scispace - formally typeset
Search or ask a question
Journal ArticleDOI

A model of the three-dimensional evolution of Arctic melt ponds on first-year and multiyear sea ice

01 Dec 2010-Journal of Geophysical Research (American Geophysical Union)-Vol. 115
TL;DR: In this article, the authors present a model that simulates the three-dimensional evolution of melt ponds on an Arctic sea ice surface and investigate the sensitivity of the melt pond cover to changes in ice topography, snow topography and vertical ice permeability.
Abstract: During winter the ocean surface in polar regions freezes over to form sea ice. In the summer the upper layers of sea ice and snow melts producing meltwater that accumulates in Arctic melt ponds on the surface of sea ice. An accurate estimate of the fraction of the sea ice surface covered in melt ponds is essential for a realistic estimate of the albedo for global climate models. We present a melt-pond-sea-ice model that simulates the three-dimensional evolution of melt ponds on an Arctic sea ice surface. The advancements of this model compared to previous models are the inclusion of snow topography; meltwater transport rates are calculated from hydraulic gradients and ice permeability; and the incorporation of a detailed one-dimensional, thermodynamic radiative balance. Results of model runs simulating first-year and multiyear sea ice are presented. Model results show good agreement with observations, with duration of pond coverage, pond area, and ice ablation comparing well for both the first-year ice and multiyear ice cases. We investigate the sensitivity of the melt pond cover to changes in ice topography, snow topography, and vertical ice permeability. Snow was found to have an important impact mainly at the start of the melt season, whereas initial ice topography strongly controlled pond size and pond fraction throughout the melt season. A reduction in ice permeability allowed surface flooding of relatively flat, first-year ice but had little impact on the pond coverage of rougher, multiyear ice. We discuss our results, including model shortcomings and areas of experimental uncertainty.

Summary (4 min read)

1. Introduction

  • The rate of decline of Arctic summer sea ice extent has increased dramatically in recent years.
  • The inability of GCMs to simulate the rapid reduction in Arctic summer sea ice extent, combined with satellite and field observations demonstrating the importance of sea ice melt, indicate the need for a more realistic representation of sea ice melt processes.
  • Horizontal water transport rates in their model vary from cell to cell depending on the solid fraction in the ice.
  • In section 3 the authors present the results of two model runs that simulate the evolution of melt ponds on first‐year ice and multiyear ice, which are compared with field data and the results of Lüthje et al. [2006].

2. Model Description

  • The automaton grid consists of cells that evolve largely independently of each other, interacting through the transport of water between cells, see Figure 1.
  • One‐dimensional thermodynamic equations following Taylor and Feltham [2004] are solved in the vertical direction in every cell to calculate the heat flux through ice, snow and meltwater (if it exists).
  • Water is driven between adjacent cells by differences in hydraulic head between the cells.
  • A higher spatial (and temporal) resolution calculation was found to have no substantial impact on the results. [19].

2.1. Calculation of Meltwater Transport and Drainage

  • The surface of sea ice is deformed by mechanical processes such as ridging, or thermodynamic processes such as the formation and drainage of melt ponds and the freezing over of partially drained ponds [Fetterer and Untersteiner, 1998] and therefore in places the sea ice surface is likely to have a negative freeboard.
  • The area covered in melt ponds is affected by horizontal and vertical water transport [Eicken et al., 2002].
  • In their model horizontal water 3 of 37 flux is calculated before vertical water flux in a given time step and vertical water flux is calculated if there is any water remaining in the cell. [24].
  • In the model described here vertical flow is limited by the lowest permeability in a vertical column.

2.2. Heat Transport Model

  • The vertical heat transport model is the same as the melt‐pond–sea‐ice model described by Taylor and Feltham [2004].
  • A simple snow model was utilized, following Maykut and Untersteiner [1971].
  • The thermodynamic model of the sea ice component (including the ice lid) was described using the equations describing a mushy layer [Feltham et al., 2006], i.e., the sea ice is assumed to consist of a solid matrix of pure ice surrounded by brine (with no air pockets).
  • The albedo depends (through the upwelling irradiance) on the presence and saturation of snow, the presence and depth of meltwater, the presence and depth of an ice lid on top of the melt pond, the depth of the sea ice beneath the melt pond, and the scattering and absorption coefficients of these media. [29].

2.3. Topography Model and Standard First‐Year and Multiyear Sea Ice Topographies

  • Due to the limited availability of sea ice topography and snow topography data, and in particular combined data, some assumptions have been made in modeling the snow and ice topographies.
  • The mean and variance of snow or ice depth is needed, which can be recovered from field data, and a covariance model is used to determine the correlation between snow or ice thicknesses at separate locations as the distance between locations increases.
  • To create an ice topography two ice topographies were initially generated, one representing ice draft below a putative sea level and one representing freeboard ice height above a putative sea level.
  • A mean ice thickness of 1.70 m was selected for the first‐year ice standard case, this is a mean ice thickness of first‐year ice with a thin snow cover as observed by Perovich et al. [2002b].
  • This is most likely unrealistic and indicates a deficiency in using a simulated rather than measured topography.

