Vogt et al 2001

Transcription

Vogt et al 2001
Global and Planetary Change 31 Ž2001. 23–44
www.elsevier.comrlocatergloplacha
Detailed mineralogical evidence for two nearly identical
glacialrdeglacial cycles and Atlantic water advection to the
Arctic Ocean during the last 90,000 years
Christoph Vogt a,) , Jochen Knies b,1, Robert F. Spielhagen c,2 , Ruediger Stein b,1
a
FB Geowissenschaftenr Geo Sciences, UniÕersitat
¨ Bremen, Post Box 330440, 28334 Bremen, Germany
Alfred Wegener Institute for Polar and Marine Research, Columbusstr., D-27568 BremerhaÕen, Germany
GEOMAR Research Center for Marine Geosciences, Kiel UniÕersity, Wischhofstr. 1-3, D-24148 Kiel, Germany
b
c
Received 5 December 1999; accepted 23 May 2001
Abstract
Three cores recovered off the northwest of Svalbard were studied with respect to glacialrinterglacial changes of clay and
bulk mineralogy, lithology and organic geochemistry. The cores cover the Late Quaternary Marine Isotope Stages ŽMIS. 6–1
Žca. 170,000 years. and are located in the vicinity of the Polar Front which separates the warm Atlantic water of the
Westspitsbergen Current and the cold Polar Water of the Transpolar Drift. Globally driven changes in the paleoenvironment
like the variable advection of warm Atlantic water into the Arctic Ocean can be distinguished from regional events by means
of source mineral signatures and organic geochemistry data. In particular, a combination of high organic carbon and low
carbonate contents, high CrN-ratios, a particular lithology and a distinct bulk and clay mineral assemblage can be related to
Svalbard ice sheet developments between 23,000 and 19,500 14C years. This complex sediment pattern has been traced to
the northwest of Spitsbergen as far north as 828N. Additionally, the same signature has been recognized in detail in upper
MIS 5 sediments. The striking similarity of the history of the SvalbardrBarents Sea Ice Sheet during the late and
earlyrmiddle Weichselian is elaborated. Both sediment horizons are intercalated between biogenic calcite rich core
sequences which contain the so-called AHigh Productivity ZonesB or ANordway EventsB related to the increased advection of
warm Atlantic water to the Arctic Ocean. This study provides further evidence that the meridional circulation pattern has
been present during most of the Weichselian and that the ice cover was often reduced in the northeastern Fram Strait and
above the Yermak Plateau. Our findings contradict the widely used reconstructions in modelling of the last glaciation cycle
and reveal a much more dynamic system in the Fram Strait and southwestern Eurasian Basin of the Arctic Ocean. q 2001
Elsevier Science B.V. All rights reserved.
Keywords: Quaternary; Fram Strait; Yermak Plateau; Atlantic water advection; lithology; mineralogy; ice sheet developments
1. Introduction
)
Corresponding author. Tel.: q49-421-218-9007; fax: q49421-218-7123.
E-mail addresses: [email protected] ŽC. Vogt.,
[email protected] ŽR.F. Spielhagen.,
[email protected] ŽR. Stein..
1
Fax: q49-471-4831-1580.
2
Fax.: q49-431-600-2941.
The Arctic Ocean is very sensitive to climatic
change and might even drive global oceanographic
and climatic changes ŽAlley, 1995.. It is adjacent to
the northern boundaries of the northern hemisphere
ice sheets. Thus, sediments of the Arctic Ocean have
0921-8181r01r$ - see front matter q 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 1 - 8 1 8 1 Ž 0 1 . 0 0 1 1 1 - 4
24
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
great potential to archivate the history of northern
hemisphere ice sheets, the fate of the produced icebergs, and changes in the sea-ice cover.
In the NW of Svalbard, relative warm Atlantic
Water of the northward flowing Westspitsbergen
Current ŽWSC. submerges at the Polar Front beneath
cold Polar Water ŽFig. 1.. The positions of the Polar
Front and the sea-ice edge depends on the strength of
the WSC and the outflow of the Polar Water, which
is strongly connected to the fresh water influx into
the Arctic Ocean Že.g. Anderson et al., 1994; Aagaard and Carmack, 1994.. Previous studies on sediment cores in the Fram Strait and NE of Svalbard
revealed rapid changes in the Atlantic Water influx
and its influence on the built-up and decay of the
SvalbardrBarents Sea Ice Sheet ŽSBIS. ŽHebbeln et
al., 1994; Dokken and Hald, 1996; Lubinski et al.,
1996; Knies et al., 1999.. Here, we present a detailed
lithological, mineralogical, and geochemical dataset
of undisturbed late Quaternary sediments NW of
Svalbard to elucidate paleoceanographic changes
during the Weichselian and to enlarge the information on rapid oceanographical and climatological
changes in the Polar Front region NW of Svalbard.
The sediment cores are well positioned to the north
and the south of the Polar Front to record changes in
its development through time ŽFig. 1..
2. Materials and methods
Gravity cores PS2122-1 and PS2123-2 and the
Kastenlot core PS2212-3 were recovered from the
NW continental margin of Svalbard and the Yermak
Plateau during expeditions ARK-VIIIr2 and 3 in
summer 1991 with RV Polarstern ŽFutterer,
1992;
¨
Rachor, 1992.. Short sediment cores containing the
Fig. 1. Surface currents in the European sector of the Arctic Ocean and locations of investigated cores PS2122-1, PS2123-2, PS2212-3
Žpositions and water depth are listed in Fig. 2. and cores PS1533-3, PS2138-1, and NP90-39 for comparison ŽTPD: Transpolar Drift;
TPD Sib : Siberian Branch; BG: Beaufort Gyre; EGC: East Greenland Current; WSC: West Spitsbergen Current; WSC s : West Spitsbergen
Current Žsubmerging.; ESC: East Spitsbergen Current; RAC: Return Atlantic Current; JMPC: Jan Mayen Polar Current, compiled from
Manley et al., 1992; Hebbeln et al., 1994; Nowaczyk et al., 1994.. Position of Spitsbergenbanken with Jurassic Shale Source rock and
proposed transport path of fine fraction material Žafter Andersen et al., 1996..
