Oceano

You might also like

You are on page 1of 19

Available online at www.sciencedirect.

com

Polar Science 7 (2013) 53e71


http://ees.elsevier.com/polar/

Structure of the Fram Strait branch of the boundary current in the


Eurasian Basin of the Arctic Ocean
Andrey V. Pnyushkov a,*, Igor V. Polyakov a, Vladimir V. Ivanov a,b, Takashi Kikuchi c
a
International Arctic Research Center, University of Alaska Fairbanks, 930 Koyukuk Dr., Fairbanks, AK 99775, USA
b
Arctic and Antarctic Research Institute, St. Petersburg, Russia
c
Japan Agency for Marine-Earth Science Technology (JAMSTEC), Yokosuka, Japan

Received 29 June 2012; revised 18 December 2012; accepted 12 February 2013


Available online 26 February 2013

Abstract

Recent mooring-based observations at several locations along the continental slope of the Arctic Ocean’s Eurasian Basin
showed a transformation of the Boundary Current (BC) from a mostly barotropic flow in Fram Strait to a jet-like baroclinic current
northeast of Svalbard, and the reemergence of the barotropic structure of the flow in the eastern Eurasian Basin. This transformation
is accompanied by a weakening of the flow from w24 cm/s in Fram Strait to w5 cm/s at the Lomonosov Ridge. The maximum of
the baroclinic component of the BC at an intermediate depth (w200e370 m) is associated with the Atlantic Water core. The depth
range of the baroclinic current maximum is controlled by cross-slope density gradients above and below the baroclinic velocity
maximum as follows from the geostrophic balance of forces. According to the model simulations, the BC splits into shallow and
deep branches in the proximity of Svalbard due to a divergence of isobaths, confirming topographically-controlled BC behavior.
The shallow branch is located at a shelf break with a typical bottom depth of w200 m and current speed of up to w24 cm/s. The
discussed results, which provide insight on some basic aspects of the dynamics of the BC (the major oceanic heat source for the
Arctic Ocean), may be of importance for understanding of the ocean’s role in shaping the arctic climate system state.
Ó 2013 Elsevier B.V. and NIPR. All rights reserved.

Keywords: Arctic Ocean; Boundary current; Current observations; Numerical model

1. Introduction 1987; Blindheim, 1989; Rudels et al., 1994). Both


AW branches converge over the continental slope north
The inflow of intermediate depth (150e800 m) of the Kara Sea and jointly propagate further, traveling
water of Atlantic origin (the so-called Atlantic Water, in a cyclonic flow along the deep Arctic Ocean Basins
AW) propagating along the continental slope as a (Aagaard, 1989; Rudels et al., 1994, 2000; Jones,
Boundary Current (BC) is the major oceanic heat 2001). The rate of mixing between these two AW
source for the Arctic Ocean (Aagaard and Greisman, branches, however, is still under debate (Rudels, 2012).
1975). AW enters the Arctic Ocean through Fram Recent studies using both observations (Schauer et al.,
Strait and the Barents Sea (Nansen, 1902; Rudels, 2002; Rudels, 2012) and modeling (e.g. Aksenov et al.,
2011) suggest that the Barents Sea branch of the AW
* Corresponding author. Tel.: þ1 907 4742683. propagates in the EB interior as a separate filament,
E-mail address: andrey@iarc.uaf.edu (A.V. Pnyushkov). located above the upper continental slope. Conversely,

1873-9652/$ - see front matter Ó 2013 Elsevier B.V. and NIPR. All rights reserved.
http://dx.doi.org/10.1016/j.polar.2013.02.001
54 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

the Fram Strait branch typically propagates above the Beaufort Sea at depths deeper than 150 m (Pickart,
lower continental slope, where the bottom depth is 2004; Nikolopoulos et al., 2009).
deeper than 1800 m (Schauer et al., 2002). An insight on the mechanisms governing the cir-
An earlier attempt to reconstruct the circulation of culation in the Eurasian Basin of the Arctic Ocean is
the AW in the Polar Basin by Timofeev (1960) using provided using a theory of buoyant geostrophic flow
water temperature observations suggested a broad AW over a sloping topography. Combining numerical ex-
spread in a generally cyclonic direction. The cyclonic periments with theoretical considerations Walin et al.
pattern of AW circulation was confirmed later by (2004) analyzed behavior of the BC, when it loses
Coachman and Barnes (1963), Nikitin (1969), Rudels buoyancy propagating toward the interior of the Arctic
et al. (1994), and McLaughlin et al. (1996). Ocean from the Nordic Seas. The authors found that
Aagaard’s (1989) analysis of summer 1980 records conservation of potential vorticity over a sloping
derived from two moorings deployed at the continental topography results in the formation of strong
slope of the Barents Sea revealed a strong current topographically-controlled BC with barotropic cross-
enveloping the continental slope (i.e. increasing with shelf transport due to along-slope variations in bot-
depth, with a maximum of 15 cm/s at the bottom). tom density. In their numerical experiment initially
Synthesizing all measurements that were available at baroclinic flow is transformed to the deep baroclinic
the time, Aagaard concluded that large-scale transport and shallow barotropic branches, while the BC prop-
of AW in the Arctic Ocean occurs as a narrow BC agates along the idealized continental slope. Similar
along the basins’ margins. The propagation of AW approach based on depth-integrated vorticity equation
along the continental slope as a topographically- over a closed depth contour was used by Nøst and
controlled, narrow boundary flow is confirmed by Isachsen (2003), who examined an impact of wind
recent observations (e.g. Rudels et al., 1994; stress divergence in the surface Ekman layer on the
McLaughlin et al., 1996; Woodgate et al., 2001; large-scale circulation in the Nordic Seas and the
Dmitrenko et al., 2008, 2009, 2012; Ivanov et al., Arctic Ocean. They found that the bottom geostrophic
2009; Rudels, 2012) and numerical models (e.g. flow is mainly driven by the surface Ekman pumping in
Gerdes et al., 2003; Karcher et al., 2003; Maslowski the upper ocean layer. Schlichtholz (2007), investi-
et al., 2004; Hunke and Holland, 2007; Zhang and gating dynamics of the East Greenland Current in Fram
Steele, 2007; Aksenov et al., 2010, 2011; Karcher Strait, supported Walin et al. (2004) conclusion about
et al., 2012). However, there are large discrepancies an importance of the bottom density gradient to the
in how the models simulate the intensity, pathways, geostrophic circulation. The author combined oceano-
and even direction of AW circulation (for discussion, graphic data analysis with theoretical consideration of
see Karcher et al., 2007). the depth-integrated vorticity equation and found
Modern observations have added important details that downstream bottom density increase of 0.01 kg/m3
that document the structure of the BC. They suggested is responsible for bottom geostrophic velocity increase
that the AW entering the Arctic Ocean through Fram of w3 cm/s. Aaboe and Nøst (2008) using linear
Strait as the West Spitsbergen Current (WSC) has a diagnostic model and climatological temperature and
primary barotropic vertical structure (Fahrbach et al., salinity as a forcing reproduced the large-scale circu-
2001; Schauer et al., 2004, 2008, Fig. 1a). Contrary lation of the BC in the Nordic Seas and the Arctic
to these findings, Walczowski et al. (2005), using Ocean. The model is based on the vorticity balance
geostrophic balance considerations, pointed out that equation integrated within the areas bounded by
the baroclinic and barotropic velocity components of enclosed depth contours. Special attention in this study
the WSC may be comparable. Mooring observations is given to the role of decreasing along-isobath bottom
only w500 km downstream from Fram Strait at the density, which changes lead to decrease of barotropic
northeast slope of Svalbard revealed the maximum transport and to increase of baroclinic transport asso-
currents at an intermediate depth of w300e400 m ciated with the BC. The authors found good agreement
(Ivanov et al., 2009, Fig. 1b), which is typical for between velocity observations and simulated currents
baroclinic currents. At the same time, the BC in the and concluded that the along-slope bottom density is
proximity of the Lomonosov Ridge was found to be important for the slope-basin water exchange in the
primarily barotropic (Woodgate et al., 2001, Fig. 1e). polar basins.
In the Canadian Basin the shelfbreak circulation also One of the assumptions used in the models by Walin
demonstrates the presence of a narrow (w20 km) et al. (2004), Aaboe and Nøst (2008), and Aaboe et al.
eastward current transporting the AW in the Alaskan (2009) is that the geostrophic bottom flow always
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 55