3.1. Standard Case Simulation Results

  • Day 140 is several days before the snow begins to melt.
  • Mean pond area reaches its maximum at the same time, however mean pond depth continues to increase due to enhanced melting beneath ponds and water being transported across the surface to the cells with the smallest ice surface height.
  • The percentage decrease in mass over the modeled melt season is 62.2% and the total ice ablation is 1.01 m. [41] (b) Change in mean snow depth (light blue), mean pond depth (red), and mean ice thickness with time for the standard first‐year ice case, with dashed lines representing the corresponding values for the alternative first‐year ice case.
  • In the multiyear ice case the mean surface albedo decreases rapidly as snow melt produces ponds and continues to decrease more slowly until the ponds freeze over.

3.2.1. Comparison of Standard Case Simulations With Observations

  • There is typically substantial variation between observations of melt ponds since observations are made at different points in the melt season and at different locations.
  • The maximum pond fraction observed by Perovich et al. [2002a] was around day 174,early in the season, after this pond fraction decreased, which is a pattern seen in both the first‐year ice and multiyear ice modeled standard cases where there is a peak in pond coverage after snow cover is removed.
  • The mean albedo of the ice surface over the melt season in the first‐year ice case is 0.64 and over the entire domain (which includes open ocean) is 0.56.
  • The authors briefly compare the results of the model presented in this paper with that of Lüthje et al. [2006] to 14 of 37 indicate the impact of the more realistic physics.
  • The total ablation for first‐year ice in the Lüthje et al. [2006] model of 0.75 m, was much less than the total ablation of 1.33 m in the model described here, this is due to the greater mass of water available in their model, due to the separate snow layer and the greater ice mean ice thickness. [55].

4. Sensitivity Studies: Results and Discussion

  • Below the authors present sensitivity studies that examine the impact on pond cover and ice and snow ablation of changes in snow topography, ice topography, and vertical perme- ability for first‐year ice and multiyear ice.
  • Unless otherwise indicated, all parameters are the same as for the appropriate standard case.
  • Tables 2 and 3 summarize the important results for the sensitivity studies (and the standard cases for reference) for first‐year ice and multiyear ice, respectively.
  • FYI denotes first‐year sea ice and MYI denotes multiyear sea ice.

4.1.1. Sensitivity of FYI Pond Coverage to the Snow Cover

  • The following studies examine the sensitivity of pond evolution and ablation to snow depth and roughness.
  • The snow topography used here represents the snow cover that would be expected on hummocky ice [Sturm et al., 2002], the standard deviation is increased from 0.15 m for the standard case to 0.25 m.
  • There was an increase in maximum mean pond area from 219 m2 in the standard case to 315 m2 in the rough snow case and mean pond area exceeds the standard case pond area between days 185 and 195, when pond fraction is at its greatest.
  • The smooth snow case differs from the standard first‐year ice case most obviously at the start of the season as the smaller variability in snow depth results in snow melting at the same rate across the grid causing initial pond fraction (not shown) and mean pond area to be greater than the standard case.
  • The dashed lines represent the corresponding values for the standard first‐year ice case.

4.3. Discussion of Sensitivity Studies

  • Tables 2 and 3 summarize the important results for the standard case runs and the sensitivity studies presented in this paper. [83].
  • The dashed lines represent the corresponding values for the standard multiyear ice case.
  • Such surface flooding was not observed in the multiyear ice case due to the deeper depressions in the ice, limiting the horizontal transport of water. [84].
  • Change in mean snow depth (light blue), mean pond depth (red), and mean ice thickness with time for the smooth ice case.

5. Conclusion and Further Work

  • The authors model uses the cellular automaton concept described by Lüthje et al. [2006], with significant improvements, and the one‐dimensional vertical heat transport model described by Taylor and Feltham [2004].
  • The area‐averaged surface albedo of the summer ice cover is largely determined by the area of the sea ice covered in melt ponds and the open ocean fraction.
  • (top) Variation in the fractional distribution of surface area with time for the low‐permeability case, where mean ice thickness is 2.50 m, standard deviation in ice thickness is 1.10 m, mean snow thickness is 0.30 m, and standard deviation in snow thickness is 0.25 m.
  • The dashed lines represent the corresponding values for the standard multiyear ice case.
  • Since the underlying ice topography plays such a central role in determining the location and extent of pond formation, detailed measurements of topography, in conjunction with observations of the pond evolution and surface forcing, will enable a stricter test of the melt‐pond–sea‐ice model. [92].