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
uppermost undisturbed 30–45 cm were taken from
the same sites ŽFig. 1..
We routinely sampled every 5 cm Ž1-cm-thick
sediment slice.. Each sample was split into two
parts. One part was dried and pulverized, and then
divided into subsamples. One of these subsamples
Žapproximately 30 mg. was examined for carbonate,
organic carbon, and nitrogen using a Heraeus CHNelemental-analyser. Carbonate contents were calculated as CaCO 3 Ž%. s Žtotal carbony total organic
carbon ŽTOC.. = 8.333. The carbon and nitrogen
measurements have a standard deviation of 0.06%
and 0.02%, respectively. The CrN weight ratios
were calculated, and hydrogen-index ŽHI. and Tmaxtemperature were determined by Rock–Eval pyrolysis ŽKnies and Stein, 1998; reproducibility "8%..
Another subsample Žabout 3 g. of the dried bulk
sample was used for the evaluation of bulk mineralogy by means of XRD-measurements with a Philips
PW 3020 diffractometer equipped with cobalt k aradiation, automatic divergence slit, graphite monochromator, and automatic sample changer ŽTable 1..
Individual bulk mineral contents were expressed as
percentages of bulk sediment. The quartz content
was determined by using the QUAX software package ŽEmmermann and Lauterjung, 1990.. The dolo-
25
mite content was inferred from the peak intensity
ratio of dolomite and calcite multiplied with the
carbonate content from the elemental analysis and,
additionally, controlled using QUAX Žstandard deviation "2% for quartz and "1% for the carbonates;
cf. Vogt, 1997.. Peak intensity ratios were used for
the evaluation of the quartz and feldspar mineralogy.
The second part of the original sample Ž2–3 g.
was treated with 3–10% H 2 O 2 to oxidize the organic
matter, disaggregated, and finally wet sieved through
a 63-mm mesh. The coarse fraction Ž) 63 mm. was
studied under the light microscope to gain an
overview of the terrigenous components. Five to ten
specimens of Neogloboquadrina pachyderma sinistral Žsize: 125–250 mm. were picked for stable
isotope measurements. The - 63-mm fraction was
separated into silt Ž2–63 mm. and clay Ž- 2 mm. by
the Atterberg settling tubes method Žaccording to
Stoke’s law; Muller,
1967.. The amount of ice-rafted
¨
debris ŽIRD. was estimated by counting terriginous
particles ) 2 mm in each centimeter of an X-ray
radiography according to the method of Grobe
Ž1987.. The clay mineral assemblage was determined
by standard preparation and analysis techniques as
outlined by Petschick et al. Ž1996, Table 1.. The
peak areas of the clay mineral groups ŽTable 1. were
Table 1
Ža. Running conditions of XRD measurements for bulk and clay mineral analysis Žrange in 82 u , stepsize in 82 ur1 s, slow scan 82 ur2 s.. Co
k alpha-radiation
Measurement
Sample preparation
Range
Stepsize
Bulk sample
powder unoriented
pressed pellets ŽPS2212.
2–100
0.02
Clay fraction
) 18 h glycolated
slow scan
textural oriented
textural oriented
textural oriented
2–18
2–40
28.5–30.5
0.02
0.02
0.005
˚ . for the quantitative evaluation of bulk and clay mineralogy Žthe smectite group includes all expandable
Žb. Used XRD peaks Ž d-value in A
˚ OLEMs ordered layered expandable minerals ŽReynolds, 1970., illite includes
minerals with a peak of the glycolated sample near 17 A,
non expandable mixed layers.. Peaks were recognized graphically by using the MacDiff program ŽPetschick et al., 1996.
Bulk and clay fraction sample
Clay fraction sample
quartz Ž3.34, 4.26., feldspar Ž3.24, 3.18.,
QzrFsp Ž4.26rŽ3.24 and 3.18..,
calcite Ž3.035., dolomite Ž2.89.,
APyroxene indexB Ž2.995–2.92.
OLEM Ž30–22 glyc, 11–12.5.,
smectite Ž17 glyc., illite Ž10, 5, 4.5.,
kaolinite Ž7, 3.58., chlorite Ž7, 3.54.
26
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
calculated and transformed into relative clay mineral
percentages by means of Biscaye factors based on
the assumption that the clay fraction consists only of
clay minerals Žcf. Wahsner et al., 1999; reproducibility: "3%..
3. Stratigraphy
The stratigraphic framework of cores PS2122-1
and PS2123-2 was deduced from their oxygen and
carbon isotope records ŽFig. 2., which are, in general, fairly correlateable to the global isotope curve
ŽSPECMAP stack; Martinson et al., 1987.. An AMS
14
C date in PS2122-1 at 321 cm bsf Žcm below core
surface. helps to pinpoint Marine Isotope Stage ŽMIS.
3. The stratigraphic framework is also corroborated
by the occurrence of the benthic foraminifera Pullenia bulloides, a stratigraphic marker for the MIS
5r4 boundary in the Nordic Seas Žcf. Haake and
Plaumann, 1989.. Additionally, a typical decrease of
Fig. 2. Chronostratigraphies and d18 O- and d13 C-records of cores PS2122-1rPS2123-2 and PS2212-3: Marine Isotope Stage ŽMIS.
assignments are based on Martinson et al. Ž1987., an AMS-14 C date at 321 cm in PS2122-1 Žbivalve shell, sample KIA367 measured at the
Leibniz Laboratory, Kiel University: 36800" 3030 14 C-years; reservoir correction: 440 years; Mangerud and Gulliksen, 1975., and the
occurence of the benthic foraminifera P. bulloides at the MIS 5r4 boundary. Additionally, the chronostratigraphy of core PS2212-3 is
based on the correlation of lithological, geochemical and mineralogical data between NP90-39, PS2138-1 and PS2212-3 Žcompare Figs. 1
and 3: AMS 14 C-dated Event I., the occurrence of S. rolshauseni ŽWollenburg et al., 2001. and coccolith abundances and paleomagnetic
data of Nowaczyk et al. Ž1994. Žincluding a hiatus in MIS 5; NGS—Norwegian–Greenland-Sea-Event..
