net deposition flux of nitrate of 2.6×108 molecules cm-2 s-1 were derived. In addition, nitrate and bromide in the morning samples are significantly higher than the afternoon samples, indicating a strong photochemistry effect. However, the mean bromide loss rate (0.027–0.040 μM) is smaller than the nitrate loss rate (0.23–0.362 μM) by an order of magnitude, implying the reactive bromine emission flux from snowpack is significantly smaller than the reactive nitrogen emission flux, which is consistent with the large difference between their derived net deposition fluxes. After considering the photochemical loss effect, the corrected bromide deposition flux at sea level is 2.73×107 molecules cm-2 s-1; for nitrate, the corrected deposition flux is 5.98×108 molecules cm-2 s-1. In addition, the surface snow nitrate and bromide at inland sites were found to be significantly correlated (R=0.48–0.76), and the [NO3-]/[Br-] ratio of 4–7 indicates a possible acceleration effect of reactive bromine in atmospheric NOX-to-nitrate conversion. This is the first time such an effect has been seen in snow chemistry data obtained with a sampling frequency as short as one day.
... However, the dominant source bromine and the underlying processes involved remain unclear, with more than half a dozen different candidates proposed. These include frost flowers (Kaleschke et al., 2004;Piot and von Glasow, 2008), first-year sea ice surface 2007b), open leads/polynyas (e.g. Peterson et al., 2016;Kirpes et al., 2019;Criscitiello et al., 2021), snowpack on tundra (Pratt et al., 2013), snowpack on sea ice (Custard et al., 2017;Peterson et al., 2019), snowpack on ice sheets (Thomas et al., 2011), and sea salt aerosols from blowing snow (Yang et al., 2008(Yang et al., , 2010(Yang et al., , 2019Frey et al., 2020;Huang et al., 2020). ...
Article
Full-text available
This study explores the role of snowpack in polar boundary layer chemistry, especially as a direct source of reactive bromine (BrOx = BrO + Br) and nitrogen (NOx = NO + NO2) in the Arctic springtime. Surface snow samples were collected daily from a Canadian high Arctic location at Eureka, Nunavut (80° N, 86° W) from the end of February to the end of March in 2018 and 2019. The snow was sampled at several sites representing distinct environments: sea ice, inland close to sea level, and a hilltop ∼ 600 m above sea level (a.s.l.). At the inland sites, surface snow salinity has a double-peak distribution with the first and lowest peak at 0.001–0.002 practical salinity unit (psu), which corresponds to the precipitation effect, and the second peak at 0.01–0.04 psu, which is likely related to the salt accumulation effect (due to loss of water vapour by sublimation). Snow salinity on sea ice has a triple-peak distribution; its first and second peaks overlap with the inland peaks, and the third peak at 0.2–0.4 psu is likely due to the sea water effect (a result of upward migration of brine). At all sites, snow sodium and chloride concentrations increase by almost 10-fold from the top 0.2 to ∼ 1.5 cm. Surface snow bromide at sea level is significantly enriched, indicating a net sink of atmospheric bromine. Moreover, surface snow bromide at sea level has an increasing trend over the measurement period, with mean slopes of 0.024 µM d−1 in the 0–0.2 cm layer and 0.016 µM d−1 in the 0.2–0.5 cm layer. Surface snow nitrate at sea level also shows a significant increasing trend, with mean slopes of 0.27, 0.20, and 0.07 µM d−1 in the top 0.2, 0.2–0.5, and 0.5–1.5 cm layers, respectively. Using these trends, an integrated net deposition flux of bromide of (1.01 ± 0.48) × 107 molec.cm-2s-1 and an integrated net deposition flux of nitrate of (2.6 ± 0.37) × 108 molec.cm-2s-1 were derived. In addition, the surface snow nitrate and bromide at inland sites were found to be significantly correlated (R = 0.48–0.76) with the [NO3-]/[Br-] ratio of 4–7 indicating a possible acceleration effect of reactive bromine in atmospheric NOx-to-nitrate conversion. This is the first time such an effect has been seen in snow chemistry data obtained with a sampling frequency as short as 1 d. BrO partial column (0–4 km) data measured by MAX-DOAS show a decreasing trend in early spring, which generally agrees with the derived surface snow bromide deposition flux indicating that bromine in Eureka atmosphere and surface snow did not reach a photochemical equilibrium state. Through mass balance analysis, we conclude that the average release flux of reactive bromine from snow over the campaign period must be smaller than the derived bromide deposition flux of ∼ 1 × 107 molec.cm-2s-1. Note that the net mean fluxes observed do not completely rule out larger bidirectional fluxes over shorter timescales.
