ArticlePDF Available

Abstract and Figures

The upper continental slope of the Storfjorden-Kveithola Trough Mouth Fans (NW Barents Sea) contains a several m-thick late Pleistocene sequence of plumites composed of laminated mud interbedded with sand/silt layers. Radiocarbon ages revealed that deposition occurred during about 130 years at a very high sedimentation rate of 3.4 cm a−1, at about 7 km from the present shelf break. Palaeomagnetic and rock magnetic analyses confirm the existence of a prominent, short-living sedimentary event. The plumites appear laterally continuous and were correlated with the sedimentary sequences described west of Svalbard and neighboring glacial depositional systems representing a major event at regional scale appointed to correspond to the deep-sea sedimentary record of Meltwater Pulse-1a. We also present new sedimentological and geochemical insights, and multi-beam data adding information on the palaeoenvironmental characteristics during MWP-1a and ice sheet decay in the NW Barents Sea.
This content is subject to copyright. Terms and conditions apply.
ORIGINAL ARTICLE
Marine sedimentary record of Meltwater Pulse 1a along the NW
Barents Sea continental margin
Renata Giulia Lucchi
1
Leonardo Sagnotti
2
Angelo Camerlenghi
1
Patrizia Macrı
`
2
Michele Rebesco
1
Maria Teresa Pedrosa
3
Giovanna Giorgetti
4
Received: 11 September 2015 / Accepted: 14 October 2015 / Published online: 20 November 2015
ÓThe Author(s) 2015. This article is published with open access at Springerlink.com
Abstract The upper continental slope of the Storfjorden-
Kveithola Trough Mouth Fans (NW Barents Sea) contains
a several m-thick late Pleistocene sequence of plumites
composed of laminated mud interbedded with sand/silt
layers. Radiocarbon ages revealed that deposition occurred
during about 130 years at a very high sedimentation rate of
3.4 cm a
-1
, at about 7 km from the present shelf break.
Palaeomagnetic and rock magnetic analyses confirm the
existence of a prominent, short-living sedimentary event.
The plumites appear laterally continuous and were corre-
lated with the sedimentary sequences described west of
Svalbard and neighboring glacial depositional systems
representing a major event at regional scale appointed to
correspond to the deep-sea sedimentary record of Melt-
water Pulse-1a. We also present new sedimentological and
geochemical insights, and multi-beam data adding infor-
mation on the palaeoenvironmental characteristics during
MWP-1a and ice sheet decay in the NW Barents Sea.
Keywords Meltwater Pulse 1a Plumites NW Barents
Sea Arctic
Introduction
Meltwater Pulse 1a
A meltwater pulse (MWP) is a short-lived, global accel-
eration in sea-level rise resulting from intense front- and/or
subglacial meltwater release, and/or surging ice streams
into oceans and iceberg discharge during ice sheets disin-
tegration [9]. It has been calculated that the rate of global
sea-level rise during meltwater pulses could have been as
high as 60 mm a
-1
during less than 500 years [29].
The existence of a pulsing mode for global sea-level rise
after LGM, was discovered through the study of drowned
late Quaternary reef-crest sequences cored in tropical areas
(Barbados, [29]; Caribbean-Atlantic reef province, [10];
Tahiti, [3,63]; Sunda Shelf, [32], Hawaii, [92]; among
others). Although high-resolution radiocarbon and
230
Th
dating of coral skeletons permit precise dating of the
events, diffuse neo-tectonic uplift of these areas often led
to controversial correlating results and difficult calculation
of the events’ magnitude.
Four main meltwater pulses have been pointed to have
occurred during the last deglaciation phase: MWP-19 ka,
also known as MWP-1a
0
, about 19 cal ka BP [15,93];
MWP-1a, 14.650–14.310 cal a BP [24]; MWP-1b,
11.500–11.000 cal a B.P. [4]; and MWP-1c at about
8.000 cal a BP [11,33]. Of these events, MWP-1a was
possibly the most prominent leading to a global sea-level
rise of about 20 m in the course of 340 a (5.9 cm a
-1
,
[24]).
Evidences of the existence MWP-1a have been found in
many low-latitude areas but straightforward evidence is
still lacking in polar areas where the event is thought to
have originated.
&Renata Giulia Lucchi
rglucchi@ogs.trieste.it
1
OGS (Istituto Nazionale di Oceanografia e di Geofisica
Sperimentale), Borgo Grotta Gigante, 42C, 34010 Sgonico,
TS, Italy
2
Istituto Nazionale di Geofisica e Vulcanologia (INGV), Via
di Vigna Murata 605, 00143 Rome, Italy
3
Instituto Andaluz de Ciencias de la Tierra, CSIC-Universidad
de Granada, Avda de las Palmeras 4,
18100 Armilla, Granada, Spain
4
Dipartimento di Scienze della Terra, Universita
`di Siena, via
Laterina 8, 53100 Siena, Italy
123
Arktos (2015) 1:7
DOI 10.1007/s41063-015-0008-6
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
An on-going controversy concerns the identification of
the trigger area and mechanisms responsible for MWP-1a
([1,5,6,1618,20,30,31,47,65,71,81,84,89,91],
among others). Four ice sheets are considered as possible
candidates responsible for shaping the global sea-level
curve: the Laurentide, Fennoscandian, Barents and
Antarctic ice sheets. Simulation of ocean-ice sheet inter-
actions indicates that different meltwater sources should
leave a geographically distinctive ‘sea-level fingerprint’
due to ice unloading histories and gravitational pull
between the shrinking ice masses and oceanic water-mas-
ses variations [18,22,62].
Scarce and ambiguous evidences of MWP-1a imprints
in the polar areas are puzzling and do not help resolve the
controversy on its origin.
In this study, we present marine sedimentary evidence
of MWP-1a recorded in the deep-sea sedimentary sequence
of the Trough Mouth Fan (TMF) offshore the Storfjorden
and Kveithola palaeo ice streams (Fig. 1). The deep-sea
areas of polar continental margins, beyond the continental
shelf brake controlled by ice grounding, are characterized
by low-energy continuous sediment accumulation unaf-
fected by glacial erosion, thus providing a valuable marine
sedimentary record of ice sheet dynamics. While Lucchi
et al. [58] gave a complete account of all the recovered
sedimentary units in the studied area, we here focus on the
interlaminated plumite unit and its significance as sedi-
mentological signature of Meltwater Pulse 1a.
Geological and oceanographic characteristics
of the study area
The Storfjorden and Kveithola TMF depositional systems
are located south of the Spitsbergen Island in the NW
Barents Sea [38,51], (Fig. 1). Seismo-stratigraphic infor-
mation in this area suggests that the onset of glacially
influenced sedimentation occurred since about 1.8 Ma,
when the Barents Sea Ice Sheet reached the continental
shelf edge [39,45]. Reconstruction of the late Pleistocene
and Holocene depositional processes of the marine sedi-
mentary record of the Storfjorden and Kveithola TMFs has
been described in Lucchi et al. [58] and it is briefly sum-
marized in the following (Fig. 2).
The older parts of the studied sediment cores contain
mass transport deposits (MTD) emplaced during LGM,
consisting of highly consolidated glacigenic diamicton
(Fig. 2a) and normally consolidated debris flows derived
Fig. 1 Location map of the
studied area with indication of
the main palaeo ice streams and
sediment cores location. Gray
hued bathymetry was acquired
during the SVAIS and
EGLACOM cruises that is
superimposed on IBCAO data
[41]. Adapted from
Camerlenghi et al. [12]
7Page 2 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
from the down-slope transport of glacigenic sediments
locally incorporating older stratigraphic intervals.
Above the LGM deposits, are proximal glacimarine
sediments containing abundant Ice Rafted Debris (IRD)
deposited between 20–19 and 17–16 cal ka BP (Fig. 2b).
The former interval contains an oxidized layer at its base
holding evidences of sea-ice free/seasonal conditions
(abundant planktonic foraminifera).
The glacimarine sediments are overlain by a progres-
sively fining-up sequence of interlaminated sediments
consisting of laminated mud interbedded with sand/silt
layers (Fig. 2c). The interlaminated sediments are several
m-thick on the upper slope, and only a few cm-thick on the
middle slope. Sedimentological and compositional char-
acteristics suggest that deposition occurred under the
effects of extensive subglacial meltwaters (plumites, [58]).
On the upper slope, the top boundary of the interlami-
nated sediments has a sharp contact with bioturbated and
slightly laminated distal glacimarine sediments (Fig. 2d)
indicating the presence of bottom currents and palaeo-en-
vironmental conditions favorable to bioactivity. On the
contrary, a massive IRD layer and distal glacimarine sed-
iments overlie the interlaminated facies on the middle
slope. The Holocene sequence on the middle slope consists
of diatom-bearing crudely laminated sediments (Fig. 2e),
and foraminifera/nannofossils-rich, heavily bioturbated
sediments both deposited under the effect of bottom cur-
rents (contourites, Fig. 2f) associated with the warm,
moderately saline Western Spitsbergen Current (WSC)
deriving from the North Atlantic Current. This current
flows northward along the continental slope transporting
heat to the Arctic area.
Materials and methods
Bathymetric, sub-bottom profiling data, and sediment cores
are the result of the merging of two data sets collected
during two cruises.
BIO Hespe
´rides IPY Cruise SVAIS (Longyearbyen,
July 29–August 17, 2007). Bathymetry was acquired with
multi-beam echo-sounder Simrad EM1002S. Data were
logged using Simrad’s Mermaid system and processed
with Caris HIPS and SIPS V 6.1.
Sub-bottom profiling employed a hull-mounted Kongs-
berg TOPAS PS 18 parametric profiler using two pri-
mary frequencies ranging between 16 and 22 kHz. Cores
were obtained with piston coring.
Fig. 2 Sediment facies and
related depositional processes
on the Storfjorden and
Kveithola TMFs after LGM.
