You are on page 1of 33

Progress in Oceanography 139 (2015) 89121

Contents lists available at ScienceDirect

Progress in Oceanography
journal homepage: www.elsevier.com/locate/pocean

A tale of two basins: An integrated physical and biological perspective


of the deep Arctic Ocean
B.A. Bluhm a,b,, K.N. Kosobokova c, E.C. Carmack d
a

UiT - The Arctic University of Troms, Department of Arctic and Marine Biology, 9037 Troms, Norway
School of Fisheries and Ocean Sciences, Institute of Marine Science, University of Alaska Fairbanks, 905 Koyukuk Drive, Fairbanks, AK 99775-7220, USA
c
P.P. Shirshov Institute of Oceanology, Russian Academy of Sciences, Nakhimovsky Prospect 36, Moscow 117218, Russia
d
Fisheries and Oceans Canada, Institute of Ocean Sciences, 9860 West Saanich Road, Sidney, British Columbia V8L 4B2, Canada
b

a r t i c l e

i n f o

Article history:
Available online 8 August 2015

a b s t r a c t
This review paper integrates the current knowledge, based on available literature, on the physical and
biological conditions of the Amerasian and Eurasian basins (AB, EB) of the deep Arctic Ocean (AO) in a
comparative fashion. The present day (Holocene) AO is a mediterranean sea that is roughly half continental shelf and half basin and ridge complex. Even more recently it is roughly two thirds seasonally and one
third perennially ice-covered, thus now exposing a portion of basin waters to sunlight and wind. Basin
boundaries and submarine ridges steer circulation pathways in overlying waters and limit free exchange
in deeper waters. The AO is made integral to the global ocean by the Northern Hemisphere Thermohaline
Circulation (NHTC) which drives Pacic-origin water (PW) through Bering Strait into the Canada Basin,
and counter-owing Atlantic-origin water (AW) through Fram Strait and across the Barents Sea into
the Nansen Basin. As a framework for biogeography within the AO, four basic, large-scale circulation systems (with L > 1000 km) are noted; these are: (1) the large scale wind-driven circulation which forces the
cyclonic Trans-Polar Drift from Siberia to the Fram Strait and the anticyclonic Beaufort Gyre in the southern Canada Basin; (2) the circulation of waters that comprise the halocline complex, composed largely of
waters of Pacic and Atlantic origin that are modied during passage over the Bering/Chukchi and
Barents/Siberian shelves, respectively; (3) the topographically-trapped Arctic Circumpolar Boundary
Current (ACBC) which carries AW cyclonically around the boundaries of the entire suite of basins, and
(4) the very slow exchange of Arctic Ocean Deep Waters. Within the basin domain two basic water mass
assemblies are observed, the difference between them being the absence or presence of PW sandwiched
between Arctic Surface Waters (ASW) above and the AW complex below; the boundary between these
domains is the Atlantic/Pacic halocline front. Both domains have vertical stratication that constrains
the transfer of nutrients to the surface layer (euphotic zone), thus leading to their oligotrophic state, particularly in the more strongly stratied Pacic Arctic where, despite high nutrient values in the inow,
convective reset of surface layer nutrients by haline convection in winter is virtually absent. First and
multi-year sea ice drastically alters albedo and insulates the underlying water column from extreme winter heat loss while its mechanical properties (thickness, concentration, roughness, etc.) greatly affect the
efciency of momentum transfer from the wind to the underlying water. Biologically, sea ice algal growth
in the basins is proportionally almost equal to or exceeding phytoplankton production, and is a habitat
and transport platform for sympagic (ice-associated) fauna. Owing to nutrient limitation due to strong
stratication and light limitation due to snow and ice cover and extreme sun angle, primary production
in the two basin domains is very low compared to the adjacent shelves. Severe nutrient limitation and
complete euphotic zone drawdown in the AB favors small phytoplankton, a ubiquitous deep chlorophyll
maximum layer, a low f-ratio of new to recycled carbon xation, and a low energy food web. In contrast,
nutrients persist albeit in low levels in the western EB, even in summer, suggesting light limitation,
heavy grazing or both. The higher stocks of nutrients in the EB are more conducive to marginal ice blooms
than in the AB. The large-scale ocean currents (NHTC and ACBC) import substantial expatriate, not locally
reproducing zooplankton biomass especially from the adjoining subarctic Atlantic (primarily Calanus nmarchicus), but also from the Pacic (e.g., Pseudocalanus spp., Neocalanus spp. and Metridia pacica). These
advective inputs serve both as source of food to resident pelagic and benthic biota within the basins, and

Corresponding author at: UiT - The Arctic University of Troms, Department of Arctic and Marine Biology, 9037 Troms, Norway. Tel.: +47 776 44382.
E-mail address: bodil.bluhm@uit.no (B.A. Bluhm).
http://dx.doi.org/10.1016/j.pocean.2015.07.011
0079-6611/ 2015 Elsevier Ltd. All rights reserved.

90

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Nomenclature
Acronyms
AB
Amerasian Basin
AO
Arctic Ocean
AOI
Arctic Oscillation Index
ACBC
Arctic Circumpolar Boundary Current
ASW
Arctic Surface Waters

AW
BSB
EB
FSB
NCP
PW

Atlantic-origin Water
Barents Sea Branch
Eurasian Basin
Fram Strait Branch
net community production
Pacic-origin water

as potential grazers exerting top down control on limited phytoplankton resources. Benthic organisms
within the AO basin show previously unappreciated biodiversity and surprising dispersion of species
given the isolation of individual basins and low vertical carbon ux and resulting biomass. Larval dispersion is aided by the large-scale ows and perhaps, we hypothesize in the deep benthos by convective
updrafts driven by geothermal heating. Zooplankton diversity, in contrast, is low, but again faunal assemblages are equally distributed between the EB and AB. Species pools of both pelagic and benthic communities change more with water depth rather than laterally, with the exception of expatriates and rare
species, with close ties to todays North Atlantic biogeographic region. Climate related change in the
AO is thus manifest at signicantly differing time scales. Throughout 90% of the Pleistocene the AO
has existed in glacial mode, with narrow continental shelves, greatly restricted river inow, thicker and
perhaps immobile sea ice, and total blockage of exchange with the Pacic Ocean. During the Holocene,
on shorter time scales of 1000100 years, signicant changes in high latitude climate are tied to changes
in temperature and perhaps moisture delivery patterns. The Arctic also experiences signicant
multi-decadal variability; however, the pace of change over the past three decades has been without
precedent. Within the basin interior the ice is now thinner and less compact, and thus more responsive
to wind stress (forcing and mixing). Concurrent with sea ice melt and increased river ow, the accumulation of fresh water and the stratication have increased, thus constraining vertical nutrient ux affecting phytoplankton size distributions, limiting primary production in parts of the basins now and likely in
the future, and increasing vulnerability to acidication. In addition, sea ice is now retreating on an annual
basis past the shelf break, exposing basin waters directly to sunlight and wind forcing. Thus, upwelling
favorable winds (generally from east to west) can now directly and efciently drive shelf-break upwelling, and draw nutrients from subsurface basin waters onto the shelf; at the same time upwelling favorable winds will also create onshore pressure gradients over the slope and basin which will act to slow or
block the ow of waters in the ACBC, and thus alter advective pathways of both abiotic and biotic materials. Given the opening of a new ocean to multiple user groups, we expect that the central AO will play an
increasing larger role both in the research and political arenas in the future, and we encourage pan-Arctic
international collaboration over focus on territorial boundaries.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Following earlier descriptions we here refer to the panarctic
domain north of Bering Strait in the Pacic and
Greenland-Scotland Ridge in the Atlantic as the Arctic Mediterranean; the Norwegian, Iceland and Greenland seas south of Fram
Strait and west of the Barents Sea comprise the Nordic Seas, and
the remainder, including both shelf and basin domains, is the Arctic Ocean (AO). We refer to the central basins at the Amerasian
Basin (AB) and the Eurasian Basin (EB). Our view of this ocean
has changed over time. It has changed because we have increased
our understanding of the area through painstaking scientic measurements since the time of the early Arctic explorers, to todays
increasingly diverse suite of measurements by icebreakers, ice
camps, moorings, ice-tethered prolers and satellites (Fig. 1). It
has also changed because the system itself is in a rapid state of
transition (ACIA, 2005; Wassmann and Lenton, 2012; Polyakov
et al., 2013; Bhatt et al., 2014). And central to understanding the
AO system now and into the future is to understand its least studied component: the deep basins.
At the onset of his historic voyage of 18931896 Nansen
believed the entire AO to be a shallow sea (Nansen, 1902). And
while Nansen discovered that a deep basin occupied the central

AO, its trans-arctic ridge systems were unsuspected until the middle of the 20th century (cf. Fig. 2a) which initially were deduced
from tide and temperature data (Fjeldstad, 1936; Worthington,
1953). Modern charts (Fig. 2b) now reveal two main basins, the
Eurasian (EB) and Amerasian basins (AB), separated by the Lomonosov Ridge. The EB is further separated into the Nansen and
Amundsen basins by the Gakkel Ridge, and the AB into the
Makarov and Canada Basin by the AlphaMendeleyev Ridge.
Though the job of accurately mapping the AO bathymetry is far
from complete, we recognize the important fact that more than
half (53%) the area of the present day AO is continental shelf, and
the remainder is the slope/rise/basin/ridge complex; (Fig. 2c;
Aagaard et al., 1985; Jakobsson et al., 2012). To simplify the complicated bathymetry of the Arctic Mediterranean, and to link physical and morphometric properties with biological distributions and
functions, we follow Carmack and Wassmann (2006) and divide
the Arctic Mediterranean into fundamental hydro-morphological
domains. These are the inow shelves, the interior shelves, the
outow shelves, the Riverine Coastal Domain (Carmack et al.,
2015), the two main basins, and the ridge and borderland features
(cf. Fig. 2d). This distinction is critical to the understanding of basin
dynamics and biogeography owing to the importance of
shelf-basin interactions and inter-domain advection.

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Fig. 1. Bar graph illustrating the number of ice camp and ship-based expeditions
taking place in the basins of the Arctic Oceans over the past six decades.

91

The large ratio of shelf to basin area is one of the truly unique
features of the modern AO; it is, however, a signature that has
changed often and rapidly in geologic time. The Arctic Basins began
to develop with the AB rst ca. 140150 Myrs ago, and then the EB
ca. 5070 Myrs ago (cf. Backman and Moran, 2009). The AOs deep
connection to the Pacic closed ca. 80 Myrs ago (Marincovich et al.,
1990), leaving Fram Strait (sill depth 2600 m) as its only deep connection to the global ocean. The AO became the modern Arctic ca.
4 Myrs ago with the closure of the Central American Seaway
(which began to freshen the Pacic relative to the Atlantic)
(Zauker et al., 1994; Haug et al., 2001; Hasumi, 2002) and the
opening Beringia Seaway (which allows the fresh Pacic water to
ow back into the Atlantic via the AO pathways) (Coachman and
Barnes, 1961; Carmack and McLaughlin, 2011). Ice core records
show that during this time, the Earths climate system began to
resonate with orbital cycles, leading perhaps to the
glacial-interglacial periods extant today (cf. Fedorov et al., 2006).
In glacial mode the AO is almost all deep basin (except for the Barents Sea) and is fed only by inow from the Atlantic. In contrast,

Fig. 2. Evolution of understanding of the Arctic basins bathymetry and hydro-morphometry: (a) bathymetric map of the Arctic Ocean at the mid-20th Century (Shirshov,
1944); (b) most-recent IBCAO map of the Arctic Ocean (Jakobsson et al., 2012); (c) simplied bathymetry and place names used in this paper; and (d) highly idealized
typology of the Arctic Mediterranean based on its hydro-morphological domains: AW is Atlantic Water; PW is Pacic Water; RCD is Riverine Coastal Domain (see Carmack
et al., 2015); AO is Arctic Outows; and arrows denote component ow directions.

92

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

during interglacial mode, as exists today, the Arctic is about 50%


shelf and 50% basin, and functions as an Atlantic-Pacic estuary
(Stigebrandt, 1981; Carmack and Wassmann, 2006). The abrupt
transition to the present day Holocene conditions  10 K years
ago thus marks an almost complete shift in variables and drivers
governing the Arctic marine system. The point here is that the
modern (Pleistocene) AO spends 90% of its time in glacial mode,
with narrow continental shelves, greatly restricted river inow,
thicker and perhaps immobile sea ice, total blockage of exchange
with the Pacic Ocean, and shelf-basin interactions relatively
unimportant. It is also likely that the functions of its bordering
marginal seas vary in geologic time. For example, water mass modications that currently take place on shelves would be absent during glacial mode, and the deep convection that currently takes
place in the Greenland Sea may move north into the EB in warm
times and south into the Norwegian Sea in colder times (cf.
Aagaard and Carmack, 1994). In this paper we focus on the Holocene AO, but recognize this as a transient phase to which the AO
basin ecosystems may still be adjusting.
A succinct and interdisciplinary overview of the status and current understanding of the Arctic Basin domain remains elusive, in
part because there had been insufcient time and effort to explore,
model and comprehensively describe basin properties and corresponding biological populations before climate change began to
blur and confuse our work-in-progress picture. This paper thus
addresses but a small part of this larger challenge by presenting
an overview of physical and biological conditions in Arctic basins
with a persistent focus on the role of advection in shaping ecosystems. First, a brief summary of AO physiography affecting advection and exchange is presented (Section 2). Next, a simple
typology of the large-scale circulation which denes the contiguous domains (c.f. Wassmann et al., 2015), the water mass assemblies, the sea ice climatology and freshwater sources relevant to
biogeography is offered (Section 3). Then, patterns in fuel for the
food web, including nutrient supply and primary production are
described (Section 4), and connections to zooplankton and benthos
distributions are drawn (Section 5). After that, special topics affecting biodiversity and carbon ows are addressed (Section 6) followed by an overview of physical and biological changes now
underway (Section 7). Finally, a perspective on ongoing and future
change is given (Section 8).

2. Basins, ridges, sills and gateways the ever-changing


geometry of the Arctic Basins
History: The bathymetry and physiography of the AO form the
backdrop for ocean circulation, the vertical partitioning of water
masses, and the life contained within them. The major milestones
in mapping the bathymetry of the basin were not reached until
19481950 when Russian expeditions nally mapped the Lomonosov, Gakkel and Mendeleev Ridges (cf. Frolov et al., 2005) and with
the discovery of the Alpha Ridge soon following in 1963. To this
day, gaps still exist in our knowledge of the basins bathymetry
and details of sill depths. Activities under the UN Convention on
Law of the Seas (UNCLOS) have renewed the interest in precision
seaoor mapping to resolve questions on territorial claims, and
massive bathymetric mapping efforts with a substantial focus on
the basins and shelf breaks are underway (Mayer et al., 2010;
Jakobsson et al., 2012) and in 2012 culminated in a new version
of the International Bathymetric Chart of the AO (Jakobsson
et al., 2012). In the central Arctic, one of the major updates includes
much higher resolution of canyons along the continental slopes
(Jakobsson et al., 2012), areas highly relevant for basin-shelf interactions such as dense water drainage and upwelling events (e.g.
Aagaard et al., 1981; Melling and Lewis, 1982; Williams et al.,

2006; Williams and Carmack, 2008; Tremblay et al., 2011;


Pickart et al., 2013).
Gateways, Basins, Ridges and Sills: Four fundamental gateways
are recognized: Fram Strait, the Barents Sea Opening, Bering Strait
and the Davis Strait; through these passages ow most of the
Arctic/sub-Arctic exchanges. Critically important for the interaction with the world ocean, Fram Strait (sill depth 2600 m) between
Greenland and Svalbard forms the only present deep-water connection between the high Arctic and the North Atlantic. Exchange
of deep water and organic matter between the Arctic Basins and
Greenland and Norwegian Basins is in both directions (e.g.
Fahrbach et al., 2001; Schauer et al., 2004). Generally, inow of
warm and saline Atlantic water into the Arctic is along the eastern
side as the West Spitsbergen Current and through the Barents Sea
opening, and outow southward is on the western side as the East
Greenland Current. Shallow and narrow Bering Strait on the Pacic
side provides a needle eye for Pacic-origin water to enter the AO
with a generally northward ow direction (Woodgate and Aagaard,
2005). After crossing the Chukchi Shelf and undergoing seasonally
varying modications, Pacic-origin water primarily exits the
Chukchi Sea shelf through Barrow Canyon, Central Channel and
Herald Canyon (Pickart et al., 2005, 2010; Weingartner et al.,
2005, 2013).
The discoveries of the under-sea ridge systems, current ows
and vertical structure of the Arctic Basin waters combined with
the history of primarily national expeditions has shaped our current perception of the central Arctic as a tale of two main basins
and four sub-basins. The massive ridges across the central basin
not only divide the sea oor, but provide boundaries for water
mass exchange and dispersal of biota (but see Section 6.2). Occupying roughly 50% of the Arctic area combined, the EB and AB are
inuenced by their intimate connection to the Atlantic and Pacic
Oceans, respectively, and by the often broad shelves around them.
The very different characteristics of the Atlantic and Pacic water
masses entering these basins as well as the different depths and
widths of the inow areas, Fram Strait, the Barents Sea Opening
and Bering Strait, set up the vertical and horizontal water mass
structure in the basins.
The central basins are classied into continental slopes and
rises leading to rather at abyssal plains that include some highlands, and into large ridges (Jakobsson et al., 2003). The massive
ridges form physical barriers between the major Arctic
sub-basins. The ultra-slow spreading Gakkel Ridge (Dick et al.,
2003) extends the North Atlantic Mid-Ocean Ridge system into
the Eurasian Arctic where it separates the Nansen and Amundsen
basins. Within its graben system is the deepest known location
in the AO (5243 m; Jakobsson, 2002). The Nansen Basin reaches
depths of 4000 m while in the Amundsen Basin extensive depths
of over 4500 m form the deepest plain of the central Arctic, the
Pole Abyssal Plain. The Lomonosov Ridge (sill depth 1870 m near
88.7N, 156.0E; see Bjrk et al., 2007), stretches continuously all
the way from the Greenland continental slope to the Russian shelf
off the New Siberian Islands separating the Eurasian from the AB
(Jakobsson et al., 2004). Within the AB, the massive AlphaMendeleev Ridge system (sill depth 2600 m at Cooperation Gap),
which occupies almost 8% of the entire AO, separates the Canada
Basin (maximum depth 3800 m) from the Makarov Basin (maximum depth 4000 m). This ridge, discovered during the drift of Ice
station Alpha, is generally less steep and much more fragmented
that the other two ridges. Several other large bathymetric features
such as the Yermak Plateau north of Svalbard and the Chukchi
Borderland in the Canada Basin add additional complexity to the
Nansen and Canada basins. The continental slopes and the ridges
play a major role in steering ocean circulation (Aagaard, 1989)
and the distribution of biological communities (Deubel, 2000;
Kosobokova and Hirche, 2009) as we will explore below.

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

93

Fig. 3. Historical perspective on Arctic Ocean connectivity and modern ice cover changes: (a) map from the mid-19th century illustrating the then prevailing notion of a
central Arctic Ocean kept ice-free by inowing ocean currents (from Overland et al., 2011); (b) highly idealized schematic of ows linking arctic and subtropical waters
according to the requirement to materials from regions where there is excess to regions where there is decit (from Sandstrm, 1919); (c) map showing present day estimates
of transports and average T and S properties of the inows and outows based on an inverse model applied to boundary uxes in summer 2005 (Tsubouchi et al., 2012), and
mass balances estimates at Fram Strait (2.0 2.7 Sv, 19972006, Schauer et al., 2008), at the Barents Sea Opening (2.0 Sv, 19972007, Smedsrud et al., 2010), at Bering Strait
(0.8 0.2 Sv, 19912004, Woodgate et al., 2005; Melling et al., 2008) and Davis Strait (2.3 0.7 Sv for 20042005; Curry et al., 2011); and (d) map comparing the Arctic
Ocean ice cover (September minimum) for the 19792000 mean versus 2012 and overlaid on ocean bathymetry.

3. Of roundabouts, intersections and elevators disentangling


the physical oceanography of the Arctic Basin
History: During the early debates of the 19th Century on the
character of the Arctic Basin, the curious notion of an ice free central Polar Sea maintained by the warm inows from the Atlantic
and Pacic oceans was proposed. For example, Captain Silas Bent
(1872) hypothesized that the Gulf Stream and Kuroshio, penetrate
the AO and carry with them warmth enough not only to dissolve all
the ice and snow they encounter in their paths, but enough also to
keep an open sea about the Pole at all seasons of the year (cf.
Overland et al., 2011; Fig. 3a). Sandstrm (1919) later discussed
the climatological requirement for arctic/subtropical exchange
and wrote, The effects of currents originating in physical change
in the water is simply that of transporting water from the region

where such water abounds to where it is rare . . . the complimentary physical processes whereby water is added to or drawn from
a certain layer adapt themselves so as to balance one another in
rate (Fig. 3b). While the expeditions of the late 1800s and early
1900s, for example Nansens Fram expedition in 18931896
(Nansen, 1902), provided proof against the open Polar Sea concept,
they did result in early evidence of Atlantic inows to the deep
basins and of the primary direction of ice drift across the Arctic
Basin via Transpolar Drift (Rudels, 2012). These ndings were later
to be conrmed by observations from Russian ice drift stations
(called North Pole) that began in 1937 (Shirshov and Fedorov,
1938). The associated uxes of mass, heat and salt by Atlantic
and Pacic waters have recently been summarized under the banner of the Arctic and Subarctic Ocean Fluxes program (Dickson
et al., 2007, 2008; Fig. 3c). While the notion of an ice free polar

94

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Fig. 4. Horizontal maps of annual mean near-surface (20 m) (a) salinity and (b) density north of 20N for the Northern Hemisphere illustrate the basic estuarine circulation
forced by low salinity and low density waters entering from the Pacic and more saline and denser waters entering from the Atlantic. Data source: http://odv.awi.de/en/data/
ocean/world_ocean_atlas_2009/.

sea has long been dismissed, strictly, it is interesting that it does


draw attention to advective inputs, which, ironically, may be
affecting todays pattern in sea ice retreat (Fig. 3d; Shimada
et al., 2005; Carmack and Melling, 2011; Polyakov et al., 2012).
Present data, discussed next, suggest a wind-driven ice cover
and upper ocean, a strong halocline complex separating the upper
ocean from the underlying waters of Atlantic origin, a weak but
eddy-rich interior circulation rimmed by a contiguous system of
topographically-guided boundary currents, called collectively the
Arctic Circumpolar Boundary Current (ACBC), owing cyclonically
along the margins of the ocean basins, and a very slow exchange
of deep waters within the basins (e.g. Aagaard, 1989; Rudels
et al., 1994; McLaughlin et al., 2002; Timmermans et al., 2003;
Aksenov et al., 2011).
3.1. Large-scale circulation
The AO is made integral to the global ocean by exchange ows
with the subarctic Atlantic and Pacic oceans, with the Atlantic
dominating in terms of mass and heat uxes and the Pacic dominating in terms of impact of freshwater ux on the vertical stratication (Aagaard and Carmack, 1989; Dickson et al., 2008). This
climate-scale system, evident in maps of surface salinity, density
(Fig. 4a and b) and dynamic topography (Fig. 5a), is linked to the
northern hemisphere hydrological cycle (or freshwater loop), in
which zonal distillation due to the transport of water vapor from
the Atlantic to the Pacic by Trade Winds across the Isthmus of
Panama, and meridional distillation due to net evaporation in
the low latitudes and net precipitation and river run-off in the high
latitudes result in a subarctic Pacic that is fresher and less dense
than its Atlantic counterpart (Fig. 4a and b) (Aagaard et al., 2006;
Carmack, 2007; Carmack and McLaughlin, 2011). The resulting
steric height difference (Fig. 5a) drives Pacic waters through Bering Strait (sill depth 50 m) and into the Canada Basin (Coachman
and Barnes, 1961) where it encounters and overows Atlantic
Water that has entered the AO via both Fram Strait (sill depth
2600 m) and the Barents Sea (sill depth 400 m). Both inows
encounter lighter surface waters freshened by river inows and
sea ice melt. This establishes the estuarine circulation of the
AO with counter ows of warm, saline Atlantic waters below
cooler and fresher Pacic waters. Low salinity Arctic waters and
sea ice exiting the AO via Fram Strait and the Canadian Archipelago

(mainly Nares Strait) are exported into the convective gyres of the
subarctic North Atlantic and then join the global thermohaline circulation which carries low-salinity waters back to the low latitudes, and therefore contributes to closing the global freshwater
loop (cf. Carmack and McLaughlin, 2011; for summary).
For the purpose of understanding biogeographical distributions
within the AO, and to serve as a backdrop for an ecology of advection (c.f. Wassmann et al., 2015), it is useful to recognize four basic
large-scale (with L > 1000 km) circulation systems; these are: (1)
the large scale wind-driven circulation of ice and the upper ocean
(0 m to 50 or 150 m, depending on location) which forces the
cyclonic Trans-Polar Drift from Siberia to the Fram Strait and the
anticyclonic Beaufort Gyre in the southern Canada Basin
(Fig. 5a); (2) the circulation of waters that comprise the halocline
complex, composed largely of waters of Pacic and Atlantic origin
that are modied by freeze/thaw processes during passage over
the Bering/Chukchi and Barents/Siberian shelves, respectively
(Jones and Anderson, 1986; Aksenov et al., 2011); (3) the
topographically-trapped Arctic Circumpolar Boundary Current
(ACBC) which carries AW cyclonically around the boundaries of
the entire suite of basins (Aagaard, 1989; Aksenov et al., 2011;
Rudels et al., 2013; Fig. 5c); and (4) the very slow exchange of
AO deep waters which form initially in the Greenland Sea, enter
through Fram Strait, and spread within the basin interior sequentially to the Nansen, Amundsen, Makarov and Canada basins
(Macdonald et al., 1993; Schlosser et al., 1997; Rudels et al.,
2012; Fig. 5d). Exchange of the latter from basin to basin is laterally
constrained by the sequence of sills, but is aided in vertical motion
by geothermal heating (Timmermans et al., 2003; Carmack et al.,
2012). A summary of mid-water sources, pathways and associated
fronts are shown in Fig. 5d. Details of this circulation typology are
discussed next.
Surface Circulation: Much of what we know of surface circulation patterns stems from early observation collected from Russian
ice stations (cf. Coachman and Aagard, 1974; Frolov et al., 2005)
and was deduced from observations of sea ice drift, starting with
Nansen (1902) and culminating in the International Arctic Buoy
Program (IABP) (Rigor et al., 2002; Prman et al., 2004). Sea ice
and water masses lying above the cyclonically spreading Atlantic
waters (including the ACBC) are primarily wind-driven. Two basic
systems are manifest: (1) the cyclonic (Atlantic) Arctic bounded to
the east by the Siberian shelves and to the west by the Trans-Polar

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

95

Fig. 5. Schematic representations of Arctic Ocean circulation: (a) Surface circulation of the Arctic Ocean as shown by dynamic topography (20/400 dbar) (World Ocean
Database 2013, (b) summary of mid-water halocline sources, ows and associated fronts (blue shows Pacic-origin waters, maroon shows Atlantic-origin waters, thick
maroon line depicts the front between them) (after McLaughlin et al., 1996); (c) schematic representation of the Arctic Circumpolar Boundary Current system (ACBC) derived
from Atlantic water inows (after Aksenov et al., 2011; Rudels et al., 2013); and (d) schematic representation of deep water exchange (Aagaard et al., 1985). BG is the Beaufort
Gyre, BSB is the Barents Sea Branch, FSB is the Fram Strait Branch, GG is the Greenland Gyre, NAC is the Norwegian-Atlantic Current, NCC is the Norwegian Coastal Current,
TPD is the Transpolar Drift.