... FF on the frozen lead surface can draw saline brine expelled upward caused by the capillary effect, resulting in a salinity of up to 10% (Perovich and Richter-Menge, 1994). The FF produces a reservoir of sea salt ions and promotes the transport of sea salt aerosols owing to its large effective surface area (Hara et al., 2017;Kaleschke, 2004). Because of the large roughness created by the FF and sunken surface relative to the surrounding ice, blowing snow tends to accumulate on the lead surface (Perovich and Richter-Menge, 1994;Ulander et al., 1995), which gradually attenuates surface salinity but enhances thermal insulation and increases albedo at the freezing lead surface. ...
Article
Full-text available
Arctic sea ice leads are linear fractures in pack ice, which provide narrow windows for the enhanced exchange of mass, energy, and momentum between the atmosphere and ocean. However, the parameterisations of the sub-grid (< 1 km) lead distribution and its associated dynamic processes are still challenging for numerical sea ice modelling. This study explores the dynamics governing lead formation in the Arctic Ocean using multiple data sources including Sentinel-1 synthetic aperture radar (SAR) images, buoy array, and atmospheric reanalysis. Random forest (RF) models trained based on the polarisation and texture features of SAR images were compared, and an optimal RF model sensitive to narrow leads was selected with implementation of screening based on lead geometry. As a case study, the lead distribution along the trajectory of buoy array in the central Arctic Ocean during the freezing period of 2018-2019 was obtained. An enhanced lead opening was observed as the sea ice motion swerved from clockwise circulation following the Beaufort Gyre (BG) to meridional advection when assembled into the Transpolar Drift (TPD). Synoptically, active lead-opening events associated with ice divergence were driven by episodic cyclones. We noticed a dramatic change in the backscatter signal over the leads caused by the growth of thin ice and formation of frost flowers. The identification and the evolution of narrow leads during the freezing season given in this study are conducive to increasing our understanding of the response mechanism of sea ice to atmospheric dynamic processes and supporting the numerical simulation of leads.
... Since first-year sea ice is saltier than multi-year sea ice, it is re-garded as a more likely contributor to bromine release (Simpson et al., 2007). Other potential sources include frost flowers (Kaleschke et al., 2004), sea salt aerosols from blowing snow (Yang et al., 2010), sea spray from open leads (Peterson et al., 2015), and BrO transported down from the stratosphere into the troposphere (Salawitch et al., 2010). Furthermore, low temperatures favour the bromine explosion reaction, and a pH below 6.5 seems to be necessary to start the activation of the autocatalytic reaction cycle described above (Fickert et al., 1999). ...
Article
Full-text available
During polar spring, ozone depletion events (ODEs) are often observed in combination with bromine explosion events (BEEs) in Ny-Ålesund. In this study, two long-term ozone data sets (2010–2021) from ozonesonde launches and in situ ozone measurements have been evaluated between March and May of each year to study ODEs in Ny-Ålesund. Ozone concentrations below 15 ppb were marked as ODEs. We applied a composite analysis to evaluate tropospheric BrO retrieved from satellite data and the prevailing meteorological conditions during these events. During ODEs, both data sets show a blocking situation with a low-pressure anomaly over the Barents Sea and anomalously high pressure in the Icelandic Low area, leading to transport of cold polar air from the north to Ny-Ålesund with negative temperature and positive BrO anomalies found around Svalbard. In addition, a higher wind speed and a higher, less stable boundary layer are noticed, supporting the assumption that ODEs often occur in combination with polar cyclones. Applying a 20 ppb ozone threshold value to tag ODEs resulted in only a slight attenuation of the BrO and meteorological anomalies compared to the 15 ppb threshold. Monthly analysis showed that BrO and meteorological anomalies are weakening from March to May. Therefore, ODEs associated with low-pressure systems, high wind speeds, and blowing snow more likely occur in early spring, while ODEs associated with low wind speed and stable boundary layer meteorological conditions seem to occur more often in late spring. Annual evaluations showed similar weather patterns for several years, matching the overall result of the composite analysis. However, some years show different meteorological patterns deviating from the results of the mean analysis. Finally, an ODE case study from the beginning of April 2020 in Ny-Ålesund is presented, where ozone was depleted for 2 consecutive days in combination with increased BrO values. The meteorological conditions are representative of the results of the composite analysis. A low-pressure system arrived from the northeast to Svalbard, resulting in high wind speeds with blowing snow and transport of cold polar air from the north.