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
the carbon isotope values from late MIS 5 to lowest
values in early MIS 3 is observed and used as a
regional stratigraphic fixpoint Žcf. Dokken and Hald,
1996; Nørgaard-Pedersen et al., 1998.. Hence, the
cores PS2122-1 and PS2123-2 most likely represent
MIS 1–5. Well-dated melting events during Termination I Že.g. at 14.5 ka. are also recognized Žcf.
Hebbeln et al., 1994.. All ages are reported according to the SPECMAP timescale Žka: 1000 years..
A well defined sequence of sedimentological,
mineralogical and organic–geochemical parameters
occurs at the MIS 3r2 boundary in all sediment
cores ŽŽFigs. 3, 4 and 8.; cf. Andersen et al., 1996..
The central part of this layer exhibits lamination and
contains high amounts of mature terrestrial organic
material, very low carbonate Žbeing mainly dolomite.
and a very distinct Žclay. mineralogy with extremely
low smectite, but high kaolinite percentages accompanied by the occurrence of ordered layered expandable minerals ŽOLEM; Fig. 3; cf. Andersen et al.,
1996.. It is labeled Event I in accordance with Knies
and Stein Ž1998. and has been deposited between
22.5 and 19.5 14 C ka as based on several AMS
14
C-ages ŽFig. 3; approximately 26–22 ka calendar
27
years according to Voelker et al., 1998.. Here, we
assume it to be a synchronous deposit in all three
sediment cores and use it as one indicator of the
lower MIS 2.
The stable isotope records of core PS2212-3 generally agree with the outlined stratigraphic concept,
although several sediment horizons are barren of
carbonate Že.g. Termination I, Figs. 2 and 4.. Therefore, correlation of mineralogical and organic–geochemical parameters with adjacent well dated cores
help to determine the exact position of the MIS 2r1
boundary at 40 cm bsf Žcf. Pagels, 1991 for the age
of carbonate free horizons during Termination I;
Stein et al., 1994.. Similarly, we use the Event I
Ž19.5–22.5 ka. which occurs in PS2212-3 at about
110 cm bsf as one important stratigraphic marker
ŽFigs. 2–5. to position the MIS boundary 3r2 at 112
cm bsf. In addition, the small d18 O-shift to higher
values at 95 cm bsf and lower values of d13 C could
indicate early MIS 2 ŽFig. 2.. Here and between 120
and 140 cm bsf, Wollenburg et al. Ž2001. found two
occurrences of the agglutinated benthic foraminifera
Siphotextularia rolshauseni, the lower being indicative of MIS 3.2 in the Nordic Seas ŽNees and Struck,
Fig. 3. Sediment characteristics versus depth of Core 90-39 ŽAndersen et al., 1996. and radiocarbon dates. Marine Isotope Stages 1–3 are
indicated by bold numbers and grey shade Žinterglacialsrinterstadials..
28
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
Fig. 4. Sediment characteristics versus depth of Kastenlot core PS2212-3 ŽYermak Plateau., gravity core PS2122-1 and gravity core
PS2123-2 ŽSpitsbergen Coast Domain.. Marine Isotope Stages 1–6 are indicated by bold numbers and grey shade Žinterglacialsrinterstadials..
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
˚ peak.rfeldspar Ž3.24 and 3.18 A˚ peaks.-ratio and
Fig. 5. Mineralogical data of cores PS2122-1, PS2123-2, and PS2212-3 versus depth. Bulk quartz content and quartz Ž4.26 A
˚ ., and relative weight percents of clay mineral groups in the clay fraction Ž - 2 mm. are shown.
k-feldsparrplagioclase ratio Ž3.24r3.18 A
29
30
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
1994.. As in PS2122-1 and in PS2123-2, the occurrence of P. bulloides and the decrease of the carbon
isotope values delineate the MIS 5r4 and MIS 4r3
boundaries. Noticeably, the carbonate free horizon is
again situated in a deglaciation event at MIS 4r3
boundary. Based on this new data, our stratigraphy
differs in MIS 1–3 from the published framework of
Nowaczyk et al. Ž1994. derived from paleomagnetic
and coccolith evidence which is discussed in more
detail in Mathiessen et al. Ž2001.. Even with our
improvements absolute ages might still differ by up
to a few thousand years.
4. Results
4.1. Lithology, IRD and grain-size distribution
In all cores the dominant lithotypes are silty clays
to clayey silts. Brown colours dominate the MIS 1–5
in core PS2212-3 while cores PS2122-1 and PS2123-
2 show regular alternations of brown and grey sediment colours ŽKnies, 1994; Vogt, 1997.. The cores
usually contain a few layers of coarser material
which, however, are only visible in the X-ray radiographs. The coarse fraction content ranges between 0
and 25 wt.% with one exception in core PS2122-1
Ž46 wt.% in MIS 1; Figs. 4 and 6.. In the nearshore
sediment cores, PS2122-1 and PS2123-2, the coarse
fraction is dominated by inorganic terrigenous components. In PS2212-3, sediment layers with lower
coarse fraction content are mainly dominated by
foraminifera shells Žsizes from 63–500 mm.. In the
Western Eurasian Basin, high numbers of calcareous
foraminifers mainly occur in intervals with low sand
contents ŽFig. 4. where they might constitute up to
50% of the total sand fraction ŽNørgaard-Pedersen et
al., 1998.. Dominance of foraminifera in a particular
sediment layer can be traced by high carbonater
calcite contents of the bulk and silt fraction ŽFig. 4,
MIS 3r2.. In general, the pattern of terrigenous IRD
and coarse fraction content seems to be very similar
Fig. 6. Relation of gravel and sand content of bulk sediment Ž%. to oxygen isotope stratigraphy age ŽMartinson et al., 1987. in cores
PS2122-1, PS2123-2 and PS2212-3.