Did you find this useful? Give us your feedback

Figures (22)

Content maybe subject to copyright    Report

A model of the three-dimensional
evolution of Arctic melt ponds on rst-year
and multiyear sea ice
Article
Published Version
Scott, F. and Feltham, D.L. (2010) A model of the three-
dimensional evolution of Arctic melt ponds on rst-year and
multiyear sea ice. Journal of Geophysical Research, 115
(C12). C12064. ISSN 0148-0227 doi:
https://doi.org/10.1029/2010JC006156 Available at
https://centaur.reading.ac.uk/30572/
It is advisable to refer to the publishers version if you intend to cite from the
work. See Guidance on citing
.
To link to this article DOI: http://dx.doi.org/10.1029/2010JC006156
Publisher: American Geophysical Union
All outputs in CentAUR are protected by Intellectual Property Rights law,
including copyright law. Copyright and IPR is retained by the creators or other
copyright holders. Terms and conditions for use of this material are dened in
the End User Agreement
.
www.reading.ac.uk/centaur
CentAUR

Central Archive at the University of Reading
Reading’s research outputs online

A model of the threedimensional evolution of Arctic melt ponds
on firstyear and multiyear sea ice
F. Scott
1
and D. L. Feltham
1,2
Received 27 January 2010; revised 15 July 2010; accepted 20 September 2010; published 28 December 2010.
[1] During winter the ocean surface in polar regions freezes over to form sea ice.
In the summer the upper layers of sea ice and snow melts producing meltwater that
accumulates in Arctic melt ponds on the surface of sea ice. An accurate estimate of the
fraction of the sea ice surface covered in melt ponds is essential for a realistic estimate of
the albedo for global climate models. We present a meltpondseaice model that
simulates the threedimensional evolution of melt ponds on an Arctic sea ice surface. The
advancements of this model compared to previous models are the inclusion of snow
topography; meltwater transport rates are calculated from hydraulic gradients and ice
permeability; and the incorporation of a detailed onedimensional, thermodynamic
radiative balance. Results of model runs simulating firstyear and multiyear sea ice are
presented. Model results show good agreement with observations, with duration of pond
coverage, pond area, and ice ablation comparing well for both the firstyear ice and
multiyear ice cases. We investigate the sensitivity of the melt pond cover to changes in ice
topography, snow topography, and vertical ice permeability. Snow was found to have an
important impact mainly at the start of the melt season, whereas initial ice topography
strongly controlled pond size and pond fraction throughout the melt season. A reduction in
ice permeability allowed surface flooding of relatively flat, firstyear ice but had little
impact on the pond coverage of rougher, multiyear ice. We discuss our results, including
model shortcomings and areas of experimental uncertainty.
Citation: Scott, F., and D. L. Feltham (2010), A model of the threedimensional evolution of Arctic melt ponds on firstyear and
multiyear sea ice, J. Geophys. Res., 115, C12064, doi:10.1029/2010JC006156.
1. Introduction
[2] The rate of decline of Arctic summer sea ice extent has
increased dramatically in recent years. A record minimum of
ice extent was recorded in 2007, beating the previous record
minimum in 2005. The 2007 extent minimum was almost
matched again in 2008. The decrease in sea ice area has
been accompanied by a decrease in sea ice volume. For
instance, Rothrock et al. [1999] observed a 40% reduction in
average ice thickness by analyzing submarine measurements
of sea ice draft from the 1970s and 1990s. Wider area esti-
mates of sea ice thickness, based on satellite altimetry [Laxon
et al., 2003; Giles et al., 2008], also reveal a reduction in ice
thickness.
[
3] Global warming is intensified in polar regions due to
the albedo feedback mechanism [e.g., Ebert et al., 1995]
and, as a result of this, Arctic sea ice is a sensitive indicator
of climate change, as well as being an important climate
component. Climate prediction studies using Global Climate
Models (GCMs), such as the Intergovernmental Panel on
Climate Change AR4 study, are unable to simulate the
observed rapid reduction o f sea ice extent [Solomon et al.,
2007]. The inability of GCMs to simulate the rapid reduc-
tion in Arctic summer sea ice extent, combined with satellite
and field observations demonstrating the importance of sea
ice melt, indicate the need for a more realistic representation
of sea ice melt processes. In particular, GCMs do not model
melt ponds on sea ice. As the melt season progresses, part of
the surface meltwater produced accumulates to form melt
ponds that cover an increasing fraction of the surface,
reaching around 50% at the end of the melt season.
[
4] Melt ponds are a persistent feature of the summertime
sea ice surface in the Arctic [Derksen et al., 1997; Fetterer
and Untersteiner, 1998; Tucker et al., 1999; Yackel et al.,
2000; Tschudi et al., 2001]. Melt ponds have a significant
impact on the both the albedo of sea ice and the amount of
sea ice melt. The albedo of pondcovered ice is variable and
has been measured in field experiments to be between 0.1
and 0.5 [e.g., Perovich et al., 2002b; Eicken et al., 2004].
These albedo values are much lower than bare ice and
snowcovered ice, which are relatively stable at 0.60.65
and 0.840.87 [Perovich, 1996]. Since the ice concentration
in the interior Arctic is greater than 85%, melt ponds con-
tribute significantly to the areaaveraged albedo, with an
1
Centre for Polar Observation and Modelling, Department of Earth
Sciences, University College London, London, UK.
2
British Antarctic Survey, Cambridge, UK.
Copyright 2010 by the American Geophysical Union.
01480227/10/2010JC006156
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, C12064, doi:10.1029/2010JC006156, 2010
C12064 1of37