... In the Antarctic boundary layer, enhanced levels of BrO occur in spring over sea-ice-covered areas (Theys et al., 2011). The production of inorganic bromine has been proposed to be related to frost flowers on thin sea ice (Kaleschke et al., 2004) and blowing of saline snow on sea ice (Yang et al., 2010). Significant interaction with sea ice cover was experienced in spring, particularly at the ice edge transect, which could have promoted NO − 3 formation via the BrO pathway, resulting in increased values of δ 18 O-NO − 3 . ...
Article
Full-text available
Nitrogen oxides, collectively referred to as NOx (NO + NO2), are an important component of atmospheric chemistry involved in the production and destruction of various oxidants that contribute to the oxidative capacity of the troposphere. The primary sink for NOx is atmospheric nitrate, which has an influence on climate and the biogeochemical cycling of reactive nitrogen. NOx sources and NOx-to-NO3- formation pathways remain poorly constrained in the remote marine boundary layer of the Southern Ocean, particularly outside of the more frequently sampled summer months. This study presents seasonally resolved measurements of the isotopic composition (δ15N, δ18O, and Δ17O) of atmospheric nitrate in coarse-mode (> 1 µm) aerosols, collected between South Africa and the sea ice edge in summer, winter, and spring. Similar latitudinal trends in δ15N–NO3- were observed in summer and spring, suggesting similar NOx sources. Based on δ15N–NO3-, the main NOx sources were likely a combination of lightning, biomass burning, and/or soil emissions at the low latitudes, as well as oceanic alkyl nitrates and snowpack emissions from continental Antarctica or the sea ice at the mid-latitudes and high latitudes, respectively. Snowpack emissions associated with photolysis were derived from both the Antarctic snowpack and snow on sea ice. A combination of natural NOx sources, likely transported from the lower-latitude Atlantic, contribute to the background-level NO3- observed in winter, with the potential for a stratospheric NO3- source evidenced by one sample of Antarctic origin. Greater values of δ18O–NO3- in spring and winter compared to summer suggest an increased influence of oxidation pathways that incorporate oxygen atoms from O3 into the end product NO3- (i.e. N2O5, DMS, and halogen oxides (XO)). Significant linear relationships between δ18O and Δ17O suggest isotopic mixing between H2O(v) and O3 in winter and isotopic mixing between H2O(v) and O3/XO in spring. The onset of sunlight in spring, coupled with large sea ice extent, can activate chlorine chemistry with the potential to increase peroxy radical concentrations, contributing to oxidant chemistry in the marine boundary layer. As a result, isotopic mixing with an additional third end-member (atmospheric O2) occurs in spring.
Preprint
Full-text available
Ozone depletion events (ODEs) occur every spring in the Arctic and have implications for the atmospheric oxidizing capacity, radiative balance, and mercury oxidation. Here we comprehensively analyze ozone, ODEs, and their connection to meteorological and air mass history variables through statistical analyses, back-trajectories, and machine learning (ML) from observations at Villum Research Station, Station Nord, Greenland. We show that the ODE frequency and duration peak in May followed by April and March, which is likely related to air masses spending more time over sea ice and increases in radiation from March to May. Back-trajectories indicate that, as spring progresses, ODE air masses spend more time within the mixed layer and the geographic origins move closer to Villum. ODE frequency and duration are increasing during May (low confidence) and April (high confidence), respectively. Our analysis revealed that ODEs are favorable under sunny, calm conditions with air masses arriving from northerly wind directions with sea ice contact. The ML model was able to reproduce the ODE occurrence and illuminated that radiation, time over sea ice, and temperature were the most important variables for modeling ODEs during March, April, and May, respectively. Several variables displayed threshold ranges for contributing to the positive prediction of ODEs vs Non-ODEs, notably temperature, radiation, wind direction, time spent over sea ice, and snow. Our ML methodology provides a framework for investigating and comparing the environmental drivers of ODEs between different Arctic sites and can be applied to other atmospheric phenomena (e.g., atmospheric mercury depletion events).