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
Fig. 7. Compilation of the glacial history of Svalbard Žadvance and retreat. along the Isfjord basin for the Weichselian period Žfrom Mangerud et al., 1998., abundance of
ice-rafted material ŽIRD. of the cores PS2122-1, PS2123-2 and PS2212-3 and oxygen isotope stratigraphy ŽPS2122-1.. Gray shaded areas show major IRD events. Cross hatched
area is Event I.
31
32
Table 2
Brief characterization of main sedimentary environments by means of sediment data used in this paper
Bulk
mineralogy
Clay
mineralogy
Lithology
IRD, coarse
fraction
Sedimentation
rate
Oxygen and
carbon isotopes
Carbonate
Organic carbon
Open water
Quartz- 25%,
yFsp,
yyQzrFsp,
yyKfsrPlg
Ill and Chl
) 70%
mainly fine
grain sizes,
brown colors
very low
low
global ocean
values
qq,
qqcalcite,
forams,
coccoliths
yyTOC, qHIvalues Ž )100.,
yCrN-ratio,
low Tma x ,
qmarine origin
Melting at the
Marginal
Ice Zone
qŽclino-.
pyroxene
qqSmectite,
qqKaolinite,
qqQzrFsp
qqfine
fraction,
cryokonites,
fecal pellets
low
large
some lighter
oxygen
values
medium
carbonate,
some
dissolution,
large dilution
qTOC, qqHI,
qpreservation,
mixed maturity
Sea-ice
cover
medium to low
Qz and Fsp
depending on
origin of sea-ice
Ill and Chl
dominant due
to mainly
gravitational
transport from
Svalbard
qfine
fraction
sand
contains
mainly
forams
low
low
stable heavy
oxygen
values,
continously
light carbon
values
qqcalcite
due to small
dilution by
terr. material,
good
preservation
medium TOC,
low HI, qCrN,
qTma x ,
terrigenous
origin
Built-up and
deglaciation
of adjacent
ice sheets
rapid changes,
built up:
qqAmatureB
mineralogy,
qqQz, yyFsp,
yyKfs,
deglaciation;
qqQz,
qqKfs, qqAmf
qq kaolinite
and OLEM
appearance,
qqQzrFspratio, rapid
changes in
clay
mineralogy
highly
variable
lamination,
also hiatus
due to
erosional
processes
possible
qqIRD,
qqsand
fraction
largest
several
meltwater
pulse, light
oxygen and
carbon
values
ycalcite dues
to dilution
and strong
dissolution,
dolomite from
Svalbard
sources
qqTOC, qq
CrN, Ahot shaleB
organic material,
qqpreservation
of allochthonous
and
authochtonous
material
Relative scale: Žqq. strong inputrincreased content to Žyy. lowest inputrcontent.
Qz—quartz; Fsp—feldspar; Plg—plagioclase; Kfs—K-feldspar; Ill—Illite; Chl—Chlorite; Amf—amphibolesrhornblende; OLEM—ordered layered expandable minerals; IRD
—ice rafted debris, dropstones; TOC—total organic carbon; CrN—organic carbonrtotal nitrogen ratio; HI-value, Tma x —data from pyrolysis of organic material.
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
Environment
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
in all cores ŽFigs. 6 and 7.. In the deep sea ŽPS2212-3.
and lower slope cores ŽPS2122-1., however, the
abundance of IRD is lower compared to the near
coast core PS2123-2 ŽFigs. 4 and 7..
4.2. Carbonate and organic carbon contents
The carbonate content in cores PS2122-1 and
PS2123-2 is mostly less than 5% ŽFig. 4.. Only
during MIS 2 values of up to 10% occur. In core
PS2123-2 dolomite contents vary between 0% and
4% often comprising most of the carbonate content.
Only in upper MIS 5 sediments and at the MIS
boundary 3r2 calcite is the dominant carbonate mineral. In core PS2212-3 the carbonate content is highly
variable throughout the whole core ŽFig. 4.. At those
intervals without N. pachyderma sin. Žsee missing
isotope values, Fig. 2. but with significant carbonate
content, the carbonate is mainly dolomite with traces
of siderite.
In all three cores the total organic carbon content
ŽTOC. ranges between 0.4% and 1.0% with distinct
maxima from 1.5% to 2.7% in MIS 6, lower MIS 5
and at the MIS boundaries 4r3, 3r2 and Termination I ŽFig. 4.. The results of the geochemical investigations reveal a dominant terrigenous input which
is consistent with other findings in the Eurasian
Basin Že.g. Schubert and Stein, 1996; terrigenous
organic carbon: CrN-ratios larger than 15, HI-values
below 100 mg HCrg C; Tmax-values above 450 8C..
The values of the CrN-ratio range between 5 and
35. Especially core PS2212-3 displays higher CrNratios above 10, and the HI-values are entirely below
100 mg HCrg C. In all cores the upper MIS 1
sediments show low CrN-ratios and, additionally, in
PS2122-1 and PS2123-2 higher HI-values, suggesting a small increase in marine organic material in
combination with increased Žbiogenic. calcite contents ŽFig. 4..