Fig. 1. Vertical structure of speed of the boundary current (BC; cm/s) derived from mooring-based observations at six sites: (a) F4/F1
(2002e2006), (b) M4 (2004e2005), (c) M11 (2009e2011), (d) M1 (2004e2005), (e) LM1 (1995e1996), and (f) M3 (2005e2006) along the
Atlantic Water (AW) pathway in the Eurasian Basin of the Arctic Ocean. Depths in inserts a-f are in meters. The solid arrows show the mean
current speed, derived by time average of mooring-based velocity measurements. Standard errors of the mean current are undistinguished from
zero at these scales of axes. The dashed arrows in (a) and (e) were obtained as a linear vertical interpolation between available measurements at
upper and lower depth levels. Circulation of the AW is schematically shown by red arrows. Bottom topography is shown by color; black rectangle
represents the area of strong divergence of isobaths north of Svalbard.

follows the depth contours which assumes zero vertical near-bottom cross-slope exchanges, which are impor-
velocity within the entire water column. We note, that tant for the ventilation of intermediate and deep layers
this simplified approach is valid for some specific of the Arctic Ocean (Rudels, 1986; Jones et al., 1995;
tasks, however for many practical applications it is Ivanov and Golovin, 2007).
oversimplified. For example, it is not consistent with Uncertainties and gaps remain in our knowledge of
generally isopycnal propagation of the BC in the the structure and intensity of the BC. What are the
Eurasian Basin of the Arctic Ocean, which requires roles of barotropic and baroclinic factors in shaping the
vertical slope of the BC. This approach neglects BC? Does the vertical structure of the BC remain
56 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

invariable along the AW pathway? What mechanisms Current Meter (RCM)-7, RCM-8, Doppler Current
maintain it? Available observations are still too scarce Meter (DCM)-11, and 3D-Acoustic Current Meter
and the existing publications based on modeling (ACM) instruments at 50, 250, 750, and 1500 m
experiments were not specifically focused on the depths. Due to bottom depth restrictions the instru-
structure of the BC in the Eurasian Basin, so they do mental observations at the F1 mooring were limited by
not clarify the vertical structure of the AW flow beyond 250 m depth. For analysis we used mean currents at
existing observations. With these questions in mind, these two moorings derived from each instrument by
the primary objectives of this study are to analyze BC averaging measurements over the entire period (i.e.
transformation along the Eurasian Basin slope from 2002e2006) of available observations (Fig. 1a). We
Fram Strait to the Lomonosov Ridge, and to under- provide the estimates of standard error (SE) of the
stand the mechanisms that govern its transformation mean current as the measures of confidence intervals of
using direct mooring observations and results of nu- averaging, so that the only velocity differences that
merical modeling. In order to achieve these objectives, exceed the SE may be considered as statistically
we have analyzed the existing current observations in robust. The SE for every level of observations was
the Eurasian Basin (Fig. 1), which are described in estimated as, where is standard deviation of current
detail in Section 2. This same section provides the speed at the and N is number of observations; the
numerical experiment design performed to gain insight standard error does not exceed 0.8 cm/s for any levels
into the physical mechanisms responsible for BC at the F1 and F4 moorings (A. Beszczynska-Möller,
structure maintenance and its along-slope trans- pers. communication).
formation. Section 3, which complements our Current observations from the mooring deployed in
observation-based analysis, further presents results of autumn 2004 at 1010 m in a position near the AW flow
the numerical simulations and describes the structure core were used to examine the BC structure northeast
of the BC. Section 4 concludes the study with of Svalbard (81 300 N, 31 000 E, w1010 m water
discussion. depth). The mooring was equipped with five con-
ductivityetemperatureedepth (CTD) recorders and
2. Data and model description three RCM-9s (for details see Ivanov et al., 2009).
Current velocity was registered once each hour. The
2.1. Observational data mooring was recovered in autumn 2006. The length of
the current record at three depths (69, 216, and 993 m)
In this study we used mooring-based observations was approximately one year with approximately 9400
of currents at six key locations along the pathway of observations at each sampling level. Time series from
Fram Strait branch of the AW in the Eurasian Basin of the mooring were averaged over the period of obser-
the Arctic Ocean (Fig. 1). The location of moorings vations, thus providing year-long means for each in-
deployed in proximity to the AW core (defined by its strument (Fig. 1b). Measurements at this mooring
maximum temperature during the time of deployment) showed relatively low variability of current speed and
allowed us to trace changes of the BC structure from direction, indicating a lack of intensive meandering
the AW entrance through Fram Strait to the eastern- and eddy formation (Ivanov et al., 2009). The SE of the
most location of the Eurasian Basin. Additional Low- mean current does not exceed 0.1 cm/s at all three
ered Acoustic Doppler Current Profiler (LADCP) depth levels of observations.
observations at several oceanographic stations over the Current observations over the Franz-Joseph Land
Laptev Sea slope have been included in our analysis to continental slope were used to examine transformation
examine the horizontal pattern of the BC. of the BC during its eastward propagation from Sval-
Averaged multi-year (2002e2006) current-meter bard toward the EB interior along the continental
measurements from long-term moorings along a slope. The M11 mooring (83 040 N; 59 480 E, w2880 m
78 500 N section were used in order to examine the total depth; Fig. 1) was deployed at the AW core, as it
structure of the AW inflow in Fram Strait (Schauer was determined by the accompanying CTD survey in
et al., 2008). The section extended from the eastern August 2009. This mooring was recovered in
Greenland shelf break (6 510 W) to the western shelf September 2011, providing a w20 month-long series
break off Svalbard (8 390 E). The eastern part of the of current observations. The mooring was equipped
section was bounded by two moorings, F1 (w8 390 E) with a McLane Moored Profiler (MMP), which
and F4 (w6 580 E) crossing the WSC. The ocean ve- measured temperature, conductivity, pressure, and
locity was sampled at 2-h intervals by Recording current speed and direction. The MMP produced
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 57

profiles once each day with high (w0.12 m) vertical easternmost part of the EB. The M3 mooring was
resolution of sampling within a 280e1090 m layer. deployed in 2005 and recovered one year later, in
Mooring time series at each level were averaged over September 2006. The available ADCP record spans
2-m depth cells and then over the entire period of the water layer between 75 and 155 m depth, with
observations; these means are shown in Fig. 1c. The SE four-meter vertical resolution and a twice per hour time
of the mean current estimates at this mooring does not interval of current velocity registration. Two RCM in-
exceed 0.3 cm/s for all levels of observations. struments, deployed at 280 and 850 m depths, comple-
The detailed vertical structure of the AW flow from ment ADCP observations and provide measurements of
September 2004 through February 2005 was obtained current speed within the AW layer. ADCP and RCM
over the Laptev Sea slope at mooring M1 which was time series were averaged over the entire period of ob-
deployed at a total water depth of w2690 m in prox- servations (i.e. 2005e2006; Fig. 1f). The SE of the mean
imity to the AW core determined with CTD section at current at this mooring does not exceed 0.2 cm/s for the
78 260 N and 125 370 E. The MMP raw current mea- ADCP and 0.1 cm/s for the RCM instruments.
surements spanning the 80e900 m layer were averaged In order to examine cross-slope structure of the BC
within 2 m depth cells providing observations at 411 and to extend the mooring-based observational analysis
levels. The total of 160 observations with daily reso- we used LADCP data collected along a 125 E section
lution was used for time averaging at each level over the Laptev Sea slope in September 2005. The
(Fig. 1d). The SE of the mean current estimates shallowest station of the section was located at
does not exceed 0.3 cm/s for the 80-150 m layer and 77 440 N, 125 590 E, the deepest station was at 77 570 N,
0.2 cm/s for the layer below. 126 030 E, and one of the stations coincided with the
The year-long (1995e1996) current observations M1 mooring location (Fig. 2, top). Raw LADCP data
described by Woodgate et al. (2001) at the moored were processed using Lamont-Doherty Earth Obser-
station LM1 located in proximity to the AW core (i.e., vatory (LDEO) software, which is available from ftp://
maximum temperature on the hydrographic transect ftp.ldeo.columbia.edu/pub/LADCP. Vertical resolution
during the mooring deployment), near the Lomonosov of the processed LADCP current profiles is 10 m,
Ridge (78 310 N and 133 580 E), were also used in this spanning the water column from the surface to the
study. Four RCM instruments were deployed at 106, near-bottom layer. Because of low data quality due to
326, 761, and 1161 m. The mooring was also equipped the large dispersion of bottom tracking measurements
with three CTD recorders located near the first two (which is necessary to derive current speed) only the
RCMs and in the bottom layer. Mean current mea- shear, or baroclinic (Thurnherr, 2008), component of
surements from these instruments averaged over the the current was used in this study. According to the
entire length of the annual records were reported by manufacturer’s description, the accuracies of raw cur-
Woodgate et al. (2001; see their Table 1) and were used rent measurements with a 300-kHz LADCP are 0.5%
for our analysis (Fig. 1e). The SEs for the estimates of of reading for magnitude and 2 for current direction.
the annual mean current vary from w0.4 cm/s at the The maximal errors of the derived velocity profiles
two upper (106 and 326 m) levels to w0.2 cm/s at determined with LDEO software do not exceed
1161 m (R. Woodgate, pers. communication). 2.8 cm/s in the surface layer and decrease linearly to
ADCP and RCM current observations from a 2.0 cm/s at 1000 m depth. We note that the LADCP
mooring deployed at 79 550 N; 142 210 E, (M3 mooring; observations are quasi-synoptical observations, in
w1347 m total depth; Fig. 1), which is separated by which mean current structure is “contaminated” by
w230 km from the LM1 mooring, were used to extend small-scale processes such as barotropic and baroclinic
our analysis with modern observations in the tides, internal waves, and mesoscale eddies. Therefore,
special care should be exercised when interpreting
Table 1 these data. In our study, we have used them to illustrate
Contribution of terms (108, m s2) in the depth-integrated mo- the existence of a good match with the location of the
mentum equation of the along-slope barotropic velocity averaged over AW core, as defined by either temperature or current
the simulated deep BC branch. speed across the continental slope. The LADCP ob-
Along-slope Coriolis Bottom Horizontal Pressure servations were accompanied by a CTD survey,
distance, km force stress advection gradient providing information about temperature and salinity
99 1.46 0.23 0.09 1.33 distribution across the continental slope. We used the
165 3.52 0.23 0.09 3.37 latter to define the position of the AW core in this
660 2.87 0.17 0.05 2.73
particular year along the Laptev Sea slope. We also
58 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