The figure reports a synthetic
interpretative lithological log,
the radiographs and photos of
the identified sediment facies
after [58]
Arktos (2015) 1:7 Page 3 of 14 7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
R/V OGS Explora IPY Cruise EGLACOM (Kristian-
sund, July 07–August 04, 2008). Bathymetry was acquired
with multi-beam echo-sounder Reson MB8150 operating at
a mean frequency of 12 kHz, and MB8111, operating at a
mean frequency of 100 kHz. Data were logged and pro-
cessed with the Reson PDS2000V 2.5.3.2 software. Sub-
bottom profiling employed a hull-mounted Benthos CAP-
6600 profiler, using a sweep of acoustic frequencies rang-
ing between 2 and 7 kHz. Cores were obtained by gravity
coring.
The final bathymetric digital terrain model was pro-
duced from the combined data set with 75 m grid spacing.
Sub-bottom profiles from both surveys were imported in
Seismic Micro-Technology’s Kingdom Suite 8.3 software.
In addition, regional bathymetric imaging was obtained
from the Olex Ocean DTM
Ò
.
The sediment cores (Table 1) were logged through an
X-ray CT scan and an Avaatech XRF-core scan for chemical
composition of the sediments. For this contribution we
considered the down-core trends of Si (silica) content as
proxy of the quartz content (terrigenous input) rather than
amorphous silica forming diatoms frustules (assumption
made after sediments microscope investigation); the Cl
(chlorine) content was associated with the water content; the
Sr (strontium) content was used as indicator of bio-produc-
tivity in place of barium that in the Arctic areas is often of
detrital origin; and the S (sulfur) content, commonly asso-
ciated to oxygen depleted environments.
Discrete samples collected at every 5–10 cm resolution
were analyzed for water content, clay mineral analyses,
total and organic carbon (C
tot
,C
org
) and total nitrogen
(N
tot
) content. For this study, we considered only the dis-
tribution of the smectite content indicated to be, in the
studied area, a proxy of Atlantic Water strength [43],
whereas the C
org
/N
tot
ratio was used to distinguish between
marine and continental-derived organic matter according to
Meyers [61]. Details on the analytical procedure applied
for geochemistry and clay mineral analyses are indicated in
Lucchi et al. [58].
Palaeomagnetic investigations were performed on
u-channels collected from undisturbed sediments. The low-
field magnetic susceptibility (k) and the natural remanent
magnetization (NRM) were measured at 1 cm spacing. The
NRM was then stepwise demagnetized by alternating field
(AF) up to 100 mT. The characteristic remnant magneti-
zation (ChRM) has been isolated for each measured
interval and its direction determined by fitting a least-
square line on stepwise AF data with principal component
analysis, according to Kirschvink [44]. The maximum
angular deviation was then computed for each determined
ChRM direction. Since the cores were not azimuthally
oriented, the ChRM declination trends obtained for each
u-channel have been arbitrarily rotated to align their
average value with the true north.
From individual characteristic remnant magnetization
directions, we calculated the corresponding Virtual Geo-
magnetic Pole (VGP), under the assumption of a geocentric
axial dipole field.
The VGP scatter (S) during an established time interval
was estimated by the angular standard deviation of the
VGP distribution during an established interval of time. In
order to filter out large deviations that may not be repre-
sentative of genuine PSV, we applied a cut-off angle to the
computed VGPs following the iterative approach described
by Vandamme [86]. Additional details on the palaeomag-
netic analytical methodology applied are reported in [73].
The age model used for this study follows Sagnotti et al.
[73], and Lucchi et al. [58], relied on rocks palaeomagnetic
parameters and radiocarbon dating analyses (Table 2) cal-
ibrated with the calibration software program Calib 6.0
[82], using the marine09 calibration curve [70], and
applying an average marine regional reservoir effect
DR=84 ±23 (south of Svalbard). The mean values of
the calibrated age range of ±1rwere then normalized to
calendar year.
Results
Outer shelf iceberg scours
Sets of large parallel furrows are observed on the northern
flank of Spitsbergen Bank in water depths between 350 and
390 m (Fig. 3). The furrows are 10–15 m deep with respect
Table 1 Core location Core ID Lat N Lon E Depth (m) Location Lenght (m)
SV-02 75°13.707014°35.9600743 Upper-slope 6.41
SV-03 75°13.352014°37.2490761 Upper-slope 6.42
SV-04 74°57.425013°53.97201839 Middle-slope 3.03
SV-05 75°06.703015°13.3070713 Upper-slope 6.32
EG-02 75°12.907013°04.58701722 Middle-slope 3.05
EG-03 75°50.615012°58.35301432 Middle-slope 2.91
SV SVAIS project, EG EGLACOM project
7Page 4 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
to the surrounding seafloor and are bound by elongated
berms about 500 m wide and about 10 m high (Fig. 3).
Three or four parallel furrows compose each set of bottom
scours. Outer shelf furrows bear two elongation directions:
NE–SW that coincides with the main direction of the
Storfjorden glacial trough axis, and E–W. Lower resolution
Olex bathymetry suggests that NE–SW furrows extend
further landwards in water as shallow as 200 m in the
southwestern edge of the Storfjorden glacial trough mark-
ing the northwestern flank of Spitsbergenbanken (Figs. 1,
3). The largest set of furrows observed is 35 km long and
1 km wide (Fig. 3a).
Similar furrows have been identified on many polar
continental margins in water depth reaching 1000 m [25],
such as on the Yermak Plateau north of Svalbard [27],
offshore West Greenland [49], the Canadian margin [66],
in Pine Island Bay, Antarctica [40], and on the Argentine
continental margin [55]. They are produced by the seafloor
scouring of mega-scale iceberg keels (megabergs, [27,
87]).
These sets of large parallel furrows are not associated to
the sets of smaller (4–5 m deep and on average 300 m
wide), V-shaped, linear, subdued ridges and groves, with a
mean NE-SW orientation, conformable to the main Stor-
fjorden Trough axis, traceable for up to 20 km inland
(Fig. 3a, b). These sets of large furrows are rather inter-
preted as Mega Scale Glacial Lineations (MSGL) produced
by streaming ice and, like in adjacent Kveithola Trough
[69,72], loose their morphological expression near the
continental shelf edge due to the presence of moraine banks
intensely scoured by icebergs.
The large parallel furrows on the northern flank of
Spitsbergen Bank are instead interpreted being produced
by large, multi-keel icebergs released during break-off of
the Storfjorden sub-ice stream III fed by the Spitsbergen-
banken ice cap. Such icebergs were concentrated on the
southeaster part of the Storfjorden Trough because the
source of continental ice for this sub-ice stream was closes
the continental shelf break [64,58]).
Sediment geochemistry and composition
The characteristics for the sediments facies observed in the
upper and middle slope are very much consistent. For this
reason, in the following we will show only the down-core
Table 2 Radiocarbon dates (after Lucchi et al. [58])
Sample ID Lab ref. Sample type Description AMS
14
C±err. q13C cal a BP ±1r
Upper-slope
SV2-5-39/40 OS-77655 Foraminifera Benthic ?planktonic 15,050 ±50 -0.24 17,748 ±139
SV3-1-0/1 OS-77656 Foraminifera Benthic ?planktonic 4860 ±30 -0.07 5039 ±87
SV3-1-32/33 OS-82683 Mollusc Bivalve 13,000 ±45 1.09 14,929 ±141
SV3-6-21/30 OS-82684 Forams and ostracods Benthic ?plankt. ?ostrac. 13,200 ±50 -0.85 15,061 ±146
SV3-6-52/53 OS-77680 Foraminifera Benthic ?planktonic 13,300 ±50 -0.4 15,156 ±117
SV5-4-82/83 OS-82689 Foraminifera Mix plankt. mostly Nps 17,350 ±85 -0.08 20,055 ±166
Mid-slope
SV4-1-0/1 OS-77682 Foraminifera Nps 1100 ±25 0.44 594 ±36
SV4-2-11/12 OS-77683 Foraminifera Nps 4000 ±30 0.83 3896 ±56
SV4-2-48/49 OS-82685 Foraminifera Mixed planktonic 7110 ±30 0.5 7519 ±38
SV4-2-59/60 OS-77684 Foraminifera Nps 7880 ±45 0.5 8264 ±59
SV4-2-65/66 OS-77685 Foraminifera Nps 8180 ±35 0.33 8558 ±58
SV4-2-85/86 OS-82686 Foraminifera Mixed planktonic 8690 ±30 -0.44 9292 ±70
SV4-3-24/27 OS-82687 Foraminifera Bethic ?planktonic 9790 ±30 0.64 10,558 ±33
SV4-3-77/79 OS-82688 Foraminifera mixed planktonic 12,050 ±40 0.09 13,389 ±61
SV4-4-94/95 OS-77686 Foraminifera Nps 21,800 ±100 -0.07 25,438 ±241
EG2-1-30/31 OS-78387 Foraminifera Bethic ?planktonic 4570 ±130 -25 4665 ±164
EG2-1-90/91 OS-78389 Foraminifera Bethic ?planktonic 9460 ±180 0 10,235 ±234
EG2-2-60/61 OS-78383 Forams and pteropods Benthic ?plankt. ?pterop. 12,100 ±180 1.41 13,481 ±181
EG3-1-90/91 OS-78385 Foraminifera Bethic ?planktonic 4910 ±120 -25 5118 ±161
EG3-2-56/57 OS-78382 Foraminifera Bethic ?planktonic 8590 ±130 0.01 9147 ±167
EG3-3-38/39 OS-78324 Foraminifera Bethic ?planktonic 9740 ±80 0.73 10,508 ±87
Nps Neogloboquadrina pachyderma sin
Arktos (2015) 1:7 Page 5 of 14 7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
sedimentological characteristics of cores SV-02 and SV-04
as the most representative for the upper and middle slope,
respectively (Fig. 4).
Core SV-02 is characterized by a high C
org
/N
tot
ratio
always largely exceeding the boundary value of 10 (except
for the interval above the oxidized layer) indicating pre-
dominant continental-derived organic matter input. The
smectite content in the sediments located above the glaci-
genic diamicton increases gradually up-core, whereas it is
virtually absent in the LGM sediments.
The Si content is generally high throughout the core but
decrease in the upper 25 cm (distal glacimarine sediments).