Article
Full-text available
Leads and fractures in sea ice play a crucial role in the heat and gas exchange between the ocean and atmosphere, impacting atmospheric, ecological, and oceanic processes. We estimated lead fractions from high-resolution divergence obtained from satellite synthetic aperture radar (SAR) data and evaluated them against existing lead products. We derived two new lead fraction products from divergence with a spatial resolution of 700 m calculated from daily Sentinel-1 images. For the first lead product, we advected and accumulated the lead fractions of individual time instances. With those accumulated divergence-derived lead fractions, we comprehensively described the presence of up to 10 d old leads and analyzed their deformation history. For the second lead product, we used only divergence pixels that were identified as part of linear kinematic features (LKFs). Both new lead products accurately captured the formation of new leads with widths of up to a few hundred meters. We presented a Lagrangian time series of the divergence-based lead fractions along the drift of the Multidisciplinary drifting Observatory for the Study of Arctic Climate (MOSAiC) expedition in the central Arctic Ocean during winter 2019–2020. Lead activity was high in fall and spring, consistent with wind forcing and ice pack consolidation. At larger scales of 50–150 km around the MOSAiC expedition, lead activity on all scales was similar, but differences emerged at smaller scales (10 km). We compared our lead products with six others from satellite and airborne sources, including classified SAR, thermal infrared, microwave radiometer, and altimeter data. We found that the mean lead fractions varied by 1 order of magnitude across different lead products due to different physical lead and sea ice properties observed by the sensors and methodological factors such as spatial resolution. Thus, the choice of lead product should align with the specific application.
Preprint
Full-text available
Leads and fractures in sea ice play a crucial role in the heat and gas exchange between the ocean and atmosphere, impacting atmospheric, ecological, and oceanic processes. Our aim was to estimate lead fractions from high-resolution divergence obtained from satellite synthetic-aperture radar (SAR) data and to evaluate it against existing lead products. We derived two new lead-fraction products from divergence with a spatial resolution of 700 m calculated from daily Sentinel-1 images. For the first lead product, we advected and accumulated the lead fractions of individual time steps. With those accumulated divergence-derived lead fractions, we described comprehensively the presence of up to 10-day-old leads and analyzed their deformation history. For the second lead product, we used only divergence pixels that were identified as part of linear kinematic features (LKFs). Both new lead products accurately captured the formation of new leads with widths of a few hundred meters. We presented a Lagrangian time series of the divergence-based lead fractions along the drift of the MOSAiC expedition in the central Arctic Ocean during winter 2019/2020. Lead activity was high in fall and spring, consistent with wind forcing and ice pack consolidation. At larger scales of 50–150 km around the MOSAiC expedition, lead activity on all scales was similar, but differences emerged at smaller scales (10 km). We compared our lead products with 6 others from satellite and airborne sources, including classified SAR, thermal infrared, microwave radiometer, and altimeter data. We found that the mean lead fractions varied by 1 magnitude across different lead products due to different physical lead and sea ice properties observed by the sensors and methodological factors such as spatial resolution. Thus, the choice of lead product should align with the specific application.