Fig. 2. (Top) Location of LADCP/CTD stations (circles) and the M1 mooring (star) in September 2005. The color scale shows water depth
(meters). (Bottom) Cross-slope distribution of baroclinic current speed (cm/s, color) and potential temperature ( C, isolines) on a section over the
Laptev Sea slope.

include in our analysis CTD observations from two numerical model based on the Regional Oceanic
cross-slope sections through moorings M4 and M1, Modeling System, described by, e.g., Shchepetkin and
collected in 2006, in order to illustrate agreement be- McWilliams (2005). This model is a 3D, non-linear,
tween mooring-based current observations and primitive-equation, Boussinesq, hydrostatic, free-
geostrophic velocities derived using temperature and surface ocean model in an s-coordinate system. A
salinity measurements. Instrumental accuracies of in- turbulent closure we applied is the non-stationary
dividual temperature and salinity measurement are generic length-scale turbulence model proposed by
0.002  C and 0.001 psu, respectively. Umlauf and Burchard (2003). This model was formu-
lated for the variables of turbulent kinetic energy (k)
2.2. Model description and speed of viscous dissipation (e), similar to the kee
model of Rodi (1987). The horizontal mixing is
To examine the mechanisms of along-slope AW defined by a Laplace operator and is prescribed along
transformation in the Eurasian Basin we used a the s-coordinate for velocity components, temperature,
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 59

and salinity. The coefficients of turbulent horizontal conditions allows us to merge the property of the ra-
viscosity and diffusivity were taken as 30 m2/s and diation condition to eliminate outgoing high-frequency
10 m2/s, respectively. The model uses splitting to gravitational waves and the property of the nudging
barotropic and baroclinic modes and a fourth-order, condition to preserve the general observed structure of
centered scheme for horizontal advection of tracers AW inflow. Distribution of the depth-averaged velocity
and momentum. The finite-difference approximation in component used for nudging was taken from the
space is constructed using a Cartesian equally-spaced mooring observations in Fram Strait, averaged over the
C-grid. 2002e2006 period (Schauer et al., 2008, Fig. 3). A
The model domain spans the area of the Eurasian velocity maximum of 25 cm/s associated with the AW
Basin slope with horizontal resolution of 3.3 km and 28 core is located in the shallow area close to Svalbard.
s-coordinate levels. The length of the model domain is Replicating observations, the current speed decreases
1650 km in the along-slope (X-axis) direction, and linearly to 2 cm/s from the core toward the deep ocean
247 km in the cross-slope (Y-axis) direction. Bottom and remains constant starting at the 50th node
topography is idealized and varies cross-slope only; (w165 km). The sponge boundary layers were intro-
however, it reflects the major features of depth distri- duced at eastern, northern, and southern borders as
bution along the Eurasian Basin slope (Fig. 3). For lateral boundary conditions for momentum to absorb
example, the mean gradient of the modeled Eurasian the inertial waves propagating toward the boundary
Basin slope is chosen to be close to the observed one. (Chen and Beardsley, 1995; Holloway, 1996). The
The southwestern corner of the domain includes an eastern boundary layer was w33 km wide and the
island representing Svalbard. southern and northern boundaries were w13 km wide,
The model is forced by barotropic inflow imitating with horizontal viscosity coefficients of 300 m2/s. The
the observed inflow of the AW into the Eurasian Basin thicker sponge layer for the eastern boundary was used
through Fram Strait (Schauer et al., 2004, 2008). The due to more intensive generation of inertial waves over
radiation condition proposed by Orlanski (1976) was the area with varied bottom topography relative to the
implemented at all lateral open boundaries for the southern and northern boundaries with constant depths.
barotropic and baroclinic components of current, Climatology data describing Arctic Ocean temper-
temperature, and salinity. For the barotropic compo- ature and salinity from the Environmental Working
nent of current at the western flank of the domain Group (EWG) (Environmental Working Group, 1997)
where AW inflow occurs we introduced the additional were used in order to prescribe initial conditions
routine of nudging (restoring) toward an a-priori for temperature and salinity in numerical experiments
defined cross-slope distribution of barotropic velocity, and for further evaluation of numerical simulation
on a timescale of 30 days. Combining these two results.

Fig. 3. Idealized model domain. Isolines show bottom topography (meters). Arrows show velocities of the barotropic inflow across the lateral
open boundary (scale is shown in the insert). Black rectangle indicates the area of isobath divergence.
60 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