Lower values are observed also between the oxidized layer
and the interval characterized by the presence of marine-
derived organic matter (388–400 cm bsf). The down-core
trend of Cl distribution is in agreement with the water
content measured on individual samples with Cl peaks
corresponding to the sandy layers in the interlaminated
sediments. The boundary between LGM sediments and the
proximal glacimarine sediments is characterized by a
salient change in Cl content. This boundary is also char-
acterized by abrupt changes in the other parameters
including the S content with high values within LGM
sediments that contain partially pyritized carbonate rock
fragments with framboid pyrite.
Core SV-04 is characterized by predominant marine-
derived organic matter input except for the proximal gla-
cimarine sedimentation (massive IRD input) and the
interlaminated sediment facies. The trend of smectite
content is characterized by a progressive increase from the
Fig. 3 Seabed morphology evidence of multi-keel mega-icebergs
scours. aSVAIS-EGLACOM bathymetry superimposed to the Olex
Ocean DTM
Ò
outlining the regional extent of the furrows beyond the
SVAIS-EGLACOM bathymetric coverage. Large parallel furrows are
outlined by thick black lines;bDETAILED bathymetry (SVAIS-
EGLACOM data sets) of multi-keel mega-icebergs scours. c3D view
and topographic profile across the multi-keel mega-icebergs scours
7Page 6 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
base of post-LGM deposition (260 cm bsf), to the
beginning of the distal glacimarine sedimentation (ca
200 cm bsf) after which the values are maintained sta-
bles with only minor fluctuations (minimum value at
202 cm bsf can be related to local sampling of IRD). The
trend of Si content presents three main steps: the glacial,
proximal glacimarine and interlaminated sediments have
the higher Si content, whereas minor values characterize
the distal glacimarine sedimentation. In the upper part of
the core, the Si content decreases within the crudely
layered sediments and became almost stable after about
9 cal ka BP. The Sr content is very low in the proximal
glacimarine and interlaminated sediments with only
slightly higher values in the distal glacimarine sediments.
On the contrary, the Sr content increases considerably in
the upper part of the core within the heavily bioturbated
sediments having a trend very similar to that of Ca con-
tent. The S content is generally high within the
glacimarine and interlaminated facies as well as the older
interval of the crudely layered sediments. A sharp
decrease of Svalues occurs at around 9 cal ka BP after
which the values increases again towards the top of the
core. The S content in the middle slope cores is, however,
one-order of magnitude lower than the one measured in
the upper slope cores.
Sediment palaeomagnetic properties
The geomagnetic angular dispersion related to Palaeosec-
ular Variations (PSV) should increase with latitude [21,60,
23]. At fixed latitudes (i.e., at a core site), the sampled time
interval is another factor affecting the measured Virtual
Geomagnetic Pole (VGP) dispersion. In general, samples
spanning a temporally large interval (e.g., [10 ka) that
fully sample the geomagnetic PSV define a large scatter of
Fig. 4 Down-core distribution of sedimentological and composi-
tional parameters measured in the studied cores. The depositional
significance of the synthetic lithological log is explained in Fig. 2.
Black bold numbers on the down-core logs refer to ages (cal ka BP)
according to the age model of Sagnotti et al. [73] and Lucchi et al.
[58]. Seismic Units A1, A2, B according to Pedrosa et al. [64]
Arktos (2015) 1:7 Page 7 of 14 7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
the measured VGPs, whereas samples referring to tempo-
rally short events under-represent the geomagnetic PSV
resulting in a limited VGP scatter (S).
Within the interlaminated sediments, the VGP scatter
computed from characteristic remnant magnetization
directions of SV-02 is very limited, with an Svalue of
9.3°(Fig. 5a), whereas in core SV-03 the Svalueisaslow
as 8.5°(Fig. 5b). The VGP scatter in the interlaminated
sediments of core SV-05 is, instead, as large as 15.8°
(Fig. 5c).
In core SV-04, spanning about the last 18 cal ka BP (or
up to 25 cal ka BP, considering the possibly reworked
sediments at the bottom of the core), the VGP scatter is
rather large, with an Svalue of 17°(Fig. 5d), that is
comparable with the VGP scatter determined for core EG-
02 (16.9°, Fig. 5e), spanning a slightly shorter time (last
16 cal ka BP). In core EG-03, that contains an expanded
Holocene sequence containing the record of the Younger
Dryas (YD) at its base, the VGP scatter is relatively small
(12.1°, Fig. 5f) with respect to core SV-04 and EG-02, but
still much higher than the values determined for the plu-
mites of cores SV-02 and SV-03.
Discussion
The NW Barents Sea sedimentary record
of MWP-1a
The over 4.5 meter thick interlaminated plumites derived
from meltwater discharge from the Storfjorden-Kveithola
glacial system, was deposited in about 130 years at a very
high sedimentation rate of 3.4 cm a
-1
, at about 7 km from
the present shelf break. This sedimentary signature has
been correlated to the MWP-1a event by Lucchi et al. [58]
based on sedimentological, geochemical and chronostrati-
graphic analysis.
Initial uncertainties on the stratigraphic assignment of such
deposit wererelated to (a) the slightly olderages determined in
core SV-03 (15,061 ±146–14,929 ±141 cal a BP), with
respect to the establish timing for the meltwater event
(14,650–14,310 cal a BP, [24]), and (b) the age range assigned
to the plumites (about 130 a) that is comparable with the
calibrated age error (±143 a).
The slightly older ages of the Storfjorden-Kveithola
upper slope plumites can be explained with an
Fig. 5 Distribution of the
Virtual Geomagnetic Pole
(VGP) within the interlaminated
sediments (yellow area) of cores
aSV-02, bSV-03, cSV-05, and
within the whole sedimentary
sequence of cores dSV-04 (last
18 cal ka BP), eEG-02 (last
16 cal ka BP), and fEG-03 (last
12 cal ka BP). S=VGP scatter
7Page 8 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
underestimation of the local regional reservoir correction
applied to the radiocarbon calibration of mixed benthic and
planktonic foraminifera. Sarnthein [75], Rae et al. [67], and
Thornalley et al. [84], indicated that deep ocean waters
after the LGM and during deglaciation were up to 1–2 ka
older than they are today determining a higher local
reservoir age. In our case the aging effect of the benthic
and planktonic foraminifera mix to explain the correlation
to the MWP-1a is only 400–600 a.
The plumites recorded in the SVAIS-EGLACOM cores
are constrained by an excellent palaeomagnetic and
stratigraphic correlation with the sedimentary sequences
described west of Svalbard and neighboring glacial depo-
sitional systems [42] where radiocarbon ages align with the
timing of MWP-1a as defined in tropical areas.
Stratigraphic equivalent deposits have been reported
from other areas offshore the West and North Svalbard
margin including the Yermak Plateau [7,13,28,42,68,
72], and the southern Barents Sea [88], indicating a nearly
synchronous regional event responsible for a massive
sediment input likely accompanied by a huge flux of fresh
meltwater into the northern Atlantic and Arctic Oceans.
The rapid emplacement of the interlaminated plumites is
supported beyond the radiocarbon dating by sediments
palaeomagnetic characteristics. According to various
models of geomagnetic Palaeosecular Variation (e.g., [23,
60,83]), the Virtual Geomagnetic Pole (VGP) scatter
representing the full geomagnetic Palaeosecular Variation
(PSV) spectrum of variability at the geographic latitudes of
the sampled cores, should be between 15°and 20°. At fixed
latitude (i.e., at a core site), the sampled time interval is
another factor affecting the retrieved VGP dispersion. In
general, samples spanning a temporally large interval
(e.g., [10 ka) fully sample geomagnetic PSV and provide
a reliable estimate of the VGP scatter, whereas samples
referring to temporally short events under-represent geo-
magnetic PSV resulting in a limited VGP scatter. The very
low VGP scatter measured in the interlaminated sediments
of cores SV-02 and SV-03 indicates that this stratigraphic
interval spans a time period that is not long enough to fully
represent the overall spectrum of variability of geomag-
netic PSV, notwithstanding a thickness of 3–4 m (Fig. 6).
This inference is consistent with the very high sedimenta-
tion rate indicated by radiocarbon dating. The higher val-
ues measured in the plumite record of core SV-05, located
at the head-scar of a landslide, were related to minor syn-
sedimentary re-depositional events occurred during the
plumites emplacement Sagnotti et al. [74].
The plumite layer is easily tracked on in the sub-bottom
profiler record of the 300 km long Storfjorden-Kveithola
margin. The thickness is estimated, using a 1500 m s
-1
average sound velocity in such water-rich sediments, to be
over 20 m-thick in the SE end of the Storfjorden TMF. The
thickness decreases rapidly down-slope where it becomes
negligible at about 35–40 km form the present shelf break
in about 1800 m water depth. The NW end of the Stor-
fjorden TMF is characterized by a thinner plumite
sequence due to faster withdrawal of the ice stream in that
area [58].
The extent of the deglaciation seismic Unit A was
mapped on the Storfjorden-Kveithola TMFs and the de-
compacted sedimentation rate averaged over 19.5 ka was
calculated to be 0.6 kg m
-2
a
-1
[54]. If we scale these
values to the MWP-1a depositional event, volumetrically
forming about 90 % of seismic Unit A, and considering a
duration of about 130 years, as measured in our cores, the
sediment mass accumulation rate results as high as
78 kg m
-2
a
-1
, corresponding to a total sediment mass
accumulation of 1.1 910
11
tonnes on the upper conti-
nental slope.
These values further stress the extreme nature of the
MWP-1a sedimentary event recorded on the upper slope of
Storfjorden-Kveithola TMFs.
Forcing mechanisms and environmental conditions
during MWP-19 ka and MWP-1a
The presence of clustered glacigenic debrites forming the
TMF of the North-western Barents Sea continental margin
confirms fully glaciated conditions of the shelf area during
the LGM [2,50,64]. Dating of these deposits indicate the
grounded ice streams reached the shelf edges at ca.
24 cal ka BP ([26,42], Fig. 7a).