Article
Full-text available
The dynamics of tropospheric ozone variations in the Arctic (Ny-A lesund, Spitsbergen, 79°N, 12°E) and in Antarctica (Neumayer-Station, 70°S, 8°W) were investigated for the period January 1993 to June 1994. Continuous surface ozone measurements, vertical profiles of tropospheric ozone by ECC-sondes, meteorological parameters, trajectories as well as ice charts were available for analysis. Information about the origins of the advected air masses were derived from 5-days back-trajectory analyses. Seven tropospheric ozone minima were observed at Ny-A lesund in the period from March to June 1994, during which the surface ozone mixing ratios decreased from typical background concentrations around 40 ppbv to values between 1 ppbv and 17 ppbv (1 ppbv O 3 corresponds to one part of O 3 in 10 9 parts of ambient air by volume). Four surface ozone minima were detected in August and September 1993 at Neumayer-Station with absolute ozone mixing ratios between 8 ppbv and 14 ppbv throughout the minima. At both measuring stations, the ozone minima were detected during polar spring. They covered periods between 1 and 4 days (Arctic) and 1 and 2 days (Antarctica), respectively. Furthermore, it was found that in both polar regions, the ozone depletion events were confined to the planetary boundary layer with a capping temperature inversion at the upper limit of the ozone poor air mass. Inside this ozone- poor layer, a stable stratification was obvious. Back-trajectory analyses revealed that the ozone-depleted air masses were transported across the marine, ice-covered regions of the central Arctic and the South Atlantic Ocean. These comparable observations in both polar regions suggest a similar ozone destruction mechanism which is responsible for an efficient ozone decay. Nevertheless, distinct differences could be found regarding the vertical structure of the ozone depleted layers. In the Arctic, the ozone-poor layer developed from the surface up to a temperature inversion, whereas in the Antarctic, elevated ozone-depleted air masses due to the influence of catabatic surface winds, were observed. DOI: 10.1034/j.1600-0889.1998.00003.x
Article
Full-text available
During the Tropospheric Ozone Production about the Spring Equinox (TOPSE) aircraft program, ozone depletion events (ODEs) in the high latitude surface layer were investigated using lidar and in situ instruments. Flight legs of 100 km or longer distance were flown 32 times at 30 m altitude over a variety of regions north of 58° between early February and late May 2000. ODEs were found on each flight over the Arctic Ocean but their occurrence was rare at more southern latitudes. However, large area events with depletion to over 2 km altitude in one case were found as far south as Baffin Bay and Hudson Bay and as late as 22 May. There is good evidence that these more southern events did not form in situ but were the result of export of ozone-depleted air from the surface layer of the Arctic Ocean. Surprisingly, relatively intact transport of ODEs occurred over distances of 900–2000 km and in some cases over rough terrain. Accumulation of constituents in the frozen surface over the dark winter period cannot be a strong prerequisite of ozone depletion since latitudes south of the Arctic Ocean would also experience a long dark period. Some process unique to the Arctic Ocean surface or its coastal regions remains unidentified for the release of ozone-depleting halogens. There was no correspondence between coarse surface features such as solid ice/snow, open leads, or polynyas with the occurrence of or intensity of ozone depletion over the Arctic or subarctic regions. Depletion events also occurred in the absence of long-range transport of relatively fresh “pollution” within the high latitude surface layer, at least in spring 2000. Direct measurements of halogen radicals were not made. However, the flights do provide detailed information on the vertical structure of the surface layer and, during the constant 30 m altitude legs, measurements of a variety of constituents including hydroxyl and peroxy radicals. A summary of the behavior of these constituents is made. The measurements were consistent with a source of formaldehyde from the snow/ice surface. Median NOx in the surface layer was 15 pptv or less, suggesting that surface emissions were substantially converted to reservoir constituents by 30 m altitude and that ozone production rates were small (0.15–1.5 ppbv/d) at this altitude. Peroxyacetylnitrate (PAN) was by far the major constituent of NOy in the surface layer independent of the ozone mixing ratio.
Article
Full-text available
The discovery of the nature of Arctic haze in the late 1970s and early 1980s [Barrie, 1986] showed that the Arctic is not a pristine environment isolated from human activity, but rather, a region well connected to natural and anthropogenic sources of chemicals by winds, ice movement, and marine currents. Copious pollution is carried on the winds to the Arctic during winter and spring from Europe and northern Asia. The study of this phenomenon led serendipitously to the discovery of ozone depletion chemistry in the Arctic marine boundary layer (MBL) at polar sunrise [Oltmans, 1981; Bottenheim et al., 1986]. In turn, research to understand surface ozone depletion chemistry led to the discovery that it is perturbing the biogeochemical cycle of many elements such as mercury; and that ozone depletion chemistry is likely to have a significant impact on radiative transfer in the atmospheric layer near the surface, with important consequences on the air-sea exchange of biologically-mediated compounds.