Zero heat, salt, and momentum fluxes were applied After the model spin-up was done the temperature and
as surface boundary conditions. A free kinematic salinity were simulated using restoring correction for
condition was implemented at the surface model 150 days of model integration. Restoring of the
boundary for simulated vertical velocity. For the bot- simulated T, S-profiles toward the model adjusted
tom layer, we used quadratic-law momentum stress for climatology was required in order to account for
current, and zero fluxes for temperature and salinity. realistic external forcing, the effects of which would
Two sets of numerical experiments were performed not otherwise be included in our idealized experiment.
in this study. The goal of the first 250-day-long baro- A relaxation time-scale constant of 30 days was used
tropic experiment was to investigate the role of baro- for the model temperature and salinity during the
tropic factors in forming the along-slope current in the prognostic run. The total length of the baroclinic
Eurasian Basin. In this experiment model temperature experiment (including initial prognostic smoothing and
and salinity for the entire model domain were set to spin-up) was 550 days. Change of the total energy over
zero. Mean total energy was used to identify the the last 30 days was less than 1%, guaranteeing that
duration of the experiment. For example, change of the simulated currents averaged over the last 7 days of the
total energy over the last 30 days in the barotropic experiment at several cross-slope sections along the
experiment was less than 1%, attesting that the model AW pathway were quasi-stationary. Data averaged
had approached an equilibrium state. over these 7 days were used in the analysis.
The second baroclinic experiment was carried out in
order to examine impacts of density on BC structure. 3. Structure of the BC
We also examined the role of the along-slope density
gradient in shaping cross-isobath water transport, 3.1. Vertical structure of the BC from observations
providing insight on the mechanisms governing
shelfeocean exchange in the EB. Replicating the The vertical structure of AW inflow carried by WSC
barotropic experiment, the vertical structure of the AW in Fram Strait has been derived using current obser-
inflow at the western flank of the model domain was vations averaged over 2002e2006 (Fig. 1a). According
defined as barotropic inflow. For this experiment the to Gascard et al. (1995), Fahrbach et al. (2001), and
initial temperature and salinity were prescribed using Schauer et al. (2004, 2008), this structure is mostly
the Arctic Ocean winter climatology (Environmental barotropic or equivalent-barotropic (see Killworth,
Working Group, 1997). The initial temperature and 1992 for details), when the vertical distribution of
salinity within the model interior were linearly inter- current speed is vertically uniform, or the current speed
polated between EWG temperature and salinity at two decreases with depth due to enhanced friction within
cross-isobath sections, at the eastern and western flanks the near-bottom layer. Averaged multi-year
of the model domain. Whereas the initial temperature (2002e2006) observations show that the maximal
and salinity provided by the EWG dataset may not current speed at the section across Fram Strait was
agree well with the simulated current pattern, due to found at the F1 mooring which is the shallowest
use of the idealized model topography, we carried out mooring and the closest to the Spitsbergen coast
model adjustment for these fields. For this, simulated (Schauer et al., 2008). The mean current speed within
temperature and salinity were allowed to evaluate the AW layer at this mooring varied from w24 cm/s at
freely during a short-period (w25 days) prognostic run 250 m to w22 cm/s at 50 m; 50 m is the shallowest
where no restoring or any other temperature and available observational level at this site (Fig. 1a, black
salinity corrections were used. In order to reduce arrows). At F4, the westernmost mooring of this sec-
further strong drift of temperature and salinity during tion, the current speed was significantly weaker
model spin-up a 375-day-long integration with fixed compared with current speed at the F1 mooring;
temperature and salinity taken from the model-adjusted however, its vertical structure was also close to baro-
climatology was carried out. This integration time is tropic. The maximal current speed (w9 cm/s) was
rather long compared with the time necessary for observed in the upper w250 m layer, decreasing to
establishing a steady-state geostrophic current; this 6 cm/s at 750 m and with some increase with depth
length of time was necessary due to strong velocity below 750 m (Fig. 1a, blue arrows).
variations close to the western border of the model The mooring records from the area northeast of
domain. Increased variability at this border is caused Svalbard (mooring M4) show a very different BC
by slow adjustment of the current velocities forced structure with a maximum velocity of w22 cm/s at an
with the relaxation term in the boundary condition. intermediate depth of w216 m (Fig. 1b). The location
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 61

of this maximum coincides with the AW core (Ivanov LADCP current observations along 125 E demon-
et al., 2009). At this mooring the maximum current strate that the maximum baroclinic current is associ-
was observed at a deeper level than in Fram Strait, but ated with the AW core as identified by temperature
current intensity was comparable. The maximum speed observations (Fig. 2, bottom). The maximum baro-
observed in the AW core exceeds the speed of the clinic currents (w9 cm/s) at the intermediate depth and
upper ocean (w69 m) and bottom (w993 m) currents the highest water temperature were observed at the
by 4 and 5 cm/s, respectively. The formation of the KD0516/M1 station. The vertical profile of the baro-
velocity maximum at an intermediate depth is evidence clinic current speed in the core shows two local max-
of the along-slope transformation of BC structure from imums. The first maximum with the strongest
mostly barotropic in Fram Strait to baroclinic jet over baroclinic current was found at w150 m depth, 70 m
the Nansen Basin slope. The baroclinic velocity shear, above the major AW temperature core at 220 m (Fig. 2,
calculated as the difference between the intermediate- bottom). An additional local baroclinic current
depth maximum of currents and the current speed in maximum was observed at w350 m depth coinciding
the upper ocean layer, exceeds the level of SE of the with a secondary temperature maximum (w1.6  C).
mean current at the mooring (i.e., w0.1 cm/s), statis- This strong agreement between the velocity- and
tically confirming the robustness of the captured temperature-defined cores of the AW provides confi-
transformation of BC structure. dence that the quasi-synoptical observations did indeed
Mooring observations over the slopes of Franz- capture the mean structure of the baroclinic current
Joseph Land (mooring M11) and the Laptev Sea (Fig. 1d).
(mooring M1) show that the BC preserves its baro- Further eastward, at the Lomonosov Ridge
clinic structure, though traces of erosion (disappear- (mooring LM1), observations showed a barotropic flow
ance) of the intermediate-depth current maximum with weak current (w5 cm/s) in the upper layer
become evident (Fig. 1c, d). The eastward progression decreasing almost linearly to 3.4 cm/s at a depth of
of AW flow along the slope segment from Svalbard 1161 m (Fig. 1, panel d). The lack of a velocity
(mooring M4) to the Franz-Joseph Land slope and maximum at intermediate and deeper levels is evidence
further eastward to the central Laptev Sea slope is of AW BC baroclinic structure erosion; the beginning
associated with deepening of the current maximum. of this erosion was observed over the central part of the
For example, depth ranged from w217 m to w365 m Laptev Sea slope.
(i.e., by w150 m) for the M1 mooring and ranged even The vertical structure of the BCdwith no evident
wider (by w350 m) at the M11 mooring, weakening baroclinic intermediate-depth velocity maximumdwas
flow by more than four times, from 22 cm/s to 5 cm/s. observed at the mooring M3 deployed over the eastern
Confirming the complex vertical structure of the BC, flank of the Lomonosov Ridge. The BC at this mooring
the high-resolution mean velocity profile at the M1 showed equivalent-barotropic structure, under which the
mooring exhibits an additional local maximum of maximal current speed (w7 cm/s) was observed at the
w3.9 cm/s at a depth of 121 m. uppermost available observational level of 55 m.
Observation-derived baroclinic velocity shear at the The current speed linearly decreases with depth to
M4 and M1 moorings agreed well with those derived w3.5 cm/s at 856 m, replicating similar weakening of
from a geostrophic relationship (Fig. 4). The geostrophic currents observed at the LM1 mooring (Fig. 1e, f). A
currents were calculated using CTD observations at two good match of current observations at the LM1 and M3
cross-slope sections at w30 E and 125 E in reference to moorings, which were deployed very close in space but
the ocean surface, where zero geostrophic velocities are separated in time by ten years, indicates that the ree-
assumed. At M4 a mooring maximum of geostrophic mergence of the barotropic structure of the flow in this
currents was w4.7 cm/s, which agrees well with the region is a persistent feature of the BC dynamics and is
baroclinic shear of w4 cm/s derived from instrumental not subject to the interannual variability of general cir-
current observations (Fig. 4a). Geostrophic currents at culation that has been reported for the Canadian Basins
the M1 mooring also show a good match with direct by Morison et al. (2012), for example.
current-meter observations (Fig. 4b). For example, the We conclude that observational data demonstrate a
derived maximum of the geostrophic current at this transformation from the mostly barotropic structure of
mooring was 2.5 cm/s, which barely and insignificantly the BC observed in Fram Strait to a jet-like baroclinic
exceeds (comparable with the SE of the mean current) structure over the Svalbard slope. The further propa-
the vertical shear of current speed (w2 cm/s) registered gation of the AW BC along the continental slope into
using collected MMP observations. the basin interior is accompanied by erosion of this
62 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

Fig. 4. Cross-slope distribution of geostrophic velocities (color, cm/s) and water temperature (isolines,  C) along 30 E (top) and 125 E (bottom)
sections, derived using 2006 CTD observations. Positive velocities are eastward. Back triangles show the position of M4 and M1 moorings.

baroclinic structure and reemergence of the barotropic as the “deep branch” forms over the slope at a typical
structure at the junction of the Lomonosov Ridge. depth of w1300 m and is separated from the shallow
branch by w100 km (Fig. 5). Taking into consideration
3.2. Structure of the BC from model results the idealized topography used in our experiment we may
arguably attribute this simulated deep branch to the
3.2.1. Barotropic experiment Svalbard Branch of the WSC. The maximal speed of the
The role of barotropic factors in shaping BC structure simulated currents in the shallow branch is 24 cm/s,
was investigated using barotropic model simulations. close to the speed of inflow current in Fram Strait
The simulated circulation in the barotropic experiment specified in the model as a boundary condition. The
closely follows bottom topography with a shallow-to- current speed reaches w8.5 cm/s in the core of the deep
right direction. Topographically-controlled behavior of branch and decreases gradually further toward the
the BC is in accordance with previously reported results deeper basin. The vertical structure of the AW flow in
of numerical simulations and theoretical considerations both branches is close to uniform with an insignificant
(Pedlosky, 1982; Nøst and Isachsen, 2003; Walin et al., (less than 0.5 cm/s) decrease of current speed in the
2004; Schlichtholz, 2007; Aaboe and Nøst, 2008; Aaboe bottom boundary layer (Fig. 5). The weak expression of
et al., 2009). The model simulated splitting of the the bottom boundary layer in the current structure may
incoming through Fram Strait AW flow into two be due to coarse vertical resolution, especially for the
branches. One branch is located at a shelf break with a deeper part of the domain.
typical depth of w200 m; we will refer to it as the The splitting of the AW inflow into two branches
“shallow branch”. Another branch which we will refer to expressed by the two velocity maximums seen on the
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 63