approximately linear decrease in albedo with increasing
pond fraction [Eicken et al., 2004]. For example, an
uncertainty in pond fraction of 15% over the entire Arctic
Ocean is equivalent to an uncertainty of 10% in the total ice
area in the calculation of mean Arctic Ocean albedo.
[
5] Melt pond parameterizations that can be incorporated
into GCMs are now being developed [Flocco and Feltham,
2007; Pedersen et al., 2009; Flocco et al., 2010], however,
to ensure that parameterizations are realistic we need to
understand the physics that govern melt pond evolution so
that the parameterizations can be physically based.
[
6] Our objective here is to create a model of melt pond
evolution on sea ice, based on the physics believed to
govern pond formation and growth, that can be used to
determine the sensitivity of melt ponds to ice and snow
surface topography and uncertainty in sea ice permeability,
and thus improve our understanding of the evolution of melt
ponds. Our model uses the cellular automaton concept
described by Lüthje et al. [2006], with significant im-
provements described below, and the onedimensional,
vertical heat transport model described by Taylor and
Feltham [2004]. In the model the ice cover is represented
by a horizontal square grid of cells like a checker board and
each cell contains a column of ice, which may have a melt
pond or snow cover, see Figure 1.
[
7] The initial ice and snow topographies have been
generated using standard statistical methods so that first
year and multiyear ice can be modeled using the statistical
properties of necessarily limited observations. Ice surface
and base heights are generated separately leading to a sur-
face topography with some ice surface heights below sea
level init ially. The initial surface topography in the Lüthje
et al. [2006] model is based on ice freeboard measure-
ments and all cells have positive initial freeboard.
[
8] In our model the entire floe is in hydrostatic equilib-
rium, but not necessarily every cell, and sea level with
respect to the floe is recalculated every time step. This al-
lows vertical drainage to be realistically modeled using
Darcys law, rather than take place at a fixed rate as in the
Lüthje et al. [2006] model. Horizontal water transport
rates in our model vary from cell to cell depending on the
solid fraction in the ice. Therefore in the model described
in this paper there is spatial as well as temporal variation
in drainage rate.
[
9] The ice and snow melt rates in our model are calcu-
lated from the detailed thermal and radiative balances
described by Taylor a nd Feltham [2004]. In the [thje et
al., 2006] model bare ice melts at a fixed rate and melting
beneath ponds take place at an enhanced rate using an ad
hoc algorithm motivated by observations. There is no basal
melting in the Lüthje et al. [2006] model and there is no
separate representation of snow cover.
[
10] In section 2 we present the meltpondseaice model
including the model of meltwater transport, an explanation
of how the cellular approach is combined with the one
dimensional thermodynamic model, and a description of the
construction of initial ice and snow topographies. In section
3 we present the results of two model runs that simulate the
evolution of melt ponds on firstyear ice and multiyear ice,
which are compared with field data and the results of Lüthje
et al. [2006]. In section 4 we present sensitivity studies for
both firstyear and multiyear sea ice in which we vary the
initial snow cover (depth and roughness), ice topography
(roughness), and vertical ice permeability, and compare
Figure 1. A schematic diagram of the cellular automaton. Each cell has an individual ice thickness, H,
and has a horizontal surface area of 25 m
2
. Melting decreases the ice thickness in a cell and allows a pond
to form on the surface. Water can drain through a cell or can be transported to adjacent cells.
SCOTT AND FELTHAM: EVOLUTION OF MELT PONDS C12064C12064
2of37