Drift and (2) the anticyclonic Beaufort Gyre over the Canada Basin
(Fig. 5a). The cyclonic ow over the EB and some of the Makarov
Basin denes a fundamentally divergent gyre, while the anticyclonic ow of the Beaufort Gyre denes a convergent system
(Fig. 5a); an important distinction in terms of regional stratication
and nutrient availability to the euphotic zone.
In recent years our understanding has been advanced by the use
of satellite observations (sea surface height and bottom pressure)
and much expanded in situ observations by ships, submarines, aircraft, ice camps, moorings and ice-tethered prolers (Morison
et al., 2012). From this has come a much deeper appreciation of
the degree of variance from long-term mean in both velocity and
in the spreading pathways of water masses. Proshutinsky and
Johnson (1997) used a modeling approach to demonstrate two
modes of AO circulation according to patterns in atmospheric forcings. It was later suggested that such variability could inuence the
storage and release of freshwater within the Beaufort Gyre into the
convective gyres of the Nordic Seas downstream (Proshutinsky
et al., 2002). Subsequently, a regional-scale time series carried

out in the Canada Basin since 2003 has revealed a trend and large
year-to-year variability in freshwater storage and strength of surface circulation within the Beaufort Gyre (McLaughlin et al.,
2009; Proshutinsky et al., 2009; Krisheld et al., 2014).
Kwok et al. (2013) applied satellite data collected over a 28-year
period (19822009) to summarize the variance and shifts in decadal drift patterns in ice drift velocity and compared these to geostrophic ow elds using the polarity of the Arctic Oscillation as a
backdrop for atmospheric forcing. Mean circulation speeds over
the basin are of order 24 cm/s, with stronger ows along the
southern edge of the Beaufort Gyre, in the Transpolar Drift, and
especially in the Fram Strait Outow. These authors further report
a net strengthening of the Beaufort Gyre and the Transpolar Drift,
especially during the last decade, with over 90% of the AO displaying a positive trend in drift speed, and a decline in multiyear sea ice
coverage. Importantly, they note the spatially averaged trends
from 20012009 in drift speeds (winter +24%/decade, summer
+18%/decade) are not explained by the much smaller trends in
wind speeds (winter +1.5%/decade, summer: 3.4%/decade),

96

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

with positive trends in drift speed in regions with reduced multiyear sea ice coverage. The increased responsiveness of ice drift to
geostrophic wind is consistent with a thinner and weaker seasonal
ice cover and suggests large-scale changes in the air-ice-ocean
momentum balance related to reductions in ice strength and
concentration.
Halocline waters: Formation and subsequent circulation of halocline waters within the AO basins remains poorly known, and is
likely due to a combination of mechanisms. Aagaard et al. (1981)
and Melling and Lewis (1982) discussed the importance of sea
ice formation and brine drainage in modifying shelf-derived
waters, and Killworth and Smith (1984) demonstrated the importance of Pacic inows to halocline formation. Rudels et al.
(1996) proposed that wintertime convection in the Nansen Basin
followed by summertime capping by ice melt and outow of fresh
shelf water contributed to halocline formation. Kikuchi et al.
(2004) further discussed the formation of convectively-formed
waters in the EB and its spatial and temporal variability, noting
that large-scale advance and retreat into the Amundsen Basin
appeared to coincide with ocean circulation and frontal shifts associated with increased cyclonic circulation in the atmosphere. Jones
and Anderson (1986) used chemical distributions to distinguish
between what they termed the upper (Pacic-derived) and lower
(Atlantic-derived) halocline components. Bauch et al. (2014) use
hydrochemical and oxygen isotope data do document an
along-slope front between shelf, slope and central EB waters in
the Laptev Sea, and conclude that halocline waters above the slope
are derived from upstream advection. McLaughlin et al. (1996)
argued that a front, the Atlantic-Pacic halocline front, spans the
Arctic basins where the Atlantic-derived waters subduct below
the Pacic-derived waters (Fig. 5b, also Karcher et al., 2012). They
further suggest that this front roughly aligns with and shifts
between the Lomonosov and AlphaMendeleev Ridge on decadal
time scales.
The inow of PW is likewise complicated, with branches developing south of Bering Strait from Anadyr Coastal Current waters,
central Bering Shelf waters, and the Alaska Coastal Current waters,
and continuing north across the Chukchi Sea via Herald, Hanna and
Barrow canyons (Weingartner et al., 1998, 2013; Pickart et al.,
2005, 2010). Due to the considerable widths of the Bering and
Chukchi shelves, and thus the long crossing time, these inows
are modied en route to become the denser, winter and lighter,
summer varieties of PW, with each branch undergoing differing
geochemical transformations (Nishino et al., 2013). Upon entry
into the AB, the tendency for PW to circulate in the cyclonic sense
of the ACBC is opposed by the anticyclonic, wind-driven Beaufort
Gyre (cf. Shimada et al., 2006). Using modeling results, Aksenov
et al. (2011) argued that cyclonic ow along the Alaska shelf and
upper slope in the AB is forced by steric sea level differences
between the Pacic and Arctic. The conuence of Pacic waters
with halocline source waters draining from the Siberian shelves
results in the Atlantic/Pacic Halocline front, separating the two
domains (Fig. 5c).
The Arctic Circumpolar Boundary Current: The input of AW to the
AO has been studied for over a century but, because of its complexity and variability, pathways and uxes are still subject to huge
uncertainties; a recent census of uxes through the main gateways
is given by Beszczynska-Mller et al. (2011). Between 8 and 9 Sv
(1 Sv = 106 m3 s1) enter the Nordic Seas over the Greenland Scotland Ridge (sill depth 800 m) and of this roughly half continues
on to the AO. Of the AW continuing north, about half enters the
AO via Fram Strait as the Fram Strait Branch (FSB) and subducts
below Arctic Surface waters (ASW) north of Svalbard. The other
half rst crosses the Barents and westernmost Kara seas, subducts
along the Atlantic Polar Front, continues across the eastern Barents
Sea and then enters the St. Anna Trough (Rudels et al., 2013). Here,

it is strongly modied by mixing with local Barents Sea waters and


the continuing eastward ow of the FSB to become the Barents Sea
Branch (BSB) (Rudels et al., 2012, 2013; Dmitrenko et al., 2014,
Fig. 5b). BSB water, having a broader density range than FSB
waters, both interleaves laterally and subducts below the continuing FSB. A third water mass, formed locally on the eastern Barents
and western Kara seas, drains into the basin through St. Anna
Trough (Aksenov et al., 2011). Subsequently, the three branches
become the ACBC and continue cyclonically around the basin
perimeter, with bifurcations occurring where ridge and slope
topographies intersect. There is still debate as to the transports
of AW into and out of the AO, with a total amount or 48 Sv being
generally accepted (cf. Dickson et al., 2008).
Current meter measurements obtained during the Nansen and
Amundsen Basins Observational System (NABOS) show the ACBC
to ow along this pathway: from >20 cm s1 in Fram Strait to
1520 cm s1 northeast of Svalbard to 69 cm s1 north of Franz
Josef Land to 35 cm s1 at the junctures of the Gakkel and Lomonosov ridges (I. Polyakov, pers. comm.). Note that slope currents
tend to be strongest where the slope is steep (Isachsen et al.,
2003). Aksenov et al. (2011) modeled the ACBC and demonstrated
that transports along the AO margins were forced by the joint
effects of buoyancy loss and non-local winds which created high
pressure upstream in the Barents Sea. Spall (2007) applied an
eddy-resolving numerical model and noted that lateral eddy uxes
from the boundary and vertical diffusion on the interior were
important drivers.
Deep Water: The pathways and rates of spreading of AO deep
waters are poorly known. In general, there is direct deep-water
exchange between the EB and the Norwegian and Greenland seas
via Fram Strait (sill depth 2600 m). From there the ow is
thought to proceed from the Nansen Basin to the Amundsen Basin
to the Makarov Basin and nally to the Canada Basin (MacDonald
et al., 1993; Schlosser et al., 1997) (Fig. 5c). From the AB there must
be return ow back to the EB, Nordic Sea and North Atlantic
(Aagaard et al., 1985). Processes of recirculation add complexity
to the above circulation scheme, see Rudels et al. (2013) for further
discussion. In terms of lateral advection of biotic material, it is
likely that ows are strongest through gaps in ridges and with
slope currents, especially where the slope is steepest (Isachsen
et al., 2003).
The overall motion of deep water within the basins below sill
depth is sluggish. The residence times of Arctic basin was determined using 14C by Schlosser et al. (1997) who demonstrated a
large 14C gradient between the EB and AB, showing that the Lomonosov Ridge is a barrier to direct deep-water exchange between the
two basins. They calculated the mean isolation age the time that
has elapsed after a water parcel leaves the surface having acquired
its initial concentration via exchange with the atmosphere - of the
EB bottom water (below 2500 m) to be about 250 years. (This isolation age is distinct from the average time a water parcel spends
in a particular deep basin.) The deep waters of the AB (below
2500 m) are older than those of the EB, with an isolation age of
450 years. Schlosser et al. (1997) found no signicant horizontal
or vertical gradients in 14C with in the AB (Makarov and Canada
basins) below 2250 m. Thus the deep waters on the North American side of the Lomonosov Ridge are either presently not being
ventilated (Macdonald and Carmack, 1991; Macdonald et al.,
1993; Aagaard and Carmack, 1994), or are being ventilated slowly
with continuous renewal by shelf water (by freezing and brine
rejection on the shelves) or inuxes from the adjacent EB
(Aagaard et al., 1985; stlund et al., 1987; Jones et al., 1995;
Rudels et al., 2000). Sinking brine plumes, if they do exist (see discussions in Aagaard et al., 1985 and Rudels et al., 2013), may provide a transport mechanism to carry organic material and biota to
depth.

97

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

More rapid ows are expected along basin and ridge slopes, and
through narrow gaps in the ridges. For example, the Lomonosov
Ridge separates the EB and AB with average depths below sea level
between 1000 and 1400 m, but with a narrow gap of sill depth
1870 m near 88200 N, 148E. (Bjrk et al., 2007). This gap leads
from the Makarov Basin side of Intra Basin, a sub-basin atop the
Lomonosov Ridge. Timmermans et al. (2005) apply hydraulic theory to estimate a ow over the Lomonosov Ridge of
0.25  106 m3 s1, but Timmermans and Garrett (2006) speculate
that ows of this magnitude are not reaching fully to the bottom.
Bjrk et al. (2007), however, present hydrographic data giving evidence that ow is actually from the Makarov to the Amundsen
Basin, and subsequently interowing at intermediate depths
(17002000 m) as a return ow from the Canadian Basin to the
EB and the Nordic Seas. Note that this direction of the water overow is opposite to that previously proposed by Jones et al. (1995)
and Timmermans et al. (2005) and discussed as intermittent ow
by de Middag et al. (2009). Geologic observations (Bjrk et al.,
2007) and modeling results (Isachsen et al., 2003) are consistent
with strong current activity at the ridge crest, thus aiding planktonic dispersion.
Vertical motions may also play an important role in dispersion
of biota. For example, an outstanding feature of the AO deep water
is its near homogeneity over thicknesses exceeding 1200 m within
the Canada Basin and 800 in the Amundsen Basin, and capped by
a temperature minimum (Timmermans et al., 2003; Bjrk and
Winsor, 2006; Carmack et al., 2012). In the Canada Basin, for example, the deep waters vary in potential temperature (h) by less than
0.001 C between 2700 m and the bottom, a feature that
Timmermans et al. (2003) and Carmack et al. (2012) ascribe to
geothermal heating and vertical convection. Salinity in the Canada
Basin increases with depth to 2700 m, but, like h, is nearly constant below this depth, at S = 34.957 psu. Bjrk and Winsor
(2006) reported similar, near-homogenous deep waters in the EB,
which they also attributed to geothermal heating. Using a
time-series of deep water properties, Carmack et al. (2012) found
that Canada Basin deep waters below 2700 m warmed at a rate
of 0.0004 C yr1 between 1993 and 2010. This rate is slightly less
than that to be expected from the reported geothermal heat ux
(50 mW m2; cf. Langseth et al., 1990). Using this heat ux they
estimated a vertical velocity scale of  of 0.8 mm s1, and thus for
the 1000 m-thick Canada Basin Deep Water the time-scale for convection is calculated to be 15 days. This value suggests rapid vertical mixing of the deep waters, particularly in relation to its long
isolation age. The deep waters found above the lower continental
margin of the deep basin maintain higher temperatures than those
in the basin interior, consistent with geothermal heat distributed
through a shallower water column.
Lateral, inter-basin exchange of water along isopycnal surfaces
at or near sill depths may effectively disperse biota. Carmack

et al. (2012), for example, call attention to a deep temperature


minimum (hmin) layer that overlies the deep water, and that it is
also warming at approximately the same rate, suggesting that
some
geothermal
heat
escapes
vertically
through
a
multi-stepped, 300-m-thick deep transitional layer; double diffusive convection and thermobaric instabilities were suggested as
possible mechanisms governing this vertical heat transfer. These
authors conclude that the hmin layer is maintained by exchange
with the Makarov Basin, likely over the AlphaMendeleyev sill
near Cooperation Gap.

3.2. Water mass structure and distribution


Within the basin domain four basic water masses (according to
origin) exist; these are in order of increasing density: (1) The Arctic
surface waters ensemble (ASW), (2) Pacic-origin waters (PW), (3)
Atlantic-origin waters (AW) and (4) Arctic deep waters. Increased
vertical gradients in salinity, or haloclines, separate these basic
water types (McLaughlin et al., 1996; Table 1). Representative vertical proles of salinity (S) and potential temperature (h) for each of
the four sub-basins illustrate the important point that
salt-stratication increases several fold above the path of Atlantic
Water ow from Fram Strait to the Canada Basin, owing to the progressive accumulation of freshwater from various sources en route
(Fig. 6). Of importance to biogeography and a consequence of
advection, two basic water mass assemblies are observed (the term
water mass assembly refers to the vertical stacking of water
masses within the basins). The difference between them is the
absence or presence of PW sandwiched between ASW above and
AW below. We refer to these domains as the Pacic Arctic and
Atlantic Arctic, and the boundary between them as the
Atlantic/Pacic halocline front (McLaughlin et al., 1996). This front
(cf. Fig. 5d) is characterized by strongly sloping isohalines separating the strongly and weakly stratied Pacic and Atlantic domains,
and by the disappearance of the PW temperature maximum
(Fig. 7). The key biophysical difference is that the Pacic Arctic is
much more strongly stratied and is richer in nutrients, particularly silicate (Codispoti et al., 2005; Fig. 8). Both domains have
salt-stratication that constrains the vertical transfer of nutrients
to the surface layer (euphotic zone), thus leading to their characteristic oligotrophic state, particularly in the Pacic Arctic where
the winter reset of surface layer nutrients by haline convection is
virtually absent (Codispoti et al., 2013). This difference is further
demonstrated by Laney et al. (2014) who used ice-tethered prolers equipped with bio-optical sensors to compare seasonal patterns in tow regions of the basins. They show that within the
Beaufort Gyre region (Pacic Arctic) a deep chlorophyll maximum
develops in summer, while in the Transpolar Drift System (Atlantic
Arctic) chlorophyll values are higher in the near-surface layer.

Table 1
Comparison between water mass properties of and river inow into the Eurasian and Amerasian Basins. Halocline values are typical of summer conditions.

T/S of Polar mixed layer


Halocline properties
T/S of Atlantic layer
T/S of deep water
Isolation age of bottom water
Primary production (g C m2 yr1
Largest rivers (Mean discharge >100 km3 yr1)
River discharge (8 largest rivers)
Sediment load (8 largest rivers)
Atlantic/Pacic inow (Sv)

Eurasian basin

Amerasian basin

Reference

1.7 to 4 C/33.834.8
One halocline/dS = 2
13 C/34.834.9
0.9 C/34.92
250
1015
Ob, Lena, Yenisey, Pechora,
Kolyma, Severnaya Dvina
65%
17%
58

1.7 to 4 C/28.034.4
Stepwise/dS up to 10
0.60.8 C/34.8034.85
0.5 C/34.95
450
15
Mackenzie, Yukon (indirect)

JOIS
JOIS
McLaughlin et al. (2009)
Bjrk and Winsor (2006) and Carmack et al. (2012)
Macdonald et al. (1993) and Schlosser et al. (1997)
Codispoti et al. (2013)
Holmes et al. (2002)

21%
73%
1

Holmes et al. (2002)


Holmes et al. (2002)
Fahrbach et al. (2001), Schauer et al. (2008)
and Woodgate et al. (2005, 2012)

98

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Fig. 6. Representative proles of (a) potential temperature (h), (b) salinity (S) and (c) h/S correlations for the four sub-basins of the Arctic Ocean: green is Nansen Basin, red is
Amundsen Basin; yellow is Makarov Basin and blue is Canada Basin. Water mass assemblies are distinguished by the absence of presence of Pacic Waters sandwiched within
the halocline. Note that the depth axis is plotted as square root of depth to better dene the upper ocean where detail is greatest. Data from the JOIS program 200504 and
from Polarstern ARKXII-1.

Surface waters of the AO are comprised of a base of either Atlantic or Pacic origin waters (depending on basin location) and
diluted by river inputs, ice melt and net precipitation
(Yamamoto-Kawai et al., 2009). Far more river water is supplied
to the EB than to the AB (Holmes et al., 2002); however, the residence time of these waters is relatively short, of order two years,
as it is carried quickly from the basin into Fram Strait by the cyclonic ow of the EB and the Trans-Polar Drift (Anderson et al., 1989).
In contrast, the accumulation of river water is especially pronounced in the Canada Basin, where the convergent winds of the
atmospheric Beaufort High accumulate low salinity waters of both
North American and Siberia within the anticyclonic Beaufort Gyre,
making this gyre the most strongly stratied component of the AO
(Aagaard and Carmack, 1989; Proshutinsky et al., 2009). The resulting heterogeneity in the distribution of freshwater components is
reected in maps of surface salinity (Fig. 4a).
Pacic Waters enter from the Bering Sea through Bering Strait,
transit the Chukchi Sea and enter the deep basins along three main
branches: one associated with Barrow Canyon; one east of Hannah
Shoals; and one following Herald Canyon (Weingartner et al., 1998,
2005; Aagaard et al., 2006). Inowing waters include nutrient-rich
water from the Gulf of Anadyr and fresher, lower nutrient Alaskan
Coastal Cater (Walsh et al., 1989). While crossing the wide Chukchi
Shelf, is modied seasonally by biological production, heat
exchange, ice formation and melting, and interaction with

sediments. Geochemical changes that occur here, associated with


de-nitrication, impact the global nutrient cycle (Yamamoto-Kawai
et al., 2006).
Seasonal modication on the Chukchi Shelf produces two basic
forms of Pacic-origin water that comprise the upper and middle
halocline of upper layer waters within the AB (Coachman and
Barnes, 1961; Fig. 6c and 7). The upper halocline, due to summer
modication and referred to as Pacic Summer Water, is characterized by a local temperature maximum, between 31 < S < 32, intermediate nutrient and high oxygen concentrations. The middle
halocline, due to winter modication and referred to as Pacic
Winter Water, is characterized by a temperature minimum near
S = 33.1, high nutrient and lower oxygen concentrations.
Coachman et al. (1975) observed the presence of two temperature
maxima and associated the fresher one with Alaska Coastal Water
and the more saline one with Pacic Summer Water. Likewise,
Shimada et al. (2001) observed two shallow temperature maxima,
one at 31 < S < 32 and located east of the Chukchi Plateau and one
at 32 < S < 33 found west of the Chukchi Plateau. Steele et al.
(2004) proposed that the pathways of Alaska Coastal Water and
Bering Sea Shelf Water are tied to variations in the Arctic Oscillation Index (AOI), such that when the AOI is positive the Pacic
Summer Water crosses the Chukchi Plateau and follows the Transpolar Drift and when the AOI is negative it also recirculates in the
Canada Basin. An important consequence of this variability is its

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

99

Fig. 7. Section of salinity (S) colored by potential temperature (h) for an XCTD section crossing the Amerasian Basin; inset shows station locations. Note that the Atlantic/
Pacic Halocline Front is evident in the steeply-sloping isohalines in the Canada Basin leading up to the Alpha-Mendeleev Ridge, and by the warmer Pacic varieties of
halocline waters extant within the Canada Basin (from Kikuchi, Itoh, Eert and Williams, pers. comm.). Black vertical lines in bottom gure indicate locations of XCTD
deployments.

impact on the storage of freshwater in the Canada Basin, and, upon


release back into the North Atlantic, affect convection in the subarctic gyres.
Atlantic Waters, as noted above, enter the Arctic basins as either
the warmer and more saline FSB or the cooler, fresher BSB, and is
carried basin-wide by the ACBC (Anderson et al., 1989, 1994;
Rudels et al., 1994, 1999; Swift et al., 1997; Schauer et al., 1997,
2002). Lateral spreading along near-isopycnal surfaces by thermohaline intrusions also advect properties into the interior of both the
EB (Perkin and Lewis, 1984; Rudels et al., 1994, 1999) and AB
(Carmack et al., 1995; McLaughlin et al., 2004, 2009; Woodgate
et al., 2007). Carmack et al. (1997) showed that such intrusions
are double diffusive structures in h and S, aligned or nested in
hS space across the Makarov, Amundsen and Nansen basins, and
also aligned in hS space with those observed by Perkin and
Lewis (1984). They found intrusive layers to be 4050 m thick
and noted that their density increased as they spread laterally into
the basin interior. Because of their apparent robustness over such
distance and time, Carmack et al. (1998) suggested that the layers
may be self-organized and self-propelled by the conversion of
potential to kinetic energy by salt ux convergence, contraction
on mixing or both. The presence of such features is allowed by
the weak ambient turbulence characteristics of the ice-covered
AO (c.f. Padman, 1995).
3.3. Sea ice
Unprecedented changes in sea ice have taken place in the past
decades including a reduction in sea-ice extent (Cavalieri and
Parkinson, 2012) and thickness (Kwok et al., 2009), and a decline
in the amount of multi-year ice (Comiso, 2012). The satellite record
reveals that over the past three decades the average summer minimum has decreased by 11% per decade. Since the loss of areal coverage of summer cover exceeds that of winter, the area of the
seasonal sea ice zone has increased (Kinnard et al., 2008). Impor-

tantly, with regards to this paper, is the fact that ice retreat is
now exposing surface waters over the deep basins; in 2012
approximately 40% of basin area (with depth >400 m) was ice free.
While not yet quantied, the effects of increased wind and solar
forcing on the upper layers of the deep basins are likely to be
substantial.
Wind and atmospheric thermodynamic forcing were the main
causes for the record Arctic sea ice retreat in summer2007 (e.g.,
Perovich et al., 2008). However, AW heat was likely important
for preconditioning Arctic sea ice by making it thinner over several
preceding decades and thus contributing to the extreme retreat
(Polyakov et al., 2010). A similar situation exists for incoming
warm Pacic Summer Water spreading off shelf and into the Beaufort Gyre in the Canada Basin (Shimada et al., 2006). Another feature of observed decline in sea ice extent is that the trend is not
strictly linear, but instead consists of a series of punctuated
changes, e.g. in 1989, 1998 and 2007 (see Perovich et al., 2014
for time series). One plausible explanation is that ice thickness
and strength is slowly and inexorably decreased, year by year, by
internal forcing related to increased heat advection by the atmosphere and ocean, and then shocked into a new dynamical state
by external forcings in extreme years. From the perspective of
complex systems behavior, thermodynamic forcing by the atmosphere and ocean acts as the slow variable to reduce ice cover
strength and resilience to external forcing, so that when exceptional wind patterns do occur, such as in 2007 (Stroeve et al.,
2012; Arctic Council, 2013), ice is readily exported south to melt
in the Nordic Seas.
The above noted changes in sea ice are biologically signicant
since sea ice and snow play several key physical and biogeochemical roles in the Arctic marine system. From a heat budget perspective, the presence of sea ice drastically alters albedo and insulates
the underlying water column from extreme winter heat loss. The
mechanical properties of sea ice (thickness, concentration, roughness, etc.) greatly affect the efciency of momentum transfer from

100

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Fig. 8. Horizontal maps showing distributions of nutrient concentrations: (a) surface nitrate; (b) nitrate at 200 m; (c) surface phosphate; (d) phosphate at 200 m; (e) surface
silicate; and (f) silicate at 200 m. Data derived from the Hydrochemical Atlas, CARINA, Codispoti et al., 2013 and JOIS (http://catalog.data.gov/dataset/hydrochemical-atlas-ofthe-arctic-ocean-nodc-accession-0044630, http://cdiac3.ornl.gov/waves/discrete/, http://www.nodc.noaa.gov/archive/arc0034/0072133/1.1/data/0-data/).

the wind to the underlying water, with a thinner and loosely consolidated ice cover being more effective in driving ocean currents.
Freezing of sea ice within the basins and subsequent local melting
is observed to increase stratication, while the export and melting
of ice in the adjacent North Atlantic is a signicant component in

the AOs freshwater budget (Aagaard and Carmack, 1989). Sea ice
melt water also has differing geochemical properties from ASW,
such as heavier d18O values which can be used as a tracer
(Macdonald et al., 1999), and lower alkalinity, a property that
may exacerbate ocean acidication (Yamamoto-Kawai et al.,

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

2009). Biologically, sea ice is the site of ice algal growth and transport, a habitat for ice meiofauna, zooplankton and small sh, and a
platform on which marine mammals may travel, hunt and breed
(Bluhm and Gradinger, 2008).
Two aspects of sea ice reduction are worth special attention
with regard to the deep basins. First, prior to the 21st century
the sea ice margin seldom retreated beyond the shelf-break, and
thus basin waters were seldom exposed to sunlight and wind. In
2012, in contrast, roughly 40% of the area of the deep basins was
exposed (Fig. 2d). This would increase solar radiation to basin
waters, enhance wind mixing and increase shelf-basin exchange
(see Section 6.3). The full physical and ecological consequences
of this new normal state (Wood et al., 2013; Jeffries et al., 2013)
are currently a matter of debate, and on a regional basis, primary
production may either increase (Arrigo et al., 2008, for
light-limited shelf areas) or decrease (McLaughlin and Carmack,
2010, for nutrient impoverished basins; Section 7). A second
impact is that the reduction of ice in summer now allows enhanced
gas exchange with the atmosphere, notably CO2, and this has been
implicated as a factor in increasing the acidity of AO surface waters
over the basins (Yamamoto-Kawai et al., 2011).
3.4. River inow and other freshwater in the basins
Of all global oceans the AO is the most riverine, covering only
3% of the global ocean surface area but capturing 10% of global
river runoff within its mediterranean boundaries (Carmack,
2000). Further, river inputs are increasing with global warming
(Peterson et al., 2002, 2006), albeit with disproportional increase
in Eurasian river run-off (McClelland et al., 2006). River inputs to
the AO also play a role in biogeochemical processes. Rivers bring
turbidity that can counteract enhancing effects of nutrients by
blocking sunlight to primary producers and clogging
lter-feeders (Syvitski et al., 1989). Direct nutrient inputs from rivers are relatively small compared to the advective inputs from the
Atlantic and Pacic Oceans (Codispoti and Owens, 1975), and likely
support only 10% of total NCP (Gordeev et al., 1996; Gordeev,
2000; see also Section 4.1). Alkalinity values, however, are typically
lower in incoming rivers than in AO waters, and their mixing with
and inclusion in surface waters acts in increase local vulnerability
to ocean acidication (Yamamoto-Kawai et al., 2011).
Liquid fresh water is also supplied to the AO through inputs
from sea ice melt and net precipitation and, relative to an appropriate reference salinity by the Norwegian Coastal Current and
inows from the Pacic Ocean through Bering Strait. Sinks of liquid
fresh water include export through the Canadian Arctic Archipelago and the western Fram Strait, and export of sea ice. Each of
the sources contributes uniquely to halocline formation and structure within the system, and to a variety of circulation and mixing
processes that affect biological distributions.
An Arctic freshwater budget for sources, sinks and storage was
rst produced by Aagaard and Carmack (1989), and has been subsequently updated by Serreze et al. (2006), Dickson et al. (2007),
Tsubouchi et al. (2012) and Haine et al. (2015). Most estimates of
freshwater content are based on the choice of reference salinity
(for discussion, see Carmack et al., 2008). Aagaard and Carmack
(1989) and Serreze et al. (2006) used 34.8, the mean salinity of
the AO, while Dickson et al. (2007) calculated content relative to
a salinity of 35, the salinity of incoming AW. Using available historical data, Serreze et al. (2006) calculated a net AO freshwater
export of 9200 km3 yr1 and an import of 8500 km3 yr1, leaving
a net imbalance of 700 km3 yr1. Using a similar approach, but
with a different reference salinity and study area, Dickson et al.
(2007) computed an export of about 9500 km3 yr1. Tsubouchi
et al. (2012) applied an inverse model to constrain ux estimates
for volume, heat and freshwater around the AO boundary (waters

101

above 1000 m) for a 32-day period in summer 2005. They calculated mean properties for water entering the Arctic to be
h = 4.49 C and S = 34.50, and for water leaving the Arctic, including
sea ice, h = 0.25 C and S = 33.81 (Fig. 3c). They calculated a corresponding volume transport into the Arctic of 9.2 Sv, an export of
9.3 Sv, and a net oceanic and sea ice freshwater ux of
187 48 mSv (Fig. 3c).
Regional circulation processes play a huge role in the distribution and pathways of freshwater components within the basins.
Initial river inow at source creates a coastal-trapped Riverine
Coastal Domain owing from west to east and rimming much of
the inner continental shelf (cf. Carmack and McLaughlin, 2000;
Carmack et al., 2015, Fig. 2d). Upon exiting the shelves, for example
during upwelling favorable wind events, river water mixtures
serve to freshen the surface waters within the basins. In general,
wind-driven currents serve either to store fresh water, as is the
case for the convergent Beaufort Gyre, or to export fresh water,
as is the case for the divergent Trans Polar Drift. Upon export into
the convective gyres of the subarctic North Atlantic the low salinity
arctic outow acts in poorly understood ways to inuence convection and the global overturning cell.
Because of its importance in determining stratication within
the AO and its potential importance in governing deep convection
in the adjacent subarctic Atlantic, much effort has gone into
observing and modeling variability in freshwater budget of the
AO, both in terms of uxes into and out of the basin and storage.
The original estimate of Aagaard and Carmack (1989) was that
the EB held 12.2  103 km3 of liquid fresh water while the AB held
the larger volume of 45.8  103 km3, despite the fact that the sum
total of river inows to the EB is much larger than that to the
Amerasian. The total freshwater content at this time, including
shelf domains and sea ice, was found to be near 100  103 km3,
approximately equal to the freshwater stored globally in lakes.
The larger volume in the AB was attributed to the storage of PW
within the Beaufort Gyre. Subsequently, the retreat of the Pacic/Atlantic halocline front from the vicinity of the Lomonosov to
the AlphaMendeleyev Ridge in the late 1980s and early 1990s
(McLaughlin et al., 1996; Morison et al., 1998) would have necessarily reduced this volume. Morison et al. (2006) argued for a
return of the AO to pre-1900s hydrography, but did not place this
strictly in the context of freshwater storage. Proshutinsky et al.
(2002) advanced a numerical and a conceptual model to purport
that storage and subsequent release of freshwater from the Beaufort Gyre and export to the Nordic Seas could inhibit deep convention. Since then, a number of studies have discussed interannual
variability in AO freshwater content as forced by winds, increased
river inow and sea ice melt. Proshutinsky et al. (2009) examined
year to year variability in the southern Canada Basin and found an
unprecedented increase in freshwater storage of 25% during the
rst decade of the 21st Century. McPhee et al. (2009) and Giles
et al. (2012) also conrmed the increase in freshwater storage
within the basin associated with wind-driven convergence in the
Beaufort Gyre. Rabe et al. (2014) calculated a trend in freshwater
storage between 1992 and 2012 for the entire AO (for bottom
depths >500) of 600 300 km3 yr1. A decrease in salinity made
up about 2/3 of the freshwater trend and a deepening of the upper
layer the remaining 1/3. Time and space variability in storage and
pathways are discussed in papers by Newton et al. (2008), Morison
et al. (2012), and Korhonen et al. (2012).
In any discussions of the AO freshwater budget, its future state
and its role in biogeochemical processes, it is critically important to
identify the source of the fresh water, be it river, ice melt, direct
precipitation or Pacic inow, this is done through the judicious
use of geochemical tracers such as d18O, barium, N/P ratios and
alkalinity (Carmack et al., 2008; Yamamoto-Kawai et al., 2008,
2009). This is important biologically, because freshwater from

102

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

the Pacic comes with labile, marine-derived carbon while rivers


bring terrestrial hard-to-digest carbon into the basins (Brown
et al., 2014); difference that have implications for the food
web (Iken et al., 2010; Dunton et al., 2012). Also, each of these
sources will respond differently and at different rates to climate
change.