There is a large consensus on considering enhanced
summer insulation as the primary mechanism determining
the onset of the northern hemisphere deglaciation. Intense
Fig. 6 Models of VGP scatter values (S) with respect to geographic
latitude according to Vandamme [86], McElhinny and McFadden
[59], Tauxe and Kent [83]. Yellow circles indicate the Svalues from
the plumite interval of cores SV-02, SV-03 and SV-05, whereas red
circles indicate the Svalues for the sedimentary sequence of cores
SV-04, EG-02, and EG-03
Arktos (2015) 1:7 Page 9 of 14 7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
ice-melting from the northern hemisphere ice sheets and
mountain glaciers produced a large volume of meltwaters
with deposition of a several m-thick laminated sequence on
the northern continental margins (e.g., [36,37,52,85]).
This event produced a first abrupt global sea-level rise
(MWP-19 ka) and a renewal of the Atlantic Meridional
Overturning Circulation [15,19,34].
The oxidized and massive IRD layer overlying LGM
sediments in the SVAIS-EGLACOM cores, records the
Storfjorden ice stream response to MWP-19 ka responsible
for breakdown of the outer part of the grounded ice shelf
(Fig. 7b). Oxidized sediments, characterized by low sulfur
content, presence of planktonic foraminifera, and marine-
derived organic matter indicate deep ocean ventilation with
ice-free (or seasonal) surface conditions. These environ-
mental characteristics support the re-establishment of a
vigorous thermohaline circulation in the north Atlantic after
the LGM.
According to Llopart et al. [53], the rapid sea-level rise
during MWP-19 ka was also responsible for local slope
instability due to destabilized distribution of slope inter-
stitial pore pressure along the Storfjorden-Kveithola TMFs,
due to the rapid increase of sea-hydrostatic pressure
(slump/debrites involving, or among, the oxidized layer in
cores SV-05 and SV-04).
Detachment and lift-off of the marine-based grounded
ice streams on the deep Antarctic shelves of the South-
eastern Weddell Sea, Antarctic Peninsula and Amundsen
Sea in response to sea-level forcing during MWP-19 ka,
lead to the deposition of a thick laminated sequence
underneath the Antarctic ice shelves [35,90,91]. Antarctic
meltwater release was in turn responsible for (1) a further
strengthening of the Atlantic meridional overturning cir-
culation accelerating the climate warming of the North
Atlantic region (onset of the Bølling-Allerød warm inter-
val, [89]) and (2) promoted sea-level rise forcing lift-off of
shallow marine grounded ice streams including the Stor-
fjorden and Kveithola ice streams in the northern margins.
Evidence of renewed North Atlantic warm inflows at the
onset of the Bølling warm interval were reported in the
south Svalbard [42,68], west and northern Svalbard mar-
gins [13,14,46,79,80], the Franz Victoria Trough [48,56]
and the NW Yermak slope [7].
Ice shelves are more sensitive to warm sub-glacial
inflows forcing their rapid melting and retreat [76]. Re-
enhanced warm North Atlantic inflows were responsible
for fast decay of the Storfjorden and Kveithola ice streams
with extensive ice-melting and retreat to an inner ground-
ing line, contributing significantly to global sea-level rise
during MWP-1a (Fig. 7c).
Contemporaneous intense calving from the retreating ice
stream in the Storfjorden Trough is inferred to have gen-
erated the release of multi-keel megabergs responsible for
the mega-scours visible on the Storfjorden outer conti-
nental shelf in water depth as deep as 390 m below present-
day sea level. The chronology of this event fits well with
the modeled 15,000 ka peak flux of iceberg across the
Storfjorden Trough [77].
Melting of the main Storfjorden ice streams (I and II)
occurred with a rapid retreat to the shallower grounding
zone wedges located on the middle shelf (thinner plumites
on the slope facing lobe I and II, Fig. 1). Diffuse melting
and rapid retreat in this area was likely responsible for a
thinner ice shelf, producing thinner icebergs unable of deep
scouring the seafloor. On the contrary, the Kveithola and
the south-western Storfjorden ice stream III, were probably
active longer as they were fed by a closer ice catchment
Fig. 7 Schematic graphic representation of the three main deposi-
tional events recording the Storfjorden and Kveithola ice stream
dynamics in the SVAIS-EGLACOM sediment cores. aLGM,
binception of ice stream decay during MWP-19 ka; cice stream
collapse during/after MWP-1a
7Page 10 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
area (Spitsbergenbanken ice cap), producing a larger
sequence of plumites on the continental slope facing lobe
III and the Kveithola Trough.
The high sulfur content measured in the Storfjorden
interlaminated sediments indicates reduced sea bottom
ventilation likely caused by strong water stratification
under extensive fresh water release during MWP-1a. It is
also possible that the presence of surface fresh water
enhanced multi-year sea ice formation during the plumite
deposition that would explain the little occurrence of IRD
within the interlaminated sediments. Strong water stratifi-
cation would contribute to reduce nutrient availability
hampering the bio-productivity and to promote poor deep-
water ventilation affecting the preservation of calcareous
foraminifera tests in the sediments, already diluted by the
strong detrital input (barren sediments).
In addition, surface water stratification would enhance
the areal distribution and sediment dispersion of the melt-
water plumes by reducing sediment flocculation in the
fresh water and particle settling across the halocline surface
(c.f. [36,57]). This explains the wide spread occurrence of
fine-grained laminated sediments having considerable
thickness only on the upper slope of TMFs in the area close
to the sediment source output.
The interlaminated plumite sequence is sharply topped
by a massive IRD layer that records the collapse of the
outer part of the Storfjorden and Kveithola ice shelves (ca
14.2 cal ka BP, [8,72]). The abrupt disappearance of the
main TMFs source of terrigenous input is responsible for a
strong change in the sedimentation rate with starving TMFs
receiving sediments only through icebergs, wind, and
along-slope contour currents (sharp decrease in Si content
at 14.3 cal ka BP, Fig. 4). The sediment composition
changes rapidly with increasing content of smectite and a
swift change in organic matter composition being of mar-
ine origin (C
org
/N
tot
ratio \10). We think that the abrupt
‘switch-on’’ of the thermohaline circulation described to
occur after melt water pulses went together with the ‘shut-
down’ of the sediment discharge and starving of TMFs
affecting the Svalbard margins [13,42,78], the northern
Barents Sea (Franz Victoria Trough, [48]) and part of the
Arctic Ocean (NW Yermak slope, [7]) suggesting a major
sudden decay of the ice sheet in the NW Barents Sea.
Conclusions
Decay and retreat of the Storfjorden and Kveithola palaeo-
ice streams after the LGM were reconstructed through
sedimentological and geochemical analysis of deep-sea
marine sediments and new multi-beam and sub-bottom
profiler data analysis.
The sedimentary record on the upper continental slope
contains a several m-thick sequence of plumites deposited
under an extensive meltwater event associated to the
palaeo-ice streams decay. Radiometric dating and rock
palaeomagnetic characteristics indicate a very fast event
(ca. 130 years) responsible for the settling of about
1.1 910
11
tonnes of sediments on the upper slope of the
Storfjorden-Kveithola TMFs with an extreme sedimenta-
tion rate of 3.4 cm a
-1
. New data constrains confirm an
early interpretation that related the event to the Meltwater
Pulse 1-a.
New sedimentological and geochemical insights indi-
cate that strong seawater stratification during extensive
meltwater release promoted a wide dispersion of fine-
grained sediments over the continental margin. Consistent
thicknesses of the plumites are recorded only on the upper
slope of TMFs, close to the sediment source output when
the Storfjorden and Kveithola ice streams were grounded at
the continental shelf brake during LGM. This would
explain the difficulties encountered to find a substantial
‘MWP-1a deposit’’ in different settings of the Arctic, as
well as Antarctic, areas.
A major ice sheet collapse at the end, or during, the
MWP-1a produced multi-keel megabergs scouring the
outer Storfjorden continental shelf, followed by TMFs
sediment starving that signed the reprise of the biogenic
activity and strengthening of the Termohaline Circulation
(THC). The whole sequence of THC ‘swich on’’—ex-
treme short-term accumulation of fine sediments with
apparent THC ‘shut-down’’—restart of the THC and
reduction of sediment delivery to the slope, described
offshore the Storfjorden-Kveithola glacial troughs, was
observed in many other areas of the NW Barents Sea and
NW Yermak slope, suggesting the whole margin under-
went an almost synchronous depositional fate with a
major sudden decay of the ice sheet after, or during, the
MWP-1a event.
Acknowledgments This study was supported by the Spanish IPY
projects SVAIS (POL2006-07390/CGL) NICE STREAMS Spain
(CTM2009-06370-E/ANT), and DEGLABAR (CTM2010-17386),
the Italian projects OGS-EGLACOM, and PNRA-CORIBAR (PdR
2013/C2.01). We thank ENI E&P Division (Milan, Italy) for the
analysis with the X-ray CT scan, and the scientific teams of SVAIS,
EGLACOM and CORIBAR projects. We are grateful to Henning
Bauch, Christian Hass and another anonymous reviewer for com-
ments and suggestions that greatly improved the manuscript.
Open Access This article is distributed under the terms of the
Creative Commons Attribution 4.0 International License (http://crea
tivecommons.org/licenses/by/4.0/), which permits unrestricted use,
distribution, and reproduction in any medium, provided you give
appropriate credit to the original author(s) and the source, provide a
link to the Creative Commons license, and indicate if changes were
made.