Article
Full-text available
Continuous year round records of atmospheric sea-salt concentrations have been recovered at three coastal Antarctic stations (Halley, Dumont d'Urville, and Neumayer) at temporal resolutions typically between 1 day and 2 weeks. The records were evaluated in terms of their spatial and seasonal variability as well as with respect to changes in the relative ion composition of airborn sea-salt particles. Annual mean sea-salt concentrations vary between 1400 ngm-3 at Dumont d'Urville, 850 ngm-3 at Neumayer, and 200 ngm-3 at Halley, respectively. They are thus considerably lower than the mean levels previously observed at the north tip of the Antarctic Peninsula but are, at their lower end, comparable to the level previously reported from Mawson. The representativeness of the atmospheric sea-salt data appears to be weak due to their high temporal variability, strong impacts of site specific aspects (such as site topography) but also due to the nonuniform sampling techniques applied so far. In accordance with the ice core evidence, the seasonal change in the atmospheric sea-salt load is found to be clearly out of phase with the seasonal cycle of the open water fraction offshore from the station as (with the exception of Dumont D'Urville) the lowest concentrations are generally observed during the local summer months. Major ion analyses of bulk aerosol and concurrently sampled fresh snow show a strong, systematic depletion of the SO42- to Na+ (Cl-) ratios with respect to bulk sea water, which appeared to be confined to the local winter half year. During that time, sea-salt SO42- was found to be depleted typically by 60-80% along with a concurrent Na+ deficit, which is in accordance with the precipitation of mirabilite. No significant fractionation of Mg2+, K+, and Ca2+ between seawater and sea-salt particles is observed. Laboratory experiments failed to simulate the SO42- fractionation in airborne seawater droplets or in the skin of seawater bubbles at low air temperatures. They gave, however, SO42- depletion factors, similar to the field observation in air and snow, in the remaining brine of seawater which was partly frozen below -8°C to an artificial sea ice surface. It is suggested therefore that the mobilization of brine from the sea ice surface constitutes an important sea-salt source in winter which may dominate the atmospheric sea-salt load at high latitudes of coastal Antarctica.
Article
In the Alaskan Beaufort Sea, ice properties and surface conditions typical of springtime leads were monitored at three sites during the initial few days of growth. Observations indicate that once the ice thickness reached approximately 2 cm, a thin (~1 mm), highly saline skim of brine formed on the surface. After only a few hours of growth the initially smooth surface of the sea ice developed some small-scale roughness. Frost flowers, the result of ice grown from the vapor phase, quickly formed on the surface of the sea ice and continued to develop during the observations. Possible mechanisms for the development and evolution of the surface skim and frost flowers are discussed. -from Authors
Article
Size-segregated aerosol samples were collected with a cascade impactor at 2 week intervals for a year at the research station Halley, situated near the coast in the Weddell Sea region of Antarctica. Sea salt is a major component of aerosol throughout the year, and we estimate that at least 60% of the total sea salt arriving at Halley is from brine and frost flowers on the sea ice surface rather than open water. Chloride in sea-salt particles is depleted relative to sodium in summer, consistent with loss of HCl as sea-salt particles react with gaseous acidic species, but is enhanced in large particles in winter because of fractionation occurring during the production of new sea ice. Non-sea-salt sulphate peaks in the summer, with the majority being in small particles indicative of a gas phase origin. The distribution of methane sulphonic acid closely follows that of non-sea-salt sulphate. In the winter, non-sea-salt sulphate is frequently negative, especially on stages collecting large particle sizes, consistent with the source of sea salt during the winter being predominantly the sea ice surface rather than open water. Nitrate peaks in the spring and summer and shows some association with sea-salt particles.
Article
Measurements and modeling studies strongly suggest that spring time depletion of ozone in the Arctic planetary boundary layer (PBL) is due to catalytic destruction by bromine atoms. However, the source of the bromine is uncertain. In this note, we propose that the source of the bromine at polar sunrise is the snow pack on the ice covering Arctic ocean and that it is released auto-catalytically, stimulated by a bromine seed from one of the brominated organic compounds, such as CHBr3, by photolysis. In this manner ~100 pptv of bromine can be transferred to the atmosphere where it can reside in the gas phase or, by scavenging, be partitioned in the aerosol or ice crystal phase. Moreover, it appears that heterogeneous recycling of bromine may be a process that self-terminates as ozone depletes to low levels. We also have included chlorine chemistry in the model in order to simulate inferred levels of chlorine atoms. This is important as it results in the production of HCHO which acts to convert post ozone depletion active bromine into HBr which is then returned to the snow pack or scavenged by aerosols or ice crystals.