4. Fuel for the food web Nutrients, carbon and production


4.1. Nutrients
The delivery of nutrients to the Arctic basins is complex, and the
dynamics of source and sink poorly understood. The main source
ultimately derives from the multiple inowing streams of Atlantic
and Pacic waters, and to some extent, by inowing rivers (cf.
Holmes et al., 2012). Each source then follows a different pathway
and is modied en route by seasonal and regional biogeochemical
processes before subducting below the ASW and circulating within
the basin. The basins themselves comprise an oligotrophic system
with strong salt stratication, extensive ice cover, and low sun
angle, all of which limit new production (Tremblay and Gagnon,
2009; McLaughlin and Carmack, 2010; Codispoti et al., 2013;
Varela et al., 2013). Strong salt stratication and damping of wind
mixing by ice cover (when present) constrain vertical mixing and
the attendant vertical supply of nutrients to the euphotic zone;
hence the lateral advection of nutrients combined with episodic
mixing events can be of disproportionate importance in maintaining production and in determining the size spectra of phytoplankton (Li et al., 2009). In the AB and parts of the EB a prominent
subsurface chlorophyll-a maximum layer lies immediately atop
the summer nutricline which, itself, is associated with the upper
boundary of Pacic Summer Water (McLaughlin and Carmack,
2010). In this sense, the supply of nutrients by inowing subarctic
waters to the deep basins behaves in the manner of a lling box
(cf. Killworth and Smith, 1984), with the subsurface interow driving vertical ow to supply surface outow.
The distinction in nutrient concentrations between the Pacic
and Atlantic water mass assemblies is evident in surface and
200 m maps of NO3 (Fig. 8a and b), PO4 (Fig. 8c and d) and SiO4 concentrations (Fig. 8e and f). We did not lter the data sources by season, but most data were collected during the summer. Nitrate is
near-depleted in the surface waters of the anticyclonic AB while
residual values (25 lmol l1) persist in the cyclonic EB (Fig. 8a).
This implies that primary production is nitrate limited in the AB,
while in the EB it is constrained by either light, grazing or both
(cf. Codispoti et al., 2013). Nitrate levels at 200 m are dominated
by Pacic water inputs (Fig. 8b). Surface values of phosphate in
the basins are generally above zero, ruling out phosphate as a limiting nutrient (Fig. 8c). Similarly, the 200 m levels of phosphate are
much higher in the AB than in the EB, a signal of de-nitrication on
the Chukchi Shelf that is carried with the arctic outow on into the
North Atlantic where it plays a role in the global nitrogen cycle
(Fig. 8d; Devol et al., 1997; Yamamoto-Kawai et al., 2006). Surface
values of silicate are relatively low in the incoming Atlantic waters
and extending across the Barents Sea; they are also low in the central Beaufort Gyre, perhaps we speculate reecting differential
sinking rates of diatoms in a domain of low f-ratio (Fig. 8e). Silicate
levels at 200 m are very high in the AB, reecting Pacic water
inputs, and strong gradients are evident across the Atlantic/Pacic
halocline front (Fig. 8f). Silicate distributions also show that the AB
supplies outow to both the Canadian Arctic Archipelago and Bafn Bay.
Differential nutrient concentrations in the two central basins
may constrain their responses to seasonal processes. Some nitrate,
albeit in low concentrations, remains present in surface waters of

the EB after the spring bloom in summer and may be available


for ice marginal blooms into the autumn. In contrast, virtually no
nutrients remain in the upper water column after the spring drawdown in the oligotrophic Beaufort Gyre centered over the AB,
hence preventing marginal ice blooms from forming. In both basins
delayed freeze-up may result in enhanced wind-mixing in fall that
could deliver nutritients to the euphotic zone, stimulate a fall
bloom, and thus increased production (Loeng et al., 2004; Ardyna
et al., 2014).
Freshwater buoyancy uxes are very large during summer over
most AO due to ice melt, river runoff, and the low salinity (particularly during summer) of the inowing Pacic Waters (Carmack,
2000; Woodgate et al., 2006; Fig. 3c). These processes create relatively shallow mixed layers well into the central basins. Summer
mixed layers in the AO are often <10 m deep (e.g. Codispoti et al.,
2005), and throughout most of this system mixed layer depths seldom exceed 50 m, even during winter (Steele and Boyd, 1998;
Carmack, 2007). The strong stratication places an important constraint on the nutrient supply to the photic zones within the
basins. The principal exception within the AO is the southern Nansen Basin adjacent to the Barents Sea (Codispoti et al., 2013).
Within the basin interior, and removed from shelf-break processes, the vertical supply of new nutrients to the surface (euphotic) zone is carried out by two main mechanisms: (1) haline
convection in winter during ice formation and growth, which
resets the summer (vegetative season) drawdown, and (2) by vertical mixing. Haline convection, the ux of which sets up and is
then depleted by the spring bloom, is severely constrained by salt
stratication, especially in the AB (Codispoti et al., 2013; Varela
et al., 2013). The latter, which supports the deep chlorophyll maximum (cf. Tremblay et al., 2009; McLaughlin and Carmack, 2010), is
enhanced by current shear and internal wave dissipation, and is
highly variable in time and space. For example, higher vertical ux
values are to be expected above the shelf-break and submarine
ridges where currents are faster, and this increases nutrient supply
to the deep chlorophyll maximum. Tidal motions over rough and
variable topography may likewise generate vertically propagating
internal waves that enhance vertical mixing within the halocline
(Kulikov et al., 2010). Vertical mixing can also be enhanced in open
water by wind forcing in late summer and fall prior to freeze-up,
potentially resulting in small fall blooms (cf. ACIA, 2005; Ardyna
et al., 2014).
Inowing rivers also supply nutrients and organic carbon, but in
much smaller quantities than supplied by currents, and impacts
are largely conned to their adjacent shelf regions (Holmes et al.,
2012). In a recent biogeochemical budget study for the AO,
Macdonald et al. (2010a,b) estimated the AO imports an equivalent
of 42.5 kmol N s1 from the Pacic and Atlantic oceans compared
to 2.5 kmol N s1 from rivers (Macdonald et al., 2010a,b). Dissolved
organic nitrogen loads from rivers to the AO can also potentially
contribute to the dissolved inorganic nutrient pool through
photo-oxidation or bacterial degradation (Holmes et al., 2012;
Tank et al., 2012).
4.2. Primary production pelagic and sea ice
As on the Arctic shelves, algal biomass in the basins results from
phytoplankton and ice algal production and there is much discussion about the magnitude and relative contributions of those two
components and their variability related to snow and ice cover
(Gradinger, 2009; Leu et al., 2011, 2015; Matrai et al., 2013). While
the contribution of ice algal production to total primary production
has been estimated at 425% on the Arctic shelves (Legendre et al.,
1992), it can be >50% across the basins (Gosselin et al., 1997) and
even up to 90% in the Canada Basin (when including sub-ice upper
water column production; Matrai and Apollonio, 2013). In a

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

modeling study, the sea ice in the Arctic basins contributed as


much as 18% to total ice algal biomass in the Arctic based on the
sheer size of the area, while daily production rates and production
per unit area were among the lowest in the Arctic (Deal et al.,
2011). The earlier ice algal bloom (mean production from March
to May, cf. Jin et al., 2012) plays an important role as an early food
source and fuel for reproduction of sympagic and some pelagic
taxa (Runge and Ingram, 1988; Sreide et al., 2010; Leu et al.,
2011), before the subsequent pelagic bloom occurs (typically from
May to September; Jin et al., 2012). More recently, the formation,
distribution and density of ice algal aggregates in and under sea
ice has been illuminated in the central Arctic, and these aggregates
were suggested to contribute signicantly to vertical export of
ice-derived carbon (Fernandez-Mendez et al., 2014; Katlein et al.,
2014).
Generally, both sea ice algal and pelagic primary production in
the two basin domains are one or two orders of magnitude less
that on adjacent continental shelves (Arrigo et al., 2008;
Gradinger, 2009; Table 1). The underlying causes outlined above
are nutrient limitation due to strong stratication and light limitation due to snow and ice cover and extreme sun angle (Sakshaug,
2004). Severe nutrient limitation favors small phytoplankton (Li
et al., 2009), e.g. agellates over diatoms, a low f-ratio of new to
recycled carbon xation (Varela et al., 2013), and a low energy food
web that favors jellysh over sh and marine mammals (cf.
Parsons and Lalli, 2002). In the AB the virtual absence of nutrients
in the surface waters combined with relatively high values in the
underlying PW leads to a ubiquitous deep chlorophyll maximum
immediately atop the interowing PW (McLaughlin and Carmack,
2010). For this reason, estimates of phytoplankton biomass based
on remote sensing data alone may underestimate productivity.
To explore net community production (NCP) patterns on a
pan-Arctic scale, Codispoti et al. (2013) estimated pelagic NCP from
seasonal changes in nutrient concentrations. They note that regional heterogeneity in NCP is sufciently large within the AO (two
orders of magnitude) as to make meaningful comparisons. In the
AB they found very low nitrate concentrations in the upper 50 m
(0 mmol m3) even in winter and correspondingly low annual
NCP (15 g C m2 yr1), and concluded that nutrient limitation
suppresses NCP in this region. Low wintertime nitrate concentrations in the upper layers of the AB suggest that winter reset of
nutrients by convective mixing is vanishingly small, and that
NCP in these sub-regions may, therefore, be insensitive to changes
in the ice and light regimes. In this area, new estimates suggest
that sympagic primary production may contribute an unusually
high proportion, as much as 90%, to total basin net community production (Matrai and Apollonio, 2013). In the EB in contrast, pelagic
NCP is >10 to 15 g C m2 yr1, and signicant surface layer nutrient concentrations persist throughout summer (Fig. 8), suggesting
that light or grazing or both may limit NCP.
While productivity estimates are critically important for understanding food web dynamics, knowledge of which species contribute to the production and their size, function, phenology, etc.
is arguably as important. Indeed, on the pan-arctic scale, there is
a great diversity of primary producers with 1874 known species
of phytoplankton and 1027 species of ice algae documented so
far (Poulin et al., 2011). Most of these species are large cells
(>20 lm), dominated by centric and pennate diatoms, and
dinoagellates. In the basins, taxonomic composition differs
between surface and sub-surface chlorophyll maximum layers,
although biomass was similar in that study (Coupel et al., 2012).
The contribution of smaller phytoplankton and ice algae to total
microalgal diversity and their contributing to total phytoplankton
biomass is only now beginning to be appreciated (Li et al., 2009;
Collins and Deming, 2011; Lee et al., 2012).

103

5. Of pattern and process, scarcity and hotspots distributions


in faunal biomass and food webs
Much of the carbon produced in the upper water column and in
the ice is consumed by sympagic and pelagic grazers in the upper
layers, and by omnivores and detritivores throughout the water
column and at the seaoor. As a result of low in situ production
in the basins and consumption of freshly produced carbon while
particles are sinking, vertical carbon ux to the deep-sea oor is
comparatively low (Olli et al., 2006). Particles advected from productive shelves such as the Barents, Chukchi and Kara Sea shelves
and from turbidites add substantial amounts of carbon to the vertical supply (Grantz et al., 1996; Cooper et al., 1999; Soltwedel,
2000). While the areas around the basins perimeter receive this
allochthonous input, biota in the central Arctic Basins away from
the shelves do not. As a result, biomass tends to decrease with
increasing water depth and/or distance from the shelf both in the
water column and at the sea oor (Fig. 9; e.g. Clough et al., 1997;
Bluhm et al., 2011a; Kosobokova and Hirche, 2000, 2009; Wei
et al., 2010; Kosobokova, 2012).
5.1. Zooplankton
The few historical assessments of the zooplankton biomass in
the Arctic basins reported very low biomass with less than 0.2
3.0 g dry weight (DW) m2 in the 01500 m layer, but were difcult to compare due to methodological differences or incomplete
sampling of the water column (Minoda, 1967; Hopkins, 1969a,
1969b; Pautzke, 1979; Kosobokova, 1982; Conover and Huntley,
1991). Ashjian et al. (2003) suggested that the zooplankton stock
may have been signicantly underestimated in many of these early
studies, and subsequent studies undertaken in the 1990s and early
2000s indeed reported greater zooplankton biomass and production (Wheeler et al., 1996; Mumm et al., 1998; Thibault et al.,
1999; Kosobokova and Hirche, 2000; Ashjian et al., 2003;
Hopcroft et al., 2005). Recent comprehensive assessments at >80
locations scattered over the EB and AB and using consistent methods estimated zooplankton biomass integrated over the entire
water column at ca. 224 g DW m2 dry mass (Kosobokova and
Hirche, 2009; Kosobokova and Hopcroft, 2010; Kosobokova,
2012). Regional variability is strongest within the EB, which is
undoubtedly related to the circulation patterns. A prominent belt
of elevated abundance and biomass is found along the Eurasian
continental slope from north of Svalbard to where the Lomonosov
Ridge meets the continental slope (Figs. 10 and 11a). The elevated
biomass is clearly conned to the core of the Atlantic inow (ACBC)
(Kosobokova and Hirche, 2009) that suggests that the ACBC
advects plankton populations from the North Atlantic, and that
they to a large extent remain within the zone affected by it
(Wassmann et al., 2015). This also includes enhanced biomass in
recirculating branches of the Atlantic inow along mid-ocean
ridges as the NansenGakkel Ridge (Hirche and Mumm, 1992)
and the Lomonosov Ridge (Kosobokova and Hirche, 2000). Biomass
drops sharply to a third or half of the slope values toward the deep
basins (Fig. 11a, Fig. 12; 23.8 g DW m2). Elevated biomass at the
basin periphery roughly coincides with the marginal ice zone and
seasonally ice-free waters (Kosobokova and Hirche, 2001, 2009)
where primary and, subsequently, secondary production are higher
compared to the until recently permanently ice covered and hence
food-limited basin centers (Sakshaug and Slagstad, 1991;
Sakshaug, 2004; Ulsbo et al., 2014).
Almost everywhere in the Arctic basins vertical proles of zooplankton abundance and biomass in summer are characterized by
maximum concentrations in the 050 m layer and a noticeable
decrease, by several orders of magnitude, with depths (Hopkins,

104

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Fig. 9. Schematic sections illustrating the distributions of water masses and vertical biomass distributions along representative sections crossing the Arctic basins; inset
shows the location of sections: (a) from the Bering Sea across the Lomonosov Ridge and into the Greenland Sea; (b) from the Bering Sea and through the Canada Basin and
Canadian Arctic Archipelago into Bafn Bay, (c) from the Bering Sea across the Lomonosov Ridge and onto an interior shelf off Siberia. Colored circles represent average
biomass distribution across all basins of mesozooplankton (red) and macrobenthos (blue) with water depth (modied from Kosobokova, 2012 and Bluhm et al., 2011b). AW
Atlantic Water, DW Deep Water, HC Halocline, PW Pacic Water, SW Surface Water, SAT St. Anna Trough; white dotted line denotes the approximate upper boundary of the
homogeneous layer.

Fig. 10. Map showing overlay of expatriate zooplankton distributions with halocline circulation patterns and the Atlantic/Pacic halocline front (colors as in Fig. 5d). Only
records deeper than 500 m are shown. Data from Kosobokova (2012).

1969a; Kosobokova, 1982; Ashjian et al., 2003; Kosobokova and


Hopcroft, 2010). The uppermost 050 m layer, where all freshly
produced carbon is concentrated, contains 4050% of overall biomass in summer (Kosobokova and Hopcroft, 2010; Kosobokova,
2012; Fig. 9), as elsewhere in the World oceans. In principal, all
vertical zooplankton distribution patterns are alike in both Arctic

basins (Fig. 12a and b), but the depth of the maximum depends
on the seasonal state of the zooplankton community and the interplay between populations of migrating species. Regional variability
is related to the proximity of the Atlantic gateway area (Fig. 12a,
proles highlighted in blue), where advected populations of the
Atlantic copepod Calanus nmarchicus cause a biomass maximum

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

105

Fig. 11. Maps showing (a) zooplankton biomass (g DW m2) (modied from Kosobokova and Hirche, 2009; Kosobokova and Hopcroft, 2010) and (b) infauna (macrofauna)
biomass (g C m2) (modied from Bluhm et al., 2011b) for the two Arctic basins. Biomass is concentrated along the shelf breaks and in inow areas. Only records deeper than
500 m are shown.

Fig. 12. Vertical distribution of zooplankton biomass (mg m3) in (a) the Eurasian and (b) Amerasian basins. Blue are stations close to pronounced AW or PW inuence, red
indicates stations north of 85N, green shows the stations distributed elsewhere. Data from Kosobokova and Hirche (2009), Kosobokova and Hopcroft (2010), and Kosobokova
(2012).

at mesopelagic depths conned to the core of Atlantic inow.


Similarly, locations in the AB close to the Pacic inow (Fig. 12b,
blue proles) also show higher biomass at all depths.
Two major components contribute to the zooplankton biomass
and production of the AO under present conditions, an autochthonous and an allochthonous community, and their comparison in
terms of species composition, community structure and life cycle
traits helps explain regional biomass distribution patterns
(Kosobokova and Hirche, 2009, see also Pomerleau et al., 2014
for comparative community structure). The former consists of
locally reproducing species that maintain resident populations in

the AO including the basin-wide distributed key species, the large


arctic calanoid copepods, Calanus hyperboreus and C. glacialis,
Metridia longa and the chaetognath Eukrohnia hamata. Their local
production forms the biomass base in all central basins and over
their slopes. The allochthonous community, in contrast, consists
of plankton populations advected from the contiguous domains
(cf. Wassmann et al., 2015), but are not able to reproduce in arctic
waters and becoming sterile expatriates there (Kosobokova and
Hirche, 2009; Kosobokova, 2012; Wassmann et al., 2015). The most
important contributors to allochthonous stock, the expatriate
copepods C. nmarchicus advected with Atlantic inow, hardly

106

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

penetrate into the AB due to their short, one year long life spans.
The AB receives hardly any allochthonous addition to locally
produced biomass, because the majority of plankton advected from
the North Pacic already dies off on the extended shelf of the
Chukchi Sea. Consequently, oceanic communities of the AB are
characterized by signicantly lower biomass compared to the EB
(Fig. 10).
Although there are pronounced seasonal variations of zooplankton abundance and biomass in the surface water layer of the Arctic
basins (Kosobokova, 1982, 2012; Ashjian et al., 2003), the vertical
zooplankton distribution pattern shown in Fig. 12 undergoes little
changes during the annual cycle. Year-round observation in the AB
revealed that abundance and biomass per unit volume at depths do
not increase substantially below 100 m even when key Arctic zooplankton species descend to overwintering depths, because they
are distributed over a wide depth range of several hundred meters
(Geinrikh et al., 1983; Kosobokova, 1982, 1983, 2012; Ashjian et al.,
2003; Darnis and Fortier, 2014). Density maxima remain within
the upper 0100 m layer throughout the year (Kosobokova, 1982,
2012; Ashjian et al., 2003).
Unique to the Arctic basins, the lower margin of the epipelagic
zone seems to be shallower than in the rest of World oceans where
it occupies the upper 200 m of the water column (Vinogradov,
1970). An abrupt drop in zooplankton biomass below 100 m
(Fig. 12) (Kosobokova, 2012), a noticeable upward shift of vertical
ranges of common mesopelagic species compared to elsewhere
(polar emergence, Kosobokova, 1989, 2012; Kosobokova and
Hopcroft, 2010), and a boundary between statistically distinct epipelagic and upper mesopelagic zooplankton communities at
100 m (Kosobokova et al., 2011) suggest that the epipelagic rises
up to depths of about 100 m in the Arctic. This phenomenon is
probably related to a combination of limited light penetration into
the water column due to low sun angle, presence of sea ice with
snow cover, strong stratication of the upper water column, and
overall low primary production concentrating fresh food in a thin
surface water layer (Zenkevitch, 1963; Harding, 1966;
Kosobokova, 1989, 2012).
The basin zooplankton fauna uses several mechanisms to cope
with the food limitation. In the surface layers, ice-associated and
upper water column pelagic crustaceans are closely tied to fresh
algal production as reected in their relatively low d15N ratios,
indicators of low trophic level (Iken et al., 2005). The deeper plankters use different food sources, because a considerable part of the
fresh organic matter is already converted into fecal material and
marine snow by ice-associated and surface-dwelling planktonic
herbivores and recycled the microbial loop within the epipelagic
zone before it sinks to depth (Olli et al., 2006). Consequently, carnivory and omnivory/detritivory are prominent feeding modes in
meso- and bathypelagic Arctic zooplankton communities
(Harding, 1974; Laakmann et al., 2009; Kosobokova et al., 2002,
2011). A number of these deep planktonic carnivores and detritivores have modied gut passages as an adaptation for more efcient digestion and assimilation of nutritionally poor and
digestively resistant food. Sigma-shaped (e.g. in the copepod
Aetideopsis rostrata) or substantially widened guts (e.g. in the copepods Scaphocalanus acrocephalus and S. polaris) permit longer
retention of food inside the gut (Kosobokova, unpublished). One
of the most striking examples is the copepod Spinocalanus antarcticus which possesses a strongly elongated and looped midgut,
enabling it to digest marine snow and organic coating of
tiny-sized mineral particles melting out of the ice and sinking
through the water column (Kosobokova et al., 2002). Vertical niche
partitioning, whereby closely-related zooplankton species occupy
different depth ranges to reduce resource competition, is another
adaptation used by the AO deep-water plankters in their

resource-restricted environment (Kosobokova et al.,


Laakmann et al., 2009; Kosobokova and Hopcroft, 2010).

2007;

5.2. Benthos
As for zooplankton, abundance and biomass also decreases with
water depth at the Arctic (and global) deep-sea seaoor (Soltwedel,
2000; MacDonald et al., 2010a,b). This decrease tends to be greater
for benthic macrofauna (infauna) than for meiofauna (Wei et al.,
2010). This trend has been interpreted as an average decrease in
metazoan size with increasing water depth (Klages et al., 2004;
Rex and Etter, 2010) and as increased importance of smaller organisms with increasing water depth (Pfannkuche and Soltwedel,
1998; Rex et al., 2006). Generally, the range of macrofaunal densities and biomass (mostly below 4000 ind m2 and <1 g C m2,
respectively, summarized by Klages et al., 2004 and Bluhm et al.,
2011a; Fig. 11) fall within the lower end of values reported from
the North Atlantic (Levin and Gooday, 2003). Meiofaunal densities
in the central Arctic under ice-cover are either lower than in the
global deep sea (Schewe, 2001) or on the same order of magnitude
as in other oligotrophic deep areas (<100 to >3000 ind 10 cm2;
Soltwedel, 2000; Vanreusel et al., 2000; Hoste et al., 2007;
Grska et al., 2014). Despite the general trend of decreasing density
and biomass with depth, however, faunal densities and biomass
may vary substantially in areas of similar depths, depending on
vertical in situ and advective carbon ux to the seaoor that is
related to ice-edge effects and vicinity to inow areas in the AO
(Schewe and Soltwedel, 2003; Bluhm et al., 2005, see also Section 6.3). Klages et al. (2004) concluded cautiously, based on the
relatively sparse biomass and oxygen consumption rates available
for the Arctic deep-sea benthos, that consumption rates are generally both low, though regionally variable with lowest rates in the
basins, and in agreement with those from other deep-sea areas,
but that vertical organic matter supply may be higher than
expected from calculations of vertical carbon ux from sediment
traps.
The food web at the sea oor of both Arctic basins is still poorly
described. We do know, however, that it is characterized by a high
degree of reworking of organic material resulting in four to ve
trophic levels (excluding marine mammals; Iken et al., 2005;
Bergmann et al., 2009; van Oevelen et al., 2011). Benthic procaryotes and, in the larger size classes, deposit feeders play a major role
in carbon recycling as reected in the dominance of procaryotes in
a carbon ow model in the EB (van Oevelen et al., 2011) and in
enriched d15N signatures of macrofaunal biomass-dominant
deposit feeders and those of their predators and scavengers in
the AB (Iken et al., 2005). Typical for the deep sea, suspension feeders are less common in the abyss of the AB (Bluhm et al., 2005;
MacDonald et al., 2010a,b) because of extremely small currents
(Timmermans et al., 2010), although this feeding type is more
prevalent in the more rapidly moving deep waters in Fram Strait
(Bergmann et al., 2009, 2011). Interestingly (and somewhat contrary to the above concept that smaller organisms become more
important with depth), some larger members of the benthic community in both Arctic Basins are apparently capable of quickly and
efciently ingesting and utilizing fresh material when it does
become available, which has been long been documented from
other deep-sea areas (e.g. Billett et al., 2001). The rapid respond
to food pulses kilometers under the Arctic sea surface was recently
conrmed from observations of holothurian (sea cucumber) consumers of fresh ice algal matter at the seaoor (Boetius et al.,
2013) and low d15N ratios (indicating low trophic level) in some
additional benthic taxa (Iken et al., 2005, Iken and Bluhm unpublished). In distant locations in both Arctic basins, the same
surface-feeding mobile holothurians, Kolga hyalina, congregated

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

in comparatively high densities (maximum >30 ind m2) to consume


fresh organic matter from the sediment surface (MacDonald et al.,
2010a,b; Boetius et al., 2013) as known for their relatives in the
deep North Atlantic (Billett and Hansen, 1982). In contrast, the
nutrition of nematodes, the density-dominant metazoan meiofauna in Arctic deep-sea sediments, is still unclear, although microbial production are a suspected food source based on correlation of
microbial activity with meiofauna densities (Hoste et al., 2007), but
a lack of uptake of microbial isotopic markers by nematodes could
not conrm the trophic relationship (Guilini et al., 2010).
Fresh algae, Melosira arctica, in K. hyalina guts were apparently
abundant and nutritious enough to spur reproduction shortly after
the food pulse in the EB (Boetius et al., 2013). These recent observations suggest that the paradigm change from the 1980s1990s
that many deep-sea taxa in fact reproduce seasonally (triggered
by food pulses) rather than continuously (due to a perceived lack
of a zeitgeber at the time; e.g. Witte, 1996) also applies to the Arctic basins. This also holds true for the omnivorous and carnivorous
deep-sea zooplankters in the AB and EB. Although able to reproduce at low rates all year round, they show clear peaks of
egg-laying activity conned to the Arctic summer followed by
maximum abundance of offspring (Kosobokova, 2012).