Fig. 5. Vertical cross-sections of simulated current speed (color scale; cm/s) along the AW pathway into the basin interior in the barotropic
experiment.

cross-slope sections coincides with an area of strong branching in Fram Strait (Aksenov et al., 2010). The
bottom topography changes off Svalbard where the effect of varying topography on the barotropic oceanic
isobaths diverge markedly (Fig. 3, black rectangle). A currents has been examined in numerous previous
strong topographically-driven transformation of the studies (e.g. Warren, 1963, 1969; Potter and Rattray,
AW inflow in the vicinity of Svalbard, where the inflow 1964; Robinson and Niiler, 1967; Niiler and
splits into several branches and re-circulates, is Robinson, 1967, and others). All authors confirmed
confirmed by current observations and CTD surveys that topographical effects are important in determining
(Perkin and Lewis, 1984; Aagaard et al., 1987; the path of propagation, separation, and meandering of
Quadfasel et al., 1987; Aagaard, 1989; Manley et al., the BC. For example, Warren (1969) used a linear
1992; Gascard et al., 1995; Manley, 1995; Rudels analytical model to demonstrate an eastward jet split-
et al., 2000; Schauer et al., 2004, 2008). Results of ting of a barotropic flow as it travels over diverging
high-resolution model simulations also confirm WSC isobaths on an f plane (f is the Coriolis parameter),
64 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

similar to what we found in our experiments. Warren 3.2.2. Baroclinic experiment


showed that the barotropic jet may split and follow the Experiments with varying baroclinic forcing
diverged isobaths, thus conserving potential vorticity. demonstrate that the two branches of the BC simulated
In order to examine the role of BC splitting caused by in the barotropic case experienced further trans-
divergence of isobaths in shaping the sea surface height, formation caused by baroclinic forces as shown at
h, the cross-slope gradients of h at several sections along several cross-slope sections along the AW pathway
the AW pathway were analyzed (Fig. 6). The analysis (Fig. 7). In contrast to the barotropic experiment, the
showed that the maximum currents in the deep branch shallow branch is shifted to a deeper part of the domain
coincide with the local minimums of the cross-slope h and retains its maximum speed of w24 cm/s. There is
gradient (i.e. the along-slope velocity is maximal where a weak transformation of the vertical structure of the
h has an inflection, Fig. 6). The model shows a rapid shallow branch compared to the barotropic case, with a
(w10% per 100 km) along-slope erosion of the h small (w0.5 cm/s) increase of current speed at an in-
gradient minimums in the deep branch, so that at termediate (w100 to 150 m) level.
w300 km from Fram Strait the h gradient vanishes and, The transformation of the simulated deep branch of
correspondingly, the maximum in the pattern of the AW flow by baroclinic forces was substantial. An
along-slope current disappears. Distribution of h is a isolated current maximum at an intermediate depth
result of the joint impact of numerous factors (i.e. bottom (w250 to 600 m) was formed (Fig. 7). This is a typical
topography, structure of the AW inflow, horizontal and feature of current structure resulting from a baroclinic
vertical friction), and thus has a complicated structure. balance of forces. Actually, there are two current
However, analysis of forces in the depth-integrated mo- maximums in the w250 to 600 m layer, but the dif-
mentum equation shows that friction within the near- ference between the two speeds is negligible and they
bottom Ekman layer may have primary importance for are indistinguishable in Fig. 7. The current speed in-
the along-slope decrease of current speed, which is creases from the surface down to w250 m; it remains
evident in the simulated deep BC branch (Fig. 5). For close to its maximum down to a level of w600 m,
example, bottom stress contributes up to w15% of the then gradually decreases to w4 cm/s in the bottom
net along-slope acceleration, in comparison with the boundary layer. The maximum current speed
Coriolis force, which is the major factor responsible for decreases slowly along the Eurasian Basin slope, from
maintaining geostrophic balance (Table 1). w9.5 cm/s at a distance of 165 km from Fram Strait
to w7.5 cm/s at a distance of 1320 km.
In order to estimate the total AW transport associ-
ated with each branch we determined the boundaries of
the branches as follows. The shallow branch spans
the area from the southern model boundary to a dis-
tance of w60 km where the current speed decreases
to w12 cm/s (50% of the core current velocity). The
southern border of the deep branch lies at a distance
of w80 km from the southern model boundary where
the current speed has a minimal value, whereas the
northern border of the deep branch was chosen at
w130 km from the shelf slope where the current speed
has decreased to 50% of the value in the core. The
vertical boundaries of the branches are taken at the
surface and at the bottom. With this definition, the
mean water transport in the deep branch reaches
w4.5 Sv (Sv ¼ 106 m3/s); that is w1.7 times stronger
than the shallow branch transport of w2.7 Sv. This
ratio holds when the branch width definitions change;
for example, it varies by only w20% when threshold
current speed is changed by 35%e70%. In addition to
Fig. 6. (Top) Simulated cross-slope sea-level (h) gradients (m/m)
along-slope weakening of the simulated baroclinic
and (bottom) along-slope surface current (cm/s) in the deep AW currents in the deep BC branch (Fig. 7), the model
branch at several locations in the barotropic experiment. demonstrates gradual horizontal expansion of the area
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 65

Fig. 7. Vertical cross-sections of simulated current speed (color scale; cm/s) along the AW pathway into the basin interior in the baroclinic
experiment.

of intensive currents in the shallow BC branch, so that Aaboe et al., 2009), the geostrophic balance dominates
a border of strong (>20 cm/s) currents is shifted to- the balance of forces for the along-slope component of
ward the deep ocean by w20 km while the BC prop- the deep simulated BC branch. This balance is deter-
agates from 165 to 1320 km away from Fram Strait. mined by the cross-slope gradient of the total pressure
This expansion is accompanied by an increase of total and the Coriolis force (Fig. 8). Pressure gradients are
barotropic transport associated within this branch from dominated by h gradients, which contribute up to 70% of
w2.5 to w3.4 Sv. This finding is in a good agreement the deep branch variability (compared to 30% caused by
with conclusions of Walin et al. (2004), who suggested the density-induced part of the pressure gradients). The
gradual shift of water transport from baroclinic to contribution of other factors, like nonlinear terms or
barotropic BC branch with along-slope propagation. vertical and horizontal mixing, was negligible.
As expected (e.g. Nøst and Isachsen, 2003; Walin The impact of density gradients on the deep-branch
et al., 2004; Schlichtholz, 2007; Aaboe and Nøst, 2008; along-slope current is not vertically uniform and may
66 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