4.3. Mineralogy
The mineral assemblage in cores PS2123-2 and
PS2212-3 mainly consists of quartz, feldspar, clay
minerals, calcite, dolomite, and accessory heavy
minerals. Core PS2123-2 reveals quartz contents of
19% to nearly 50% Žaverage 35.5%., core PS2212-3
12% to 45% Žaverage 29.5%; Fig. 4.. The quartzr
33
feldspar-ratios ŽQzrFsp. of the bulk fraction range
between 0.19 ŽPS2212-3; average 0.37. and 1.56
ŽPS2123-2; average 0.68. showing a clear difference
in the content of feldspar relatively to quartz ŽTable
2; Fig. 5.. Generally, the ratios are closely related to
the quartz content and exhibit maxima near the MIS
boundary 3r2. In the K-feldsparrplagioclase plot
ŽKfsrPlg. PS2123-2 displays several reductions to 0
due to very low or missing K-feldspar, while only a
few core horizons in PS2212-3 miss K-feldspar. One
of the prominent minima is in Event I sediments
ŽFig. 5..
The clay fraction Ži.e. particles smaller than 2
mm. contains at least 90% clay minerals. The illite
group is the major constituent of the clay mineral
fraction. Chlorite contents are generally constant
around 20%. Quartz is the most important non-clay
mineral in the clay fraction reaching a maximum of
8% ŽVogt, 1997.. Smaller percentages of plagioclase, carbonates, and heavy minerals were observed.
Ordered–layered expandable minerals ŽOLEM; cf.
Reynolds, 1970. occur in a few Event I samples. The
QzrFsp-ratio of the clay fraction of Event I sediments is also strongly increased ŽFig. 5..
For the paleoceanographic reconstruction, smectite and kaolinite are the most important clay mineral
groups. In general, smectite contents range between
0% and 20% and kaolinite between 12% and 30%.
Only a few distinct layer show higher values ŽFig.
5.. Overall, the smectite and kaolinite records of
cores PS2122-1 and PS2123-2 display a very similar
pattern ŽFig. 5.. In core PS2212-3 sediments of
Termination I high smectite Ž) 10%. and kaolinite
Ž) 20%. values coincide, but the phasing of smectite
and kaolinite peaks differs. The most obvious other
feature is a low smectiterhigh kaolinite couple during Event I which combines with the only occurrence of OLEM.
5. Discussion
Due to the complex sedimentary environment at
the Yermak Plateau and off the NW Spitsbergen
coast it is difficult to rely on only a few parameter to
reconstruct the paleoceanographic history. Various
phenomena can influence the sedimentary processes,
34
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
including surface water and sea-ice transport from
the Siberian shelf regions, icebergs, possibly brines
from the near Spitsbergen glaciers or from winter
sea-ice production, and surface and bottom currents
from the Norwegian–Greenland Sea and the central
Arctic Ocean. Therefore, a set of distinct indicators
of sedimentary environments have been compiled
ŽTable 2..
In this paper we will concentrate on two time
intervals Ž0–30 and 50–85 ka. during which well
defined changes in the influx of Atlantic water were
related to the built-up and melt-down of the SBIS
Žcf. Hebbeln et al., 1994; Andersen et al., 1996.. The
younger time interval from 0 to 30 ka comprises the
well investigated upper MIS 3 to 1. We will show
very similar developments of the marine sedimentation west Žsee Elverhøi et al., 1995. and northwest of
Svalbard Žthis study.. More strikingly, the older time
interval Žca. 50–85 ka. exhibits the same sedimentological record. Hence, we assume a similar paleoceanographic and glaciological history for these two
Weichselian SBIS cycles.
5.1. SÕalbardr Barents sea ice sheet adÕances between 27 and 16 ka
This distinct Event I signal ŽTable 2; a AmatureB
organic–geochemical and mineralogical signature.
can be found in all three analyzed cores above the
MIS boundary 3r2 ŽFigs. 4–8.. It is intercalated
between sediments with high calcite content and
increased foraminifera and coccolith abundances
ŽFigs. 4 and 7; PS2212-3: cf. Nowacyzk et al.,
1994.. The increased abundance of planktonic species
Žthe Ahigh productivity zonesB; Dokken and Hald,
1996. suggests seasonally open-water conditions not
only at the SW coast of Spitsbergen but also to the
north up to the northern Yermak Plateau. The seasonally open water could have supported a moisture
supply for the two-step build-up of the northwestern
SBIS between 27 and 23 ka and 19 and 16 ka Žcf.
Hebbeln et al., 1994; Elverhøi et al., 1995.. Additionally, high numbers of IRD larger than 2 mm
ŽFig. 7. suggest an increased iceberg transport before
and after Event I. These high numbers of IRD occur
together with strongly increased quartz contents and
reduced feldspar contents leading to high QzrFspratios ŽFig. 5.. The near-coast core PS2123-2 reveals
the highest values ŽFigs. 5–7.. Increased amounts of
amphiboles in the bulk fraction have been attributed
to the input of crystalline rocks from Fennoscandia
and Svalbard ŽAndersen et al., 1996. and were also
recognized in our cores. One source region for the
quartz-richrfeldspar-depleted material ŽFig. 5. could
be the Paleozoic crystalline strata ŽHekla Hoek. and
the Devonian clastic wedge rocks which extensively
outcrop in northern Spitsbergen ŽWinsnes, 1988..
The Paleozoic material from Spitsbergen and also
from Fennoscandia, however, does contain some
amount of K-feldspar Žcf. Andersen et al., 1996.. In
contrast, the siliciclastic nature of the mature JurassicrCretaceous Shale material, identified as the
source material for Event I, could also produce a
high quartz content while it is extremely depleted in
ŽK-. feldspar ŽAndersen et al., 1996.. The KfsrPlgratio decreases nearly to 0 ŽFig. 5., pointing to a
dominant input from these shales during the finefraction sedimentation of Event I.