6. Who is where and why? Faunal distribution in light of history


and physiography
6.1. Pacic, Atlantic or Arctic? Origin, biodiversity and biogeography
of the Arctic Basins biota
The Arctic basins patterns in diversity and biogeography are
closely tied to the geological history and semi-isolation of the area
as well as to the connections with the Pacic and Atlantic Oceans
(Golikov and Scarlato, 1989). Originally an embayment of the
North Pacic, the Arctic deep sea was inuenced by Pacic fauna
until 80 million years ago when the deep-water connection
closed (Marincovich et al., 1990). Exchange with the deep Atlantic
began 40 million years ago, coinciding with a strong cooling period (Savin et al., 1975).
Three different mechanisms are conceivable for the colonization of the Arctic basins by benthic taxa (Mironov et al., 2013).
These authors summarized that some eurybathic benthic shelf taxa
in the Atlantic Arctic supposedly expanded their range into the
Arctic deep sea during Pleiostocene glaciations, while other shelf
faunas were eradicated. Those taxa that migrated to depths are
by some authors considered the ancestral fauna of some of todays
Arctic deep-sea fauna (Gurjanova, 1985; Nesis, 1984). Species
inventories of Arctic shelves (Piepenburg et al., 2011) and basins
(Bluhm et al., 2011a) showed that indeed 60% of the macroand megabenthic deep-sea species (>500 m) are shared with the
Arctic shelves. Mironov et al. (2013) argue that the second and
third lines of argument are that the Arctic deep-sea fauna is dominated by and originates from fauna common in other areas of the
worlds deep sea, either the North Atlantic or the North Pacic.
Multivariate analysis of Eurasian, Amerasian and GreenlandIceland-Norwegian Seas polychaete species distributions supported
this second line of argument in that it revealed a strong (recent)
Atlantic inuence and the absence of modern Pacic fauna
(Bluhm et al., 2011a). Also, Mironov et al.s (2013) synthesis documented that of over 100 species from 92 genera in 8 classes other
than polychaetes, over half of all genera were deep-sea specialists,
with closer ties to the Norwegian and Greenland Seas (albeit in
part as a transition area) than the North Pacic on the species level.
Those authors caution, however, that all possible source regions of
the Arctic Basins fauna are of multiple origins themselves, complicating the disentanglement of biogeographic origins.

107

Similarly, the deep-sea zooplankton fauna is also believed to


originate primarily from other areas of the deep Worlds oceans.
Almost half of all resident zooplankton species of the AO have wide
distribution ranges with the majority of those being cosmopolitan
(25%), bipolar (9%), or species in common with the North Atlantic
and the North Pacic (10%) (Kosobokova et al., 2011). The modern
North Atlantic fauna contributes 25%, although it was long believed
that at least half of the zooplankton species in the Arctic Ocean
were of North Atlantic origin due to free deep water exchange
through Fram Strait (Brodsky, 1956; Brodsky and Pavshtiks,
1977; Grainger, 1989). Taxa in common with the North Pacic
are very few (2%) and their inuence on the modern metazoan
planktonic fauna of the AO can be largely neglected (Kosobokova,
2012). As with the benthic fauna, faunal exchange between the
North Atlantic and the AO through Fram Strait has taken place
for the last 18 million years, whereas the shallow Bering Strait
and Chukchi Sea prevented penetration of mid- and deep-water
non-migrating plankton species from the North Pacic into the
Arctic for the past 80 million years. In contrast to the benthos,
however, hardly any shelf planktonic taxa expand their range into
the Arctic deep sea, or show submergence into the abyssal. The
underlying reasons are the large differences in ecological requirements of epipelagic neritic (shelf) inhabitants relative to the conditions available at great depths. Therefore, only two of the three
hypotheses of colonization of the Arctic basins by deep-water
fauna suggested by Mironov et al. (2013) could explain the origin
of the Arctic deep-sea zooplankton.
A recent faunal species inventory of Arctic Basins benthos
yielded more than 1100 taxa recorded deeper than 500 m and
north of 80N in Fram Strait (Bluhm et al., 2011a). Relative to an
earlier inventory (Sirenko, 2001), over 400 new taxa were added,
primarily in the speciose groups of round worms (Nematoda), bristle worms (Annelida) and crustaceans. Taxa with very few species
were also species poor globally. As elsewhere in the global deep
sea, the Arctic Basins have a high fraction of rare and endemic species, although the estimates (5080% endemism rate per
Vinogradova, 1997 and Mironov et al., 2013) may be articially
high because of the relatively low sampling effort. Several dozen
new species have been discovered in the Arctic Basins in recent
years (e.g. Rogacheva, 2007; Gagaev, 2008) although species discovery rates are by far lower than those in the Antarctic deep ocean
(Brandt et al., 2007), because of lower sampling effort in the Arctic,
and high faunistic connectivity to the North Atlantic. Dozens of
new distribution locations, however, are recorded by any
deep-sea expedition, and are thought to reect the poor knowledge
of the fauna rather than actual recent species range extensions (e.g.
MacDonald et al., 2010a; Sharma and Bluhm, 2011).
The largest recent metazoan plankton fauna inventory is based
on collections from 134 locations where sampling was conducted
from the surface to the bottom or 3000 m depth during recent icebreaker expeditions and older Russian drifting ice stations
(Kosobokova et al., 2011). The inventory of 174 species from eight
taxa (Cnidaria, Ctenophora, Mollusca, Annelida, Nemertea, Crustacea, Chaetognatha, and Larvacea) is now assumed to be nearly
complete with the exception of the deepest water layers, where
both unrecorded and new species continue to be found
(Kosobokova et al., 2011). The number of mesozooplankton species
in the basins is about half of that in the entire AO including its
shelves (the latter ca. 370 species). The epipelagic basin fauna is
low in diversity (Kosobokova et al., 2011) and consists of a few
ubiquitous forms and seasonally migrating species endemic to arctic waters, such as the copepods Calanus hyperboreus, C. glacialis,
and Metridia longa. Of the 174 species recorded in the Basins, 38
species are recent additions to earlier lists (Sirenko et al., 1996;
Kosobokova et al., 1998; Sirenko, 2001), and 18 of those were
recently described (Markhaseva, 1998, 2002; Markhaseva and

108

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Fig. 13. Map showing the distribution of some (a) pelagic and (b) benthic taxa occurring in the Eurasian and Amerasian basins. Taxa are identied in the legend. The
distributions suggest that the Lomonosov Ridge does not serve as a distribution barrier. Note that for many other species, even fewer distribution records exist for the Arctic
basins preventing a general conclusion on distribution patterns. Data from Bluhm et al. (2011b) and Kosobokova et al. (2011).

Kosobokova, 1998, 2001; Raskoff et al., 2010; Andronov and


Kosobokova, 2011) or are currently under description. Crustaceans
strongly dominate zooplankton species number (70%), and copepods are the most diverse group among them. Only 14% of all zooplankton species in the basins are endemic to the Arctic
(Kosobokova, 2012) with one-third of them ice-associated and
the remaining two-thirds bathypelagic with a clear preference
for depths below 1000 m (Kosobokova et al., 2011). The evolution
of these deep-water endemics is without doubt related to the
geological history of the AO (Kosobokova et al., 2011).
6.2. The role of the ridges steering currents, but no barrier for
pan-Arctic Basin species dispersal?
The prominent ridge systems (Section 2) do not act as strict
zoogeographical barriers for zooplankton or benthic invertebrate
fauna between the EB and AB (Koltun, 1964; Deubel, 2000;
Bluhm et al., 2011a; Kosobokova et al., 2011). The same
deep-water and endemic zooplankton species are found on both
sides of the Lomonosov Ridge (Fig. 13a) and the vertical structure
of epipelagic, mesopelagic and bathypelagic communities are consistent between basins (Kosobokova et al., 2011). Some deep water
benthic species and similar communities are also found on both
sides of the Lomonosov Ridge (Bluhm et al., 2011a,b; Mironov
et al., 2013; Fig. 13b), although the high rate of endemic and rare
species at the Arctic Basins seaoor limits generalization of this
nding (MacDonald et al., 2010a; Mironov et al., 2013). The current
patterns suggest regular and recent exchange of at least some of
the deep fauna across the ridge, and nd some support in recent
insights of the circulation patterns in the basins. For example,
deep-water ow was recently documented between the Makarov
and Amundsen basins through the deepest sill (1870 m) in the
Lomonosov Ridge allowing water and biotic exchange between
AB and EB deep waters (Bjrk et al., 2007, 2010). The distribution
of the bathypelagic zooplankton assemblage below roughly
10001500 m also suggests physical and biological exchanges
across this ridge sill are possible (Kosobokova et al., 2011),
although exact boundaries in zooplankton community structure
cannot be drawn because of limited vertical resolution in the sampling technique. If arctic deep-water zooplankton taxa undergo

large vertical migrations, this could also serve as a mechanism


for exchange between basins if horizontal exchange of fauna across
the ridge is or recently was occuring. As noted above, free exchange
from the EB to the AB is likely down to the sill depth of 2400 m
(Timmermans et al., 2005) and from the Makarov to the Canada
Basin through Cooperation Gap in the Alpaha/Mendeleyev Ridge
with sill depth of 2400 m (Carmack et al., 2012).
It is unclear, however, exactly how benthic fauna disperse
across the Arctic deep sea and over the ridges. Dispersal distance
estimates are sparse for any region, but are thought to be a function of larval period, current velocities, habitat, larval biology and
adult genetic makeup (Grantham et al., 2003; Palumbi, 2003 and
references therein). On the one hand, conditions in the deep Arctic
basins facilitate long-distance dispersal, because larval life spans
tend to be longer at low versus high temperatures (Fetzer and
Arndt, 2008) and in soft bottom versus other habitat types
(>30 d; Grantham et al., 2003). In addition, predation pressure in
the deep water column is likely low given the low zooplankton
and nekton densities in deep layers (Raskoff et al., 2010;
Kosobokova et al., 2011). On the other hand, dispersal is probably
constrained by low horizontal ow rates in near-bottom Arctic
deep water (<1 cm s1; see Timmermans et al., 2010, but also obvious from high densities of animal tracks despite low faunal densities, Macdonald et al. (2010a)), and pelagic life stages are generally
thought to be less common in high latitudes than elsewhere
(Thorson, 1950; although challenged by Fetzer and Arndt, 2008
and others). We speculate that larval dispersal in the Arctic basins
might be aided by upward convective transport by geothermal
heating that has been estimated at 1 mm/sec (Carmack et al.,
2012), followed by lateral transport over ridge sills within the
overlying water (at 1 cm s1; Timmermans et al., 2010) for horizontal dispersal in ow rates higher than in waters below sill
depth, especially near steep topography. Based on these assumptions, pelagic larvae could rise to the overlying layer in a few days,
disperse laterally across the Lomonosov Ridge if they started in the
Amundsen Basin, for example, and settle in the Makarov Basin
within a month. Taxa without pelagic larvae such as amphipods
and isopods would be expected to disperse much more slowly,
and perhaps have restricted distributions as found for, e.g. Antarctic isopods and ostractods (Brandt et al., 2007), although we found

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

no clear evidence for the latter in the literature. A comprehensive


compilation of dispersal stages and distances of Arctic fauna has
yet to be done.
In the upper water column, halocline, and Atlantic layers, Atlantic and Pacic expatriate zooplankton taxa are advected with their
respective waters either into the EB or AB (Kosobokova and Hirche,
2009; Kosobokova and Hopcroft, 2010; Kosobokova, 2012; Nelson
et al., 2009, 2014; Wassmann et al., 2015). In addition, inhabitants
of the shelves may be advected offshore as neritic expatriates
(Kosobokova, 2012; Wassmann et al., 2015). Although abundances
of expatriates in the AB are much lower than in the EB, they do
result in differences in the species composition between basins
in surface waters and midwater layers. Recent Pacic benthic
species, in contrast, are virtually limited to the shelves, in particular the Chukchi and Beaufort Sea shelves (Bilyard and Carey, 1980).
6.3. Two-way trafc: shelf to basin and basin to shelf interactions
The Arctic Basins are intricately inuenced by their surrounding
shelves, and the shelves are inuenced by the deep central Arctic,
especially along the margins. Above, we have already discussed the
role of the shallow PW inow for the water mass structure and
stratication and the role of AW inow for biogeography, biodiversity and faunal biomass patterns. In this section, we focus on physically driven shelf-basin interactions and their biological
consequences.
6.3.1. Shelf to basin
Here, we summarize key shelf-basin interactions, recognizing
that other mechanisms are concurrently operating. Particularly
important is the role of ice formation and brine drainage over
shelves in the maintenance of the halocline complex in the basins
(Aagaard et al., 1981; Melling and Lewis, 1982). While the subarctic North Pacic and North Atlantic are sources of halocline waters,
the physical and biogeochemical composition of the water masses
from these areas undergo substantial changes during transit of the
inow shelves before entering the basins (Jones and Anderson,
1986). These authors proposed that brine production during sea
ice formation creates saline bottom waters which inhibit vertical
mixing with overlying fresh riverine water entrainment of
re-mineralized nutrients into the photic layer. The constituent
nutrients are then advected horizontally into the halocline layer
of the deep basins (c.f. Popova et al., 2013). Similarly, Nitishinsky
et al. (2007) and Dmitrenko et al. (2011) have documented
enhanced near-bottom nutrient concentrations in the Laptev Sea.
The inuence of this continental shelf pump on halocline waters
has been further investigated by Anderson et al. (2010) who
showed that the Chukchi and East Siberian seas supply an excess
of carbon and nutrients to the Canada Basin; and by Alkire et al.
(2010) who used the tracer NO (NO = 9  [NO
3 ] + [O2]) to demonstrate a Siberian shelf inuence in the Makarov Basin.
Dense water drainage from shelves occurs preferentially
through cross-shelf troughs such as Barrow Canyon and the St.
Anna Trough (Garrison and Becker, 1976; Mountain et al., 1976;
Dmitrenko et al., 2011). The general trend of decreasing faunal biomass with increasing water depth from the shelf toward the basin
is interrupted in such areas, because carbon sources are concentrated and channeled from the productive shelves into the less productive basins. The western Beaufort Sea slope downstream of PW
drainage through the Barrow Canyon is one such location. Biomass
of Arctic cod, snow crab and certain epifaunal invertebrates are
higher along the shelf break of the Western Beaufort Sea than on
the adjacent shelf (Crawford et al., 2011; Logerwell et al., 2011;
Parker-Stetter et al., 2011; Ravelo et al., 2015). Belugas , whose
prey include Arctic cod (Frost and Lowry, 1984; Loseto et al.,
2009; Quakenbush et al., 2015), tend to use this area on the upper

109

continental slope of the Beaufort Sea that coincides with the comparatively warm Atlantic layer (Suydam et al., 2005; Citta et al.,
2013). The occurrence of belugas along the western Beaufort slope
is also correlated with a well-developed Alaska Coastal Current
producing strong frontal features (Stafford et al., 2013) and these
authors suggest this mechanism would provide enhanced foraging
opportunities.
Mesoscale eddy formation provides another mechanism for off
shelf transport of material properties and biota. Baroclinic eddies
have long been recognized to populate the halocline of basin interiors (Hunkins, 1974; Newton et al., 1974); an early census suggests that eddies comprise up to one quarter of the surface of the
southern Canada Basin (Manley and Hunkins, 1985). More
recently, Zhao et al. (2014) examined ice-tethered proler data
deployed between 2004 and 2013 to carry out a census of mesoscale eddies within the AO halocline in the basins. They documented 127 eddies, 95% of which were anticyclonic, the majority
of which had anomalously cold cores and were observed in the
Beaufort Gyre (AB eddies) and the Transpolar Drift (EB eddies).
Such eddies typically have horizontal length scales on the order
of 1020 km and horizontal (azimuthal) velocities of order 10
25 cm s1 (Timmermans et al., 2010) and their formation has been
attributed to instabilities related to current/topography interaction
(DAsaro, 1988), frontal zone processes (Timmermans et al., 2008)
and shelf break jets (Pickart et al., 2005, 2013; Spall et al., 2008).
Offshore transport of shelf-origin waters by eddies of Pacic origin
in the upper halocline are associated with elevated concentrations
of nutrients, organic carbon, and suspended particles (Mathis et al.,
2007; Watanabe et al., 2012; Pickart et al., 2013). These authors
attributed these features to derive from the boundary current
along the edge of the Chukchi Shelf. OBrien et al. (2011) examined
sediment trap data from the slope off the Canadian Beaufort Shelf
and argued that eddy phenomena also played a major role in transporting sediments from shelf to basin. For the Canada Basin, eddies
in three different depth domains were recognized: (1) shallow
upper halocline eddies centered at 80 m), (2) lower halocline
eddies (200 m) and (3) deep eddies (1200 m) (Carpenter and
Timmermans, 2012). Eddies observed in the deep waters at depths
between 200 and 2000 m of both the AB (Swift et al., 1997;
Carpenter and Timmermans, 2012) and EB (Schauer et al., 2002;
Aagaard et al., 2008) had water mass properties derived from adjacent shelves. Velocities within deep eddy cores were found to
range from 225 cm s1, and thus appear to have the potential to
transport material properties within and below the Atlantic layer
and re-suspend particles where Atlantic water abuts the continental slope (Carpenter and Timmermans, 2012).
Sea ice, through its drift patterns, also transports large amounts
of particles from rivers and the shelves into the central AO (Eicken
et al., 2005). This transport mechanism is in fact thought to have
critically shaped the sedimentation regime in the AO in the geological record (e.g. Nrgaard-Pedersen et al., 1998). Particle types are
comprised not only of sediments and terrigenous carbon that are in
part region-specic in terms of their composition and can, therefore, be used as tracers (Dethleff et al., 2000), but also of fresh algal
material that provides an allochthonous food source for basin
fauna (Boetius et al., 2013).
6.3.2. Basin to shelf
Wind forcing is the principal mechanism driving basin waters
onto the shelf, either via upwelling favorable winds (generally
easterlies) bringing water from depth onto the shelf, or via downwelling favorable winds (generally westerlies) driving surface
waters onto the shelf. In the arctic wind forcing may take place
in regions that are ice covered, partially ice covered or ice free
(see Wadams, 2000, for discussion). Clearly, however, as sea ice
continues to retreat, more and more of the shelf break domain is

110

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

exposed annually to wind-forced, shelf-break upwelling (Carmack


and Chapman, 2003). In the Canada Basin this trend has the potential to increase the volume of nutrient-rich Pacic Water onto the
shelf, and potentially increase primary productivity (Tremblay
et al., 2011; Spall et al., 2014). Williams et al. (2006, 2008) examined the joint roles of wind-forcing and bathymetry (e.g. canyons,
isobaths divergence) on upwelling and cross shelf transport.
The efciency of upwelling in driving water exchange across the
shelf-slope boundary is strongly affected by topography and the
presence of canyons, headlands and convergent isobaths. For
example, Williams et al. (2006) examine shelf-break exchange in
Mackenzie Canyon, a cross shelf canyon in the Beaufort Sea shelf
forced by wind- and ice-driven ocean surface-stresses. They note
that while the canyon is approximately 400 m deep and 60 km
wide, it is only 23 times the baroclinic Rossby radius at its mouth,
and thus behaves as a dynamically narrow canyon. Upwelling
events are associated with wind in the ice-free summer season
and with ice motion in winter (see also Pickart et al., 2013), but
ice motion does not necessarily correspond to wind-stress because
internal ice stresses differentially block downwelling-causing ice
motion (e.g. it is easier to move ice offshore than onshore). This
asymmetry between upwelling and downwelling ow, combined
with the regional predominance of upwelling-causing ice motion,
show that Mackenzie Trough is an effective conduit for transporting deeper, nutrient-rich water onto the shelf (Williams et al.,
2006). In the Mackenzie Trough Carmack and Kulikov (1998)
observed vertical isopleth displacements associated with Atlantic
Water exceeding 400 m and, upon cessation of wind forcing, collapsing and then relaxing back into the basin to generate a Kelvin
wave and its associated velocity eld propagating eastward along
the slope.
In addition to the large canyons such as Mackenzie, St. Anna and
Barrow canyons, there are hundreds of shallow submarine valleys
that are relict channels of lower Pleistocene sea levels during the
ice ages. Many such channels are shallow and do not extend as
far as the shelf edge, so that their role in cross-shelf exchange in
response to upwelling favorable winds is unclear and likely different from case to case. For example, Williams et al. (2008) examined
ow in Kugmallit Valley on the Canadian Beaufort Shelf and indeed
found cross shelf ows associated with upwelling and downwelling wind forcing. However, because the valley ends out before
reaching the shelf break, upwelling found within the valley must
transport water derived from across the shelf rather than across
the shelf break.
Upwelling and cross-shelf exchange can also be greatly
enhanced by isobath convergence/divergence. For example, hydrographic and remote sensing data from shelf waters north of Cape
Bathurst show consistent evidence of upwelling events (Williams
and Carmack, 2008). These authors propose that this enhanced
upwelling is forced by the vorticity adjustment of the along-shelf
ow to the isobath divergence at the cape. Benthic samples near
the cape show enhanced numbers and diversity of organisms
which suggest that nutrients brought to the surface by upwelling
allow enhanced primary production in the region that ultimately
feeds the benthos (Conlan et al., 2013).
The biological consequences of accelerated seasonal ice retreat
and enhanced upwelling of basin halocline waters onto adjacent
shelves are dramatic because increases in new production require
an upward shift in the delivery of nitrate to the euphotic zone
(Tremblay and Gagnon, 2009). To demonstrate this, Tremblay
et al. (2011) quantied the effects of shelf-break upwelling in the
Canadian Beaufort Sea where repeated episodes of ice retreat and
upwelling favourable winds resulted in 2 to 6-fold increases in primary production. They also explain that interpretations of change
in arctic productivity must distinguish between new production,
which yields a net increase in plant biomass due to increases in

nitrogen delivery, and regenerated production, which is supported


by the recycling of biological products (see also Codispoti et al.,
2013).
Upwelling and downwelling favorable winds will also affect the
along-slope and off-slope transport of the ACBC. Lien et al. (2013)
show events in which the relative strengths of the FSB and BSB are
affected by wind-forced Ekman transport in the northern Barents
Sea, such that the decrease in sea surface height resulting from
upwelling winds induces a cyclonic circulation anomaly along
the slope which, in turn, weakens the FSB owing along the basin
slope. If these events got more common we speculate that we may
observe a general, pan-arctic wide pattern involving weakening of
property transport by the ACBC, with concurrent decreased delivery of allochthonous material to far-eld portions of the basins,
counter-posed by increased production of autochthonous material
by shelf-break upwelling, a hypothesis which would need rigorous
testing.

7. Of halocline structure, primary production and vertical ux


change underway?!
No matter what is used as a clock, rapid change has become the
hallmark of modern AO research, and this, in turn, is manifest at
signicant and differing time scales. As noted before, throughout
90% of the Pleistocene the AO exists in glacial mode, with narrow
continental shelves, greatly restricted river inow, thicker and perhaps immobile sea ice, and total blockage of exchange with the
Pacic Ocean (Marincovich et al., 1990). The abrupt transition to
the present day Holocene or interglacial mode conditions
10 K years ago marks an almost complete shift in variables and
drivers governing the Arctic system. During the Holocene, on
shorter time scales of 1000100 years, signicant changes in high
latitude climate are now becoming apparent in temperature (e.g.
the medieval warm period and little ice age) and perhaps moisture
delivery patterns (e.g., Overpeck et al., 1997). The pace of change
over the past three decades has been even more impressive, with
summer ice extent in 2012 being 54% less than the 1980 benchmark and thickness diminishing by half (Cavalieri and Parkinson,
2012; Barber et al., 2015 and references therein). Associated with
the rapid retreat and thinning of sea ice the water column has
warmed at depths exceeding 800 m owing to warmer waters
entering from the Atlantic and Pacic Oceans (Polyakov et al.,
2013; Table 2).
The truly unique aspect of such change for the Arctic basins is
that now, for the rst time in the observational record, sea ice is
retreating on an annual basis past the shelf break, and thus exposing basin waters directly to sunlight and wind forcing. Our rough
calculation suggests that the summer 2012 sea ice retreat left
approximately 40% of the basin area exposed. One consequence
is that upwelling favorable winds can now directly and efciently
drive shelf-break upwelling, and draw nutrients from subsurface
basin waters onto the shelf (Section 6.3) (Carmack and Chapman,
2003). However, the same upwelling favorable winds will also create onshore pressure gradients over the slope and basin, which will
act to disrupt the ow of waters in the ACBC (see Section 6.3).
Within the basin interior remaining ice is thinner and less compact, and thus more responsive to wind stress (forcing and mixing)
(Kwok et al., 2013).
Within the EB, and later in the AB, the rst observed changes in
water column temperature were advection related (Quadfasel
et al., 1991; Carmack et al., 1995). Since then both observational
evidence and modeling results suggest that temperature changes
in AW are related to pulses in the AW inows (Polyakov et al.,
2005; Karcher et al., 2011). The rst signal of far-eld effects on
Canada Basin waters, as delivered by the ACBC, was the increased

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

111

Table 2
Observed and modeled changes in the Amerasian and Eurasian Basins in the past few decades.
Layer

Amerasian basin (mostly Pacic inuenced part of Canada basin)

Sea ice

Remaining MYI limited to north of Canadian Archipelago


(Stroeve et al., 2012; Barber et al., 2015)

Surface mixed layer

Decreasing salinity, increased stratication, deepening nutricline (McLaughlin and Carmack, 2010; Jackson et al., 2010)
Observed and modeled high variability in freshwater distributions and pathways (Newton et al., 2008, Polyakov et al. 2008,
Morison et al., 2012)
Very low nutrient concentration in Beaufort Gyre suggests no/
little increase in primary production is possible (cf. Codispoti
et al., 2013)
Small phytoplankton thrive (Li et al., 2009)
Decrease for small zooplankton dependent on authochtonous
production, increase/no change in large zooplankton beneting from allochtonous production (Hunt et al., 2014)?
Aragonite-saturation decreasing (Yamamoto-Kawai et al.,
2009)
Enhanced shelf-break upwelling (Carmack and Chapman,
2003; Tremblay et al., 2011; Pickart et al., 2013)
Interannual shifts in location of the Atlantic-Pacic halocline
front (McLaughlin et al., 1996; Karcher et al., 2012)
Warm pulse penetration into the basin (Shimada et al., 2004)
and temperature increase between 0 and <1 C (McLaughlin
et al., 2009; Polyakov et al., 2013)
More intense anticyclonic Beaufort Gyre after 2004 constraining the cyclonic ACBC that transports AW (Proshutinsky et al.,
2009; Karcher et al., 2012)
Geothermal warming (0.004 C per decade; Carmack et al.,
2012), no time series to document biological change

Halo-cline

Atlantic water

Arctic Basins Deep water

ventilation in the 1990s of Atlantic-origin BSB waters as shown by


freshening and increased levels of chlorouorocarbon and oxygen
at depths of 6001600 m (McLaughlin et al., 2002). Warmer
Atlantic-origin FSB water, rst observed in the Nansen Basin during
the early 1990s, extended to the Northwind Ridge along the western reaches of the Canada Basin in 2002 (Shimada et al., 2004) and
subsequently spread across most of the southern basin interior by
2007 (McLaughlin et al., 2009). These newer and warmer FSB
waters were also fresher and more ventilated than existing basin
waters. A second pulse was observed to begin in Fram Strait in
1999 (Schauer et al., 2004) and then appeared in the eastern EB
in 2004; Polyakov et al. (2005) used this warming pulse to calculate an advection speed of along-slope warming of 1.5 cm s1.
The warming pulse peaked in the Arctic basin interior in 2007
2008 with maximum temperature anomalies of up to 1 C, and
20082010 observations suggest that the AO interior was in transition toward a cooler state in the 1990s, but not well-correlated
with increases in the temperature of Bering Sea source waters
(Polyakov et al., 2011). Shimada et al. (2006) noted that recent
increases in Pacic summer water temperatures are
well-correlated with the onset of sea-ice reduction that began in
the late 1990s. They suggest that reduced internal ice stress in
autumn allows a more efcient coupling of anti-cyclonic wind
forcing to ice and the upper ocean, which subsequently redirects
Pacic summer water into the central basin at sufciently shallow
depths to affect winter heat loss to the atmosphere and ice.
In concert with increasing focus on tracking sea ice changes in
the Arctic over the past few decades, the role of the AO halocline,
its possible variability over time and its different structure in the
AB and EB have received increasing attention. Increased study
effort is primarily due to the climate-relevant function of the halocline as the insulating layer between the cold polar mixed layer

Eurasian basin
Remaining MYI limited to north of Greenland (Barber et al.,
2015)
Ice algal production from shelves transported to basins, where
fast vertical ux supplied food to benthic fauna (Boetius et al.,
2013)
Variability in freshwater content (Timmermans et al., 2011)
Nutrient concentrations not limiting, perhaps increase in primary production possible (Codispoti et al., 2013)
No change in primary production between 1995 and 2007
(Wassmann et al., 2010)
Increased frequency of fall blooms (ACIA, 2005; Ardyna et al.,
2014)
(modeled) increase in secondary production across both
basins, slightly higher in Eurasian Basin (Slagstad et al.,
2011, their Fig. 9)

Possible warming owing to increased vertical heat ux from


AW (Polyakov et al., 2012)

Temperature increase by 1 C with change in inow volume


(Polyakov et al., 2013)