these studies with observations. Finally, in section 5, we
summarize our results and state our main conclusions.
2. Model Description
[11] The automaton grid consists of cells that evolve
largely independently of each other, interacting through the
transport of water between cells, see Figure 1. Each cell
represents a 5 m × 5 m square area of sea ice and, within this
area, ice thickness, meltwater depth and snow cover are
assumed to be uniform. The entire grid represents an 200 m
× 200 m area of a sea ice floe (40 cells per side). The area is
constrained to this size so that it can represent an arbitrary
section of a sea ice floe without the complication of having
to take edge effects into consideration. The boundaries are
periodic so that meltwater transported out of one edge cell is
transported back into the opposite edge cell. A time step of
the model consists of the following five stages:
[
12] 1. Onedimensional thermodynamic equations fol-
lowing Taylor and Feltham [2004] are solved in the vertical
direction in every cell to calculate the heat flux through ice,
snow and meltwater (if it exists). These calculations estab-
lish the albedo, volume of meltwater produced, basal abla-
tion and the saturation of snow on a cell by cell basis.
[
13] 2. Sea level with respect to the floe is established and
used to calculate the hydraulic head in each cell.
[
14] 3. Water is driven between adjacent cells by differ-
ences in hydraulic head between the cells. The volume of
horizontal water transport is calculated using Darcys law
for flow through a porous medium.
[
15] 4. Vertical drainage through the ice in each cell is
calculated using Darcys law and the hydraulic head.
[
16] 5. The volume of water transported into and out of
the cells is updated and one cycle of the automaton model is
complete.
[
17] Note the choice of the order of operation of (2)(4),
which corresponds to the rapidity of the relevant physical
processes, is needed in order to calculate meltwater trans-
port accurately for all practical choices of model time step
(i.e., greater than about a minute). We used a model time
step of 1 h.
[
18] Each cell in the cellular model calls a separate one
dimensional thermodynamic model, as described by Taylor
and Feltham [2004]. The thermodynamic model model is
run at a lower vertical spatial and temporal resolution than
that by Taylor and Feltham [2004] (20 grid points and time
steps of 1 h, compared with 641 grid points and time steps
of 600 seconds in the original model runs), first due to time
constraints, to allow a model run to be completed on a
typical workstation in just over a week, and secondly to
ensure that the cellular automaton and thermodynamic
models are of comparable accuracy. The resolution of the
thermodynamic model was tested in isolation from the cel-
lular model to ensure that the lowerresolution results were
not significantly different from higherresolution results.
The relatively coarse grid length of 5 m was chosen because
this is the average distance water is expected to travel in a
time step length of 1 h. A higher spatial (and temporal)
resolution calculation was found to have no substantial
impact on the results.
[
19] We describe the meltpondseaice model in the
following sections: section 2.1 describes the calculation of
meltwater transport and drainage, section 2.2 briefly de-
scribes the thermodynamic and radiative model used to
calculate melt rates, and section 2.3 describes the generation
of the sea ice and snow topographies.
2.1. Calculation of Meltwater Transport and Drainage
[
20] The surface of sea ice is deformed by mechanical
processes such as ridging, or thermodynamic processes such
as the formation and drainage of melt ponds and the freezing
over of partially drained ponds [Fetterer and Untersteiner,
1998] and therefore in places the sea ice surface is likely
to have a negative freeboard. In this model the entire floe/
computational domain is in hydrostatic equilibrium but not
individual cells. Sea level with respect to the floe is estab-
lished initially using the assumption that the entire floe is in
hydrostatic equilibrium and is then updated as mass is
removed from the surface and base of the ice.
[
21] Mean draft, D, is calculated every time step from
D ¼
P
x
i
þ x
s
þ x
p

A
; ð1Þ
where x is the mass of ice, snow and water in each cell, where
index i represents ice, s represents snow, and p represents
melt pond, r is ocean density, and A is total floe area.
[
22] The area covered in melt ponds is affected by hori-
zontal and vertical water transport [Eicken et al., 2002]. In
this model water can be removed from the grid by vertical
drainage and can be transported between cells, depending on
differences in hydraulic head between cells. We model
vertical and horizontal water transport in each cell using
Darcys law and we assume for simplicity that sea ice is a
saturated porous medium. In the vertical direction the Darcy
velocity, v, reduces to
v ¼
v
g
m
y
H
; ð2Þ
where p
v
is the vertical ice permeability, g is gravitational
acceleration, m is dynamic viscosity, which, for water, is
10
3
kg m
1
s
1
, r
m
is the density of meltwater, which is
initially formed from melted snow and is taken to be
1000 kgm
3
, y is the height of the melt pond surface above
sea level, and H is ice thickness. In the horizontal direction the
Darcy velocity, u, is given by
u ¼
h
g
m
r ; ð3Þ
where p
h
is the ice permeability in the horizontal direction,
and y is the fluid surface height.
[
23] The structure of sea ice is such that the upper surface
and several centimeters below the sea ice surface is often a
highly porous, crusty layer of sea ice [Eicken et al., 2002].
We assume that most horizontal water transport is limited by
flow through this porous crust. The solid fraction in the sea
ice crust is lower than that in the ice below and therefore the
permeability will be greater here than at any other depth in
the sea ice. The permeability at the base of the ice in the
summer melt season is small enough to make horizontal
water flux greater than vertical water flux for the same
pressure gradient, and therefore is the dominant way in
which water is transported. In our model horizontal water
SCOTT AND FELTHAM: EVOLUTION OF MELT PONDS C12064C12064
3of37

Citations
More filters
01 Dec 2014
TL;DR: The authors used spring melt-pond area to forecast the Arctic sea-ice minimum in September, which proves accurate, as increasing meltponds reduce surface albedo, allowing more melt to occur, creating a positive feedback mechanism.
Abstract: Prediction of seasonal Arctic sea-ice extent is of increased interest as the region opens up due to climate change. This work uses spring melt-pond area to forecast the Arctic sea-ice minimum in September. This proves accurate, as increasing melt-ponds reduce surface albedo, allowing more melt to occur, creating a positive feedback mechanism.