in Fig. 9). The existence of a depth level with zero


density gradient is confirmed by the climatological
distribution of the density gradient near Svalbard
derived from the Environmental Working Group (1997)
climatology (Fig. 9, bottom). We note, however, that the
model generates a more complex density structure with
several depth levels where the density gradients change
signs, resulting in several velocity maximums. They
differ insignificantly, however, so that in Fig. 9 they look
like a single jet-like core.
Even in the baroclinic experiment, the simulated BC
is generally governed by barotropic factors. For
example, the distribution of the simulated sea-surface
elevation generally reflects the shape of bottom re-
lief, indicating topographically-controlled behavior of
the BC (Fig. 10). Deviations of isolines of the sea-
surface elevation from f/H (where H is the bottom
depth) contours occurs in areas of strong joint effect of
baroclinicity and relief (JEBAR; Sarkisyan and Ivanov,
Fig. 8. Balance of forces for the along-slope velocity component in 1971), which follows from conservation of potential
the core of the deep branch of AW flow in the Eurasian Basin vorticity. However, baroclinicity may play an impor-
(330 km from Fram Strait) in the baroclinic experiment. fU denotes tant role in determining, for instance, cross-slope water
Z z
the Coriolis term; Py ¼ gr0 ðvhðx; yÞ=vyÞ þ g ðvrðx; y; ~zÞ=vyÞd~z is transport between simulated deep and shallow BC
0
the cross-slope pressure gradient; Zy ¼ gr0 ðvhðx; yÞ=vyÞ is the cross- branches. Earlier Schlichtholz (2007) and Aaboe and
shelf sea-level gradient; other components are indistinguishable from Nøst (2008) suggested that along-isobath density gra-
zero with the choice of the horizontal axis scale. dients are dynamically important for the BC, as they
determine the vertical shear of the cross-isobath
be seen in Fig. 8 as a difference between acceleration component of the geostrophic flow and act as a
due to the total pressure gradient Py (black line) and driving agent for barotropic flow. Our numerical sim-
that due to the sea-level gradient Zy (red line). As it ulations show that the role of the along-isobath bottom
follows from the geostrophic relationship, a decrease density gradient in controlling BC dynamics may be
of the cross-slope pressure gradient leads to an increase not as important. For example, the results of the bar-
in the speed of the along-slope current component. oclinic experiment here show that the cross-slope water
Pressure gradients become stronger due to density in transport in the deep BC branch (Fig. 11, top), derived
the w250 to 600 m layer, which matches the layer of as an integral of the simulated cross-slope current ve-
maximal simulated current. The effect of sea-level locities (v) over the entire water column, has stronger
gradients is partially compensated by density gradi- correlation with depth-averaged density gradients
ents in the layer below w600 m, so the simulated (Fig. 11, red line at the bottom panel). The correlation
current decreases with depth. of depth-integrated water transport at the southern
It can be deduced from the geostrophic relationship border of the deep BC branch with the depth-integrated
that the maximum speed at an intermediate depth typical along-slope density gradient is 0.82; it becomes as
of the jet-like vertical structure of AW flow is reached lower as 0.6 for the near-bottom density gradient. The
when the cross-slope density gradients above and below stronger correlation of the depth-averaged cross-iso-
this depth have opposite signs. This finding is in accor- bath water transport with depth-integrated along-slope
dance with the theory which suggests that the increase in density gradient, compared to correlation with the
geostrophic velocity with depth may be determined by simulated near-bottom density gradients, is insensitive
the horizontal density gradient (Pedlosky, 1982). The to a choice of borders of the BC branches, remaining at
density distribution in the deep branch derived from the w0.2 to 0.3 higher.
simulated temperature and salinity did indeed show that We also note that temperature and salinity restoring
the depth at which the cross-slope density gradient used in the model does not have strong impact on
changes sign (black line in Fig. 9, top) is associated with behavior of the simulated BC. For example, the
the depth of the along-slope current maximum (red lines restoring term integrated over the deep BC branch in
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 67

Fig. 9. (Top) Vertical cross-section of the simulated cross-slope density gradient (105 kg/m4; positive if density increases toward the deep ocean)
and (bottom) cross-slope density gradients derived from Environmental Working Group (1997) climatology (105 kg/m4) 330 km downstream
from Fram Strait. Black lines show the depth of the cross-slope zero density gradient. Red isolines show simulated current speed (cm/s) in the deep
branch of AW flow.

the heat balance equation constitutes only w25% of from mainly barotropic in Fram Strait, to jet-like
the term, describing the horizontal advection of heat baroclinic flow with an intermediate-depth maximum
(Table 2). in the AW core along the Nansen Basin slope, to a
return to the original barotropic structure further
4. Discussion and general conclusions eastward, at the junction of the Lomonosov Ridge and
slope. Our modeling results, which captured this
Along-slope mooring-based observations showed a transformation well, suggest that the along-slope
substantial transformation of the Arctic Ocean BC, modification of the BC is controlled by the slope of

Fig. 10. Map of simulated sea-surface elevation (color, m) in the baroclinic experiment and distribution of bottom depth (isolines, m). Matching
of isolines of sea-surface elevation and bottom depth indicates topographically-controlled behavior of the BC.
68 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

segment of trajectory of the BC propagation within the


EB interior supports our conclusion that the trans-
formation of the BC shape is driven by changes of
along-flow buoyancy gradients. The simulated weak-
ening of currents in the deep branch of the BC
(w2 cm/s per 1000 km, Fig. 7) is not as strong as
determined from observations (Fig. 1), probably due to
stronger horizontal exchange with ambient waters than
was included in the model. Another reason is that the
climatological temperature and salinity used in our
simulations have coarse spatial resolution and are not
perfect in representing cross-slope density gradients.
Another possible mechanism by which barotropic BC
transport is weakened and transformed to baroclinic
transport is proposed by Aaboe and Nøst (2008) and is
based on the governing role of the along-slope gradi-
ents of bottom density.
Fig. 11. (Top) Along-slope (y ¼ 80 km) distribution of cross-isobath
We should note that, despite the results presented
depth-averaged water transport, and (bottom) along-slope gradients here and previous observational and modeling efforts,
of near-bottom (green) and depth-averaged (red) water density. our knowledge of BC spatial structure remains very
Cross-isobath water transport was derived as an integral of the limited. The baroclinic structure of the BC northeast of
simulated cross-slope current velocities (v) over the entire water Svalbard is confirmed by observations at three levels
column. The position of the section corresponds to the southern
border of the simulated deep BC branch in the baroclinic experiment.
only; thus our representation of this structure lacks
important specific features (like an additional current
speed maximum), which have been found, for example,
isopycnal surfaces with opposite cross-slope density with high-resolution observations at the M1 mooring.
gradient signs above and below the velocity maximum Quasi-synoptical LADCP-based measurements of the
(Fig. 12). Observations show that the current weakens BC are subject to high-frequency noise as discussed
by more than four times as it progresses along the above. At the same time, observations presented here
slope, from w24 cm/s in Fram Strait to w5 cm/s at complemented by modeling results clearly demonstrate
the Lomonosov Ridge. Weakening of the baroclinic the existence of the current speed maximum at inter-
component of the BC in the course of eastward BC mediate levels. We conclude that the baroclinic structure
propagation (Figs. 1 and 7) can be explained by a of the BC is evident along an extensive segment of the
sequential decrease of density contrast between the AW pathway from the north slope of Svalbard to the
AW and the ambient waters. We speculate that this northeastern part of the Laptev Sea.
decrease may also be due to horizontal exchange, so Our idealized model suggests that, in addition to the
that BC weakening should be intensified in areas of deep branch, the BC has one shallower branch. This
strong shelfebasin interaction (for example, on the shallow branch is located at a shelf break with a typical
slope of Severnaya Zemlya (Ivanov and Golovin, depth of w200 m and current speed of w24 cm/s in
2007; Walsh et al., 2007) or north of the Kara Sea, the core. The existence of the shallow AW branch in
where the Barents Sea branch of AW flow converges Fram Strait was confirmed by the results of numerical
with the Fram branch (Schauer et al., 2002). Good experiments with a high-resolution model, which
agreement between geostrophic (density-driven) and demonstrated a very complex WSC structure in the
observed mooring-based current speed over a large vicinity of Svalbard (Aksenov et al., 2010) and a dual-

Table 2
Contribution of terms (108,  C s1) in the heat transfer equation averaged over the simulated deep BC branch.
Along-slope distance, km Horizontal advection Vertical advection Horizontal diffusion Vertical diffusion Nudging term
165 1.83 0.42 0.13 1.73 0.31
660 0.82 0.18 0.09 0.7 0.22
1320 0.73 0.16 0.09 0.71 0.17
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 69