Warming more than expected from geothermal heating


(Rudels et al., 2013)
Decreased megafauna densities and sediment microbial biomass at HAUSGARTEN for 20022007 (Bergmann et al., 2011)

above and the comparatively warm Atlantic layer underneath


(e.g. Shimada et al., 2005). In this sandwiched position, the halocline controls the extent to which the underlying and warmer
basin waters affect sea ice volume and phenology. A strengthening
in the stratication of the halocline in the Canada and Makarov
Basins was observed through a decade-long series of CTD data
between 1997 and 2008, whereas the halocline was apparently
rather stable in the Amundsen Basin over that same time period
(cf. Bourgain and Gascard, 2011). Temporal changes within the
summer halocline in the Canada Basin are related to development
of a near-surface temperature maximum within the halocline and
shoaling of the summer halocline (Jackson et al., 2010). These
authors argue that heat contained within the shallow
near-surface temperature maximum and released during the
freeze-up season could delay freeze-up or reduce ice thickness.
Light regime and nutrient supply critically inuence the
amount of primary production that can take place in the basins.
How will nutrient concentrations play out in the future Arctic
Basins? In their modeling study, Zhang et al. (2010) predict an
increase in nutrient availability to the generalized upper AO
because of enhanced air-sea momentum transfer due to reduced
sea ice cover. Although these authors do not specify exactly where
or how widespread the nutrient replenishment would occur, nutrient enhancement will likely occur primarily along the shelf break
related to upwelling, rather than in the basins per se (Carmack
and Chapman, 2003; Tremblay et al., 2011; Table 2). Furthermore,
horizontal advection of nutrient rich PW and AW may contribute
the nutrients for up to 20% of the total AO primary production
(Popova et al., 2013). For the basins, model predictions of enhanced
nutrient availability do not match the observed reduced nitrate
concentrations of the Canada Basin time series of the upper ocean
(Li et al., 2009) or recent observations of extremely low nutrient

112

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

concentrations in the same area (Codispoti et al., 2013; Matrai and


Apollonio, 2013).
More numerous than observations and predictions of changing
nutrient concentrations are observations and predictions about the
trends in primary production, although some failed to separate the
Arctic basins from the shelves and some seem contradictory. Several authors agree that primary production in the AO as a whole
will increase or has increased - under current climate change scenarios, but with signicant regional variability and including areas
of no increase or decrease. Exactly where and by how much primary production may change has been estimated based on models,
in situ observations and satellite imagery (e.g. Arrigo et al., 2008;
Pabi et al., 2008; Zhang et al., 2010; Slagstad et al., 2011; Ji et al.,
2013; Vancoppenolle et al., 2013), but is debated in light of the
dependence on and variability in light and nutrient availability
and re-supply, and extent and decline in sea ice cover (Tremblay
et al., 2009; Tremblay and Gagnon, 2009; Lee et al., 2010; Zhang
et al., 2010; Brown and Arrigo, 2012; Codispoti et al., 2013). The
primary reasons brought forward to explain increases in primary
production both on the shelves and in the basins include increasing
duration of the open water period and a growing spatial extent of
open water, resulting in longer and greater exposure to solar radiation with a resulting extended phytoplankton growing season
(e.g. Arrigo et al., 2008; Zhang et al., 2010; Slagstad et al., 2011;
Ardyna et al., 2014). Estimates based on remotely sensed chlorophyll, sea ice cover and sea surface temperatures suggest that half
of the increase in primary production between 1998 and 2006
occurred in pelagic waters (Pabi et al., 2008), but no estimate
shelf versus basins waters was given. Similarly, modeled primary
production in the upper 100 m for the period of 19982007
increased mostly in the permanently and seasonally-covered
areas with decreasing ice thickness and increasing summer melt
back in recent years [versus the open waters of the GIN Seas
and southern Barents Sea] (Zhang et al., 2010), but the proportions
of increase over the basins versus the shelves were, again, not specied. A different model estimated a basin-specic 4-fold increase
of gross primary production assuming an air temperature increase
of up to 8 C (Slagstad et al., 2011). An inter-comparison between
ve different biogeochemical models for the Arctic summarized
that physical and chemical boundary conditions differed substantially between models, mainly in terms of nutrient concentrations
and winter mixing (Popova et al., 2012), and these authors caution
that predictive capabilities of models are reduced as long as these
basic boundary conditions are inconsistent.
Comparisons of modeled phenology of primary and secondary (micro-, meso-, predatory zooplankton) production at
different locations along the SHEBA drift (Ashjian et al., 2003)
showed that spatial and temporal variability in primary production
can be expected to be largest in areas with seasonal ice
cover (Zhang et al., 2010). Given that seasonal ice cover is predicted for the entire basin before the end of the century (Wang
and Overland, 2009), substantial increases in primary production
and changes in phenology might be inferred from the Zhang
et al. (2010) results. Another, basin-specic forecast, however,
actually suggests low or no increase in primary production for
the AB, particularly the Canada Basin, and northern Beaufort Sea
because of particularly strong stratication and very low nutrient
concentrations (Bluhm and Gradinger, 2008; Codispoti et al.,
2013). Although nutrient (nitrate) concentrations do not appear
to be limiting in the EB (Codispoti et al., 2013), modeled primary
production for the period 19952007 did not increase in the European Basin sector (Wassmann et al., 2010). Regardless of the trajectory of the quantity of primary production, increased light
intensity below the ice and in open water will favor a longer vegetative season (Arrigo et al., 2012; Ji et al., 2013; Ardyna et al.,
2014).

The anticipated change in ice algal production has received less


attention than phytoplankton production. In the 1990s, ice algae
contributed a larger proportion to total algal biomass in the Arctic
Basins (up to 50%) than over the shelves (Gosselin et al., 1997). The
high ratio in the Arctic Basins is mainly driven by the very low
pelagic primary production in the oligotrophic Arctic (Section 4.2).
The relatively high pelagic primary production extant on the Arctic
shelves may explain why model-estimated annual Arctic phytoplankton production in the upper ocean from 19982006 was
found to be 20 times higher than that of ice algal production
(Jin et al., 2012), and thus a distinction must be made between
shelf and basin ice algal production.
Algal production associated with sea ice is by many assumed to
decrease in the future based on the overall decrease in sea ice cover
(e.g. Zhang et al., 2010). It is worth asking, however, if the anticipated decline could be counterbalanced, in the near future at least,
by the switch from multi-year to rst-year sea ice in the Arctic
Basins (Barber et al., 2015), which has several effects. First, thinner
ice allows more light to penetrate the ice resulting in higher ice
algal production than in thicker ice, provided nutrients are available for growth (Gradinger, 2009), and snow depth does not
increase. Second, nutrient exchange across the ice-water interface
is more efcient in warmer, saltier and thinner rst-year ice than
in the multi-year ice of the past (Deal et al., 2011). Third, surface
melt ponds and melt holes producing a patchwork of small (below
satellite detection) but productive hot spots have been suggested
to become more common in the central Arctic in the future (Lee
et al., 2011). And last, massive amounts of ice algae presumably
grown on the shelves have recently been observed over the central
basin where they contributed to algal biomass (Boetius et al.,
2013). In the end, the balance of sea ice and ice algal bloom phenology with regional nutrient distribution patterns will determine the
trends in absolute and relative ice algal production (Leu et al.,
2015).
What do the outlined physical and biological changes mean for
the pelagic biota in the Arctic basins, vertical carbon ux, and the
ultimate carbon recipients, the Arctic deep-sea benthos? For zooplankton, one can reasonably anticipate that the production of
autochthonous zooplankton will closely be tied to future primary
production levels in the Arctic Basins, and allochthonous zooplankton biomass will continue to be driven by the inow of Atlantic
and Pacic Waters (Wassmann et al., 2015). Model estimates suggest increasing secondary production (based on Calanus glacialis) in
the basins from an average zero or negative rate up to 1.5 g C m2 y1 under their extreme scenario of 8 C air temperature increase
(Slagstad et al., 2011). Given the uncertainty of the trend and trajectory in primary production for the basins, however, (see disagreement between models on primary forcing factors in Popova
et al., 2012) and the resulting particulate organic matter pools
available for food webs, this trend seems uncertain. Future states
for ice-associated invertebrate fauna are equally uncertain, but
are likely tied to sea ice extent and ice algal production (Bluhm
et al., 2010). Survival strategies for low ice conditions may exists;
Berge et al. (2012) hypothesize that some ice fauna may conduct
vertical migrations to depths where deep northward currents
would take them back into sea ice-covered areas where they
ascend again and re-colonize the underside of drifting pack ice
when present. These authors suggest this mechanism would
reduce loss of ice fauna from the Arctic.
Predictions on vertical ux and resulting benthic standing
stocks currently remain largely speculative. Two opposing scenarios are: Increasing primary production linked to decreasing sea ice
cover, rising surface water temperature and enhanced zooplankton
production (cf Pabi et al., 2008; Zhang et al., 2010; Slagstad et al.,
2011) could result in increasing vertical export of particulate
organic matter to the sea oor. This scenario nds support in a

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

2-year sediment trap deployment over the Laptev Sea margin


where POC export ux increased in response to the extreme summer sea ice melt in 2007 (Lalande et al., 2009). The opposing scenario could be that vertical ux in the basins may be hampered
by increasing vertical stratication (at least in the Canada Basin;
Carmack et al., 2006; Jackson et al., 2010), low and perhaps declining nutrient concentrations in the upper layer of the Canada Basin
(Li et al., 2009; Codispoti et al., 2013), and the co-variance of light
weight, small algae with strong stratication (Li et al., 2009; Lee
et al., 2012). Presumably those small algae would have limited
penetration capability in a strengthening pycnocline compared to
that of heavy, large, chain-forming diatoms as described by
Boetius et al. (2013). Along the same lines, a shift from large diatoms to smaller coccolithoporid cells occurred at the deep-sea
observatory HAUSGARTEN in Fram Strait over the last decade in
warm years, and vertical carbon ux of biogenic particulate silica
was lower in warm than in cold years (Lalande et al., 2013).
Changes in water column processes in some fashion perpetuated
to epibenthic megafauna densities that declined at HAUSGARTEN
between 2002 and 2007 with concurrently declining sediment
organic content, sediment microbial biomass and increasing bottom water temperatures (Bergmann et al., 2011). Temporal trends
on vertical ux or of deep benthic communities in the AB are, to
our knowledge, not available.

8. Outlook
Throughout much of the 20th century the AO and its central
deep basins in particular was viewed as a small, remote, slowly
changing and relatively unimportant part of the global system. A
perceived need to catch-up for lost time now prevails, requiring
that we pay more attention to the pan-Arctic perspective and to
the importance of advection and physical/biological connectivity
and their joint roles for the basins. This paper summarizes, from
an interdisciplinary point of view, our current perception of how
the Arctic Basins are set-up and operating, and how basin oceanography and biology have recently (e.g. the last few decades) changed
(Table 2). The AO is subject to large amplitude multi-decadal variability and long-term trends (Polyakov et al., 2013), thus challenging interpretations of observed changes to climate drivers. These
authors state, however, that the exceptional magnitude of recent
high-latitude changes, both oceanic and atmospheric, implies an
irreversible shift of the AO to a new climate state (see also
Jeffries et al., 2013; Wood et al., 2013). What does this new state
hold for biota? A review of the climate change footprints in arctic
ecosystems by Wassmann et al. (2011) gives evidence that all components of the high-latitude marine ecosystem are being impacted
by global change, though most existing reports considered large
mammals and birds only, and mainly in shelf areas.
What, then, do we know? Concrete ndings from decade-long
time-series span from declines in sea ice extent and thickness
(Kwok et al., 2009; Stroeve et al., 2012, Barber et al., 2015) to
increasing river discharges (McClelland et al., 2006), with consequences that include the appearance of aragonite-undersaturated
(low pH) waters in the Canada Basin this past decade
(Yamamoto-Kawai et al., 2008, 2011; Table 2). Other changes
include warming of both Atlantic (Polyakov et al., 2011) and Pacic
(Shimada et al., 2006) inow waters and increased stratication
resulting from both ice melt and increased river inow (Jackson
et al., 2010). Primary production appears to have increased over
the shelves (Pabi et al., 2008; Zhang et al., 2010) but perhaps not
over the basins because of at least regionally nutrient limitation
(Tremblay et al., 2011; Codispoti et al., 2013; Matrai and
Appolonio, 2013) along with changes in algal community size
structure with prevalence of small taxa in nutrient-poor basin

113

areas or times (Li et al., 2009; Lee et al., 2012). Knowledge of


long-term patterns of vertical ux and biological observations
above the primary producer level, however, is rudimentary
because of very short time series (Bauerfeind et al., 2009;
Bergmann et al., 2011). The interconnections between physical,
chemical and (lower trophic) biological changes are slowly beginning to be incorporated into pan-Arctic models and clearly document that such connections exist, although they yet need to be
tied to higher tropic levels (Wassmann et al., 2010 and 2015,
Zhang et al., 2010; Slagstad et al., 2011).
One basic question remains elusive: will new (export) primary
production increase or decrease under conditions of a temporally
and spatially reduced ice cover and what role will the basins play?
And will the Arctic food web of the future provide more or less
energy to the higher trophic levels, with potential implications to
the development of commercial sheries? The response will likely
depend on region, process and scale. On the panarctic shelves,
which we include here for comparative purposes the annual
removal of ice cover beyond the shelf break will enhance wind
and ice-forced shelf-break upwelling; the resulting increases in
both nutrient uxes and solar radiation should thus increase new
production (Tremblay et al., 2011; Williams and Carmack, 2015
and Section 6.3); however, the stronger winds required to drive
such upwelling are more common in fall during diminishing light
conditions. This may result in a fall bloom setting (cf. ACIA, 2004,
Fig. 9.13, Ardyna et al., 2014) or result in delayed uptake until
the following spring. In the deep central basins the increased addition of freshwater from ice melt will increase surface layer stratication, particularly in the Pacic Arctic sector (cf. Jackson et al.,
2010). In the Beaufort Gyre, increased coupling of wind stress to
surface waters under reduced ice cover has increased Ekman convergence of surface waters, and thus further increased stratication (Proshutinsky et al., 2009; McLaughlin and Carmack, 2010).
In this case the resulting decrease in vertical nutrient ux to the
euphotic zone should thus decrease new production, but perhaps
increase the relative contribution of ice algae.
The current distribution of faunal biomass (zooplankton, benthos, sh) suggests the energy transfer from primary production
into secondary production in the basins is (1) generally concentrated at the shelf break, i.e. along the basin perimeter, with
decreasing trends associated with increasing depth, and (2) peaks
in areas receiving inux of Atlantic and Pacic -source nutrients,
new production and both allochthonous and autochthonous fauna.
Abutment of water masses along the slope (e.g. the Atlantic layer in
contact with the bottom) may provide preferential habitat for
some species during all or some portion of their life cycle, and
mechanisms of pelagic/benthic coupling (Wassmann, 1998;
Grebmeier, 2012) are thus of paramount importance. For example,
concentrations of Arctic/polar cod (Boreogadus saida), appear to
peak in areas where the (upper portion of) the AW meets the sea
oor (Logerwell et al., 2011; Parker-Stetter et al., 2011; Crawford
et al., 2011). Assuming this distribution pattern is found to hold
true around the basin perimeter, one needs to be concerned about
the potential for these sh providing an easy and concentrated target for future sheries in the Arctic Basins, thus requiring new
management policies in multiple nations (Christiansen et al.,
2014). With various sh species changing their core concentrations
northward in the northern hemisphere (Mueter and Litzow, 2008;
Perry et al., 2005), sheries eets are certain to follow suit; on the
Barents Sea shelf a B. saida shery already goes back to the 1960s
(summarized by Hop and Gjster, 2013). Easy access of a eet to
harvestable grounds would have cascading effects throughout the
food web, reaching to vertebrate predators given that B. saida has
been a stable and lipid-rich staple (c.f. Hop and Gjster, 2013)
for those subsistence-harvested, poster-child Arctic marine
mammals.

114

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

In summary, for the rst time in recent history, a new deep


ocean appears to be opening (cf. Kinnard et al., 2011) and it is likely
that within a few decades or less the Arctic will see mostly ice-free
summers extending fully across its basins. The rapidly changing
Arctic, leading the globe through poorly understood climate amplication processes, offers much in return to allow in-depth study of
social and ecological processes that will be impacted in a warmer
world (ARR, 2013). Prediction of future states is challenging.
Rampal et al. (2011) showed that models underestimate the
observed thinning trend by almost a factor 4 on average, in a
way even more spectacular than they do for the sea ice extent
decline and this critical underestimation of the thinning trend cannot be explained entirely by an underestimation of the decline of
the perennial, thicker, ice-covered portion of the Arctic. Its
mediterranean conguration, with connections to the subarctic
Atlantic and Pacic, provides an ideal setting in which to address
ecology in an advective system (Wassmann et al., 2015). Its deep
basins, in particular, are experiencing conditions that have not
occurred in hundreds if not thousands if not hundreds of thousand
years. As a leading indicator of change it is, in fact, a window to the
future, providing an opportunity to unite science and policy in
ways that can be tested in the high-latitudes and the later applied
with more condence to the lower latitudes (cf. Carmack et al.,
2012). Although uniting science and policy is challenging given
the different agendas and interests between policy makers and scientists, it was in part the political information needs that motivated numerous basin-focused research investigations in the past
decade such as mapping activities that were driven by the desire
to clarify national boundaries. Building on the achievements of
IPY and the International Conference on Arctic Research Planning
II-outcomes (ICARP II Science Plan 4, 2005) and developing ICARP
III, the Arctic research community has shown and will continue
to show pathways forward on how international collaboration is
more productive than focusing on national boundaries. This time
around, the central AO will play an increasing larger role both in
the research and political arenas.
Acknowledgments
This article resulted from the 3rd pan-Arctic Symposium entitled Overarching perspectives of contemporary and future ecosystems in the AO held in Motuvun, Croatia, in October of 2012. We
thank the workshop organizers P. F. Wassmann and M. Reigstad for
creating an inspiring atmosphere for fruitful discussion, and for
supporting the authors travel. We appreciate P.F. Wassmanns persistent encouragement as our guest editor. T. Kikuchi and collaborators are thanked for letting us use their data in Fig. 7. The
manuscript beneted from discussion with B. Bjrk, R. Crawford,
R. Gradinger, I. Polyakov, R. Rember, and K. Wood. We gratefully
acknowledge the preparation of graphics by Falk Huettmann, Patricia Kimber and Liusen Xie. Mette Kaufman and Flora Grabowska
helped with the reference list. KK was partially supported by the
Russian Foundation for Basic Research under Grant 13-04-00551.
This review has beneted from BBs work funded by NOAAs Ofce
of Ocean Exploration (grant #NA 16RP2627), The Bureau of Ocean
Energy Management (Agreement #M12AC00011), and the Alfred.
P. Sloan Foundations Arctic Ocean Diversity Census of Marine Life
project.
References
Aagaard, K., Coachman, L., Carmack, E., 1981. On the halocline of the Arctic Ocean.
Deep-Sea Research Part A: Oceanographic Research Papers 28, 529547.
Aagaard, K., Swift, J.H., Carmack, E.C., 1985. Thermohaline circulation in the Arctic
Mediterranean Seas. Journal of Geophysical Research 90, 48334846. http://
dx.doi.org/10.1029/JC090iC03p04833.

Aagaard, K., 1989. A synthesis of the Arctic Ocean circulation. Rapports et Procesverbaux des Runions. Conseil International pour I0 xploration de la Mer 188,
1122.
Aagaard, K., Carmack, E.C., 1989. The role of sea ice and other fresh water in the
Arctic circulation. Journal of Geophysical Research 94 (C10), 1448514498.
http://dx.doi.org/10.1029/JC094iC10p14485.
Aagaard, K., Carmack, E.C., 1994. The Arctic Ocean and climate: a perspective. In:
Johannessen, O.M., Muench, R.D., Overland, J.E. (Eds.), The Polar Oceans and
Their Role in Shaping the Global Environment, Geophysical Monograph Series
85. American Geophysical Union, Washington, D.C. http://dx.doi.org/10.1029/
GM085p0005, 85. pp. 520.
Aagaard, K., Weingartner, T.J., Danielson, S.L., Woodgate, R.A., Johnson, G.C.,
Whitledge, T.E., 2006. Some controls on ow and salinity in Bering Strait.
Geophysical Research Letters, 3319.
Aagaard, K., Andersen, R., Swift, J., Johnson, J., 2008. A large eddy in the central
Arctic Ocean. Geophysical Research Letters 35, L09601. http://dx.doi.org/
10.1029/2008GL033461.
ACIA, 2005. Arctic Climate Impact Assessment. Cambridge University Press,
Cambridge, 1042p.
Alkire, M.B., Falkner, K.K., Morison, J., Collier, R.W., Guay, C.K., Desiderio, R.A., Rigor,
I.G., McPhee, M., 2010. Sensor-based proles of the NO parameter in the central
Arctic and southern Canada Basin: New insights regarding the cold halocline.
Deep Sea Research I 57, 14321443.
Aksenov, Y., Ivanov, V.V., Nurser, A.J.G., Bacon, S., Polyakov, I.V., Coward, A.C.,
Naveira-Garabato, A.C., Beszczynska-Moeller, A., 2011. The Arctic circumpolar
boundary current. Journal of Geophysical Research 116, C09017.
Anderson, L.G., Jones, E.P., Koltermann, K.P., Schlosser, P., Swift, J.H., Wallace, D.W.R.,
1989. The rst oceanographic section across the Nansen Basin in the Arctic
Ocean. Deep-Sea Research 36, 475482.
Anderson, L.G., Bjrk, G., Holby, O., Jones, E.P., Kattner, G., Koltermann, K.P.,
Liljeblad, B., Lindegren, R., Rudels, B., Swift, J., 1994. Water masses and
circulation in the Eurasian Basin: results from the Oden 91 expedition. Journal
of Geophysical Research 99 (C2), 32733283. http://dx.doi.org/10.1029/
93JC02977.
Anderson, L.G., Tanhua, T., Bjork, G., Hjalmarsson, S., Jones, E.P., Jutterstrom, S.,
Rudels, B., Swift, J.H., Wahlstom, I., 2010. Arctic Ocean shelf-basin interaction:
an active continental shelf CO2 pump and its impact on the degree of calcium
carbonate solubility. Deep-Sea Research I 57, 869879. http://dx.doi.org/
10.1016/j.dsr.2010.03.012.
Andronov, V.N., Kosobokova, K.N., 2011. New species of small, bathypelagic
calanoid copepods from the Arctic Ocean: Brodskius arcticus sp. nov.
(Tharybidae) and three new species of Pertsovius gen. nov. (Discoidae).
Zootaxa 2809, 3346.
Arctic Council, 2013. Arctic Resilience Interim Report 2013. Stockholm Environment
Institute and Stockholm Resilience Centre, Stockholm, ISBN 978-91-86125-431117 page.
Ardyna, M., Babin, M., Gosselin, M., Devred, E., Rainville, L., Tremblay, J.., 2014.
Recent Arctic Ocean sea ice loss triggers novel fall phytoplankton blooms.
Geophysical Research Letters 41 (17), 62076212.
Arrigo, K.R., Dijken, G.v., Pabi, S., 2008. Impact of a shrinking Arctic ice cover on
marine primary production. Geophysical Research Letters 35, L19603. http://
dx.doi.org/10.1029/2008GL035028.
Arrigo, K.R., Perovich, D.K., Pickart, R.S., Brown, Z.W., van Dijken, G.L., Lowry, K.E.,
Mills, M., Palmer, M.A., Balch, W.M., Bahr, F., Bates, N.R., Benitez-Nelson, C.,
Bowler, B., Brownlee, E., Ehn, J.K., Frey, K.E., Garley, R., Laney, S.R., Lubelczyk, L.,
Mathis, J., Matsuoka, A., Mitchell, B.G., Schieber, B., Sosik, H.M., Stephens, M.,
Swift, J.H., 2012. Massive phytoplankton blooms under Arctic sea ice. Science
336 (6087), 1408-1408.
Ashjian, C.J., Campbell, R.G., Welch, H.E., Butler, M., Van Keuren, D., 2003. Annual
cycle in abundance, distribution, and size in relation to hydrography of
important copepod species in the western Arctic Ocean. Deep-Sea Research I
50, 12351261.
Backman, J., Moran, K., 2009. Expanding the Cenozoic paleoceanographic record in
the Central Arctic Ocean: IODP Expedition 302 Synthesis. Central European
Journal of Geosciences 1 (2), 157175. http://dx.doi.org/10.2478/v10085-0090015-6.
Barber, D.G., Hop, H., Mundy, C.J., Else, B., Dmitrenko, I.A., Tremblay, J.-E., Ehn, J.K.,
Assmy, P., Daase, M., Candlish, L.M., Rysgaard, S., 2015. Selected physical,
biological and biogeochemical implicaitons of a rapidly changing Arctic
Marginal Ice Zone. Progress in Oceanography 139, 122150.
Bauch, D., Torres-Valdes, S., Polyakov, I., Novikhin, A., Dmitrenko, I., McKay, J., Mix,
A., 2014. Halocline water modication and along-slope advection at the Laptev
Sea continental margin. Ocean Science 10, 141154. http://dx.doi.org/10.5194/
os-10-141-2014, www.ocean-sci.net/10/141/2014/.
Bauerfeind, E., Nthig, E.-M., Beszczynska-Mller, A., Fahl, K., Kaleschke, L., Kreker,
K., Klages, M., Soltwedel, T., Lorenzen, C., Wegner, J., 2009. Particle
sedimentation patterns in the eastern Fram Strait during 20002005: results
from the Arctic long-term observatory HAUSGARTEN. Deep-Sea Research I 56,
14061417. http://dx.doi.org/10.1016/j.dsr.2009.04.011.
Bent, S., 1872. Thermal paths to the pole. R.P. Studley Co., Saint Louis, 40 pp.
Berge, J., Varpe, ., Moline, M.A., Wold, A., Renaud, P.E., Daase, M., Falk-Petersen, S.,
2012. Retention of ice-associated amphipods: possible consequences for an icefree Arctic Ocean. Biology Letters. http://dx.doi.org/10.1098/rsbl.2012.0517.
Bergmann, M., Dannheim, J., Bauerfeind, E., Klages, M., 2009. Trophic relationships
along a bathymetric gradient at the deepsea observatory HAUSGARTEN. DeepSea Research I 56, 408424.