As Event I is developed in all three cores similarly, we regard it as a synchronous regional event
representing the advanced SBIS and a time of northward transport of fine fraction material along the
western Svalbard continental slope. The reduction of
the horizons thickness from S to N Žtens of centimeter W of Spitsbergen to a few at the Yermak Plateau.
supports the assumption that currents transported the
fine fraction from Ža. reworking by the advanced
SBIS at the Spitsbergenbanken to Žb. injection into
the intermediate waters of a Paleo-Westspitsbergen
Current through dense, suspension-rich bottom-water
currents from the Storfjorden Trough Žcf. Hebbeln et
al., 1994; Andersen et al., 1996., and Žc. finally
reaching the northern Yermak Plateau at 828N and
water depth of 2500 m Žcompare Fig. 1..
On Spitsbergenbanken and in the area SE of
Spitsbergen only thin Quaternary sediment blankets
cover outcrops of Late Triassic to Early Cretaceous
sedimentary rocks ŽFig. 1.. An Early Cretaceous
Ahot shaleB member of the Mesozoic formations
which is rich in mature organic material and possesses a diagenetically mature mineral assemblage
ŽFig. 3: OLEM clays, no smectite, low feldspar and
no K-feldspar. is of particular importance ŽAndersen
et al., 1996.. This set of AmatureB material was
found in sediment cores south of our study area
ŽFigs. 1 and 3: NP 90-39., additionally manifested
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
Fig. 8. Age versus parameter plot of PS2212-3 from the NE of the Yermak Plateau. Hatched area—TOCrOLEMrkaolinite event; grey shaded areas are MIS 1 and 3.
35
36
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
by the identification of Mesozoic palynomorphs
ŽElverhøi et al., 1995; Andersen et al., 1996.. Event I
is also documented in cores to the south and east of
our cores combined with high abundances of Mesozoic sediment clasts in the IRD ŽFig. 1; Andersen et
al., 1996; Knies and Stein, 1998..
As our cores are rather distal to the source of the
Event I material, the thickness of the horizon is
strongly reduced Ži.e. more than 1 m in NP9039 ŽFig.
3. to 0.02 m in PS2212-3 ŽFigs. 4 and 5... Therefore,
it is difficult to deduce in more detail the developments of the SBIS’s glacier fronts in the SE or NE of
Svalbard on the basis of our records. According to
Dokken Ž1995. the Event I sediments could also
pinpoint a deglazial phase at the southeastern extension of the SBIS while Knies et al. Ž2000. and
Kleiber et al. Ž2000. reconstructed the largest extension of the northeastern SBIS in the vicinity of the
Franz Victoria Trough at the upper continental slope
at approximately 23 ka ŽFig. 1: PS2138-1.. Between
23 and 15.4 ka small undulations of the SBIS front
but no larger deglaciations are reported. Partly laminated fine fraction sediments are deposited in the
Franz Victoria Trough between approximately 22
and 20.5 ka and interpreted as meltwater plumes in
front of the ice sheet ŽKleiber et al. 2000.. Leirdal
Ž1997. reconstructed the same environment north of
the Hinlopenstrait, northern Spitsbergen. A hold of
the ice sheet advance is most probable and the fine
fraction material records the meltwater plumes in
front of the undulating ice sheet. This fine material
would then be redistributed as outlined above. The
strong similarities of the cores from the Western
Fram Strait and PS2212-3 on the northeastern Yermak Plateau slope leads us to assume a Paleo-WSC
very similar to today including a strong intermediate
water component. The position on the Yermak
Plateau slope excludes the influence of Nansen Basin
bottom currents. Sea-ice or icebergs from the Franz
Victoria Trough Žcompare Fig. 1. should have transported more coarse fraction and poorly sorted material.
5.2. Early MIS 2
The deposition of dolomite, in particular the input
of bulk and clay-fraction dolomite at the northeastern
Yermak Plateau site PS2212 was high during early
MIS2 ŽFig. 5; cf. Vogt, 1997.. As dolomite derives
from the northern Svalbard Paleozoic carbonate
rocks, increased iceberg production from the advanced SBIS on the northwestern and northern coast
of Spitsbergen can be assumed including glacial rock
flour Žclay fraction dolomite.. While IRD and sand
sedimentation continued through the entire MIS 2 in
the W-Spitsbergen cores, coarse fraction input diminished at about 17 to 15.5 ka in the northern Yermak
Plateau core PS2212-3, and fine fraction dominated
ŽFig. 8.. Icebergs might be blocked by sea ice from
moving to the Yermak Plateau site or meltout of
particles might be prevented by cold ŽPolar. water
conditions. Core positions closer to the Northern
Barents Sea slope yield strongly increased sedimentation of IRD ŽKubisch, 1992; Knies et al., 2000,
2001..
To the west of Spitsbergen, seasonally open water
conditions prevailed as indicated by increased calcite
contents, low TOC contents, low CrN ratios and
slightly increased HI-values as well as coccolith and
subpolar planktic foraminifera abundances ŽFig. 4;
cf. Hebbeln et al., 1994; Dokken, 1995..
5.3. Intermediate water carbonate dissolution before
and during early Termination I
Core PS2212-3 ŽWD: 2550 m. was affected by
complete carbonate dissolution during Termination I
as outlined by the absence of carbonate ŽFig. 8;
17–14 ka. and the benthic foraminifera assemblage
ŽWollenburg et al., 2001.. During the same time
interval, the NW-Spitsbergen cores PS2122-1 and
PS2123-2, and the neighboring Fram Strait, Nansen
Basin and Gakkel Ridge sediments reveal increased
biogenic calcite contents ŽFig. 4; Stein et al., 1994;
Andersen et al., 1996.. Nearly the complete sand
fraction in Nansen Basin and Gakkel Ridge cores is
comprised of foraminifera Ž90–100%; Markussen et
al., 1985; Nørgaard-Pedersen et al., 1998.. Hence,
the water depth of core PS2212-3 seems to be most
affected. Later, only dolomite was preserved delineating reduced dissolution of carbonate at that time.