Arktos (2015) 1:7 Page 11 of 14 7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
References
1. Alley RB, Clack PU, Huybrechts P, Jounghin I (2005) Ice-sheet
and sea-level changes. Science 310:456–460
2. Andreassen K, Nilssen LC, Rafaelsen B, Kuilman L (2004)
Three-dimensional seismic data from the Barents Sea margin
reveal evidence of past ice streams and their dynamics. Geology
32(8):729. doi:10.1130/G20497.1
3. Bard E, Hamelin B, Arnold M, Montaggioni L, Cabioch G, Faure
G, Rougerie F (1996) Deglacial sea-level record from Tahiti corals
and the timing of global meltwater discharge. Nature 382:241–244
4. Bard E, Hamelin B, Delanghe-Sabatier D (2010) Deglacial
Meltwater Pulse 1B and Younger Dryas Sea Levels Revisited
with Boreholes at Tahiti. Science 327:1235–1237
5. Bentley MJ, Fogwill CJ, Le Brocq AM, Hubbard AL, Sugden
DE, Dunai TJ, Freeman S (2010) Deglacial history of the West
Antarctic Ice Sheet in the Weddell Sea embayment: constraints
on past ice volume change. Geology 38:411–414
6. Bentley MJ, O’Cofaigh C, Anderson JB et al (2014) A commu-
nity-based geological reconstruction of Antarctic Ice Sheet
deglaciation since the Last Glacial Maximum. Quat Sci Rev
100:1–9. doi:10.1016/j.quascirev.2014.06.025
7. Birgel D, Hass HC (2004) Oceanic and atmospheric variations
during the last deglaciation in the Fram Strait (Arctic Ocean): a
coupled high-resolution organic-geochemical and sedimentolog-
ical study. Quat Sci Rev 23:29–47
8. Bjarnado
´ttir LR, Ru
¨ther DC, Winsborrow MCM, Andreassen K
(2013) Grounding-line dynamics during the last deglaciation of
Kveithola, W Barents Sea, as revealed by seabed geomorphology
and shallow seismic stratigraphy. Boreas 42:84–107
9. Blanchon P (2011) Meltwater pulses. In: Hopley David (ed)
Encyclopedia of Modern Coral Reefs: Structure, form and pro-
cess. Springer, Encyclopedia of earth science series, pp 683–690
10. Blanchon P, Shaw J (1995) Reef drowning during the last
deglaciation: evidence for catastrophic sea-level rise and ice-
sheet collapse. Geology 23:4–8
11. Blanchon P, Jones B, Ford DC (2002) Discovery of a submerged
relic reef and shoreline off Grand Cayman: further support for an
early Holocene jump in sea level. Sediment Geol 147:253–270
12. Camerlenghi A, Rebesco M, Accettella D (2015) Trough-mouth
fan, Storfjorden. In: Dowdeswell JA, Canals M, Jakobsson M,
Todd BJ, Dowdeswell EK, Hogan KA (eds) Atlas of Submarine
Glacial Landforms, Mem Geol Soc London (in press)
13. Chauhan T, Rasmussen TL, Noormets R (2015) Palaeoceanog-
raphy of the Barents Sea continental margin, north of Nordaust-
landet, Svalbard, during the last 74 ka. Boreas. doi:10.1111/bor.
12135
14. Chauhan T, Rasmussen TL, Noormets R, Jakobsson M, Hogan
KA (2014) Glacial history and paleoceanography of the southern
Yermak Plateau since 132 ka BP. Quat Sci Rev 92:155–169.
doi:10.1016/j.quascirev.2013.10.023
15. Clark PU, McCabe AM, Mix AC, Weaver AJ (2004) Rapid rise
of sea level 19,000 years ago and its global implications. Science
304:1141–1144
16. Clark PU, Alley RB, Keigwin LD, Licciardi JM, Johnsen SJ,
Wang H (1996) Origin of the first global meltwater pulse fol-
lowing the Last Glacial Maximum. Paleoceanography
11:563–577
17. Clark PU, Arthur SD, Shakun JD, Carlson AE, Clark J, Wohlfarth
B, Mitrovica JX, Hostetler SW, McCabe AM (2009) The Last
Glacial Maximum. Science 325:710–714
18. Clark PU, Mitrovica JX, Milne GA, Tamisiea ME (2002) Sea
level finger printing as a direct test for the source of global
meltwater pulse 1a. Science 295:2438–2441
19. Clark PU, Pisias NG, Stocker TF, Weaver AJ (2002) The role of
the thermohaline circulation in abrupt climate change. Nature
415:863–869
20. Clark PU, Shakun JD, Baker PA et al (2012) Global climate
evolution during the last deglaciation. P Natl Acad Sci USA
109(19):E1134–E1142
21. Cox A (1970) Latitude dependence of the angular dispersion of
the geomagnetic field. Geophys J Roy Astr S 20:253–269
22. Crucifix M, Berger A (2002) Simulation of ocean-ice sheet
interactions during the last deglaciation. Paleoceanography
17(4):1054
23. Deenen MH, Langereis CG, van Hinsbergen DJ, Biggin AJ
(2011) Geomagnetic secular variation and the statistics of
palaeomagnetic directions. Geophys J Int 186(2):509–520
24. Deschamps P, Durand N, Bard E, Hamelin B, Camoin G, Thomas
AL, Henderson GM, Okuno J, Yokoyama Y (2012) Ice-sheet
collapse and sea-level rise at the Bølling warming 14,600 years
ago. Nature 483:559–564
25. Dowdeswell JA, Bamber JL (2007) Keel depths of modern
Antarctic icebergs and implications for sea-floor scouring in the
geological record. Mar Geol 243:120–131
26. Dowdeswell JA, Elverhøi A (2002) The timing of initiation of
fast-flowing ice streams during a glacial cycle inferred from
glacimarine sedimentation. Mar Geol 188(1–2):3–14
27. Dowdeswell JA, Jakobsson M, Hogan KA, O’Regan M, Back-
manb J, Evans J, Hell B, Lo
¨wemark L, Marcussen C, Noormets
R, Cofaigh CO
´, Selle
´nE,So
¨lvsten M (2010) High-resolution
geophysical observations of the Yermak Plateau and northern
Svalbard margin: implications for ice-sheet grounding and deep-
keeled icebergs. Quat Sci Rev 29:3518–3531
28. Elverhøi A, Andersen ES, Dokken T, Hebbeln D, Spielhagen R,
Svendsen JI, Sørflaten M, Rørnes A, Hald M, Forsberg CF (1995)
The growth and decay of the Late Weichselian ice sheet in
western Svalbard and adjacent areas based on provenance studies
of marine sediments. Quat Res 44:303–316
29. Fairbanks RG (1989) A 17.000-year glacio-eustatic sea level
record: influence of glacial melting rates on the Younger Dryas
event and deep-ocean circulation. Nature 342:637–642
30. Golledge NR, Menviel L, Carter L, Fogwill CJ, England MH,
Cortese G, Levy RH (2014) Nat Comm 5:5107. doi:10.1038/
ncomms6107
31. Gregoire LJ, Payne AJ, Valdes PJ (2012) Deglacial rapid sea
level rises caused by ice-sheet saddle collapses. Nature
487:219–222
32. Hanebuth T, Stattegger K, Grootes PM (2000) Rapid flooding of
the Sunda Shelf: a late-glacial sea-level record. Science
288:1033–1035
33. Harris PT, Heap AD, Marshall JF, McCulloch M (2008) A new
coral reef province in the Gulf of Carpentaria, Australia:
colonisation, growth and submergence during the early Holocene.
Mar Geol 251:85–97
34. He F, Shakun JD, Clark PU, Carlson AE, Liu Z, Otto-Bliesner
BL, Kutzbach JE (2013) Northern Hemisphere forcing of
Southern Hemisphere climate during the last deglaciation. Nature
494:81–85
35. Heroy DC, Anderson JB (2007) Radiocarbon constraints on
Antarctic Peninsula Ice Sheet retreat following the Last Glacial
Maximum (LGM). Quat Sci Rev 26:3286–3297
36. Hesse R, Khodabakhsh S, Klauck I, Ryan WBF (1997) Asym-
metrical turbid surface plume deposition near ice-outlets of the
PleistoceneLaurentide ice sheet in the Labrador Sea. Geo-Mar
Lett 17:179–187
37. Hesse R, Klauck I, Khodabakhsh S, Piper D (1999) Continental
slope sedimentation adjacent to an ice margin. III. The upper
Labrador Slope. Mar Geol 155:249–276
7Page 12 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
38. Hjelstuen BO, Elverhøi A, Faleide JI (1996) Cenozoic erosion
and sedimentary yield in the dranaige area of the Storfjorden Fan.
Global Planet Change 12:95–116
39. Hjelstuen BO, Sejrup HP, Haflidason H, Nyga
˚rd A, Ceramicola
S, Bryn P (2005) Late Cenozoic glacial history and evolution of
the Storegga Slide area and adjacent slide flank regions, Nor-
wegian continental margin. Mar Petrol Geol 22:57–69
40. Jakobsson M, Anderson JB, Nitsche FO et al (2011) Geological
record of ice shelf break-up and grounding line retreat. Geology,
Pine Island Bay. doi:10.1130/G32153.1
41. Jakobsson M, Mayer L, Coakley B et al (2012) The international
bathymetric chart of the Arctic ocean (IBCAO) version 3.0.
Geophys Res Lett 39:L12609
42. Jessen SP, Rasmussen TL, Nielsen T, Solheim A (2010) A new
Late Weischselian and Holocene marine chronology for the
western Svalbard slope 30,000–0 cal years BP. Quat Sci Rev
29:1301–1312
43. Junttila J, Aagaard-Sørensen S, Husum K, Hald M (2010) Late
Glacial-Holocene clay minerals elucidating glacial history in the
SW Barren Sea. Mar Geol 276:71–85
44. Kirschvink JL (1980) The least-squares line and plane and the
analysis of palaeomagnetic data. Geophys J Int 62:699–718
45. Knies J, Matthiesson J, Vogt C, Laberg JS, Hjelstuen BO,
Smelror M, Larsen E, Andreassen K, Eidvin T, Vorren TO (2009)
The Pliocene glaciation of the Barents Sea-Svalbard region: a
new model based on revised chronostragigraphy. Quat Sci Rev
28:812–829
46. Koc¸ N, Klitgaard-Kristensen D, Hasle K, Forsberg CF, Solheim
A (2002) Late glacial palaeoceanography of Hinlopen Strait,
northern Svalbard. Polar Res 21:307–314
47. Kopp RE (2012) Tahitian record suggests Antarctic collapse.
Nature 483:549–550
48. Kleiber H, Knies J, Niessen F (2000) The Late Weichselian
glaciation of the Franz Victoria Trough, northern Barents Sea: ice
sheet extent and timing. Mar Geol 168:25–44
49. Kuijpers A, Dalhoff F, Brandt MP, Hu
¨mbs P, Schott T, Zotova A
(2008) Giant iceberg plow marks at more than 1 km water depth
offshore West Greenland. Mar Geol 246:60–64
50. Laberg JS, Vorren TO (1995) Late Weichselian submarine debris
flow deposits on the Bear Island Trough Mouth Fan. Mar Geol
127:45–72
51. Laberg JS, Vorren TO (1996) The glacier-fed fan at the mouth of
Storfjorden trough, western Barren Sea: a comparative study.