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121


Bergmann, M., Soltwedel, T., Klages, M., 2011. The interannual variability of
megafaunal assemblages in the Arctic deep sea: Preliminary results
from the HAUSGARTEN observatory (79 N). Deep-Sea Research I 58,
711722.
Beszczynska-Mller, A., Woodgate, R.A., Lee, C., Melling, H., Karcher, M., 2011. A
synthesis of exchanges through the main oceanic gateways to the Arctic Ocean.
Oceanography 24, 8299.
Bhatt, U.S., Walker, D.A., Walsh, J.E., Carmack, E.C., Frey, K.E., Meier, W., Moore, S.,
Parmentier, F.J.W., Post, E., Romanovsky, V., Simpson, W., 2014. Implications of
Arctic sea ice decline for the earth system. Annual Review of Environment and
Resources 39, 5789. http://dx.doi.org/10.1146/annurev-122012-094357.
Billett, D.S.M., Hansen, B., 1982. Abyssal aggregations of Kolga hyalina Danielssen
and Koren (Echinodermata: Holothurioidea) in the northeast Atlantic Ocean: a
preliminary report. Deep-Sea Research 29, 799818.
Billett, D.S.M., Bett, B.J., Rice, A.L., Thurston, M.H., Galeron, J., Sibuet, M., Wolff, G.A.,
2001. Long-term change in the megabenthos of the Porcupine Abyssal Plain (NE
Atlantic). Progress in Oceanography 325, 325348.
Bilyard, G.R., Carey, A.G., 1980. Zoogeography of western Beaufort Sea Polychaeta
(Annelida). Sarsia 65, 1926.
Bjrk, G., Winsor, P., 2006. The deep waters of the Eurasian Basin, Arctic Ocean:
Geothermal heat ow, mixing and renewal. Deep-Sea Research I 53, 1253
1271. http://dx.doi.org/10.1016/j.dsr.2006.05.006.
Bjrk, G., Jakobsson, M., Rudels, B., Swift, J.H., Anderson, L., Darby, D.A., Backman, J.,
Coakley, B., Winsor, P., Polyak, L., Edwards, M., 2007. Bathymetry and deepwater exchange across the central Lomonosov Ridge at 8889 N. Deep-Sea
Research I 54, 11971208.
Bjrk, G., Anderson, L.G., Jakobsson, M., Antony, D., Eriksson, B., Eriksson, P.B., Hell,
B., Hjalmarsson, S., Janzen, T., Jutterstrom, S., Linders, J., Lowemark, L.,
Marcussen, C., Olsson, K.A., Rudels, B., Sellen, E., Slvsten, M., 2010. Flow of
Canadian Basin deep water in the Western Eurasian Basin of the Arctic Ocean.
Deep Sea Research Part I 574, 577586.
Bluhm, B.A., MacDonald, I.R., Debenham, C., Iken, K., 2005. Macro- and megabenthic
communities in the high Arctic Canada Basin: initial ndings. Polar Biology 28,
218231.
Bluhm, B.A., Gradinger, R., 2008. Regional variability in food availability for Arctic
marine mammals. Ecological Applications 18 (suppl. 2), 7796.
Bluhm, B.A., Gradinger, R.R., Schnack-Schiel, S.B., 2010. Sea ice meio- and
macrofauna. In: Thomas, D., Dieckmann, G. (Eds.), Sea ice, second ed. WileyBlackwell, Oxford, pp. 357394.
Bluhm, B.A., Gebruk, A.V., Gradinger, R., Hopcroft, R.R., Huettmann, F., Kosobokova,
K.N., Sirenko, S.I., Weslawski, J.M., 2011a. Arctic marine biodiversity an update
of species richness and examples of biodiversity change. Oceanography 24,
232248.
Bluhm, B.A., Ambrose Jr., W.G., Bergmann, M., Clough, L.M., Gebruk, A.V., Hasemann,
C., Iken, K., Klages, M., MacDonald, I.R., Renaud, P.E., Schewe, I., Soltwedel, T.,
Wlodarska-Kowalczuk, M., 2011b. Diversity of the Arctic deep-sea benthos.
Marine Biodiversity 41, 87107.
Boetius, A., Albrecht, S., Bakker, K., Bienhold, C., Felden, J., Fernandez-Mendez, M.,
Hendricks, S., Katlein, C., Lalande, C., Krumpen, T., Nicolaus, M., Peeken, I., Rabe,
B., Rogacheva, A., Rybakova, E., Somavilla, R., Wenzhfer, F., Polarstern, R.V.ARK
27-3-Shipboard Science Party,, 2013. Export of algal biomass from the melting
Arctic sea ice. Science 339, 14301432.
Bourgain, P., Gascard, J.C., 2011. The Arctic Ocean halocline and its interannual
variability from 1997 to 2008. Deep-Sea Research I 58, 745756.
Brandt, A., Gooday, A.J., Brando, S.N., Brix, S., Brkeland, W., Cedhagen, T.,
Choudhury, M., Cornelius, N., Danis, B., De Mesel, I., Diaz, R.J., Gillan, D.C.,
Ebbe, B., Howe, J.A., Janussen, D., Kaiser, S., Linse, K., Malyutina, M., Pawlowski,
J., Raupach, M., Vanreusel, A., 2007. First insights into the biodiversity and
biogeography of the Southern Ocean deep sea. Nature 447 (7142), 307311.
http://dx.doi.org/10.1038/nature05827.
Brodsky, K.A., 1956. Deep-sea life in the Arctic Basin. Priroda 5, 4148 (Translation
No. 242, American Meteorological Society).
Brodsky, K.A., Pavshtiks, E.A., 1977. Plankton of the central part of the Arctic Basin.
Polar Geography 1, 143161 (Translated from Voprosy Geographii, 101, 148
157; 1976).
Brown, K.A., McLaughlin, F., Tortell, P.D., Varela, D.E., Yamamoto-Kawai, M., Hunt, B.,
Francois, R., 2014. Determination of particulate organic carbon sources to the
surface mixed layer of the Canada Basin, Arctic Ocean. Journal of Geophysical
Research: Oceans 119, 10841102. http://dx.doi.org/10.1002/2013JC009197.
Brown, Z.W., Arrigo, K.R., 2012. Contrasting trends in sea ice and primary
production in the Bering Sea and Arctic Ocean. ICES Journal of Marine
Research 69, 11801193.
Carmack, E.C., 2000. The Freshwater Budget of the Arctic Ocean: Sources, Storage
and Sinks. In: Lewis, E.L. (Ed.), The Freshwater Budget of the Arctic Ocean. NATO
Advanced Research Series, pp. 91126.
Carmack, E.C., 2007. The alpha/beta ocean distinction: a perspective on freshwater
uxes, convection, nutrients and productivity in high-latitude seas. Deep-Sea
Research
II
54
(2326),
25782598.
http://dx.doi.org/10.1016/
j.dsr2.2007.08.018.
Carmack, E.C., Macdonald, R.W., Perkin, R.G., McLaughlin, F.A., Pearson, R.J., 1995.
Evidence for warming of Atlantic water in the Southern Canadian Basin of the
Arctic Ocean: Results from the Larsen-93 expedition. Geophysical Research
Letters 22, 10611064.
Carmack, E.C., Aagaard, K., Swift, J.H., Perkin, R.G., McLaughlin, F.A., Macdonald,
R.W., Jones, E.P., Smith, J.N., Ellis, K.M., Kilius, L.R., 1997. Changes in temperature

115

and tracer distributions within the Arctic Ocean: results from the 1994 Arctic
Ocean Section. Deep-Sea Research 44, 14871502.
Carmack, E.C., Kulikov, E.A., 1998. Wind-forced upwelling and internal Kelvin wave
generation in Mackenzie Canyon, Beaufort Sea. Journal of Geophysical Research
103, 1844718458. http://dx.doi.org/10.1029/98JC00113.
Carmack, E.C., Aagaard, K., Swift, J.H., Perkin, R.G., McLaughlin, F.A., Macdonald,
R.W., Jones, E.P., 1998. Thermohaline transitions. In: Imberger, J. (Ed.), Physical
Processes in Lakes and Oceans. Coastal and Estuarine Studies, 54, American
Geophysical Union, pp. 179186.
Carmack, E.C., McLaughlin, F.A., 2000. Arctic Ocean Change and consequences to
biodiversity: a perspective on linkage and scale. Memoirs of National Institute
of Polar Research 54, 365375.
Carmack, E., Chapman, D.C., 2003. Wind-driven shelf/basin exchange on an Arctic
shelf: the joint roles of ice cover extent and shelf-break bathymetry.
Geophysical Research Letters 30 (14). http://dx.doi.org/10.1029/2003GL017526.
Carmack, E., Barber, D., Christensen, J., Macdonald, R., Rudels, B., Sakshaug, E., 2006.
Climate variability and physical forcing of the food webs and the carbon budget
on pan-arctic shelves. Progress in Oceanography 71, 145181. http://dx.doi.org/
10.1016/j.pocean.2006.10.005.
Carmack, E., Wassmann, P., 2006. Food webs and physical-biological coupling on
pan-Arctic shelves: unifying concepts and comprehensive perspectives.
Progress in Oceanography 71 (24), 446477. http://dx.doi.org/10.1016/
j.pocean.2006.10.004.
Carmack, E.C., McLaughlin, F.A., Yamamoto-Kawai, M., Itoh, M., Shimada, K.,
Krisheld, R., Proshutinsky, A., 2008. Freshwater storage in the Northern
Ocean and the special role of the Beaufort Gyre. In: Dickson, R.R., Meincke, J.,
Phines, P. (Eds.), ArcticSubarctic Ocean Fluxes, Dening the Role of the
Northern Seas in Climate. Springer, pp. 145169.
Carmack, E.C., Melling, H., 2011. Warmth from the deep. Nature Geoscience 4, 78.
Carmack, E.C., McLaughlin, F.A., 2011. Towards recognition of physical and
geochemical change in subarctic and arctic seas. Progress in Oceanography
90, 90104. http://dx.doi.org/10.1016/j.pocean.2011.02.007.
Carmack, E.C., Williams, W.J., Zimmermann, S.L., McLaughlin, F.A., 2012. The Arctic
Ocean warms from below. Geophysical Research Letters 39, L07604. http://
dx.doi.org/10.1029/2012GL050890.
Carmack, E.C., Winsor, P., Williams, W., 2015. The contiguous Riverine Coastal
Domain. Progress in Oceanography 139, 1323.
Carpenter, J.R., Timmermans, M.-L., 2012. Deep eddies in the Canada Basin.
Geophysical Research Letters 39, L20602. http://dx.doi.org/10.1029/
2012GL053025.
Cavalieri, D.J., Parkinson, C.L., 2012. Arctic sea ice variability and trends, 19792010.
The Cryosphere 6, 881889.
Christiansen, J.S., Mecklenburg, C.W., Karamushko, O.V., 2014. Arctic marine shes
and their sheries in light of global change. Global Change Biology 20 (2), 352
359.
Citta, J.J., Suydam, R.S., Quakenbush, L.T., Frost, K.J., OCorry-Crowe, G.M., 2013. Dive
behavior of Eastern Chukchi beluga whales (Delphinapterus leucas), 1998
2008. Arctic 66, 389406.
Clough, L.M., Ambrose Jr., W.G., Cochran, J.K., Barnes, C., Renaud, P.E., Aller, R.C.,
1997. Infaunal density, biomass and bioturbation in the sediments of the Arctic
Ocean. Deep-Sea Research II 44, 16831704.
Coachman, L.K., Barnes, C.A., 1961. The contribution of Bering Sea water to the
Arctic Ocean. Arctic 14 (3), 147161.
Coachman, L.K., Aagard, K., 1974. Physical oceanography of Arctic and Subarctic
Seas. In: Herman, Y. (Ed.), Marine Geology and Oceanography of Arctic Seas.
Springer Verlag, New York, pp. 172.
Coachman, L.K., Aagaard, K., Tripp, R.B., 1975. Bering Strait. The Regional and
Physical Oceanography. University of Washington Press, Seattle, Washington,
172p.
Codispoti, L.A., Owens, T.G., 1975. Nutrient transports through Lancaster Sound in
relation to the Arctic Oceans reactive silicate budget and the outow of Bering
Strait waters. Limnology and Oceanography 20 (1), 115119.
Codispoti, L.A., Flagg, C., Kelly, V., Swift, J.H., 2005. Hydrographic conditions during
the 2002 SBI process experiments. Deep-Sea Research Part II 52 (24), 3199
3226.
Codispoti, A., Kelly, V., Thessen, A., Matrai, P., Suttles, S., Hill, V., Light, B., 2013.
Synthesis of primary production in the Arctic Ocean: III. Nitrate and phosphate
based estimates of net community production. Progress in Oceanography 110,
126150. http://dx.doi.org/10.1016/j.pocean.2012.11.006.
Collins, R.E., Deming, J.W., 2011. Abundant dissolved genetic material in Arctic sea
ice Part I: Extracellular DNA. Polar Biology 34 (12), 18191830.
Comiso, J.C., 2012. Large decadal decline of the Arctic multiyear ice cover. Journal of
Climate 25, 11761193.
Conlan, K., Hendrycks, E., Aitken, A., Williams, B., Blasco, S., Crawford, E., 2013.
Macrofaunal biomass distribution on the Canadian Beaufort Shelf. Journal of
Marine Systems 127, 7687.
Conover, R.J., Huntley, M., 1991. Copepods in ice-covered seasdistribution,
adaptations to seasonally limited food, metabolism, growth patterns and life
cycle strategies in polar seas. Journal of Marine Systems 2 (1), 141.
Cooper, L.W., Cota, G.F., Pomeroy, L.R., Grebmeier, J.M., Whitledge, T., 1999.
Modication of NO, PO, and NO/PO during ow across the Bering and
Chukchi shelves: implications for use as Arctic water mass tracers. Journal of
Geophysical Research 104 (C4), 78277836.
Coupel, P., Jin, H.Y., Joo, M., Horner, R., Bouvet, H.A., Garon, V., Sicre, M.-A., Gascard,
J.-C., Chen, J.F., Ruiz-Pino, D., 2012. Phytoplankton distribution in unusually low

116

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

sea ice cover over the Pacic Arctic. Biogeosciences Discussions 9 (2), 2055
2093.
Crawford, R.E., Vagle, S., Carmack, E.C., 2011. Water mass and bathymetric
characteristics of Arctic cod habitat along the continental shelf and slope of
the Beaufort and Chukchi seas. Polar Biology 35, 179190. http://dx.doi.org/
10.1007/s00300-011-1051-9.
Curry, B., Lee, C.M., Petrie, B., 2011. Volume, freshwater, and heat uxes through
Davis Strait, 200405. Journal of Physical Oceanography 41 (3), 429436.
Darnis, G., Fortier, L., 2014. Food, temperature, and the seasonal vertical migration
of key arctic copepods in the thermally stratied Amundsen Gulf (Beaufort Sea,
Arctic Ocean). Journal of Plankton Research. http://dx.doi.org/10.1093/plankt/
fbu035.
DAsaro, E.A., 1988. Observations of small eddies in the Beaufort Sea. Journal of
Geophysical Research: Oceans (19782012) 93 (C6), 66696684.
Deal, C., Jin, M., Elliott, S., Hunke, E., Maltrud, M., Jeffery, N., 2011. Large-scale
modeling of primary production and ice algal biomass within arctic sea ice in
1992. Journal of Geophysical Research 116, C07004. http://dx.doi.org/10.1029/
2010JC006409.
Dethleff, D., Rachold, V., Tintelnot, M., Antonow, M., 2000. Sea-ice transport of
riverine particles from the Laptev Sea to Fram Strait based on clay mineral
studies. International Journal of Earth Sciences 89 (3), 496502.
Deubel, H., 2000. Structures and nutrition requirements of macrozoobenthic
communities in the area of the Lomonossov Ridge in the Arctic Ocean.
Reports on Polar Research 370, 1147 (in German).
Devol, A.H., Codispoti, L.A., Christenen, J.P., 1997. Summer and winter denitrication
rates in Arctic shelf sediments. Continental Shelf Research 17, 10291050.
Dick, H.J.B., Lin, J., Schouten, H., 2003. An ultraslow-spreading class of ridge. Nature
426, 405412.
Dickson, R., Rudels, B., Dye, S., Karcher, M., Meincke, J., Yashayaev, I., 2007. Current
estimates of freshwater ux through Arctic and subarctic seas. Progress in
Oceanography 73, 210230. http://dx.doi.org/10.1016/j.pocean.2006.12.003.
Dickson, R.R., Meincke, J., Rhines, P., 2008. Arctic-Subarctic Ocean Fluxes: Dening
the Role of the Northern Seas in Climate A general introduction. In: Dickson,
R.R., Meincke, J., Rhines, P. (Eds.), Arctic-Subarctic Ocean Fluxes: Dening the
Role of the Northern Seas in Climate. Springer, Netherlands, Dordrecht, pp. 113.
Dmitrenko, I.A., Ivanov, V.V., Kirillov, S.A., Vinogradova, E.L., Torres-Valdes, S.,
Bauch, D., 2011. Properties of the Atlantic derived halocline waters over the
Laptev Sea continental margin: evidence from 2002 to 2009. Journal of
Geophysical Research: Oceans 116. http://dx.doi.org/10.1029/2011jc007269.
Dmitrenko, I.A., Kirillov, S.A., Serra, N., Koldunov, N.V., Ivanov, V.V., Schauer, U.,
Polyakov, I.V., Barber, D., Janout, M., Lien, V.S., Makhotin, M., Aksenov, Y., 2014.
Heat loss from the Atlantic water layer in the northern Kara Sea: causes and
consequences. Ocean Sciences 10, 719730. http://dx.doi.org/10.5194/os-10719-2014.
Dunton, K.H., Schonberg, S.V., Cooper, L.W., 2012. Food web structure of the Alaskan
nearshore shelf and estuarine lagoons of the Beaufort Sea. Estuaries and Coasts
35 (2), 416435.
Eicken, H., Gradinger, R., Gaylord, A., Mahoney, A., Rigor, I., Melling, H., 2005.
Sediment transport by sea ice in the Chukchi and Beaufort Seas: increasing
importance due to changing ice conditions? Deep-Sea Research Part II 52 (24),
32813302.
Fahrbach, E., Meincke, J., sterhus, S., Rohardt, G., Schauer, U., Tverberg, V., Verduin,
J., 2001. Direct measurements of volume transports through Fram Strait. Polar
Research 20, 217224.
Fedorov, A.V., Dekens, P.S., McCarthy, M., Ravelo, A.C., deMenocal, P.B., Barreiro, M.,
Pacanowski, R.C., Philander, S.G., 2006. The Pliocene paradox (mechanisms for a
permanent El Nio). Science 312, 14851489.
Fernndez-Mndez, M., Wenzhfer, F., Peeken, I., Srensen, H.L., Glud, R.N., Boetius,
A., 2014. Composition, buoyancy regulation and fate of ice algal aggregates in
the central Arctic Ocean. PloS One 9 (9), e107452.
Fetzer, I., Arndt, W.E., 2008. Reproductive strategies of benthic invertebrates in the
Kara Sea (Russian Arctic): adaptation of reproduction modes to cold water.
Marine Ecology Progress Series 356, 189202.
Fjeldstad, J.E., 1936. Results of tidal observations. The Norwegian North Polar
Research Expedition with the Maud, 19181925, Scientic Results, 4, 188, A
S John Griegs Boktrykkeri, Bergen.
Frolov, I.E., Gudkovich, Z.M., Radionov, V.F., Shirochkov, A.V., Timokhov, L.A., 2005.
The Arctic Basin. Results from the Russian drifting stations. Springer Verlag,
Berlin, 276p.
Frost, K.J., Lowry, L.F., 1984. Trophic relationships of vertebrate consumers in the
Alaskan Beaufort Sea. In: Barnes, P.W., Schell, D.M., Reimnitz, E. (Eds.), The
Alaskan Beaufort Sea ecosystems and environments. Academic Press, New York,
pp. 381401.
Gagaev, S.Y., 2008. Sigambra healyae sp. n., a new species of polychaete (Polychaeta:
Pilargidae) from the Canada Basin of the Arctic Ocean. Russian Journal of Marine
Biology 34, 7375.
Garrison, G.R., Becker, P., 1976. Barrow canyon: A drain for the Chukchi Shelf.
Journal of Geophysical Research 81, 44454453.
Geinrikh, A.K., Kosobokova, K.N., Rudyakov, Y.A., 1983. Seasonal variations in the
vertical distribution of some prolic copepods of the Arctic basin. In:
Vinogradov, M.E. (Ed.), Biologiya Tsentralnogo Arkicheskogo basseyna
(Biology of the central Arctic Basin). Nauka, Moscow, 1980. Canadian
Translation Fisheries and Aquatic Science, vol. 4925, pp. 122.
Giles, K.A.S., Laxon, W.A., Ridout, L., Wingham, D.J., Bacon, S., 2012. Western Arctic
Ocean freshwater storage increased by wind-driven spin-up of the Beaufort
Gyre. Nature Geoscience 5, 194197.

Golikov, A.N., Scarlato, O.A., 1989. Evolution of arctic ecosystems during the
Neogene period. In: Herman, Y. (Ed.), The Arctic Seas: Climatology,
Oceanography and Biology. Van Nostrand, Reinhold, New York, pp. 257259.
Gordeev, V.V., Martin, J.M., Sidorov, I.S., Sidorova, M.V., 1996. A reassessment of the
Eurasian river input of water, sediment, major elements, and nutrients to the
Arctic Ocean. American Journal of Science 296 (6), 664691.
Gordeev, V.V., 2000. River input of water, sediment, major ions, nutrients and trace
metals from Russian territory to the Arctic Ocean. In: Lewis, E.L., Jones, E.P.,
Lemke, P., Prowse, T.D., Wadhams, P. (Eds.), The Freshwater Budget of the Arctic
Ocean. Klewar Academic Publishers, Dordrecht, pp. 297322.
Grska, B., Grzelak, K., Kotwicki, L., Hasemann, C., Schewe, I., Soltwedel, T.,
Wodarska-Kowalczuk, M., 2014. Bathymetric variations in vertical
distribution patterns of meiofauna in the surface sediments of the deep Arctic
ocean (HAUSGARTEN, Fram Strait). Deep-Sea Research Part I 91, 3649.
Gosselin, M., Levasseur, M., Wheeler, P.A., Horner, R.A., Booth, B.C., 1997. New
measurements of phytoplankton and ice algal production in the Arctic Ocean.
Deep-Sea Research II 44, 16231644.
Gradinger, R., 2009. Sea ice algae: major contributors to primary production and
algal biomass in the Chukchi and Beaufort Sea during May/June 2002. Deep-Sea
Research II 56, 12011212.
Grainger, E.H., 1989. Vertical distribution of zooplankton in the central Arctic
Ocean. In: Rey, L., Alexander, V. (Eds.), Proceedings 6th Conference Comit
Arctique International, pp. 4860.
Grantham, B.A., Eckert, G.L., Shanks, A.L., 2003. Dispersal potential of marine
invertebrates in diverse habitats. Ecological Applications 13, S108S116.
Grantz, A.G., Phillips, R.L., Mullen, M.W., Starratt, S.W., Jones, G.A., Naidu, A.S.,
Finney, B.P., 1996. Character, paleoenvironment, rate of accumulation, and
evidence for seismic triggering of Holocene turbidites, Canada Abysal Plain,
Arctic Ocean. Marine Geology 133, 5173.
Grebmeier, J.M., 2012. Shifting patterns of life in the Pacic Arctic and Sub-Arctic
seas. Annual Review of Marine Science 4, 6378.
Guilini, K., van Oevelen, D., Soetaert, K., Middelburg, J.J., Vanreusel, A., 2010.
Nutritional importance of benthic bacteria for deep-sea nematodes from the
Arctic ice margin: results of an isotope tracer experiment. Limnology and
Oceanography 55 (5), 19771989.
Gurjanova, E.F., 1985. Deep water species of Arctic endemic genera of echinoderms
and species of autochtonic arctic isopod genus. In: Treshnikov, A.F. (Ed.), Atlas of
the Arctic. Moscow, Main Department of geodesy and cartography of the
Council of Ministers, 131p.
Haine, T.W.N., Curry, B., Gerdes, R., Hansen, E., Karcher, M., Lee, C., Rudels, B., Spreen,
G., de Steur, L., Stewart, K.D., Woodgate, R., 2015. Arctic freshwater export:
Status, mechanisms and prospects. Global and Planetary Change 125, 1335.
Harding, G.C.H., 1966. Zooplankton distribution I the Arctic Ocean with notes of life
cycles. MS Thesis, McGill University, Montreal.
Harding, G.C.H., 1974. The food of deep-sea copepods. Journal Marine Biology
Association UK 54, 141155.
Hasumi, H., 2002. Sensitivity of the global thermohaline circulation to interbasin
freshwater transport by the atmosphere and the Bering Strait throughow.
Journal of Climate 15, 25162526.
Haug, G.H., Tiedemann, R., Zahn, R., Ravelo, A.C., 2001. The role of Panama uplift on
oceanic freshwater balance. Geology 29, 207210.
Hirche, H.J., Mumm, N., 1992. Distribution of dominant copepods in the Nansen
Basin, Arctic Ocean, in summer. Deep-Sea Research Part A 39 (2), S485S505.
Holmes, R.M., McClelland, J.W., Peterson, B.J., Tank, S.E., Bulygina, E., Eglinton, T.,
Gordeev, V.V., Gurtovaya, T.Y., Raymond, P.A., Repeta, D.J., Staples, R., Striegl, R.,
Zhulidov, A.V., Zimov, S.A., 2012. Seasonal and annual uxes of nutrients and
organic matter from large rivers to the Arctic Ocean and surrounding seas.
Estuaries and Coasts 35, 369382.
Holmes, R.M., McClelland, J.W., Peterson, B.J., Shiklomanov, I.A., Shiklomanov, A.I.,
Zhulidov, A.V., Gordeev, V.V., Bobrovitskaya, N.N., 2002. A circumpolar
perspective on uvial sediment ux to the Arctic Ocean. Global
Biogeochemical Cycles 16 (4), 45-1.
Hop, H., Gjster, H., 2013. Polar cod (Boreogadus saida) and capelin (Mallotus
villosus) as key species in marine food webs of the Arctic and the Barents Sea.
Marine Biology Research 99, 878894.
Hopcroft, R.R., Clarke, C., Nelson, R.J., Raskoff, K.A., 2005. Zooplankton communities
of the Arctics Canada Basin: the contribution by smaller taxa. Polar Biology 283,
198206.
Hopkins, T.L., 1969a. Zooplankton standing crop in the Arctic Basin. Limnology and
Oceanography 14, 8095.
Hopkins, T.L., 1969b. Zooplankton biomass related to hydrography along the drift
track of Arlis II in the Arctic Basin and the East Greenland Current. Journal of
Fisheries Research Board Canada 26, 305310.
Hoste, E., Vanhove, S., Schewe, I., Soltwedel, T., Vanreusel, A., 2007. Spatial and
temporal variation in deep-sea meiofauna assemblages in the Marginal Ice Zone
of the Arctic Ocean. Deep-Sea Research I 54, 109129.
Hunkins, K.L., 1974. Subsurface eddies in the Arctic Ocean. Deep Sea Research and
Oceanographic Abstracts 21 (12), 10171033.
Hunt, B.P.V., Nelson, J., Williams, W., Carmack, E.C., McLaughlin, F.A., Vagle, S.,
Young, K., Brown, K.A., 2014. Zooplankton community structure and dynamics
in the Arctic Canada Basin during a period of intense environmental
change (20042009). Journal of Geophysical Research: Oceans 119 (4),
25182538.
ICARP II, 2005. Science Plan 4 Deep Central Basin of the Arctic Ocean. Second
International Conference on Arctic Research Planning (ICARP II). Copenhagen,
Denmark, 1012 November 2005.

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121


Iken, K., Bluhm, B.A., Gradinger, R., 2005. Food web structure in the high
Arctic Canada Basin: evidence from d13C and d15N analysis. Polar Biology 28,
238249.
Iken, K., Bluhm, B.A., Dunton, K., 2010. Benthic food web structure serves as
indicator of water mass properties in the southern Chukchi Sea. Deep-Sea
Research II 57, 7185.
Isachsen, P.E., LaCasce, J.H., Mauritzen, C., Hkkinen, S., 2003. Wind-driven
variability of the large-scale recirculating ow in the Nordic Seas and Arctic
Ocean. Journal of Physical Oceanography 33 (12), 25342550.
Jakobsson, M., 2002. Hypsometry and volume of the Arctic Ocean and its
constituent seas. Geochemistry, Geophysics, Geosystems 35, 118.
Jackson, J.M., Carmack, E.C., McLaughlin, F.A., Allen, S.E., Ingram, R.G., 2010.
Identication, characterization and change of the near-surface temperature
maximum in the Canada Basin, 19932008. Journal of Geophysical Research
115. http://dx.doi.org/10.1029/2009JC005265.
Jakobsson, M., Grantz, A., Kristoffersen, Y., Macnab, R., 2004. Bathymetry and
physiography of the Arctic Ocean and its constituent seas. In: Stein, R.,
Macdonald, R.W. (Eds.), The organic carbon cycle in the Arctic Ocean. Springer,
Heidelberg, pp. 16.
Jakobsson, M., Mayer, L., Coakley, B., Dowdeswell, J.A., Forbes, S., Fridman, B.,
Hodnesdal, H., Noormets, R., Pedersen, R., Rebesco, M., Schenke, H., Zarayskaya,
Y., Accettella, D., Armstrong, A., Anderson, R.M., Bienhoff, P., Camerlenghi, A.,
Church, I., Edwards, M., Gardner, J.V., Hall, J.K., Hell, B., Hestvik, O., Kristoffersen,
Y., Marcussen, C., Mohammad, R., Mosher, D., Nghiem, S.V., Pedrosa, M.T.,
Travaglini, P.G., Weatherall, P., 2012. The International Bathymetric Chart of the
Arctic Ocean (IBCAO) Version 3.0. Geophysical Research Letters 39, L12609.
http://dx.doi.org/10.1029/2012GL052219.
Jacobsson, M., Polyak, L., Edwards, M., Kleman, J., Coakley, B., 2003. Glacial
geomorphology of the Central Arctic Ocean: the Chukchi Borderland and the
Lomonosov Ridge. Earth Surface Processes and Landforms 33, 526545.
Jeffries, M.O., Overland, J.E., Perovich, D.K., 2013. The Arctic shifts to a new normal.
Physics Today 66, 3540. http://dx.doi.org/10.1063/PT.3.2147.
Ji, R., Jin, M., Varpe, ., 2013. Sea ice phenology and timing of primary production
pulses in the Arctic Ocean. Global Change Biology 193, 734741.
Jin, M., Deal, C., Lee, S.H., Elliot, S., Hunke, E., Maltrud, M., Jeffery, N., 2012.
Investigation of Arctic sea ice and ocean primary production for the period
19922007 using a 3-D global ice-ocean ecosystem model. Deep-Sea Research II
8184, 2835.
Jones, E.P., Anderson, L.G., 1986. On the origin of the chemical properties of the
Arctic Ocean halocline. Journal of Geophysical Research: Oceans 91 (C9), 759
767. http://dx.doi.org/10.1029/JC091iC09p10759.
Jones, E.P., Rudels, B., Anderson, L.G., 1995. Deep waters of the Arctic Ocean: origins
and circulation. Deep-Sea Research I 42 (5), 737760.
Karcher, M., Beszczynska-Mller, A., Kauker, F., Gerdes, R., Heyen, S., Rudels, B.,
Schauer, U., 2011. Arctic Ocean warming and its consequences for the Denmark
Strait overow. Journal of Geophysical Research: Oceans (19782012), 116
(C2).
Karcher, M., Smith, J.N., Kauker, F., Gerdes, R., Smethie, W.M., 2012. Recent changes
in Arctic Ocean circulation revealed by iodine-129 observations and modeling.
Journal of Geophysical Research: Oceans (19782012), 117 (C8).
Katlein, C., Fernndez-Mndez, M., Wenzhfer, F., Nicolaus, M., 2014. Distribution of
algal aggregates under summer sea ice in the Central Arctic. Polar Biology.
http://dx.doi.org/10.1007/s00300-014-1634-3.
Kikuchi, T., Hatakeyama, K., Morison, J.H., 2004. Distribution of convective lower
halocline water in the eastern Arctic Ocean. Journal of Geophysical Research
109, C12030. http://dx.doi.org/10.2029/2003JC002223.
Killworth, P.D., Smith, J.M., 1984. A one-and-a-half dimensional model of the Arctic
Halocline. Deep-Sea Research 31, 271293.
Kinnard, C., Zdanowicz, C.M., Koerner, R.M., Fisher, D.A., 2008. A changing Arctic
seasonal ice zone: Observations from 18702003 and possible oceanographic
consequences. Geophysical Research Letters 35 (2), L02507.
Kinnard, C., Zdanowicz, C.M., Fisher, D.A., Isaksson, E., de Vernal, A., Thompson, L.G.,
2011. Reconstructed changes in Arctic sea ice over the past 1,450 years. Nature
479 (7374), 509512.
Klages, M., Boetius, A., Christensen, J.P., Deubel, H., Piepenburg, D., Schewe, I.,
Soltwedel, T., 2004. The benthos of Arctic seas and its role for the organic carbon
cycle at the seaoor. In: Stein, R., Macdonald, R.W. (Eds.), The organic carbon
cycle in the Arctic Ocean. Springer, Heidelberg, pp. 139167.
Korhonen, M., Rudels, B., Marnela, M., Wisotzki, A., Zhao, J.P., 2012. Time and space
variability of freshwater content, heat content and seasonal ice melt in the
Arctic Ocean from 1991 to 2011. Ocean Science Discussions 9, 26212677.
http://dx.doi.org/10.5194/osd-9-2621-2012.
Koltun, V.M., 1964. About the study of the bottom fauna of the Greenland Sea and
the central part of the Arctic basin. Proceedings of the Arctic and Antarctic
Institute 259, 1378 (in Russian).
Kosobokova, K.N., 1982. Composition and distribution of the biomass of
zooplankton in the central Arctic Basin. Oceanology 22, 744750.
Kosobokova, K.N., 1983. Seasonal variations of the vertical distribution and age
composition of Microcalanus pygmaeus, Oithona similis, Oncaea borealis and O.
notopus populations in the central Arctic basin. In: Vinogradov, M.E. (Ed.),
Biologiya Tsentralnogo Arkicheskogo basseyna (Biology of the central Arctic
Basin). Nauka, Moscow, 1980, Canadian Translation Fisheries and Aquatic
Science 4926, pp. 77108.
Kosobokova, K.N., 1989. Vertical distribution of plankton animals in the eastern part
of the central Arctic Basin. Explorations of the Fauna of the Seas, Marine
Plankton, Leningrad 41 (49), 2431 (in Russian).