Walin et al., 2004 for details), our barotropic experi-


ment demonstrates the leading role of topography in
this splitting. Nevertheless, our model simulations
show that the deep core of the BC in the barotropic
experiment vanishes more rapidly in comparison with
baroclinic case, so that the two-core structure of the
BC cannot be identified at w660 km away from Fram
Strait (Fig. 4). Thus, we speculate that baroclinic
forcing incorporated via climatological temperature
and salinity distributions enhances the strength of the
BC and, probably, plays a crucial role in maintaining
of the two-core structure of the BC along the conti-
nental slope. However, we should note, that the prop-
erties of both these branches should be verified using
future observations.
LADCP observations across the Laptev Sea slope
showed that the maximum baroclinic currents corre-
sponds closely to the AW temperature core, suggesting
that cross-slope temperature observations may be used
to identify the BC position in the Eurasian Basin. We
understand that the idealized topography which is used
in the model experiments may lead to the absence of
some interesting and important features (like topo-
Fig. 12. Schematic of boundary current structure for (top) barotropic graphic eddies and local meandering) in the structure
case and (bottom) baroclinic case.
of the Arctic Ocean BC. However, we think that the
results of our simulations suit the purposes of this
study by successfully reproducing important properties
core structure of the AW flow along the EB shelf. of the BC structure.
However, the physical mechanisms, which are Knowledge of the pan-Arctic BC structure is vital
responsible for the generation of the shelf BC branch, for understanding rapid climate changes and the speed
are different. Aksenov et al. (2010, 2011) suggested with which these changes may propagate into the
that an influx of potential vorticity through the St. Arctic Ocean interior. However, our knowledge about
Anna Trough is the major driver of the generation of the structure of the BC and its branches in the Eurasian
the shelf BC branch. The splitting of two branches of Basin is still sketchy. For example, we still have no
AW inflow in our simulations is controlled by con- direct observational confirmation of the existence of
servation of the potential vorticity of the barotropic the shallow BC branch in the Eurasian Basin of the
flow and is governed by a strong divergence of Arctic Ocean. There is also a lack of observational
topography which occurs northeast of Svalbard (Fig. 1, information about along-slope heat and water trans-
black rectangle). In contrast to the deep branch, the ports. Thus, carefully-orchestrated large-scale sus-
shallow branch is mostly barotropic all the way along tained observations are critically important for further
the Eurasian slope. Even though the simulated shallow improving our understanding of BC structure and for
branch water transport is weaker than that of the deep assessing the ocean’s role in shaping the processes that
branch by w70%, it still constitutes a substantial water occur in the high-latitude regions.
transport which should be taken into account when
evaluating AW inflow into the Polar Basin interior. In Acknowledgments
contrast to Walin et al. (2004), who also found splitting
of the BC for two branches in their idealized model of This study was supported by NASA grant
the Nordic Seas and the Arctic Ocean, the mechanism #NNX08A050G (AP, IP, VI) and JAMSTEC (IP, VI)
of this splitting found in our simulations is different. grant. The authors thank A. Beszczynska-Möller,
Instead of water transport redistribution between the U. Schauer (both AWI, Germany), and R. Woodgate
baroclinic deep branch and barotropic shelf branch due (APL, USA) for help with current observations in Fram
to loss of buoyancy by the BC along its pathway (see Strait and at the Lomonosov Ridge.
70 A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71

References Gerdes, R., Karcher, M., Kauker, F., Schauer, U., 2003. Causes and
development of repeated Arctic Ocean warming events. Geophys.
Aaboe, S., Nøst, O.-A., 2008. A diagnostic model of the Nordic Seas Res. Lett. 30 (19). http://dx.doi.org/10.1029/2003GL018080.
and Arctic Ocean circulation: quantifying the effects of a variable Holloway, P.E., 1996. A numerical model of internal tides with
bottom density along a sloping topography. J. Phys. Oceanogr. application to the Australian North West Shelf. J. Phys. Ocean-
38, 2685e2703. ogr. 26, 21e37.
Aaboe, S., Nøst, O.A., Hansen, E., 2009. Along-slope variability of Hunke, E.C., Holland, M.M., 2007. Global atmospheric forcing data
barotropic transport in the Nordic Seas: simplified dynamics for Arctic ice-ocean modeling. J. Geophys. Res. 112, C04S14.
tested against observations. J. Geophys. Res. C 114, C03009. http://dx.doi.org/10.1029/2006JC003640.
http://dx.doi.org/10.1029/2008JC005094. Ivanov, V.V., Golovin, P.N., 2007. Observations and modeling of dense
Aagaard, K., Greisman, P., 1975. Toward new mass and heat budgets water cascading from the northwestern Laptev Sea shelf. J. Geo-
for the Arctic Ocean. J. Geophys. Res. 80, 3821e3827. phys. Res. 112, C09003. http://dx.doi.org/10.1029/2006JC003882.
Aagaard, K., Foldvik, A., Hillman, S.R., 1987. The West Spitsbergen Ivanov, V.V., Polyakov, I.V., Dmitrenko, I.A., Hansen, E.,
current: disposition and water mass transformation. J. Geophys. Repina, I.A., Kirillov, S.A., Mauritzen, C., Simmons, H.,
Res. 92, 3778e3784. Timokhov, L.A., 2009. Seasonal variability in Atlantic Water off
Aagaard, K., 1989. A synthesis of the Arctic Ocean circulation. Spitsbergen. Deep-Sea Res. I 56, 1e14.
Rapports et Proces-verbaux Réunions Conseil permanent Inter- Jones, E.P., 2001. Circulation in the Arctic Ocean. Polar Res. 20 (2),
national pour l’Exploration de la Mer 188, 11e22. 139e146.
Aksenov, Y., Bacon, S., Coward, A., Nurser, G., 2010. The North Jones, E.P., Rudels, B., Anderson, L.G., 1995. Deep waters of the
Atlantic inflow to the Arctic Ocean: high-resolution model study. Arctic Ocean: origins and circulation. Deep-Sea Res. I 42 (5),
J. Mar. Syst. 79, 1e22. 737e760.
Aksenov, Y., Ivanov, V.V., Nurser, A.J.G., Bacon, S., Polyakov, I.V., Karcher, M.J., Gerdes, R., Kauker, F., Köberle, C., 2003. Arctic
Coward, A.C., Naveira-Garabato, A.C., Beszczynska-Moeller, A., warming: evolution and spreading of the 1990s warm event in the
2011. The arctic circumpolar boundary current. J. Geophys. Res. Nordic seas and the Arctic Ocean. J. Geophys. Res. 108. http://
116, C09017. http://dx.doi.org/10.1029/2010JC006637. dx.doi.org/10.1029/2001JC001265.
Blindheim, J., 1989. Cascading of Barents Sea bottom water into the Karcher, M., Kauker, F., Gerdes, R., Hunke, E., Zhang, J., 2007. On
Norwegian Sea. Rapports et Proces-verbaux Réunions Conseil the dynamics of Atlantic Water circulation in the Arctic Ocean. J.
permanent International pour l’Exploration de la Mer 188, Geophys. Res. 112, C04S02. http://dx.doi.org/10.1029/
49e58. 2006JC003630.
Chen, C., Beardsley, R., 1995. A numerical study of stratified tidal Karcher, M., Smith, J.N., Kauker, F., Gerdes, R., Smethie Jr., W.M.,
rectification over finite-amplitude banks, part I: symmetric banks. 2012. Recent changes in Arctic Ocean circulation revealed by
J. Phys. Oceanogr. 25, 2090e2110. iodine-129 observations and modeling. J. Geophys. Res. 117,
Coachman, L.K., Barnes, C.A., 1963. The movement of Atlantic C08007. http://dx.doi.org/10.1029/2011JC007513.
Water in the Arctic Ocean. Arctic 16, 9e16. Killworth, P., 1992. An equivalent-barotropic mode in the fine res-
Dmitrenko, I.A., Polyakov, I.V., Kirillov, S.A., Timokhov, L.A., olution Antarctic model. J. Phys. Oceanogr. 22, 1379e1387.
Frolov, I.E., Sokolov, V.T., Simmons, H.L., Ivanov, V.V., Manley, T.O., Bourke, R.H., Hunkins, K.L., 1992. Near-surface
Walsh, D., 2008. Toward a warmer Arctic Ocean: spreading of circulation over the Yermak plateau in northern Fram Strait. J.
the early 21st century Atlantic Water warm anomaly along the Mar. Syst. 3, 107e125. http://dx.doi.org/10.1016/0924-7963(92)
Eurasian Basin margins. J. Geophys. Res. 113, C05023. http:// 90033-5.
dx.doi.org/10.1029/2007JC004158. Manley, T.O., 1995. Branching of Atlantic Water within the
Dmitrenko, I.A., Bauch, D., Kirillov, S.A., Koldunov, N., Greenland-Spitsbergen passage: an estimate of recirculation. J.
Minnette, P.J., Ivanov, V.V., Hölemann, J.A., Timokhov, L.A., Geophys. Res. 100, 20627e20634.
2009. Barents Sea upstream events impact the properties of Maslowski, W., Marble, D., Walczowski, W., Schauer, U.,
Atlantic Water inflow into the Arctic Ocean: evidence from 2005 Clement, J.L., Semtner, A.J., 2004. On climatological mass, heat,
to 2006 downstream observations. Deep-Sea Res. I 58, 513e527. and salt transports through the Barents Sea and Fram Strait from
Dmitrenko, I.A., Kirillov, S.A., Ivanov, V.V., Rudels, B., Serra, N., a pan-Arctic coupled ice-ocean model simulation. J. Geophys.
Koldunov, N.V., 2012. Modified halocline water over the Laptev Res. 109, C03032. http://dx.doi.org/10.1029/2001JC001039.
sea continental margin: historical data analysis. J. Clim. 25, McLaughlin, F.A., Carmack, E.C., Macdonald, R.W., Bishop, J.K.B.,
5556e5565. 1996. Physical and geochemical properties across the Atlantic/
Environmental Working Group (EWG), 1997. Joint U.S. e Russian Pacific water mass front in the Southern Canadian Basin. J.
Atlas of the Arctic Ocean [CD-ROM]. Natl. Snow and Ice Data Geophys. Res. C 101, 1183e1197.
Cent., Boulder, Colorado. Morison, J., Kwok, R., Peralta-Ferriz, C., Alkire, M., Rigor, I.,
Fahrbach, E., Meincke, J., Osterhus, S., Rohardt, G., Schauer, U., Andersen, R., Steele, M., 2012. Changing Arctic Ocean fresh-
Tverberg, V., Verduin, J., 2001. Direct measurements of volume water pathways. Science 481, 66e70. http://dx.doi.org/10.1038/
transport through Fram Strait. Polar Res. 20, 217e224. nature10705.
Gascard, J.-C., Richez, C., Rouault, C., 1995. New insights on large- Nansen, F., 1902. Oceanography of the North Polar Basin. The
scale oceanography in Fram Strait: the West Spitsbergen current. Norwegian Polar Expedition 1893-1896. Sci. Result 9, 427.
In: Smith, W. (Ed.), Oceanography of the Arctic: Marginal Ice Niiler, P.P., Robinson, A.R., 1967. Theory of free inertial currents. II.
Zones and Continental Shelves. In: Grebmeier, J. (Ed.), Coastal A numerical experiment for the path of the Gulf Stream. Tellus
and Estuarine Studies Series. American Geophysical Union, 19 (4), 601e619.
Washington, DC, pp. 131e182. Nikitin, M.M., 1969. Aperiodic currents of the Arctic Basin. Trudy
AANII 117, 101 (in Russian).
A.V. Pnyushkov et al. / Polar Science 7 (2013) 53e71 71