Geol Rundsch 85:338–349
52. Lekens WAH, Sejrup HP, Haflidason H, Petersen GØ, Hjelstuen
B, Knorr G (2005) Laminated sediments preceding Heinrich
event 1 in the Northern North Sea and Southern Norwegian Sea:
origin, processes and regional linkage. Mar Geol 216:27–50
53. Llopart J, Urgeles R, Camerlenghi A, Lucchi RG, De Mol B,
Rebesco M, Pedrosa MT (2014) Slope Instability of Glaciated
Continental Margins: Constraints from Permeability-Compress-
ibility Tests and Hydrogeological Modeling Off Storfjorden, NW
Barents Sea. In: Krastel S et al (eds) Submarine Mass Movements
and Their Consequences, Advances in Natural and Technological
Hazards Research, vol 37. Springer Science book series,
pp 95–104
54. Llopart J, Urgeles R, Camerlenghi A, Lucchi, RG, Rebesco M,
De Mol B (2015) Development of the Storfjorden and Kveithola
Trough Mouth Fans, North-Western Barents Sea. Quat Sci Rev
129:68–84
55. Lo
´pez-Martı
´nez J, Mun
˜oz A, Dowdeswell JA, Line
´s C, Acosta J
(2011) Relict sea-floor ploughmarks record deep-keeled Antarctic
icebergs to 45°S on the Argentine margin. Mar Geol 288:43–48
56. Lubinski DJ, Korsun S, Polyak L, Forman SL, Lehman SJ,
Herlihy FA, Miller GH (1996) The last deglaciation of the Franz
Victoria Trough, northern Barents Sea. Boreas 25:89–100
57. Lucchi RG, Rebesco M (2007) Glacial contourites on the
Antarctic Peninsula margins: insight for palaeoenvironmental and
palaeoclimatic conditions. In: Viana AR Rebesco M (eds), Eco-
nomic and Palaeosignificance of Contourite Deposits. Geol Soc
London Special Publ 276:111–127
58. Lucchi RG, Camerlenghi A, Rebesco M et al (2013) Postglacial
sedimentary processes on the Storfjorden and Kveithola TMFs:
impact of extreme glacimarine sedimentation. Global Planet
Change 111:309–326. doi:10.1016/j.gloplacha.2013.10.008
59. McElhinny MW, McFadden PL (1997) Palaeosecular variation
over the past 5Myr based on a new generalized database. Geo-
phys J Int 131(2):240–252
60. McFadden PL, Merrill RT, McElhinny MW (1988) Dipole/
quadrupole family modeling of palaeosecular variation. J Geo-
phys Res 93(B10):11583–11588
61. Meyers PA (1994) Preservation of elemental and isotopic source
identification of sedimentary organic matter. Chem Geol
114:289–302
62. Mitrovica JX, Tamisiea ME, Davis JL, Milne GA (2001) Recent
mass balance of polar ice sheets inferred from patterns of global
sea level change. Nature 409:1026–1029
63. Montaggioni LF, Cabioch G, Camoinau GF, Bard E, Ribnaud-
Laurenti A, Faure G, De
´jardin P, Re
´cy J (1997) Continuous
record of reef growth over the past 14 ky on the mid-Pacific
island of Tahiti. Geology 25:555–558
64. Pedrosa M, Camerlenghi A, De Mol B, Urgeles R, Rebesco M,
Lucchi RG, Shipboard participants of the SVAIS and EGLA-
COM Cruises (2011) Seabed Morphology and Shallow Sedi-
mentary Structure of the Storfjorden and Kveitehola Trough-
Mouth Fans (north west Barents Sea). Mar Geol 286(1–4):65–81.
doi:10.1016/j.margeo.2011.05.009
65. Peltier WR (2005) On the hemispheric origins ofmeltwater pulse
1a. Quat Sci Rev 24:1655–1671
66. Piper DJW, Pereira CPG (1992) Late Quaternary sedimentation
in central Flemish Pass. Can J Earth Sci 29:535–550
67. Rae JWB, Sarnthein M, Foster GL, Ridgwell A, Grootes PM,
Elliott T (2014) Deep water formationin the North Pacific and
deglacial CO2 rise. Paleoceanography 29:645–667
68. Rasmussen TL, Thomsen E, Slubowska MA, Jessen S, Solheim
A, Koc¸ N (2007) Paleoceanographic evolution of the SW Sval-
bard margin (76°N) since 20,000 14C yr BP. Quat Res
67:100–114
69. Rebesco M, Liu Y, Camerlenghi A, Winsborrow M et al (2011)
Deglaciation of the Barents Sea Ice Sheet - a swath bathymetric
and sub-bottom seismic study from the Kveithola Trough. Mar
Geol 279:141–147. doi:10.1016/j.margeo.2010.10.018
70. Reimer PJ, Baillie MGL, Bard E et al (2009) IntCal09 and
Marine09 radiocarbon age calibration curves, 0–50,000 years cal
BP. Radiocarbon 51:1111–1115
71. Rinterknecht VR, Clark PU, Raisbeck GM, Yiou F, Bitinas A,
Brook EJ, Marks L (2006) The Last Deglaciation of the South-
eastern Sector of the Scandinavian Ice Sheet. Science
311:1449–1452
72. Ru
¨ther DC, Bjarnado
´ttir LR, Junttila J, Husum K, Rasmussen TL,
Lucchi RG, Andreassen K (2012) Pattern and timing of the north-
western Barents Sea Ice Sheet deglaciation and indications of
episodic Holocene deposition. Boreas 41:494–512
73. Sagnotti L, Macrı
`P, Lucchi RG, Rebesco M, Camerlenghi A
(2011) A Holocene paleosecular variation record from the north-
western Barents Sea continental margin. Geochem Geophy Geosy
12(11):Q11Z33. doi:10.1029/2011GC003810
74. Sagnotti L, Macrı
`P, Lucchi RG (2015) Geomagnetic
palaeosecular variation around 15 ka ago from NW Barents Sea
cores (south of Svalbard). Geophys J Int (in press)
75. Sarnthein M (2011) Northern Meltwater Pulse, CO2, and changes
in Atlantic convection. Science 331:156–158
Arktos (2015) 1:7 Page 13 of 14 7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
76. Shepherd A, Wingham D, Payne T, Skvarca P (2003) Larsen Ice
Shelf has progressively thinned. Science 302:856–859
77. Siegert MJ, Dowdeswell JA (2002) Late Weichselian iceberg,
surface-melt and sediment production from the Eurasian Ice
Sheet: results from numerical ice-sheet modelling. Marine Geol
188:109–127
78. Slubowska-Woldengen M, Koc¸ N, Rasmussen TL, Klitgaard-
Kristensen D, Hald M, Jennings AE (2008) Time-slice recon-
structions of ocean circulation changes on the continental shelf in
the Nordic and Barents Seas during the last 16,000 cal yr B.P.
Quat Sci Rev 27:1476–1492
79. Slubowska-Woldengen M, Rasmussen TL, Koc¸ N, Klitgaard-
Kristensenc D, Nilsen F, Solheim A (2007) Advection of Atlantic
Water to the western and northern Svalbard shelf since
17,500 cal yr BP. Quat Sci Rev 26:463–478
80. Slubowska MA, Koc N, Rasmussen TL, Klitgaard-Kristensen D
(2005) Changes in the flow of Atlantic water into the Arctic Ocean
since the last deglaciation: evidence from the northern Svalbard
continental margin, 801N. Paleoceanography 20:PA4014
81. Stanford JD, Rohling EJ, Hunter SH, Roberts AP, Rasmussen SO,
Bard E, McManus J, Fairbanks RG (2006) Timing of meltwater
pulse 1a and climate responses to meltwater injections. Paleo-
ceanography 21(4):PA4103. doi:10.1029/2006PA001340
82. Stuiver M, Reimer PJ (1993) Extended 14C database and revised
CALIB radiocarbon calibration program. Radiocarbon 35:215–230
83. Tauxe L, Kent DV (2004) A simplified statistical model for the
geomagnetic field and the detection of shallow bias in paleo-
magnetic inclinations: was the ancient magnetic field dipolar.
Geophy Monogr Series 145:101–116
84. Thornalley DJR, Bauc HA, Gebbie G, Guo W, Ziegler M, Ber-
nasconi SM, Barker S, Skinner LC, Yu J (2015) A warm and
poorly ventilated deep Arctic Mediterranean during the last
glacial period. Science 349(6249):706–710. doi:10.1126/science.
aaa9554
85. Tripsanas EK, Piper DJW (2008) Late Quaternary stratigraphy
and sedimentology of Orphane Basin: implications for meltwater
dispersal in the southern Labrador Sea. Palaeogeogr Palaeocli-
matol Palaeoecol 260:521–539
86. Vandamme D (1994) A new method to determine palaeosecular
variation. Phys Earth Interiors 85:131–142
87. Vogt PR, Crane K, Sundvor E (1994) Deep Pleistocene iceberg
ploughmarks on the Yermak Plateau: sidescan and 3.5 kHz evi-
dence for thick calving ice fronts and a possible marine ice sheet
in the Arctic Ocean. Geology 22:403–406
88. Vorren TO, Hald M, Thomsen E (1984) Quaternary sediments
and environments on the continental shelf off northern Norway.