18 citations

Journal Article
TL;DR: In the summer of 1998, a program of aerial photography was carried out at the main site of the Surface Heat Budget of the Arctic Ocean (SHEBA) program at altitudes ranging from 1220 to 1830 m as discussed by the authors.
Abstract: [1] During spring and summer, the Arctic pack ice cover undergoes a dramatic change in surface conditions, evolving from a uniform, reflective surface to a heterogeneous mixture of bare ice, melt ponds, and leads. This transformation is accompanied by a significant decrease in areally averaged, integrated albedo. The key factors contributing to this reduction in albedo are the melting of the snow cover, the formation and growth of the melt ponds, and the increase in the open water fraction. To document these changes and enable quantification of the evolution of the ponds throughout the melt season, a program of aerial photography was carried out at the main site of the Surface Heat Budget of the Arctic Ocean (SHEBA) program. A modified square pattern, 50 km on a side, surrounding the SHEBA site was flown at altitudes ranging from 1220 to 1830 m. Twelve of these aerial survey photography flights were completed between 20 May and 4 October 1998. The flights took place at approximately weekly intervals at the height of the melt season, with occasional gaps as long as 3 weeks during August and September due to persistent low clouds and fog. In addition, flights on 17 May and 25 July were flown in a closely spaced pattern designed to provide complete photo coverage of a 10-km square centered on the SHEBA main site. Images from all flights were scanned at high resolution and archived on CD-ROMs. Using personal computer image processing software, we have measured ice concentration, melt pond coverage, statistics on size and shape of melt ponds, lead fraction, and lead perimeter for the summer melt season. The ponds began forming in early June, and by the height of the melt season in early August the pond fraction exceeded 0.20. The temporal evolution of pond fraction displayed a rapid increase in mid-June, followed by a sharp decline 1 week later. After the decline, the pond fraction gradually increased until mid-August when the ponds began to freeze. By mid-September the surface of virtually all of the ponds had frozen. The open water fraction varied between 0.02 and 0.05 from May through the end of July. In early August the open water fraction jumped to 0.20 in just a few days owing to ice divergence. Melt ponds were ubiquitous during summer, with number densities increasing from 1000 to 5000 ponds per square kilometer between June and August.

16 citations

Journal ArticleDOI
01 Jan 2022-Elementa
TL;DR: In this article , the authors characterize the seasonal behavior and variability in the snow, surface scattering layer, and melt ponds from spring melt to autumn freeze-up using in situ surveys and auxiliary observations and compare the results to satellite retrievals and output from two models: the Community Earth System Model (CESM2) and the Marginal Ice Zone Modeling and Assimilation System (MIZMAS).
Abstract: Melt ponds on sea ice play an important role in the Arctic climate system. Their presence alters the partitioning of solar radiation: decreasing reflection, increasing absorption and transmission to the ice and ocean, and enhancing melt. The spatiotemporal properties of melt ponds thus modify ice albedo feedbacks and the mass balance of Arctic sea ice. The Multidisciplinary drifting Observatory for the Study of Arctic Climate (MOSAiC) expedition presented a valuable opportunity to investigate the seasonal evolution of melt ponds through a rich array of atmosphere-ice-ocean measurements across spatial and temporal scales. In this study, we characterize the seasonal behavior and variability in the snow, surface scattering layer, and melt ponds from spring melt to autumn freeze-up using in situ surveys and auxiliary observations. We compare the results to satellite retrievals and output from two models: the Community Earth System Model (CESM2) and the Marginal Ice Zone Modeling and Assimilation System (MIZMAS). During the melt season, the maximum pond coverage and depth were 21% and 22 ± 13 cm, respectively, with distribution and depth corresponding to surface roughness and ice thickness. Compared to observations, both models overestimate melt pond coverage in summer, with maximum values of approximately 41% (MIZMAS) and 51% (CESM2). This overestimation has important implications for accurately simulating albedo feedbacks. During the observed freeze-up, weather events, including rain on snow, caused high-frequency variability in snow depth, while pond coverage and depth remained relatively constant until continuous freezing ensued. Both models accurately simulate the abrupt cessation of melt ponds during freeze-up, but the dates of freeze-up differ. MIZMAS accurately simulates the observed date of freeze-up, while CESM2 simulates freeze-up one-to-two weeks earlier. This work demonstrates areas that warrant future observation-model synthesis for improving the representation of sea-ice processes and properties, which can aid accurate simulations of albedo feedbacks in a warming climate.