Nikolopoulos, A., Pickart, R.S., Fratantoni, P.S., Shimada, K., Confluence and redistribution of Atlantic Water in the Nansen,
Torres, D.J., Jones, E.P., 2009. The western Arctic boundary Amundsen and Makarov basins. Ann. Geophys. 20, 257e273.
current at 152 W: structure, variability, and transport. Deep-Sea Schauer, U., Fahrbach, E., Osterhus, S., Rohardt, G., 2004. Arctic
Res. II 56, 1164e1181. warming through the Fram Strait: oceanic heat transport from 3
Nøst, O.A., Isachsen, P.E., 2003. The large-scale time-mean ocean years of measurements. J. Geophys. Res. 109, C06026. http://
circulation in the Nordic Seas and Arctic Ocean estimated from dx.doi.org/10.1029/2003JC001823.
simplified dynamics. J. Mar. Res. 61, 175e210. Schauer, U., Beszczynska-Möller, A., Walczowski, W., Fahrbach, E.,
Orlanski, I., 1976. A simple boundary condition for unbounded Piechura, J., Hansen, E., 2008. Variation of flow through the
hyperbolic flows. J. Comp. Phys. 21, 251e269. Fram Strait to the Arctic Ocean between 1997 and 2006. In:
Pedlosky, J., 1982. Geophysical Fluid Dynamics. Springer-Verlag, Dickson, B. (Ed.), ArcticeSubarctic Ocean Fluxes. Springer,
New York. p. 624. Dordrecht, pp. 65e85.
Perkin, R.G., Lewis, E.L., 1984. Mixing in the West Spitsbergen Schlichtholz, P., 2007. Density-dependent variations of the along-
current. J. Phys. Oceanogr. 14, 1315e1325. isobath flow in the East Greenland Current from Fram Strait to
Pickart, R.S., 2004. Shelfbreak circulation in the Alaskan Beaufort Denmark Strait. J. Geophys. Res. 112, C12022. http://dx.doi.org/
Sea: mean structure and variability. J. Geophys. Res. 109 (C4), 10.1029/2006JC003987.
C04024. http://dx.doi.org/10.1029/2003JC001912. Shchepetkin, A.F., McWilliams, J.C., 2005. The Regional Ocean
Potter, G.H., Rattray, M., 1964. The influence of variable depth on Modeling System: a split-explicit, free-surface, topography
steady zonal barotropic flow. Dtsh. Hydrogr. Z. 17 (4), 164e174. following coordinates ocean model. Ocean Model. 9, 347e404.
Quadfasel, D., Gascard, J.-C., Koltermann, K.-P., 1987. Large-scale Thurnherr, A.M., 2008. How to Process LADCP Data with the LDEO
oceanography in Fram Strait during the 1984 Marginal Ice Zone Software. Available online at: ftp://ftp.ldeo.columbia.edu/pub/
experiment. J. Geophys. Res. 92, 6719e6728. LADCP/HOWTO.
Robinson, A.R., Niiler, P.P., 1967. Theory of free inertial currents. I. Timofeev, V.T., 1960. Water Masses of the Arctic Basin. Gidrome-
Path and structure. Tellus 19 (2), 269e291. teoizdat, Leningrad. p. 190 (in Russian).
Rodi, W., 1987. Examples of calculation methods for flow and Umlauf, L., Burchard, H., 2003. A generic length-scale equation for
mixing in stratified fluid. J. Geophys. Res. 92, 5305e5328. geophysical turbulence models. J. Mar. Res. 61, 235e265.
Rudels, B., 1986. The theta-S relations in the northern seas: impli- Walczowski, W., Piechura, J., Osinski, R., Wieczorek, P., 2005. The
cations for the deep circulation. Polar Res. 4, 133e159. West Spitsbergen Current volume and heat transport from syn-
Rudels, B., 1987. On the mass balance of the Polar Ocean with optic observations in summer. Deep-Sea Res. I 52 (8),
special emphasis on the Fram Strait. Nor. Polarinst. Skr. 188, 1374e1391.
1e53. Walin, G., Brostr, G., Nilsson, J., Dahl, O., 2004. Baroclinic boundary
Rudels, B., Jones, E.P., Anderson, L.G., Kattner, G., 1994. On the currents with downstream decreasing buoyancy; a study of an
intermediate depth waters of the Arctic Ocean. In: idealized Nordic Sea system. J. Mar. Res. 62, 517e543.
Johannessen, O.M., Muench, R.D., Overland, J.E. (Eds.), The Walsh, D., Polyakov, I.V., Timokhov, L.A., Carmack, E., 2007.
Polar Oceans and Their Role in Shaping the Global Environment. Thermohaline structure and variability in the eastern Nansen
Geophysical Monograph, vol. 85, pp. 33e46. Basin as seen from historical data. J. Mar. Res. 65, 685e714.
Rudels, B., Muench, R.D., Gunn, J., Shauer, U., Friedrich, H.J., 2000. Warren, B.A., 1963. Topographic influences on the path of the Gulf
Evolution of the Arctic Ocean boundary current north of the Stream. Tellus 15 (2), 167e183.
Siberian shelves. J. Mar. Syst. 25, 77e99. Warren, B.A., 1969. Divergence of isobaths as a cause of current
Rudels, B., 2012. Arctic Ocean circulation and variability e advec- branching. Deep-Sea. Res. 6, 339e355.
tion and external forcing encounter constraints and local Woodgate, R.A., Aagaard, K., Muench, R.D., Gunn, J., Björk, G.,
processes. Ocean Sci. 8, 261e286. http://dx.doi.org/10.5194/os- Rudels, B., Roach, A.T., Schauer, U., 2001. The Arctic Ocean
8-261-2012. boundary current along the Eurasian slope and the adjacent
Sarkisyan, A.S., Ivanov, V.F., 1971. Joint effect of baroclinicity and Lomonosov Ridge: water mass properties, transports and trans-
bottom relief as an important factor in the dynamics of sea cur- formations from moored instruments. Deep-Sea Res. I 48,
rents (in Russian). Izv. Acad. Sci. USSR, Atmos. Ocean. Phys. 7, 1757e1792.
116e124. Zhang, J., Steele, M., 2007. Effect of vertical mixing on the Atlantic
Schauer, U., Rudels, B., Jones, E.P., Anderson, L.G., Muench, R.D., Water layer circulation in the Arctic Ocean. J. Geophys. Res. 112,
Björk, G., Swift, J.H., Ivanov, V., Larsson, A.-M., 2002. C04S04. http://dx.doi.org/10.1029/2006JC003732.

You might also like