Mar Geol 57:229–257
89. Weaver AJ, Saenko OA, Clark PU, Mitrovica JX (2003) Melt-
water Pulse 1A from Antarctica as a trigger of the Bølling-
Allerød warm interval. Science 299:709–1713
90. Weber ME, Clark PU, Kuhn G, Timmermann A, Sprenk D,
Gladstone R, Zhang X, Lohmann G, Menviel L (2014) Millen-
nial-scale variability in Antarctic ice-sheet discharge during the
last deglaciation. Nature 510:134–138
91. Weber ME, Clark PU, Ricken W, Mitrovica JX, Hostetler SW,
Kuhn G (2011) Interhemispheric ice-sheet synchronicity during
the Last Glacial Maximum. Science 334:1265–1269
92. Webster JM, Clague DA, Riker-Coleman K, Gallup C, Braga JC,
Potts D, Moore JG, Winterer EL, Paull CK (2004) Drowning of
the -150 m reef off Hawaii: a casualty of global meltwater pulse
1A? Geology 32:249–252
93. Yokoyama Y, Lambeck K, De Deckker P, Johnston P, Fifield LK
(2000) Timing of the last glacial maximum from observed sea-
level minima. Nature 406:713–716
7Page 14 of 14 Arktos (2015) 1:7
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Shallow sediment cores at or close to the drill sites provide initial sediment properties and paleoceanographic interpretations (Caricchi et al., 2019(Caricchi et al., , 2020Carbonara et al., 2016;Jessen et al., 2010;Lucchi et al., 2013Lucchi et al., , 2014Lucchi et al., , 2015Lucchi et al., , 2018Melis et al., 2018;Rigual Hernández et al., 2017;Rasmussen et al., 2007Rasmussen et al., , 2014Ślubowska et al., 2005;Ślubowska-Woldengen et al., 2007Ślubowska-Woldengen et al., , 2008Torricella et al., 2022; among others). The existing geologic data set includes multicores, box cores, gravity cores, piston cores (including long Calypso piston cores), and MeBo drilling. ...
... A robust age model was defined through paleomagnetic and biostratigraphic analyses, identification of tephras, and radiocarbon dating of the abundant biogenic carbonate fraction (Caricchi et al., 2019(Caricchi et al., , 2020. The sedimentologic analyses indicated the consistent presence of contouritic deposition and the existence of short-lived, abrupt depositional events associated to prominent meltwater events, like the MWP-1A (Lucchi et al., , 2015, and Heinrich-like events indicating a highly dynamic SBSIS during last 60 ky Caricchi et al., 2019). The drill sites on the Bellsund and Isfjorden drifts are designed to recover the most expanded sequence down to Seismic Reflector R4A to specifically address suborbital oscillations, MBE, MPT, and the onset of shelf edge glaciation in this area. ...
... They show a consistent pattern in magnetic susceptibility (MS), lithology, and color of the sediments as described in 11 cores from the western Svalbard margin spanning the last 30 ky (Jessen et al., 2010): a dark unsorted layer of low MS consisting of coarse material of black and brown shales dating 24 ky and a laminated clay layer of low MS dating c. 15 ky ( Figure 2). The pattern of MS typical for the western margin has later been shown to be consistent both at Vestnesa Ridge (e.g., Howe et al., 2008;Consolaro et al., 2015;Sztybor & Rasmussen, 2017), southwest of Svalbard (Caricchi et al., 2019;Lucchi et al., 2013Lucchi et al., , 2015 and north of Svalbard (Chauhan et al., 2016). The event Meltwater Pulse MWP-1A documented in (e.g., Lucchi et al., 2013) is also recognized as dark, finely laminated sediments of low MS. ...
... The sediment sequences at Vestnesa Ridge thus consist of alternating bioturbated fine-grained and ice rafted debris (IRD) rich deposits associated with bottom current transport and iceberg release of detritus, respectively; and laminated sediments coupled with massive IRD intervals associated with deglaciations at the MIS 4-3 and MIS 2-1 transitions (e.g., Caricchi et al., 2019;Consolaro et al., 2015;Jessen et al., 2010;Jessen & Rasmussen, 2015;Lucchi et al., 2013;Lucchi et al., 2015;Sztybor & Rasmussen, 2017) (Figure 2). ...
Article
Full-text available
Gas transport through sediments to the seabed and seepage occurs via advection through pores, faults, and fractures, and as solubility driven gas diffusion. The pore pressure gradient is a key factor in these processes. Yet, in situ measurements for quantitative studies of fluid dynamics and sediment deformation in deep ocean environments remain scarce. In this study, we integrate piezometer data, geotechnical tests, and sediment core analyses to study the pressure regime that controls gas transport along the Vestnesa Ridge in the eastern Fram Strait. The data show a progressive westward decrease in induced pore pressure (i.e., from c. 180 to c. 50 kPa) upon piezometer penetration and undrained shear strength of the sediments, interpreted as a decrease in sediment stiffness. In addition, the data suggest that the upper c. 6 m of sediments may be mechanically damaged due to variations in gas diffusion rates and exsolution. Background pore pressures are mostly at hydrostatic conditions, but localized excess pore pressures (i.e., up to 10 kPa) exist and point toward external controls. When analyzed in conjunction with observations from geophysical data and sediment core analyses, the pore pressure data suggest a spatial change from an advection dominated to a diffusion dominated fluid flow system, influenced by the behavior of sedimentary faults. Understanding gas transport mechanisms and their effect on fine‐grained sediments of deep ocean settings is critical for constraining gas hydrate inventories, seepage phenomena and sub‐seabed sediment deformations and instabilities.
... ka BP, possibly with a duration of only 150-300 years, mainly centred around 14.5 to 14.4 cal. ka BP (Jessen et al. 2010;Hormes et al. 2013;Lucchi et al. 2013Lucchi et al. , 2015. The GZW is the only evidence of a major stillstand of the ice margin before the ice retreated into the inner Kongsfjorden fjord area (Landvik et al. 2005; The geological record on Svalbard has been exposed to glacial erosion repeatedly and is fragmentary, and does not therefore preserve a continuous record of the Weichselian and preceding glacial periods Svendsen et al. 1996;Forwick & Vorren 2009). ...
... Laberg & Vorren 1995;King et al. 1996;Dimakis et al. 2000). At the ice-contact zone there may also be different types of outwash materials such as coarse-grained bedload sediments and finer sediments from meltwater plumes (Lucchi et al. 2015). The finer sediments are thought to have emerged as a plume of sediment-laden subglacial meltwater at the glacial grounding line, further dispersed as either high-density underflows or low-density over/inflows depending on grain size . ...
Article
Full-text available
The Arctic is a climate‐sensitive area, responding rapidly to present changes, but for the past changes, the record is still incomplete. For instance, the Weichselian glacial history of the Svalbard–Barents Sea Ice Sheet (SBIS) has largely been reconstructed based on studies of the fragmentary Spitsbergen terrestrial and shelf records. However, the sediments removed from the land and shelf areas during peak glacials were deposited on trough mouth fans located along the continental slope. By studying the stratigraphy and processes of the trough mouth fans, comprising a more complete sediment archive, our new data have allowed gaps in the Weichselian glacial history of the SBIS to be refined and filled. Here we present new lithological and geochronological data from the Kongsfjorden Trough Mouth Fan, closely linked to the advance and decay of the SBIS. High‐resolution TOPAS seismic profiles reveal three distinct packages of glacigenic debris flows (GDFs) within its upper stratigraphy, each interpreted to represent an advance of the SBIS to the shelf edge. A radiocarbon dated, 12.6‐m‐long core from the southern flank of the Kongsfjorden Trough Mouth Fan penetrates trough sediments directly linked to the youngest GDF package and terminates in the second GDF, allowing us to study the last two Kongsfjorden ice‐stream advances in greater detail than was previously possible. The age model of core GS10‐164‐09PC, based on combining 14C‐, 18O‐stable isotope and magnetic susceptibility data, spans the last ~54 ka. An Early Weichselian glacial advance is tentatively dated to have ended at ~90 ka. A second peak glaciation is estimated at ~70 ka, followed by a deglaciation from ~54 ka. An ice rafted debris‐rich unit (U7) dated between 38 and 34 ka, followed by a plumite (U6), indicates an advance of unknown extent. The Last Glacial Maximum advance is dated to before 24 ka BP, followed by a rapid deglaciation at ~15 ka. The presence of coarser‐grained sorted sediments at the present seafloor is attributed to the influence of the West‐Spitsbergen Current, acting on water depths of at least 846 m, and is thought to have worked in the vicinity of the coring site since ~14 ka BP.
... Overlying laminated siltstones, lacking ice-rafted clasts-with, however, the notable exception of the sandstone boulder-, lacking storm-generated facies and bioturbation, but showing rhythmical stratification pattern are tentatively interpreted as the settling of hypopycnal meltwater plumes beneath a perennial sea-ice cover [Rüther et al., 2012]. In this case, and accounting for high ice-marginal accumulation rates, the time span represented by the fine-grained Unit 5 might be relatively short [Lucchi et al., 2015]. The sea-ice cover hypothesis is favored relative to an iceshelf model because of the shallow water depths, the lack of ploughmarks from large tabular icebergs, and the lack of facies marking the retreating calving-line [Smith et al., 2019]. ...
... Selected elements were normalized versus titanium (Ti) following a standard procedure described in Croudace and Rothwell (2015) and references therein. In this study, we consider the following ratios as paleoenvironmental proxies: Zr/Rb used as a grain size proxy (Wu et al. 2020) with positive peaks indicating higher percentages of coarse silt and sand, Si/Al used to determine the variation between the biogenic silica (Si) versus the detrital (Al) fraction (Dickson et al. 2010), Ca/Ti used to infer variations between the biogenic carbonate (Ca) and detrital (Ti) fractions (Olsen, Anderson, and Knudsen 2012;Lucchi et al. 2013), and Si/ Ti and K/Ti used as proxies for the content of quartz and K-feldspars delivered to the area mainly by meltwaters and/ or sea ice, therefore representing a detrital input (Marsh et al. 2007;Agnihotri et al. 2008;Diekmann et al. 2008;Piva et al. 2008;Lucchi et al. 2013Lucchi et al. , 2015Shala et al. 2014). Additional information on the compositional proxies is reported in Table 1. ...