16 citations

Journal ArticleDOI
04 Jun 2018
TL;DR: In this article, the authors constructed a simple model of melt pond boundaries as the intersection of a horizontal plane, representing the water level, with a random surface representing the topography and showed that an autoregressive class of anisotropic random Fourier surfaces provides topographies that yield the observed fractal dimension transition, with the ponds evolving and growing as the plane rises.
Abstract: During the late spring, most of the Arctic Ocean is covered by sea ice with a layer of snow on top. As the snow and sea ice begin to melt, water collects on the surface to form melt ponds. As melting progresses, sparse, disconnected ponds coalesce to form complex, self-similar structures which are connected over large length scales. The boundaries of the ponds undergo a transition in fractal dimension from 1 to about 2 around a critical length scale of 100 square meters, as found previously from area–perimeter data. Melt pond geometry depends strongly on sea ice and snow topography. Here we construct a rather simple model of melt pond boundaries as the intersection of a horizontal plane, representing the water level, with a random surface representing the topography. We show that an autoregressive class of anisotropic random Fourier surfaces provides topographies that yield the observed fractal dimension transition, with the ponds evolving and growing as the plane rises. The results are compared with a partial differential equation model of melt pond evolution that includes much of the physics of the system. Properties of the shift in fractal dimension, such as its amplitude, phase and rate, are shown to depend on the surface anisotropy and autocorrelation length scales in the models. Melting-driven differences between the two models are highlighted. Mathematics Subject Classification (2010). 51, 35, 42, 86.

15 citations


Cites background from "A model of the three-dimensional ev..."

  • ...The albedo of sea ice floes, which is the ratio of reflected sunlight to incident sunlight, is determined in late spring and summer primarily by the evolution of melt pond geometry [23, 20]....

    [...]

  • ...Small and medium scale models of melt ponds that include some of these mechanisms have been developed [5, 25, 23], and melt pond parameterizations are being incorporated into global climate models [6, 10, 15]....

    [...]

  • ...Sea ice albedo has been a significant source of uncertainty in climate projections and remains a fundamental challenge in climate modeling [6, 23, 15, 20]....

    [...]

Journal ArticleDOI
TL;DR: In this article, a 1D numerical energy-balance and phase transition supraglacial lake model, GlacierLake, is presented, which incorporates snowfall, in situ snow and ice melt, incoming water from the surrounding catchment, ice lid formation, basal freeze-up and thermal stratification.
Abstract: We present a newly developed 1-D numerical energy-balance and phase transition supraglacial lake model: GlacierLake. GlacierLake incorporates snowfall, in situ snow and ice melt, incoming water from the surrounding catchment, ice lid formation, basal freeze-up and thermal stratification. Snow cover and temperature are varied to test lake development through winter and the maximum lid thickness is recorded. Average wintertime temperatures of −2 to and total snowfall of 0 to 3.45 m lead to a range of the maximum lid thickness from 1.2 to 2.8 m after days, with snow cover exerting the dominant control. An initial ice temperature of with simulated advection of cold ice from upstream results in 0.6 m of basal freeze-up. This suggests that lakes with water depths above 1.3 to 3.4 m (dependent on winter snowfall and temperature) upon lid formation will persist through winter. These buried lakes can provide a sizeable water store at the start of the melt season, expedite future lake formation and warm underlying ice even in winter.

15 citations

References
More filters
01 Jan 2007
TL;DR: The first volume of the IPCC's Fourth Assessment Report as mentioned in this paper was published in 2007 and covers several topics including the extensive range of observations now available for the atmosphere and surface, changes in sea level, assesses the paleoclimatic perspective, climate change causes both natural and anthropogenic, and climate models for projections of global climate.
Abstract: This report is the first volume of the IPCC's Fourth Assessment Report. It covers several topics including the extensive range of observations now available for the atmosphere and surface, changes in sea level, assesses the paleoclimatic perspective, climate change causes both natural and anthropogenic, and climate models for projections of global climate.

32,826 citations

Journal ArticleDOI
TL;DR: In this paper, a one-dimensional thermodynamic model of sea ice is presented that includes the effects of snow cover, ice salinity, and internal heating due to penetration of solar radiation.
Abstract: A one-dimensional thermodynamic model of sea ice is presented that includes the effects of snow cover, ice salinity, and internal heating due to penetration of solar radiation. The incoming radiative and turbulent fluxes, oceanic heat flux, ice salinity, snow accumulation, and surface albedo are specified as functions of time. The model is applied to the central Arctic.