All sediment cores on the deeper slope of the northwestern Barents Sea and the Yermak Plateau show
decreased carbonate contents during early Termination I Žaround 15 ka. and are affected by dissolution
and in particular dilution of Žbiogenic. carbonate
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
sedimentation Žcf. Pagels, 1991; Elverhøi et al., 1995;
Hebbeln and Wefer, 1997; Bauch et al., 1999; Knies
et al., 1999..
To explain the differences between PS2212-3 and
the adjacent cores we assume the production of
carbonate aggressive brines in the vicinity of the
ice-sheet andror sea-ice edge north of Spitsbergen
following the model of Steinsund and Hald Ž1994..
Since the adjacent Nansen Basin cores show only
small dissolution of carbonate during this time span
Žcf. Markussen et al., 1985; Pagels, 1991; Vogt,
1997., we regard brine injection as the most probable
process leading to complete carbonate dissolution in
about 2500 m water depth at the NE Yermak Plateau.
Similar processes have been assumed for the Northern Barents Sea Shelf Žcf. Pagels, 1991; Kohler,
¨
1992; Knies and Stein, 1998.. Brine production could
have been caused by processes at the ice-sheet front
or during freeze-up of sea-ice in autumn. In slope
sediment cores west of Spitsbergen benthic foraminifera register oxygen and carbon isotope values which
are indicative of brines ŽLloyd et al., 1996a.. Today,
such brines reach water depth of 2000 m ŽQuadfasel
et al., 1988; Schauer et al., 1997.. A lowered sea-level
and differences in brine andror intermediate water
composition might have led to the increased water
depths of corrosive shelf waters.
5.4. Surface water enÕironment at N Yermak Plateau
during early Termination I
If the dissolution of carbonate is due to brine
ejection during autumn freeze-up, seasonally open
water must have existed near Site PS2212. As biogenic carbonate is absent due to dissolution, other
tracers of open water conditions have to be used for
reconstruction. Sea-ice and icebergs could have been
melted during summer. In this case a sediment of
mixed origin would be dominant, indicating strong
sea-ice detritus deposition together with IRD from
icebergs derived from adjacent Spitsbergen Žcf. Knies
et al., 1999, 2000.. Sea-ice sediment input can be
deduced from increased smectite contents above 15%
combined with increased pyroxene contents ŽFig. 7;
Letzig, 1995; Vogt, 1997; Behrends, 1999.. Increased organic carbon contents with a terrigenous
ŽCrN-ratio ) 10. and partly marine signature ŽHI-
37
index increase., low IRD counts but increased sand
contents and high amounts of the fine fraction Žsilt
and clay. are observed ŽFig. 8: ca. 16 and 14 ka.. All
this evidence indicates sedimentation from melting
sea-ice, entrained on the Siberian shelves and transported to the Yermak Plateau position by the Transpolar Drift ŽNurnberg
et al., 1994; Hebbeln and
¨
Wefer, 1997; Knies and Stein, 1998.. The benthic
foraminifera assemblage strongly supports the idea
of seasonally open water above PS2212-3 ŽWollenburg et al., 2001..
5.5. The Termination I deglaciation (16–9 ka)
In the western Fram Strait high IRD and sand
deposition continued ŽFigs. 4, 6, 7–8.. The d18 O
plots show light values conspicuous of melting events
ŽFigs. 2 and 6.. The mineralogy is typical for Svalbard sources ŽTable 2., which indicates that icebergs
from the Svalbard fjords could drift northward in the
seasonally open water. Thus, we conclude that seasonally open waters prevailed in western Fram Strait
and north of Spitsbergen near the SBIS ice edge
around 16–17 ka. During summer the Polar Front
reached this region. To the north the Nansen Basin
was covered by perennial sea-ice as indicated by low
sedimentation rates of mainly fine grained material
Žcf. Stein et al., 1994; Vogt, 1997; Nørgaard-Pedersen et al., 1998..
In core PS2212-3 the carbonate free horizon
smoothly passes over to sediments with increased
HI-indices Žca. 13 ka.. The same development has
been observed in other cores from the Fram Strait
and the adjacent Nansen Basin ŽAndersen et al.,
1996; Vogt, 1997.. This suggests small increases in
the input of marine organic carbon ŽFig. 8; Knies et
al., 1998.. Higher calcite contents in the NW Spitsbergen cores might also indicate higher productivity
during this second phase of early Termination I.
High fine fraction, kaolinite and maximum smectite
contents in the NE Yermak Plateau core PS2212-3
are probably the result of biologically enhanced Žfecal pellets. sedimentation at the summer ice edge
during the ice-melt induced phytoplankton bloom,
similar to the recent summer situation in the Fram
Strait described by Berner and Wefer Ž1994.. Other
productivity proxies do reflect the increased productivity during this time ŽŽe.g. opal, benthic foramini-
38
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
fera; Nurnberg
et al., 1995; Wollenburg et al., 2001..
¨
The WSC seems to have reached further north,
propelling seasonally open water as far north as
828N.
This occurred contemporary with an increased
transport of sea-ice sediments due to extensive entrainment of Žsmectite rich. shelf sediments during
the flooding of western Siberian shelves by the rising
sea-level Žcf. Forman et al., 1996; Kassens et al.,
1999 and references therein.. Additionally, starting
meltdown of the Kara Sea Ice Sheet might have
enhanced the remobilization of fine-grained material
in the Kara Sea ŽVogt, 1997; Polyak et al., 1997..