Article
Full-text available
A reconstruction of the last 2,000 years BP of environmental and oceanographic changes on the western margin of Spitsbergen was performed using a multidisciplinary approach including the fossil assemblages of diatoms, planktic and benthic foraminifera and calcareous nannofossils and the use of geochemistry (X-ray fluorescence spectroscopy, X-ray diffraction). We identified two warm periods (2,000–1,600 years BP and 1,300–700 years BP) that were associated with the Roman Warm Period and the Medieval Warm Period that alternate with colder oceanic conditions and sea ice coverage occurred during the Dark Ages (1,600–1,300 years BP) and the beginning of the Little Ice Age. During the Medieval Warm Period the occurrence of ice-rafted debris and Aulocoseira spp., a specific diatom genus commonly associated with continental freshwater, suggests significant runoff of meltwaters from local glaciers
... The cores were previously analyzed using a multidisciplinary approach (Lucchi et al., 2012(Lucchi et al., , 2015Sagnotti et al., 2011a including Accelerator Mass Spectrometry (AMS) 14 C dating, lithofacies analysis, paleomagnetic and rock magnetic analyses. ...
Article
Reconstruction of geomagnetic field changes has a strong potential to complement geodynamo modeling and improve the understanding of Earth's core dynamics. Recent works based on geomagnetic measurements pointed out that over the last two decades the position of the north magnetic pole has been largely determined by the influence of two competing flux lobes under Canada and Siberia. In order to understand if the waxing and waning of magnetic flux lobes have driven the path of geomagnetic paleopoles in the past, we present an augmented and updated record of the chronology and paleosecular variation of geomagnetic field for the last 22 kyr derived from sedimentary cores collected along the north-western margin of Barents Sea and western margin of Spitsbergen (Arctic). The path of the virtual geomagnetic pole (VGP) has been reconstructed over this time period and compared with the maps of the radial component of the geomagnetic field at the core-mantle boundary, obtained from the most recent models. The VGP path includes centuries during which the VGP position is stable and centuries during which its motion accelerates. We recognize both clockwise and counterclockwise VGP paths, mostly developing inside the surface projection of the inner core tangent cylinder in the Arctic region. The VGP path seems to follow the appearance of Br patches of normal magnetic flux, especially those located under Siberia and Canada areas, but also those that may cause peculiar paleomagnetic features such as the Levantine Iron Age Anomaly.
Article
Full-text available
Vestnesa Ridge is built-up of thick contourites mainly deposited during the last ∼5 million years. Methane leaks from deep gas reservoirs creating pockmarks on its crest, and which have been the focus of numerous studies. Sedimentation patterns in relation to the pronounced changes in oceanography and climate of the last glacial-interglacial cycles and its possible impact of seepage of gas have rarely been studied. Here, we present a detailed history of contourite development covering the last ∼130,000 years with most details for the last 60,000 years. The study is based on 43 marine sediment cores and 1,430 km of shallow seismic lines covering the ridge including methane seep sites, with the purpose of reconstructing changes in depositional patterns in relation to paleoceanographical changes on glacial, interglacial, and millennial time scale in relation to activity of seepage of gas. The results show that thick Holocene deposits occurred below ∼1,250 m water depth in the western part of the ridge. Both in pockmarks at western and eastern Vestnesa Ridge, seepage decreased at ∼10–9 ka in the early Holocene. The fine Holocene mud likely reduced seepage to a slow diffusion of gas and microbial oxidation probably prevented escape from the seafloor. Results also showed that seepage of gas was highly variable during the glacial, and low to moderate during the cold Heinrich stadial H1 (19–15 ka) and Younger Dryas stadial (13–12 ka). Seepage reached a maximum during the deglaciation in the Bølling and Allerød interstadials 15–13 ka and early Holocene 12–10 ka. The deglaciation was a period of rapid climatic, oceanographic, and environmental changes. Seepage of gas varied closely with these events indicating that slower tectonic/isostatic movements probably played a minor role in these millennial scale rapid fluctuations in gas emission.
Chapter
During the Bølling–Allerød Interstadial, the extensive European Ice Sheet Complex (EISC) persisted over much of its domain, albeit split into independent ice sheets and remnant masses and in retreat. Here we give an overview of the Bølling–Allerød development of the EISC and its remnant bodies, and we introduce and synthesise the landscapes and characteristic landforms produced by these ice sheets, which are described in Chapters 29–33. The geomorphological record of the EISC gives an increasingly detailed picture of the pattern of ice sheet retreat during this period as well as powerful indicators of retreat behaviour, mechanisms and, qualitatively, rates of change, and we explore contrasts and commonalities in these elements between the marine and terrestrial ice sheet sectors. Conversely, chronological uncertainties are significant, particularly in the marine sectors but also around smaller remnant ice masses at the EISC periphery. These lend doubt to whether or not ice existed during the Bølling–Allerød in Britain and Ireland, and have profound implications for the role of the EISC in major sea level rise events and, consequently, the function of this ice sheet complex within the global climate system. Furthermore, the magnitude of retreat during the Bølling–Allerød is challenging to determine inboard of the well-defined Younger Dryas positions, and was likely variable across different sectors of the ice sheet complex. There is abundant evidence for substantial meltwater production during the Bølling–Allerød, and here we consider some of the wider ice–ocean–climate implications of meltwater discharge as well as how meltwater landform records from deglaciated ice sheet domains, including the EISC, continue to challenge and to stimulate important advances in understanding subglacial hydrology.
Chapter
The rapid atmospheric warming of the Bølling–Allerød Interstadials marked the final demise of the marine-based ice sheet across the Barents Shelf, with the collapse further accelerated by a marked subsurface warming at the ice–ocean interface. The glacial geomorphology records a period of rapid environmental change, characterised by collapsing ice streams, large subglacial meltwater features indicative of intense surface melting, and the onset of widespread isostatic recovery of the shelf. This depressurisation of the shelf in turn triggered subsurface gas hydrate destabilisation across the hydrocarbon-rich Barents Sea, affecting glacial dynamics and deglacial methane seepage. By the Younger Dryas stadial at 12.9 ka the Barents Sea ice sheet had fully deglaciated, leaving only remnant ice caps across the High Arctic archipelagos.
Article
Full-text available
A robust understanding of Antarctic Ice Sheet deglacial history since the Last Glacial Maximum is important in order to constrain ice sheet and glacial-isostatic adjustment models, and to explore the forcing mechanisms responsible for ice sheet retreat. Such understanding can be derived from a broad range of geological and glaciological datasets and recent decades have seen an upsurge in such data gathering around the continent and Sub-Antarctic islands. Here, we report a new synthesis of those datasets, based on an accompanying series of reviews of the geological data, organised by sector. We present a series of timeslice maps for 20 ka, 15 ka, 10 ka and 5 ka, including grounding line position and ice sheet thickness changes, along with a clear assessment of levels of confidence. The reconstruction shows that the Antarctic Ice sheet did not everywhere reach the continental shelf edge at its maximum, that initial retreat was asynchronous, and that the spatial pattern of deglaciation was highly variable, particularly on the inner shelf. The deglacial reconstruction is consistent with a moderate overall excess ice volume and with a relatively small Antarctic contribution to meltwater pulse 1a. We discuss key areas of uncertainty both around the continent and by time interval, and we highlight potential priorities for future work. The synthesis is intended to be a resource for the modelling and glacial geological community.
Article
Full-text available
The sedimentary sequence deposited during the deglaciation phase following the last glacial maximum in the Storfjorden trough, on the northwestern Barents Sea south of Svalbard, was sampled with 10 piston and gravity cores during the SVAIS and EGLACOM cruises. Three cores (SV-02, SV-03 and SV-05) collected on the upper continental slope are characterized by a thin (20-40 cm) Holocene interval and a thick (up to 4.5 m in core SV-03) late Pleistocene sequence of finely laminated fine-grained sediments that have been interpreted as plumites deposited during the Melt Water Pulse 1a (MWP-1a). Radiocarbon ages obtained at the top and bottom of this stratigraphic interval revealed that deposition occurred during less than two centuries at around 15 ka ago, with a very high sedimentary rate exceeding 3 cm a(-1). We studied the palaeomagnetic and rock magnetic properties of this interval, by taking magnetic measurements at 1 cm spacing on u-channel samples collected from the three cores. The data show that this sequence is characterized by good palaeomagnetic properties and the palaeomagnetic and rock magnetic trends may be correlated at high resolution from core to core. The obtained palaeomagnetic data therefore offer the unique opportunity to investigate in detail the rate of geomagnetic palaeosecular variation (PSV) in the high northern latitudes at a decadal scale. Notwithstanding the palaeomagnetic trends of the three cores may be closely matched, the amplitude of directional PSV and the consequent virtual geomagnetic pole (VGP) scatter (S) is distinctly higher in one core (SV-05) than in the other two cores (SV-02 and SV-03). This might result from a variable proportion of two distinct populations of magnetic minerals in core SV-05, as suggested by the variable tendency to acquire a gyromagnetic remanent magnetization at high fields during the AF demagnetization treatment. For the plumite interval of cores SV-02 and SV-03, where the magnetic mineralogy is uniform and magnetite is the main magnetic carrier, a S value of about 9 degrees is obtained. We consider this value as a reliable approximation of palaeomagnetic secular variation at a latitude of 75 degrees N over a time interval spanning a couple of centuries around 15 ka ago.
Article
Full-text available
LETTERS TO NATURE example, ~50-40 mmyr1 is roughly equivalent to discharge rates of 16,000 km3 yr~' for MWP-1 A). A third meltwater pulse, smaller than the two other, was identified at ~7,600cal. yr bp in a compilation of Caribbean corals together with the Barbados curve . ...
Article
Full-text available
Changes in the formation of dense water in the Arctic Ocean and Nordic Seas [the "Arctic Mediterranean" (AM)] probably contributed to the altered climate of the last glacial period. We examined past changes in AM circulation by reconstructing radiocarbon ventilation ages of the deep Nordic Seas over the past 30,000 years. Our results show that the glacial deep AM was extremely poorly ventilated (ventilation ages of up to 10,000 years). Subsequent episodic overflow of aged water into the mid-depth North Atlantic occurred during deglaciation. Proxy data also suggest that the deep glacial AM was ~2° to 3°C warmer than modern temperatures; deglacial mixing of the deep AM with the upper ocean thus potentially contributed to the melting of sea ice, icebergs, and terminal ice-sheet margins. Copyright © 2015, American Association for the Advancement of Science.