1,058 citations

Journal ArticleDOI
TL;DR: In this paper, a comparison of sea-ice draft data acquired on submarine cruises between 1993 and 1997 with similar data acquired between 1958 and 1976 indicates that the mean ice draft at the end of the melt season has decreased by about 1.3 m in most of the deep water portion of the Arctic Ocean, from 3.1 m in 1958-1976 to 1.8 m in the 1990s.
Abstract: Comparison of sea-ice draft data acquired on submarine cruises between 1993 and 1997 with similar data acquired between 1958 and 1976 indicates that the mean ice draft at the end of the melt season has decreased by about 1.3 m in most of the deep water portion of the Arctic Ocean, from 3.1 m in 1958–1976 to 1.8 m in the 1990s. The decrease is greater in the central and eastern Arctic than in the Beaufort and Chukchi seas. Preliminary evidence is that the ice cover has continued to become thinner in some regions during the 1990s.

995 citations

Journal ArticleDOI
30 Oct 2003-Nature
TL;DR: An eight-year time-series of Arctic ice thickness is used, derived from satellite altimeter measurements of ice freeboard, to determine the mean thickness field and its variability from 65° N to 81.5° N, which reveals a high-frequency interannual variability in mean ArcticIce thickness that is dominated by changes in the amount of summer melt, rather than byChanges in circulation.
Abstract: Possible future changes in Arctic sea ice cover and thickness, and consequent changes in the ice-albedo feedback, represent one of the largest uncertainties in the prediction of future temperature rise1,2. Knowledge of the natural variability of sea ice thickness is therefore critical for its representation in global climate models3,4. Numerical simulations suggest that Arctic ice thickness varies primarily on decadal timescales3,5,6 owing to changes in wind and ocean stresses on the ice7,8,9,10, but observations have been unable to provide a synoptic view of sea ice thickness, which is required to validate the model results3,6,9. Here we use an eight-year time-series of Arctic ice thickness, derived from satellite altimeter measurements of ice freeboard, to determine the mean thickness field and its variability from 65° N to 81.5° N. Our data reveal a high-frequency interannual variability in mean Arctic ice thickness that is dominated by changes in the amount of summer melt11, rather than by changes in circulation. Our results suggest that a continued increase in melt season length would lead to further thinning of Arctic sea ice.

537 citations

Journal ArticleDOI
TL;DR: In this article, the spectral and wavelength-integrated albedo on multi-year sea ice was measured every 2.5 m along a 200m survey line from April through October.
Abstract: [1] As part of ice albedo feedback studies during the Surface Heat Budget of the Arctic Ocean (SHEBA) field experiment, we measured spectral and wavelength-integrated albedo on multiyear sea ice. Measurements were made every 2.5 m along a 200-m survey line from April through October. Initially, this line was completely snow covered, but as the melt season progressed, it became a mixture of bare ice and melt ponds. Observed changes in albedo were a combination of a gradual evolution due to seasonal transitions and abrupt shifts resulting from synoptic weather events. There were five distinct phases in the evolution of albedo: dry snow, melting snow, pond formation, pond evolution, and fall freeze-up. In April the surface albedo was high (0.8–0.9) and spatially uniform. By the end of July the average albedo along the line was 0.4, and there was significant spatial variability, with values ranging from 0.1 for deep, dark ponds to 0.65 for bare, white ice. There was good agreement between surface-based albedos and measurements made from the University of Washington's Convair-580 research aircraft. A comparison between net solar irradiance computed using observed albedos and a simplified model of seasonal evolution shows good agreement as long as the timing of the transitions is accurately determined.

444 citations


"A model of the three-dimensional ev..." refers background or methods in this paper

  • ...This behavior has been observed on first‐year sea ice [Tucker et al., 1999; Perovich et al., 2002a]....

    [...]

  • ...The albedo of pond‐covered ice is variable and has been measured in field experiments to be between 0.1 and 0.5 [e.g., Perovich et al., 2002b; Eicken et al., 2004]....

    [...]

  • ...All forcing data is identical to that described by Taylor and Feltham [2004], is diurnally averaged, and is based on measurements made during the Surface and HEat Budget of the Arctic (SHEBA) field study [Perovich et al....

    [...]

  • ...As part of the SHEBA field study, Sturm et al. [2002] found that the Arctic snow cover distribution could be modeled by the spherical covariance model with a range of 20 m. Mean snow depth was 33.7 cm with a standard deviation of 19.3 cm. Snow depths ranged from 0 to 1.50 m. During the SHEBA field study, measurements of ice thickness were taken at intervals of 5 m along a series of straight lines of between 200 m and 500m in length across the sea ice surface. From these ice thickness measurements the mean and standard deviation of each ice type were evaluated. The sea ice range was taken to be 10 m, following Sturm et al. [2002]. Table 1 shows the mean and standard deviation in ice thickness and snow depth that were used to initialize the standard model runs....

    [...]

  • ...As part of the SHEBA field study, Sturm et al. [2002] found that the Arctic snow cover distribution could be modeled by the spherical covariance model with a range of 20 m....

    [...]