Article
This paper describes a laboratory study of frost flower growth on young sea ice at different temperatures and the effect of these flowers on the surface temperature observed with an infrared radiometer. The flowers grew on sea ice which formed in a salt water tank at room temperatures of -20, -24, and -30°C, with an additional experiment at -16°C, where no flowers appeared. The growth habit and height of the observed crystals depended on the existence of a region of supersaturated vapor adjacent to the surface and on the range of temperatures in the surface boundary layer. The source of the surface brine from which the flowers grew was probably brine transport within the ice toward the cold upper surface driven by the thermomolecular pressure gradient. The evaporation of vapor from this liquid into the atmospheric boundary layer provided the supersaturated region adjacent to the ice surface. Two kinds of flowers were observed; at -20 and -24°C, dendritic crystals grew approximately between the -12 and -16°C isotherms, and at -30°C, rod-like flowers appeared between -16 and -25°C. These limits correspond to earlier work on crystal growth from the vapor. In each case, the maximum flower height approximately equaled the height of the isotherm corresponding to the colder temperature limit for each crystal type, -16°C for the dendrites and -25°C for the rods. The effect of the flowers on the radiometer surface temperature was as follows: because the flowers protrude 10-20 mm above the surface into the boundary layer, the infrared temperature of the flower-covered ice was about 4-6°C colder than that of the same ice cleared of flowers. We also found that the insulating effect of the flowers caused the ice surface temperature beneath the flowers to be 1-2°C warmer than the surrounding bare ice. The importance of the flower growth is that infrared satellite observations of thin ice in winter will be colder than the actual surface temperature, which may account for the absence of warm young ice in infrared satellite images.
Article
Frost flowers are fragile ice crystals containing salt which grow to a height of 10-30 mm on the surface of young sea ice. Such flowers are observed all over the Arctic. The importance of the flowers and their accompanying slush layer is that they provide a rapid way to change the surface albedo and increase the surface roughness young of sea ice. This paper describes a laboratory technique for growing frost flowers and the physical processes which accompany the growth. The study was carried out in a a saltwater tank located in a cold room. To grow frost flowers, we alternately cool the surface of the growing sea ice with a fan, then supply it with water vapor from a vaporizer. For these conditions and a room temperature of -22°C, the frost flowers begin to grow when the ice thickness reaches 5-8 mm. The flowers form at random locations on the ice and grow vertically to a height of 10-15 mm while spreading laterally from their original sites.Beneath the flowers, the surface is initially dry; then as the flowers spread laterally, a high-salinity slush layer forms beneath them. This layer, which forms only under the flowers, grows to a thickness of 5 mm in 48 hours and has a characteristic lateral scale of 100-200 nm. The salinity of the slush layer is about 80 psu, compared with a frost flower salinity of 100 psu. Within 24 hours of their appearance, the flowers grow to cover 75-90% of the surface. A surface water budget for the flowers and slush layer show that most of the water in the flowers and slush layer comes from the ice interior, not from the vaporizer. This implies that an external vapor source may be important in determining the initial growth of the flowers but not in their subsequent development.
Article
Formation of a sea ice cover produces profound changes in the state of the atmosphere and ocean, primarily by altering the surface heat exchange. The high albedo of the ice and its insulation of the atmosphere from the underlying water give rise to a climate over the polar oceans that is more characteristic of the continental ice sheets than of a marine environment. Unlike the ice sheets, however, sea ice is only a thin veneer whose thickness and areal extent are sensitive to small changes in heat input. Variations in sea ice extent have the potential to amplify small changes in climate through a variety of positive feedback mechanisms (Kellogg, 1975), leading to their central role in several ice age theories (Brooks, 1949; Ewing and Donn, 1956, 1958; Budyko, 1966).