117

Kosobokova, K.N., 2012. Zooplankton of the Arctic Ocean: Community structure,


Ecology, Spatial distribution. GEOS, Moscow, 272p (in Russian).
Kosobokova, K., Hirche, H.-J., 2000. Zooplankton distribution across the Lomonosov
Ridge, Arctic Ocean: species inventory, biomass and vertical structure. Deep-Sea
Research I 47, 20292060.
Kosobokova, K.N., Hirche, H.J., 2001. Reproduction of Calanus glacialis in the Laptev
Sea, Arctic Ocean. Polar Biology 24, 3343.
Kosobokova, K.N., Hanssen, H., Hirche, H.J., Knickmeier, K., 1998. Composition and
distribution of zooplankton in the Laptev Sea and adjacent Nansen Basin during
summer, 1993. Polar Biology 19, 6376.
Kosobokova, K.N., Hirche, H.-J., Scherzinger, T., 2002. Feeding ecology of
Spinocalanus antarcticus, a mesopelagic copepod with a looped gut. Marine
Biology 141, 503511.
Kosobokova, K.N., Hirche, H.-J., Hopcroft, R.R., 2007. Reproductive biology of deepwater calanoid copepods from the Arctic Ocean. Marine Biology 151, 919934.
Kosobokova, K., Hirche, H.-J., 2009. Biomass of zooplankton in the eastern Arctic
Ocean a base line study. Progress in Oceanography 82, 265280.
Kosobokova, K.N., Hopcroft, R.R., 2010. Diversity and vertical distribution of
mesozooplankton in the Arctics Canada Basin. Deep-Sea Research II 57, 96110.
Kosobokova, K.N., Hopcroft, R.R., Hirche, H.-J., 2011. Patterns of zooplankton
diversity through the depths of the Arctics central basins. Marine Biodiversity
41, 2950.
Krisheld, R.A., Proshutinsky, A., Tateyama, K., Williams, W.J., Carmack, E.C.,
McLaughlin, F.A., Timmermans, M.L., 2014. Deterioration of perennial sea ice
in the Beaufort Gyre from 2003 to 2012 and its impact on the oceanic
freshwater cycle. Journal of Geophysical Research: Oceans 1192, 12711305.
Kulikov, E.A., Rabinovich, A.B., Carmack, E.C., 2010. Variability of baroclinic tidal
currents on the Mackenzie Shelf, the Southeastern Beaufort Sea. Continental
Shelf Research 30, 656667.
Kwok, R., Cunningham, G.F., Wensnahan, M., Rigor, I., Zwally, H.J., Yi, D., 2009.
Thinning and volume loss of the Arctic Ocean sea ice cover: 20032008. Journal
of Geophysical Research 114 (C7). http://dx.doi.org/10.1029/2009jc005312.
Kwok, R., Spreen, G., Pang, S., 2013. Arctic sea ice circulation and drift speed:
Decadal trends and ocean currents. Journal of Geophysical Research 118, 2408
2425.
Laakmann, S., Stump, S.M., Auel, H., 2009. Vertical distribution and dietary
preferences of deep-sea copepods (Euchaetidae and Aetideidae; Calanoida) in
the vicinity of the Antarctic Polar Front. Polar Biology 32, 679689.
Lalande, C., Forest, A., Barber, D.G., Gratton, Y., Fortier, L., 2009. Variability in the
annual cycle of vertical particulate organic carbon export on Arctic shelves:
contrasting the Laptev Sea, Northern Bafn Bay and the Beaufort Sea.
Continental Shelf Research 29, 21572165.
Lalande, C., Bauerfeind, E., Nthig, E.-M., Beszczynska-Mller, A., 2013. Impact of a
warm anomaly on export uxes of biogenic matter in the eastern Fram Strait.
Progress in Oceanography 1090, 7077.
Langseth, M.G., Lachenbruch, A.H., Marshall, B.V., 1990. Geothermal observations in
the Arctic region. The Geology of North America 50, 133151.
Laney, S.R., Krisheld, R.A., Toole, J.M., Hammar, T.R., Ashjian, C.J., Timmermans,
M.L., 2014. Assessing algal biomass and bio-optical distributions in perennially
ice-covered polar ocean ecosystems. Polar Science 8, 7385.
Lee, S., Stockwell, D., Whitledge, T., 2010. Uptake rates of dissolved inorganic
carbon and nitrogen by under-ice phytoplankton in the Canada Basin in
summer 2005. Polar Biology 33, 10271036. http://dx.doi.org/10.1007/
s00300-010-0781-4.
Lee, S.H., McRoy, C.P., Joo, H.M., Gradinger, R., Cui, X.H., Yun, M.S., Chung, K.H., Kang,
S.-H., Kang, C.-K., Choy, E.J., Son, S.H., Carmack, E.C., Whitledge, T.E., 2011. Holes
in progressively thinning Arctic sea ice lead to new ice algae habitat.
Oceanography 24, 3020308.
Lee, S.H., Stockwell, D.A., Joo, H.M., Son, Y.B., Kang, C.K., Whitledge, T.E., 2012.
Phytoplankton production from melting ponds on Arctic sea ice. Journal of
Geophysical Research 117, C04030. http://dx.doi.org/10.1029/2011JC007717.
Legendre, L., Ackley, S.F., Dieckmann, G.S., Gulliksen, B., Horner, R., Hoshiai, T.,
Melnikov, I.A., Reeburgh, W.S., Spindler, M., Sullivan, C.W., 1992. Ecology of sea
ice biota. Polar Biology 1234, 429444.
Leu, E., Sreide, J.E., Hessen, D.O., Falk-Petersen, S., Berge, J., 2011. Consequences of
changing sea-ice cover for primary and secondary producers in the European
Arctic shelf seas: timing, quantity, and quality. Progress in Oceanography 901,
1832.
Leu, E., Mundy, C.J., Assmy, P., Campbell, K., Gabrielsen, T.M., Gosselin, M., JuulPedersen, T., Gradinger, R., 2015. Arctic spring awakening steering principles
behind the phenology of vernal ice algae blooms. Progress in Oceanography
139, 151170.
Levin, L.A., Gooday, A., 2003. The Deep Atlantic Ocean. In: Tyler, P.A. (Ed.),
Ecosystems of the World: The Deep Sea. Elsevier, Amsterdam, pp. 111178.
Li, W.K.W., McLaughlin, F.A., Lovejoy, C., Carmack, E.C., 2009. Smallest algae thrive
as the Arctic Ocean freshens. Science 326 (5952), 539. http://dx.doi.org/
10.1126/science.1179798.
Lien, V.S., Vikeb, F.B., Skagseth, ., 2013. One mechanism contributing to covariability of the Atlantic inow branches to the Arctic. Nature Communications
4, 1488.
Loeng, H., Brander, K., Carmack, E., Denisenko, S., Drinkwater K., Hansen B., Kovacs
K., Livingston P., McLaughlin F., Sakshaug E., 2004. Marine Systems, Chapter 9,
454538. In: Arctic Climate Impact Assessment, Arctic Council, 1042 p.
Logerwell, E., Rand, K., Weingartner, T.J., 2011. Oceanographic characteristics of the
habitat of benthic sh and invertebrates in the Beaufort Sea. Polar Biology 34,
17831796. http://dx.doi.org/10.1007/s00300-011-1028-8.

118

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Loseto, L.L., Stern, G.A., Connelly, T.L., Deibel, D., Gemmill, B., Prokopowicz, A.,
Fortier, L., Ferguson, S.H., 2009. Summer diet of beluga whales inferred by fatty
acid analysis of the eastern Beaufort Sea food web. Journal of Experimental
Marine Biology and Ecology 374, 1218.
MacDonald, I.R., Bluhm, B.A., Iken, K., Gagaev, S., Strong, S., 2010a. Benthic
macrofaunal and megafaunal assemblages in the Arctic deep-sea Canada
Basin. Deep-Sea Research II 57, 136152.
Macdonald, R.W., Carmack, E.C., 1991. Age of Canada Basin Deep Waters: A way to
estimate primary production for the Arctic Ocean. Science 254, 13481350.
Macdonald, R.W., Carmack, E.C., Wallace, D.W.R., 1993. Tritium and radiocarbon
dating of Canada basin deep waters. Science 259, 103104.
Macdonald, R.W., Carmack, E.C., McLaughlin, F.A., Falkner, K.K., Swift, J.H., 1999.
Connections among ice, runoff and atmospheric forcing in the Beaufort Gyre.
Geophysical Research Letters 26, 22232226.
Macdonald, R.W., Anderson, L.G., Christensen, J.P., Miller, L.A., Semiletov, I.P., Stein,
R., 2010b. Polar margins: the Arctic Ocean. In: Liu, K.K., Atkinson, L., Quiones,
R., Talaue-McManu, L. (Eds.), Carbon and Nutrient Fluxes in Continental
Margins: A Global Synthesis. Springer, New York, pp. 291303.
Manley, T.O., Hunkins, K., 1985. Mesoscale eddies of the Arctic Ocean. Journal of
Geophysical
Research
90,
49114930.
http://dx.doi.org/10.1029/
JC090iC03p04911.
Markhaseva, E.L., 1998. New species of the genus Xanthocalanus (Copepoda,
Calanoida, Phaennidae) from the Laptev Sea. Journal of Marine Systems 15,
413419.
Markhaseva, E.L., 2002. Phaennocalanus unispinosus (Copepoda, Calanoida,
Phaennidae): new genus, and new species from the bathypelagial Arctic
Basin. Sarsia 87, 312318.
Markhaseva, E.L., Kosobokova, K.N., 1998. New and rare species of calanoid
copepods from the central Arctic Basin (Crustacea, Copepoda). Zoosystematica
Rossica 7, 4553.
Markhaseva, E.L., Kosobokova, K.N., 2001. Arctokonstantinus hardingi (Copepoda,
Calanoida: Arctokonstantinidae): new family, new genus and new species of
from the bathypelagial Arctic Basin. Sarsia 86, 319324.
Marincovich Jr., L., Brouwers, E.M., Hopkins, D.M., McKenna, M.C., 1990. Late
Mesozoic and Cenozoic paleogeographic and paleoclimatic history of the Arctic
Ocean Basin, based on shallow-water faunas and terrestrial vertebrates. In:
Geological Society of America, The Geology of North America, v. L, The Arctic
Ocean Region, pp. 403426.
Matrai, P., Apollonio, S., 2013. New estimates of microalgae production based upon
nitrate reductions under sea ice in Canadian shelf seas and the Canada Basin of
the Arctic Ocean. Marine Biology 160, 12971309.
Matrai, P.A., Olson, E., Suttles, S., Hill, V., Codispoti, L.A., Light, B., Steele, M., 2013.
Synthesis of primary production in the Arctic Ocean: I. Surface waters, 1954
2007. Progress in Oceanography 110, 93106. http://dx.doi.org/10.1016/
j.pocean.2012.11.004.
Mathis, J.T., Pickart, R.S., Hansell, D.A., Kadko, D., Bates, N.R., 2007. Eddy transport of
organic carbon and nutrients from the Chukchi Shelf: impact on the upper
halocline of the western Arctic Ocean. Journal of Geophysical Research 112,
C05011. http://dx.doi.org/10.1029/2006JC003899.
Mayer, L.A., Armstrong, A.A., Calder, B.R., Gardner, J.V., 2010. Seaoor mapping in
the Arctic: support for a potential US extended continental shelf. International
Hydrographic Review 3, 1423.
McClelland, J.W., Dery, S.J., Peterson, J., Holmes, R.M., Woods, E.F., 2006. A pan-arctic
evaluation of changes in river discharge during the latter half of the 20th
century. Geophysical Research Letters 33, L06715. http://dx.doi.org/10.01029/
2006GL025753.
McLaughlin, F.A., Carmack, E.C., 2010. Nutricline deepening in the Canada Basin,
20032009. Geophysical Research Letters 37, L24602. http://dx.doi.org/
10.1029/2010GL045459.
McLaughlin, F.A., Carmack, E.C., Macdonald, R.W., Bishop, J.K.B., 1996. Physical and
geochemical properties across the Atlantic/Pacic water mass front in the
southern Canadian Basin. Journal of Geophysical Research 101 (C1), 11831197.
http://dx.doi.org/10.1029/95JC02634.
McLaughlin, F., Carmack, E., Macdonald, R., Weaver, A.J., Smith, J., 2002. The Canada
Basin, 19891995: upstream events and far-eld effects of the Barents Sea.
Journal of Geophysical Research 107. http://dx.doi.org/10.1029/2001JC000904.
McLaughlin, F.A., Carmack, E.C., Macdonald, R.W., Melling, H., Swift, J.H., Wheeler, P.A.,
Sherr, B.F., Sherr, E.B., 2004. The joint roles of Pacic and Atlantic-origin waters in
the Canada Basin, 19971998. Deep-Sea Research I: Oceanographic Research
Papers 51 (1), 107128. http://dx.doi.org/10.1016/j.dsr.2003.09.010.
McLaughlin, F.A., Carmack, E.C., Williams, W.J., Zimmermann, S., Shimada, K., Itoh,
M., 2009. Joint effects of boundary currents and thermohaline intrusions on the
warming of Atlantic water in the Canada Basin, 19932007. Journal of
Geophysical Research 114. http://dx.doi.org/10.1029/2008jc005001.
McPhee, M.G., Proshutinsky, A., Morison, J.H., Steele, M., Alkire, M.B., 2009. Rapid
change in freshwater content of the Arctic Ocean. Geophysical Research Letters
36 (10), L10602.
Melling, H., Lewis, E.L., 1982. Shelf drainage ows in the Beaufort Sea and their
effect on the Arctic pycnocline. Deep-Sea Research 29, 967989.
Melling, H., Agnew, T.A., Falkner, K.K., Greenberg, D.A., Lee, C.M., Mnchow, A.,
Petrie, B., Prinsenberg, S.J., Samelson, R.M., Woodgate, R.A., 2008. Fresh-water
uxes via Pacic and Arctic outows across the Canadian polar shelf. In:
Dickson, R.R., Meinke, J., Rhines, P. (Eds.), Arctic-Subarctic Ocean Fluxes.
Springer, Netherlands, pp. 193247.

de Middag, R., Baar, H.J.W., Laan, P., Bakker, K., 2009. Dissolved aluminum and the
silicon cycle in the Arctic Ocean. Marine Chemistry 115, 176195.
Minoda, T., 1967. Seasonal distribution of Copepoda in the Arctic Ocean from June
to December, 1964. Records of Oceanographic Works in Japan 9, 161168.
Mironov, A.N., Dilman, A.B., Krylova, E.M., 2013. Global distribution patterns of
genera occurring in the Arctic Ocean deeper than 2000 m. Invertebrate Zoology
10 (1), 167194.
Mountain, D.G., Coachman, L.K., Aagaard, K., 1976. On the ow through Barrow
Canyon. Journal of Physical Oceanography 6, 461470.
Morison, J., Steele, M., Andersen, R., 1998. Hydrography of the upper Arctic Ocean
measured from the nuclear submarine USS Pargo. Deep-Sea Research Part I 451,
1538.
Morison, J., Steele, M., Kikuchi, T., Falkner, K., Smethie, W., 2006. Relaxation of
central Arctic Ocean hydrography to pre-1990s climatology. Geophysical
Research Letters 33, L17604. http://dx.doi.org/10.1029/2006GL026826.
Morison, J., Kwok, R., Peralta-Ferriz, C., Alkire, M., Rigor, I., Andersen, R., Steele, M.,
2012. Changing Arctic Ocean freshwater pathways. Nature 481, 6670.
Mueter, F.J., Litzow, M.A., 2008. Sea ice retreat alters the biogeography of the Bering
Sea continental shelf. Ecological Applications 18 (2), 309320.
Mumm, N., Auel, H., Hanssen, H., Hagen, W., Richter, C., Hirche, H.J., 1998. Breaking
the ice: large-scale distribution of mesozooplankton after a decade of Arctic and
transpolar cruises. Polar Biology 20, 189197.
Nansen, F., 1902. Oceanography of the North Polar Basin. Scientic results.
Norwegian North Polar Expedition 3 (9), 18931896.
Nelson, R.J., Carmack, E.C., McLaughlin, F.A., Cooper, G.A., 2009. Penetration of
Pacic zooplankton into the western Arctic Ocean tracked with molecular
population genetics. Marine Ecology Progress Series 381, 129138.
Nelson, R.J., Ashjian, C., Bluhm, B., Conlan, K., Gradinger, R., Grebmeier, J., Hill, V.,
Hopcroft, R., Hunt, B., Joo, H., Kirchman, D., Kosobokova, K., Lee, S., Li, W.,
Lovejoy, C., Poulin, M., Sherr, E., Young, K., 2014. Biodiversity and biogeography
of the lower trophic fauna of the Pacic Arctic region sensitivities to climate
change. In: Grebmeier, J.M., Maslowski, W., Zhao, J. (Eds.), The Pacic Arctic
sector: ecosystem status and trends in a rapidly changing environment.
Springer, Dordrecht, pp. 269336.
Nesis, K.N., 1984. A hypothesis on the origin of western and eastern Arctic
distribution of areas of marine bottom animals. Soviet Journal of Marine Biology
9, 235243.
Newton, J.L., Aagaard, K., Coachman, L.K., 1974. Baroclinic eddies in the Arctic
Ocean. Deep Sea Research and Oceanographic Abstracts 1 (9), 707719.
Newton, R., Schlosser, P., Martinson, D.G., Maslowski, W., 2008. Freshwater
distribution in the Arctic Ocean: Simulation with a high resolution model and
model-data comparison. Journal of Geophysical Research 113, C05024. http://
dx.doi.org/10.1029/2007JC004111.
Nitishinsky, M., Anderson, L.G., Hlemann, J.A., 2007. Inorganic carbon and nutrient
uxes on the Arctic Shelf. Contintental Shelf Research 27, 15841599. http://
dx.doi.org/10.1016/j.csr.2007.01.019.
Nishino, S., Itoh, M., Williams, W.J., Semiletov, I., 2013. Shoaling of the nutricline
with an increase in near-freezing temperature water in the Makarov Basin.
Journal of Geophysical Research 118. http://dx.doi.org/10.1029/2012/JC008234.
Nrgaard-Pedersen, N., Spielhagen, R.F., Thiede, J., Kassens, H., 1998. Central Arctic
surface ocean environment during the past 80,000 years. Paleoceanography 13,
193204.
OBrien, M.C., Melling, H., Pedersen, T.F., Macdonald, R., 2011. The role of eddies and
energetic ocean phenomena in the transport of sediment from shelf to basin in
the Arctic. Journal of Geophysical Research 116, C08001. http://dx.doi.org/
10.1029/2010JC006890.
Olli, K., Wassmann, P., Reigstad, M., Ratkova, T.N., Arashkevich, E., Pasternak, A.,
Matrai, P., Knulst, J., Tranvik, L., Klais, R., Jakobsen, A., 2006. The fate of
production in the central Arctic Ocean top-down regulation by zooplankton
expatriates. Progress in Oceanography 72, 84113.
stlund, H.G., Possnert, G., Swift, J.H., 1987. Ventilation rate of the deep Arctic
Ocean from carbon 14 data. Journal of Geophysical Research: Oceans 92 (C4),
37693777.
Overland, J.E., Wood, K.R., Wang, M., 2011. Warm Arctic-cold continents: Climate
impacts of the newly open Arctic Sea. Polar Research 30. http://dx.doi.org/
10.3402/polar.v30i0.15787.
Overpeck, J., Hughen, K., Hardy, D., Bradley, R., Case, R., Douglas, M., Finney, B., et al.,
1997. Arctic environmental change of the last four centuries. Science 278
(5341), 12511256.
Pabi, S., van Dijken, G.L., Arrigo, K.R., 2008. Primary production in the Arctic Ocean,
19982006. Journal of Geophysical Research 113, C08005. http://dx.doi.org/
10.1029/2007jc004578.
Padman, L., 1995. Small-scale Physical Processes in the Arctic Ocean. Marginal Ice
Zones and Contintal Shelves, Arctic Oceanography, pp. 97129.
Palumbi, S.R., 2003. Population genetics, demographic connectivity, and the design
of marine reserves. Ecological Applications 13, S146S158.
Parker-Stetter, S.L., Horne, J.K., Weingartner, T.J., 2011. Distribution of polar cod and
age-0 sh in the U.S. Beaufort Sea. Polar Biology 43, 15431557. http://
dx.doi.org/10.1007/s00300-011-1014-1.
Parsons, T.R., Lalli, C.M., 2002. Jellysh population explosions: revisiting a
hypothesis of possible causes. La mer, 40, 111121, Tokyo: Societe francojaponaise d oceanographie.
Pautzke, C.G., 1979. Phytoplankton primary production below the Arctic Ocean pack
ice, an ecosystems analysis. Ph.D. Thesis, University of Washington.

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121


Perkin, R.G., Lewis, E.L., 1984. Mixing in the West Spitsbergen current. Journal of
Physical Oceanography 148, 13151325.
Perovich, D.K., Richter-Menge, J.A., Jones, K.F., Light, B., 2008. Sunlight, water, and
ice: Extreme Arctic sea ice melt during the summer of 2007. Geophysical
Research Letters 35, L11501. http://dx.doi.org/10.1029/2008GL034007.
Perovich, D., Gerland, S., Hendricks, S., Meier, W., Nikolaus, M., Tschudi, M., 2014.
Sea ice. In: Arctic report card, <http://www.arctic.noaa.gov/reportcard>.
Perry, A.L., Low, P.J., Ellis, J.R., Reynolds, J.D., 2005. Climate change and distribution
shifts in marine shes. Science 308 (5730), 19121915.
Peterson, B.J., Holmes, R.M., McClelland, J.W., Vorosmarty, C.J., Lammers, R.B.,
Shiklomanov, A.I., Shiklomanov, I.A., Rahmstorf, S., 2002. Increasing river
discharge to the Arctic Ocean. Science 298, 21712173.
Peterson, B.J., McClelland, J., Curry, R., Holmes, R.M., Walsh, J.E., Aagaard, K., 2006.
Trajectory shifts in the Arctic and subarctic freshwater cycle. Science 313,
10611066.
Pfannkuche, O., Soltwedel, T., 1998. Small benthic size classes along the N.W.
European Continental Margin: spatial and temporal variability in activity and
biomass. Progress in Oceanography 42, 189207.
Prman, S., Haxby, W.F., Colony, R., Rigor, I., 2004. Variability in Arctic sea ice drift.
Geophysical Research Letters 31, L16402. http://dx.doi.org/10.1029/
2004GL020063.
Pickart, R.S., Pratt, L.J., Zimmermann, S., Torres, D.J., 2005. Flow of wintertransformed water into the western Arctic. Deep-Sea Research II 52, 3175
3198.
Pickart, R.S., Pratt, L.J., Torres, D.J., Whitledge, T.E., Proshutinsky, A.Y., Aagaard, K.,
Agnew, T.A., Moore, G.W.K., Dail, H.J., 2010. Evolution and dynamics of the ow
through Herald Canyon in the western Chukchi Sea. Deep-Sea Research II 57 (1),
526.
Pickart, R.S., Spall, M.A., Mathis, J.T., 2013. Dynamics of upwelling in the Alaskan
Beaufort Sea and associated shelfbasin uxes. Deep-Sea Research I 176, 3551.
Piepenburg, D., Archambault, P., Ambrose Jr., W.G., Blanchard, A., Bluhm, B.A.,
Carroll, M.L., Conlan, K., Cusson, M., Feder, H.M., Grebmeier, J.M., Jewett, S.C.,
Lvesque, M., Petryashev, V.V., Sejr, M.K., Sirenko, B., Wodarska-Kowalczuk, M.,
2011. Towards a pan-Arctic inventory of the species diversity of the macro- and
megabenthic fauna of the Arctic shelf seas. Marine Biodiversity 41, 5170.
Polyakov, I.V., Beszczynska, A., Carmack, E., Dmitrenko, I., Fahrbach, E., Frolov, I.E.,
Gerdes, R., Hansen, E., Holfort, J., Ivanov, V.V., Johnson, M.A., Karcher, M., Kauker,
F., Morison, J., Orvik, K.A., Schauer, U., Simmons, H.L., Skagseth, ., Sokolov, V.T.,
Steele, M., Timokhov, L.A., Walsh, D., Walsh, J.E., 2005. One more step toward a
warmer Arctic. Geophysical Research Letters 32, L17605. http://dx.doi.org/
10.1029/2005GL023740.
Polyakov, I.V., Timokhov, L.A., Alexeev, V.A., Bacon, S., Dmitrenko, I.A., Fortier, L.,
Frolov, I.E., Gascard, J-.C., Hansen, E., Ivanov, V.V., Laxon, S., Mauritzen, C.,
Perovich, D., Shimada, K., Simmons, H.L., Sokolov, V.T., Steele, M., Toole, J., 2010.
Arctic Ocean warming contributes to reduced polar ice cap. Journal of Physical
Oceanography 40, 27432756.
Polyakov, I.V., Alekseev, V.A., Ashik, I.M., Bacon, S., Beszczynska-Mller, A.,
Dmitrenko, I., Fortier, L., Gascard, J.-C., Hansen, E., Hlemann, J., Ivanov, V.V.,
Kikuchi, T., Krillov, S., Lenn, Y.-D., Piechura, J., Repina, I., Timokhov, L.A.,
Walczowski, W., Woodgate, R., 2011. Fate of the early- 2000s Arctic warm pulse.
Bulletin of the American Meteorological Society 92, 561566. http://dx.doi.org/
10.1175/2010BAMS2921.1.
Polyakov, I.V., Pnyushkov, A.V., Rember, R., Ivanov, V.V., Lenn, Y.D., Padman, L.,
Carmack, E.C., 2012. Mooring-based observations of double-diffusive staircases
over the Laptev Sea slope. Journal of Physical Oceanography 42 (1), 95109.
Polyakov, I.V., Bhatt, U.S., Walsh, J.E., Abrahamsen, E.P., Pnyushkov, A.V., Wassmann,
P.F., 2013. Recent oceanic changes in the Arctic in the context of long-term
observations. Ecological Applications 23 (8), 17451764.
Pomerleau, C., Nelson, R.J., Hunt, B.P.V., Sastri, A.R., Williams, W.J., 2014. Spatial
patterns in zooplankton communities and stable isotope ratios (d13C and d15N)
in relation to oceanographic conditions in the sub-Arctic Pacic and western
Arctic regions during the summer of 2008. Journal of Plankton Research 36 (3),
757775.
Popova, E.E., Yool, A., Coward, A.C., Dupont, F., Deal, C., Elliott, S., Hunke, E., Jin, M.,
Steele, M., Zhang, J., 2012. What controls primary production in the Arctic
Ocean? Results from an intercomparison of ve general circulation models with
biogeochemistry. Journal of Geophysical Research: Oceans 117. http://
dx.doi.org/10.1029/2011jc007112.
Popova, E.E., Yool, A., Aksenov, Y., Coward, A.C., 2013. Role of advection in Arctic
Ocean lower trophic dynamics: a modelling perspective. Journal of Geophysical
Research 118 (3), 15711586. http://dx.doi.org/10.1002/jgrc.20126.
Poulin, M., Daugbjerg, N., Gradinger, R., Ilyash, L., Ratkova, T., von Quillfeldt, C.,
2011. The pan-Arctic biodiversity of marine pelagic and sea-ice unicellular
eukaryotes: a rst-attempt assessment. Marine Biodiversity 41, 1328.
Proshutinsky, A.Y., Johnson, M.A., 1997. Two circulation regimes of the wind driven
Arctic Ocean. Journal of Geophysical Research: Oceans 102 (C6), 1249312514.
http://dx.doi.org/10.1029/97jc00738.
Proshutinsky, A., Bourke, R.H., McLaughlin, F.A., 2002. The role of the Beaufort Gyre
in Arctic climate variability: seasonal to decadal climate scales. Geophysical
Research Letters 29 (23), 21002105. http://dx.doi.org/10.1029/2002GL015847.
Proshutinsky, A., Krisheld, R., Timmermans, M.-L., Toole, J., Carmack, E.,
McLaughlin, F., Williams, W.J., Zimmermann, S., Itoh, M., Shimada, K., 2009.
Beaufort Gyre freshwater reservoir: state and variability from observations.
Journal of Geophysical Research 114. http://dx.doi.org/10.1029/2008jc005104.
Quadfasel, D.A., Sy, A., Wells, D., Tunik, A., 1991. Warming in the Arctic. Nature 350,
385.

119

Quakenbush, L.T., Suydam, R.S., Bryan, A.L., Lowry, L.F., Frost, K., Mahoney, B.A,
2015. Diet of beluga whales (Delphinapterus leucas) in Alaska from stomach
contents, MarchNovember, in press.
Rabe, B., Karcher, M., Kauker, F., Schauer, U., Toole, J.M., Krisheld, R.A., Pisarev, S.,
Kikuchi, T., Su, J., 2014. Arctic Ocean basin liquid freshwater storage trend
19912012. Geophysical Research Letters 41, 961968. http://dx.doi.org/
10.1002/2013GL058121.
Rampal, P., Weiss, J., Dubois, C., Campin, J.M., 2011. IPCC climate models do not
capture Arctic sea ice drift acceleration: consequences in terms of projected sea
ice thinning and decline. Journal of Geophysical Research 116 (C8), C00D07.
Raskoff, K.A., Hopcroft, R.R., Kosobokova, K.N., Purcell, J.E., Youngbluth, M., 2010.
Jellies under ice: ROV observations from the Arctic 2005 Hidden Ocean
expedition. Deep-Sea Research II 57, 111126.
Ravelo, A.M., Konar, B.H., Bluhm, B.A., 2015. Spatial variability in epibenthic
communities on the Alaskan Beaufort Sea shelf. Polar Biology. http://dx.doi.org/
10.1007/s00300-015-1741-9.
Rex, M., Salawitch, R.J., Deckelmann, H., von der Gathen, P., Harris, N.R.P.,
Chippereld, M.P., Naujokat, B., Reimer, E., Allaart, M., Andersen, S.B.,
Bevilacqua, R., Braathen, G.O., Claude, H., Davies, J., De Backer, H., Dier, H.,
Dorokhov, V., Fast, H., Gerding, M., Godin-Beekmann, S., Hoppel, K., Johnson, B.,
Kyro, E., Litynska, Z., Moore, D., Nakane, H., Parrondo, M.C., Risley, A.D.,
Skrivankova, P., Stubi, R., Viatte, P., Yushkov, V., Zerefos, C., 2006. Arctic winter
2005: Implications for stratospheric ozone loss and climate change.
Geophysical Research Letters 33, L23808. http://dx.doi.org/10.1029/
2006GL026731.
Rex, M.A., Etter, R.J., 2010. Deep-sea Biodiversity: Pattern and Scale. Harvard
University Press, Cambridge, MA, 368p.
Rigor, I.G., Wallace, J.M., Colony, R.L., 2002. Response of sea ice to the Arctic
Oscillation. Journal of Climate 1518, 26482663.
Rogacheva, A.V., 2007. Revision of the Arctic group of species of the family
Elipidiidae (Elasipodida, Holothuroidea). Marine Biology Research 3,
376396.
Rudels, B., 2012. Arctic Ocean circulation and variability advection and external
forcing encounter constraints and local processes. Ocean Sciences 8, 261286.
http://dx.doi.org/10.5194/os/8-261-2012.
Rudels, B., Jones, E.P., Anderson, L.G., Kattner, G., 1994. On the intermediate depth
waters of the Arctic Ocean. The Polar Oceans and Their Role in Shaping the
Global Environment: The Nansen Centennial Volume. Geophysical Monographs,
vol. 85. American Geophysical Union, pp. 3346.
Rudels, B., Anderson, L.G., Jones, E.P., 1996. Formation and evolution of the surface
mixed layer and halocline of the Arctic Ocean. Journal of Geophysical Research:
Oceans 19782012 101C4, 88078821.
Rudels, B., Friedrich, H.J., Quadfasel, D., 1999. The Arctic circumpolar boundary
current. Deep-Sea Research II 46, 10231062.
Rudels, B., Muench, R.D., Gunn, J., Schauer, U., Friedrich, H.J., 2000. Evolution of the
Arctic boundary current north of the Siberian Shelves. Journal of Marine
Systems 25, 7799.
Rudels, B., Anderson, L., Eriksson, P., Fahrbach, E, Jakobsson, M., Jones, E.P., Melling,
H., Prinsenberg, S., Schauer, U., Yao, T., 2012. In: Lemke, P., Jacobi, H.-W. (Eds.),
Observations in the ocean, in Arctic Climate Change: The ACSYS Decade and
Beyond, Atmospheric and Oceanographic Sciences Library, vol. 43, Chap. 4,
Springer, Dordrecht, Netherlands, pp. 117198.
Rudels, B., Schauer, U., Bjrk, G., Korhonen, M., Pisarev, S., Rabe, B., Wisotzki, A.,
2013. Observations of water masses and circulation in the Eurasian Basin of the
Arctic Ocean from the 1990s to the late 2000s. OS Special Issue: IceAtmosphere-Ocean interactions in the Arctic Ocean during IPY: The Damocles
Project 9 (1), 147169. http://dx.doi.org/10.5194/os-9-147-2013.
Runge, J.A., Ingram, R.G., 1988. Under ice grazing by planktonic, calanoid copepods
in relation to a bloom of ice microalgae in southeastern Hudson Bay. Limnology
and Oceanography 33, 280286.
Sandstrm, J.W., 1919. The Hydrodynamics of the Canadian Atlantic Waters. In:
Hjort, J.H. (Ed.), Canadian Fisheries Expedition, 19141915: Investigations in
the Gulf of St Lawrence and Atlantic Waters of Canada, Kings Press, Ottawa, pp.
221248.
Sakshaug, E., 2004. Primary and secondary production in the Arctic seas. In: Stein,
R., Macdonald, R.W. (Eds.), The Organic Carbon Cycle in the Arctic Ocean.
Springer, New York, pp. 5781.
Sakshaug, E., Slagstad, D., 1991. Light and productivity of phytoplankton in polar
marine ecosystems: a physiological view. Polar Research 10 (1), 6986.
Savin, S.M., Douglas, R.C., Stehli, F.G., 1975. Tertiary marine paleo-temperatures.
Geological Society American Bulletin 86, 14991510.
Schauer, U., Muench, R.D., Rudels, B., Timokhov, L., 1997. Impact of eastern shelf
waters on the Nansen Basin intermediate layers. Journal of Geophysical
Research 102 (C2), 33713382.
Schauer, U., Rudels, B., Jones, E.P., Anderson, L.G., Muench, R.D., Bjrk, G., Swift, J.H.,
Ivanov, V., Larsson, A.M., 2002. Conuence and redistribution of Atlantic water
in the Nansen, Amundsen and Makarov basins. Annales Geophysicae 20, 257
273.
Schauer, U., Fahrbach, E., Osterhus, S., Rohardt, G., 2004. Arctic warming through the
Fram Strait: Oceanic heat transport from 3 years of measurements. Journal of
Geophysical Research 109, C06026. http://dx.doi.org/10.1029/2003JC001823.
Schauer, U., Beszczynska-Mller, A., Walczowski, W., Fahrbach, E., Piechura, J.,
Hansen, E., 2008. Variation of measured heat ow through the Fram Strait
between 1997 and 2006. In: Dickson, R.R., Meincke, J.M., Rhines, P. (Eds.),
Arctic-Subarctic Ocean Fluxes Dening the role of the Northern Seas in
Climate. Springer, Netherlands, pp. 6585.

120

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121

Schewe, I., 2001. Small-sized benthic organisms of the Alpha Ridge, Central Arctic
Ocean. International Review of Hydrobiology 86, 317335.
Schewe, I., Soltwedel, T., 2003. Benthic response to ice-edge-induced particle ux in
the Arctic Ocean. Polar Biology 26, 610620.
Schlosser, P., Kromer, B., Ekwurzel, B., Bnisch, G., McNichol, A., Schneider, R., von
Reden, K., Oestlund, H.G., Swift, J.H., 1997. The rst trans-Arctic 14C section:
comparison of the mean ages of the deep waters in the Eurasian and Canadian
basins of the Arctic Ocean. Nuclear Instruments and Methods in Physics Research
Section B: Beam Interactions with Materials and Atoms 123 (1), 431437.
Serreze, M.C., Barrett, A.P., Slater, A.G., Woodgate, R.A., Aagaard, K., Lammers, R.B.,
Steele, M., Moritz, R., Meredith, M., Lee, C.M., 2006. The large-scale freshwater
cycle of the Arctic. Journal of Geophysical Research: Oceans 111 (C11), C11010.
http://dx.doi.org/10.1029/2005JC003424.
Sharma, J., Bluhm, B.A., 2011. Diversity of larger free-living nematodes from
macrobenthos (>250 lm) in the Arctic deep-sea Canada Basin. Marine
Biodiversity 41 (3), 455465.
Shimada, K., Carmack, E.C., Hatakeyama, K., Takizawa, T., 2001. Varieties of shallow
temperature maximum waters in the western Canadian Basin of the Arctic
Ocean. Geophysical Research Letters 2818, 34413444.
Shimada, K., McLaughlin, F., Carmack, E., Proshutinsky, A., Nishino, S., Itoh, M., 2004.
Penetration of the 1990s warm temperature anomaly of Atlantic Water in the
Canada Basin. Journal of Geophysical Research 31, L20301. http://dx.doi.org/
10.1029/2004GL020860.
Shimada, K., Itoh, M., Nishino, S., McLaughlin, F., Carmack, E., Proshutinsky, A., 2005.
Halocline structure in the Canada Basin of the Arctic Ocean. Geophysical
Research Letters 32 (L03605), 2004G. http://dx.doi.org/10.1029/L021358.
Shimada, K., Kamoshida, T., Itoh, M., Nishino, S., Carmack, E., McLaughlin, F.,
Zimmermann, S., Proshutinsky, A., 2006. Pacic Ocean inow: inuence on
catastrophic reduction of sea ice cover in the Arctic Ocean. Geophysical
Research Letters 33. http://dx.doi.org/10.1029/2005GL025624.
Shirshov, P.P., 1944. Scientic results of the drift of the Severnyi Polyus station.
Report delivered at the General Assembly of the USSR Academy of Sciences on
1417 Feb. 1944, Moscow, Publishing house of the Academy of Sciences of the
USSR, 110140 (in Russian).
Shirshov, P., Fedorov, E., 1938. Scientic work of the drifting north polar station.
Nature 141, 629632.
Sirenko, B.I., 2001. List of species of free-living invertebrates of Eurasian Arctic seas
and adjacent deep waters. Russian Academy of Sciences, Zoological Institute, St
Petersburg, Explorations of the Fauna of the Seas 51 (59), 131p.
Sirenko, B.I., Markhaseva, E.L., Buzhinskaya, G.N., Golikov, A.A., Menshutkina, T.V.,
Petryashov, V.V., Semenova, T.N., Stepanjants, S.D., Vassilenko, S.V., 1996.
Preliminary data on suprabenthic invertebrates collected during RV Polarstern
cruise in the Laptev Sea. Polar Biology 16, 345352.
Slagstad, D., Ellingsen, I.H., Wassmann, P., 2011. Evaluating primary and secondary
production in an Arctic Ocean void of summer sea ice: an experimental
simulation approach. Progress in Oceanography 90, 117131.
Smedsrud, L.H., Ingvaldsen, R., Nilsen, J.E., Skagseth, ., 2010. Heat in the Barents
Sea: Transport, storage, and surface uxes. Ocean Science 6 (1), 219234.
Soltwedel, T., 2000. Metazoan meiobenthos along continental margins: a review.
Progress in Oceanography 46, 5984.
Sreide, J.E., Leu, E.V.A., Berge, J., Graeve, M., Falk-Peteresen, S., 2010. Timing of
blooms, algal food quality and Calanus glacialis reproduction and growth in a
changing Arctic. Global Change Biology 1611, 31543163.
Spall, M.A., 2007. Circulation and water mass transformation in a model of the
Chukchi Sea. Journal of Geophysical Research: Oceans 19782012, 112C5.
Spall, M.A., Pickart, R.S., Fratantoni, P.S., Plueddemann, A.J., 2008. Western Arctic
shelfbreak eddies: Formation and transport. Journal of Physical Oceanography
388, 16441668.
Spall, M.A., Pickart, R.S., Brugler, E.T., Moore, G.W.K., Thomas, L., Arrigo, K.R., 2014.
Role of shelfbreak upwelling in the formation of a massive under-ice bloom in
the Chukchi Sea. Deep-Sea Research II 105, 1729.
Stafford, K.M., Okkonen, S.R., Clarke, J.T., 2013. Correlation of a strong Alaska Coastal
Current with the presence of beluga whales Delphinapterus leucas near Barrow,
Alaska. Marine Ecology Progress Series 474, 287297. http://dx.doi.org/
10.3354/meps10076.
Steele, M., Boyd, T., 1998. Retreat of the cold halocline layer in the Arctic Ocean.
Journal of Geophysical Research 103, 1041910435.
Steele, M., Morison, J., Ermold, W., Rigor, I., Ortmeyer, M., Shimada, K., 2004.
Circulation of summer Pacic halocline water in the Arctic Ocean. Journal of
Geophysical Research 109. http://dx.doi.org/10.1029/2003jc002009.
Stigebrandt, A., 1981. A model for the thickness and salinity of the upper layer in the
Arctic Ocean and the relationship between the ice thickness and some external
parameters. Journal of Physical Oceanography 11, 14071422.
Stroeve, J.C., Kattsov, V., Barrett, A.P., Serreze, M.C., Pavlova, T., Holland, M.M., Meier,
W.N., 2012. Trends in Arctic sea ice extent from CMIP5, CMIP3 and
observations. Geophysical Research Letters 39 (16). http://dx.doi.org/10.1029/
2012GL052676.
Suydam, R.S., Lowry, L.F., Frost, K.J., 2005. Distribution and movements of beluga
whales from the eastern Chukchi Sea stock during summer and early autumn.
OCS Study MMS 2005035 Final Report, 48 p.
Swift, J.H., Jones, E.P., Aagaard, K., Carmack, E.C., Hingston, M., Macdonald, R.W.,
McLaughlin, F.A., Perkin, R.G., 1997. Waters of the Makarov and Canada basins.
Deep-Sea Research II 44 (8), 15031529.
Syvitski, J.P.M., Farrow, G.E., Atkinson, R.J.A., Moore, P.G., Andrews, J.T., 1989. Bafn
Island fjord macrobenthos: bottom communities and environmental
signicance. Arctic 42, 232247.

Tank, S.E., Manizza, M., Holmes, R.M., McClelland, J.W., Peterson, B.J., 2012. The
processing and impact of dissolved riverine nitrogen in the Arctic Ocean.
Estuaries and Coasts 352, 401415.
Thibault, D., Head, E.J.H., Wheeler, P.A., 1999. Mesozooplankton in the Arctic Ocean
in summer. Deep-Sea Research I 46, 13911415.
Thorson, G., 1950. Reproductive and larval ecology of marine bottom invertebrates.
Biological Reviews 25, 145.
Timmermans, M.-L., Garrett, C., Carmack, E., 2003. The thermohaline structure
and evolution of the deep waters in the Canada Basin, Arctic Ocean. Deep-Sea
Research I 50, 13051321. http://dx.doi.org/10.1016/S0967-0637(03)
00125-0.
Timmermans, M.-L., Winsor, P., Whitehead, J.A., 2005. Deep-Water ow over the
Lomonosov Ridge in the Arctic Ocean. Journal of Physical Oceanography 35,
14891493.
Timmermans, M.-L., Garrett, C., 2006. Evolution of the Deep Water in the Canadian
Basin in the Arctic Ocean. Journal of Physical Oceanography 36, 866874.
Timmermans, M.-L., Toole, J., Proshutinsky, A., Krisheld, R., Plueddemann, A., 2008.
Eddies in the Canada Basin, Arctic Ocean, observed from ice-tethered prolers.
Journal of Physical Oceanography 38 (1), 133145.
Timmermans, M.L., Rainville, L., Thomas, L., Proshutinsky, A., 2010. Moored
observations of bottom-intensied motions in the deep Canada Basin, Arctic
Ocean. Journal of Marine Research 68, 625641. http://dx.doi.org/10.1357/
002224010794657137.
Timmermans, M.L., Proshutinsky, A., Krisheld, R.A., Perovich, D.K., Richter-Menge,
J.A., Stanton, T.P., Toole, J.M., 2011. Surface freshening in the Arctic Oceans
Eurasian Basin: An apparent consequence of recent change in the wind-driven
circulation. Journal of Geophysical Research: Oceans (19782012) 116 (C8).
Tremblay, J.E., Gagnon, J., 2009. The effects of irradiance and nutrient supply on the
productivity of Arctic waters: a perspective on climate change. In: Nihoul, C.J.,
Kostianoy, A.G. (Eds.), Inuence of Climate Change on the Changing Arctic and
Subarctic Conditions. Springer Science, pp. 7392.
Tremblay, G., Belzile, C., Gosselin, M., Poulin, M., Roy, S., Tremblay, J.., 2009. Late
summer phytoplankton distribution along a 3500 km transect in Canadian
Arctic waters: strong numerical dominance by picoeukaryotes. Aquatic
Microbial Ecology 541, 55.
Tremblay, J-., Blanger, S., Barber, D.G., Asplin, M., Martin, J., Darnis, G., Fortier, L.,
Gratton, Y., Link, H., Archambault, P., Sallon, A., Michel, C., Williams, W.J.,
Philippe, B., Gosselin, M., 2011. Climate forcing multiplies biological
productivity in the coastal Arctic Ocean. Geophysical Research Letters 38,
L18604. http://dx.doi.org/10.1029/2011GL048825.
Tsubouchi, T., Bacon, S., Naveira Garabato, A.C., Aksenov, Y., Laxon, S.W., Fahrbach,
E., Beszczynska-Mller, A., Hansen, E., Lee, C.M., Ingvaldsen, R.B., 2012. The
Arctic Ocean in summer: A quasi-synoptic inverse estimate of boundary uxes
and water mass transformation. Journal of Geophysical Research: Oceans
(1978-2012) 117, C1.
Ulsbo, A., Cassar, N., Korhonen, M., van Heuven, S., Hoppema, M., Kattner, G.,
Anderson, L.G., 2014. Late summer net community production in the central
Arctic Ocean using multiple approaches. Global Biogeochemical Cycles 28,
11291148, doi. 10.1002/2014GB004833.
Vancoppenolle, M., Bopp, L., Madec, G., Dunne, J., Ilyina, T., Halloran, P.R., Steiner, N.,
2013. Future Arctic Ocean primary productivity from CMIP5 simulations:
Uncertain outcome, but consistent mechanisms. Global Biogeochemical Cycles
27 (3), 605619.
Vanreusel, A., Clough, L., Jacobsen, K., Ambrose, W., Jivaluk, J., Ryheul, V., Herman, R.,
Vincx, M., 2000. Meiobenthos of the central Arctic Ocean with special emphasis
on the nematode community structure. Deep Sea Research I 47, 18551879.
Varela, D.E., Crawford, D.W., Wrohan, I.H., Wyatt, S.N., Carmack, E.C., 2013. Pelagic
primary productivity and upper ocean nutrient dynamics across arctic and
subarctic seas. Journal of Geophysical Research 118 (12), 71327152.
Vinogradov, M.Y., 1970. Vertical Distribution of Oceanic Zooplankton. Israel
Program for Scientic Translations, Jerusalem, 339 pp.
Vinogradova, N.G., 1997. Zoogeography of the abyssal and hadal zones. Advances in
Marine Biology 32, 326387.
van Oevelen, D., Bergmann, M., Soetaert, K., Bauerfeind, E., Hasemann, C., Klages, M.,
Schewe, I., Soltwedel, T., Budaeva, N.E., 2011. Carbon ows in the benthic food
web at the deep-sea observatory HAUSGARTEN (Fram Strait). Deep Sea
Research Part I 58 (11), 10691083.
Wadams, P., 2000. Ice in the Ocean, Taylor-Francis, 357p.
Walsh, J., McRoy, C.P., Coachman, L.K., Goering, J.J., Nihoul, J.J., Whitledge, T.E.,
Blackburn, T.H., Blackburn, T., Parker, P., Wirick, C., Shuert, P., 1989. Carbon and
nitrogen cycling within the Bering/Chukchi Seas: Source regions for organic
matter effecting AOU demands of the Arctic Ocean. Progress in Oceanography
22 (4), 277359.
Wang, M., Overland, J.E., 2009. A sea ice free summer within 30 years? Geophysical
Research Letters 36 (7). http://dx.doi.org/10.1029/2009GL037820.
Wassmann, P., 1998. Retention versus export food chains: processes controlling
sinking loss from marine pelagic systems. Hydrobiologia 363 (13), 2957.
http://dx.doi.org/10.1023/A:1003113403096.
Wassmann, P., Slagstad, D., Ellingsen, I., 2010. Primary production and climatic
variability in the European sector of the Arctic Ocean prior to 2007: preliminary
results. Polar Biology 33, 16411650. http://dx.doi.org/10.1007/s00300-0100839-3.
Wassmann, P.F., Duarte, C.M., Agusti, S., Sejr, M.K., 2011. Footprints of climate
change in the Arctic marine ecosystem. Global Change Biology 17, 12351249.
Wassmann, P., Lenton, T., 2012. Arctic tipping points in the Earth System
perspective. AMBIO 41 (1), 19.

B.A. Bluhm et al. / Progress in Oceanography 139 (2015) 89121


Wassmann, P., Carmack, E., Kosobokova, K.N., Slagstad, D., Drinkwater, K., Hopcroft,
R.R., Moore, S.E., Ellingsen, I., Nelson, R.J., Popva, E., Berge, J., 2015. The
contiguous domains of Arctic Ocean advection: trails of life and death. Progress
in Oceanography 139, 4265.
Watanabe, E., Kishi, M.J., Ishida, A., Aita, M.N., 2012. Western Arctic primary
productivity regulated by shelf-break warm eddies. Journal of Oceangraphy 68,
703718. http://dx.doi.org/10.107/s10872-012-0128-6.
Wei, C-.K., Rowe, G.T., Escobar-Briones, E., Boetius, A., Soltwedel, T., Caley, M.J.,
Soliman, Y., Huettmann, F., Qu, F., Yu, Z., Pitcher, C.R., Haedrich, R.L., Wicksten,
M.K., Rex, M.A., Baguley, J.G., Sharma, J., Danovaro, R., MacDonald, I.R., Nunnally,
C.C., Deming, J.W., Montagna, P., Lvesque, M., Weslawski, J.M., WlodarskaKowalczuk, M., Ingole, B.S., Bett, B.J., Yool, A., Bluhm, B.A., Iken, K.,
Narayanaswamy, B.E., 2010. Global patterns and predictions of seaoor
biomass using random forests. PLoS ONE 5 (12), e15323.
Weingartner, T.J., Cavalieri, D.J., Aagaard, K., Sasaki, Y., 1998. Circulation, dense
water formation, and outow on the northeast Chukchi shelf. Journal of
Geophysical Research 103 (C4), 76477661. http://dx.doi.org/10.1029/
98jc00374.
Weingartner, T., Aagaard, K., Woodgate, R., Danielson, S., Sasaki, Y., Cavalieri, D.,
2005. Circulation on the north central Chukchi Sea shelf. Deep-Sea Research II
52 (2426), 31503174. http://dx.doi.org/10.1016/j.dsr2.2005.10.015.
Weingartner, T., Dobbins, E., Danielson, S., Winsor, P., Potter, R., Statscewich, H.,
2013. Hydrographic variability over the northeastern Chukchi Sea shelf in
summer-fall 20082010. Continental Shelf Research 67, 522.
Wheeler, P.A., Gosselin, M., Sherr, E., Thibault, D., Kirchman, D.L., Benner, R.,
Whitledge, T.E., 1996. Active cycling of organic carbon in the central Arctic
Ocean. Nature 380, 607699.
Williams, W.J., Carmack, E.C., 2008. Combined effect of wind-forcing and isobath
divergence on upwelling at Cape Bathurst, Beaufort Sea. Journal of Marine
Research 66, 645663.
Williams, W., Carmack, E.C., 2015. The Interior shelves of the Arctic Ocean:
physical oceanographic setting and effects of summertime sea-ice retreat on
nutrient supply. Progress in Oceanography 139, 2441.
Williams, W.J., Carmack, E.C., Shimada, K., Melling, H., Aagaard, K., Macdonald, R.W.,
Ingram, R.G., 2006. Joint effects of wind and ice motion in forcing upwelling in
Mackenzie Trough, Beaufort Sea. Continental Shelf Research 26, 23522366.
Williams, W.J., Melling, H., Carmack, E.C., Ingram, R.G., 2008. Kugmallit Valley as a
conduit for cross-shelf exchange on the Mackenzie Shelf in the Beaufort Sea.
Journal of Geophysical Research: Oceans 113 (C2).
Witte, U., 1996. Seasonal reproduction in deep-sea sponges triggered by vertical
particle ux? Marine Biology 124, 571581.
Wood, K.R., Overland, J.E., Salo, S.A., Bond, N.A., Williams, W.J., Dong, X., 2013. Is
there a new normal climate in the Beaufort Sea? Polar Research 32, doi.org/
10.3402/polar.v32i0.19552.

121

Woodgate, R.A., Aagaard, K., 2005. Revising the Bering Strait freshwater ux into the
Arctic Ocean. Geophysical Research Letters 32, L02602.
Woodgate, R.A., Aagaard, K., Weingartner, T.J., 2005. Monthly temperature, salinity,
and transport variability of the Bering Strait through ow. Geophysical
Research Letters 32. http://dx.doi.org/10.1029/2004GL021880.
Woodgate, R.A., Aagaard, K., Weingartner, T.J., 2006. Interannual changes in the
Bering Strait uxes of volume, heat and freshwater between 1991 and 2004.
Geophysical Research Letters 33 (15), L15609.
Woodgate, R.A., Aagaard, K., Swift, J.H., Smethie Jr., W.M., Falkner, K.K., 2007.
Atlantic water circulation over the Mendeleev Ridge and Chukchi Borderland
from thermohaline intrusions and water mass properties. Journal of
Geophysical Research 112, C02005.
Woodgate, R.A., Weingartner, T.J., Lindsay, R., 2012. Observed increases in Bering
Strait oceanic uxes from the Pacic to the Arctic from 2001 to 2011 and their
impacts on the Arctic Ocean water column. Geophysical Research Letters 39,
L24603. http://dx.doi.org/10.1029/2012GL054092.
Worthington, V.L., 1953. Oceanographic results of Operation Skijump I and II.
Transactions of the American Geophysical Union 34, 543551.
Yamamoto-Kawai, M., Carmack, E., McLaughlin, F., 2006. Nitrogen balance and
Arctic through ow. Nature 443, 43. http://dx.doi.org/10.1038/443043a.
Yamamoto-Kawai, M., McLaughlin, F.A., Carmack, E.C., Nishino, S., Shimada, K.,
2008. Freshwater budget of the Canada Basin, Arctic Ocean, from salinity, d18O,
and nutrients. Journal of Geophysical Research 113, C01007. http://dx.doi.org/
10.1029/2006JC003858.
Yamamoto-Kawai, M., McLaughlin, F.A., Carmack, E.C., Nishino, S., Shimada, K.,
Kurita, N., 2009. Surface freshening of the Canada Basin, 20032007: River
runoff versus sea ice meltwater. Journal of Geophysical Research 114 (C1).
http://dx.doi.org/10.1029/2008JC005000.
Yamamoto-Kawai, M., McLaughlin, F.A., Carmack, E.C., 2011. Effects of ocean
acidication, warming and melting of sea ice on aragonite saturation of the
Canada Basin surface water. Geophysical Research Letters 38. http://dx.doi.org/
10.1029/2010GL045501.
Zauker, F., Stocker, T.M., Broecker, W.S., 1994. Atmospheric freshwater uxes and
their effect on the global thermohaline circulation. Journal of Geophysical
Research 99 (C6), 1244312457.
Zenkevitch, L., 1963. Biology of the seas of the USSR. Allen & Unwin, London, pp.
955.
Zhang, J., Spitz, Y.H., Steele, M., Ashjian, C., Campbell, R., Berline, L., Matrai, P., 2010.
Modeling the impact of declining sea ice on the Arctic marine planktonic
ecosystem. Journal of Geophysical Research 115, C10015. http://dx.doi.org/
10.1029/2009JC005387.
Zhao, M., Timmermans, M.L., Cole, S., Krisheld, R., Proshutinsky, A., Toole, J., 2014.
Characterizing the eddy eld in the Arctic Ocean halocline. Journal of
Geophysical Research: Oceans 119. http://dx.doi.org/10.1002/2014JCO104888.

You might also like