A stronger warm water influx with the WSC
system andror the sea-level rise triggered glacial
retreats in Kara SearSt. Anna Trough area and
increased the melting of icebergs above the Yermak
Plateau Žcf. Hebbeln and Wefer, 1997; Polyak et al.,
1997.. The origin of the PS2212-3 sediment during
this time can be clearly related to Siberian shelf
sources and the Franz Josef Land region ŽTable 2;
i.e. high kaolinite, smectite and pyroxene contents,
weak smectite crystallinity, high KfsrPlg-ratio of
bulk sample, high sand contents but small gravel
contents; cf. Vogt, 1997.. Most of the material has
probably been transported by sea-ice and the related
surface waters as well as some icebergs from the St.
Anna Trough and Franz Josef Land Žhigh kaolinite
and K-feldspar contents.. A first IRD peak combined
with strongly increased kaolinite contents has been
recognized by Knies et al. Ž1999. at the northern
Barents Sea slope between 16.8 and 15.4 ka and has
been related to the early melting of the northern
SBIS between Franz Josef Land and Svalbard.
The cores off NW Spitsbergen display highest
IRD-counts near the MIS 2r1 stage boundary with
material originating mainly from Spitsbergen ŽFig.
7.. Elverhøi et al. Ž1995. and Hebbeln et al. Ž1994.
report a distinct melting event of the western SBIS at
14.5 ka in sediment cores which are located south of
the study area ŽFig. 1.. This melting event can also
be recognized in all cores of this study ŽFigs. 6–8,
grey shades. and in cores to the northeast Že.g.
PS2138, Knies et al., 1999, 2001. as well as a
second melting event around 13 ka ŽNørgaard-Pedersen et al., 1998.. Highest coarse fraction contents
during the 13 ka melting occur on the NE Yermak
Plateau site ŽFig. 6., which are accompanied by the
highest single smectite peak within the whole core
ŽFig. 8.. Solely smectite in the fine fraction indicates
input from more eastern regions than Franz Josef
Land and its adjacent troughs, which would have
increased the kaolinite content ŽWahsner et al., 1999..
Our records correspond very well to that of Birks
et al. Ž1994., who summarized the climatological
development of the Svalbard region and concluded
that the deglaciation of the Barents Sea started probably as early as 15 ka, but certainly before 13.3 ka in
the central and southern parts. In the north, the
earliest deglaciation sediments in the Franz Victoria
Trough west of Franz Josef Land are dated to 15.4
ka ŽKleiber et al., 2000..
Between 13 and 12 ka, a rapid retreat of the
western SBIS glaciers signalled the fast decrease of
the ice sheet ŽMangerud et al., 1998., which is also
recorded in the NW Spitsbergen cores by high IRD
signals, increased coarse-grained fraction and high
quartz contents ŽFigs. 4–7.. The early Termination I
sequence in core PS2212-3 yields high quartz contents but low QzrFsp- and KfsrPlg-ratios in the
bulk fraction ŽFig. 5.. This is rather indicative of
crystalline rocks like the Hekla Hoek basement of
Svalbard. In contrast, PS2123-2 exhibits fairly low
quartz contents, the third high carbonatercalcite peak
in MIS 2, lowest TOC contents, CrN ratios below 5,
and the beginning increase of HI values which are all
indicative of Žseasonally. open water ŽTable 2, Figs.
4 and 5.. A drastic increase in sediment accumulation rates is observed Žfrom about 12 to ) 20 grcm2
ka; Knies, 1994.. As the siltrclay-ratio increases
ŽFig. 4., an increase in bottom currents can be
assumed, which is today related to a strong WSC
current activity Žcf. Boulton, 1990.. Finally, high
IRD-input and coarse fraction contents mark the
complete breakdown of the SBIS after 10 ka.
5.6. SÕalbard ice sheet adÕance during stage 5 similar to the 23 ka eÕent?
A comparison of the sedimentological, organic–
geochemical and mineralogical signature of Event I
in core PS2212-3 ŽFig. 8. with the climatic records
from the NW Spitsbergen cores PS2122-1 and
PS2123-2 during the middle to upper MIS 5 reveals
striking similarities ŽFigs. 4–7 and 9: ca. 85 to 75
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
Fig. 9. Age versus parameter plot of PS2122-1 NW of Svalbard. Hatched area—TOCrOLEMrkaolinite event; grey shaded areas are MIS 3 and 5.
39
40
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
ka.. Assuming the same development as during MIS
3 and 2, the occurrence of the Event I-like sediment
layer indicates another SBIS advance to the SE in
late MIS 5. During this time interval core PS2123-2
shows a peak in the IRD record ŽFig. 7..
Unfortunately, parts of this time interval are missing in core PS2212-3 because of a hiatus. The hiatus
might have been caused by increased bottom currents at the NE slope of the Yermak Plateau. Currents might indicate a stronger and deeper WSC or
dense bottom water formed at the front of the advanced northern SBIS andror brine formation during
the winter sea-ice built-up. Alternatively, a debris
flow or turbidity current could have eroded the sediments on the NE slope of the Yermak Plateau, which
was also observed with sub-bottom echosounder investigations in that region ŽPARASOUND data;
Bergmann, 1996.. At the adjacent site PS1533 Žca.
2000 m water depth, 11 km to the west of PS2212-3.
the sediments point to seasonally open-water conditions and winter sea-ice cover with low iceberg
occurrences during MIS 5a ŽPagels, 1991; Kubisch,
1992.. Some seasonally open water is also indicated
by the increased calcite and foraminifera contents,
low CrN-ratios and mainly fine fraction sediment in
PS2122-1 and PS2123-2 before and after the event
ŽFigs. 4 and 9.. Open water would provide moisture
for the build-up of the SBIS as during MIS 2 and 3,
and outlined by Hebbeln et al. Ž1994..
The younger IRD-peak in PS2122-1rPS2123-2
and the Event I-like sediment layer point to another
SBIS advance in MIS 5a. It is again intercalated
between sediments with increased calcite contents,
one indicator for seasonally open water conditions
ŽFig. 9.. This is in accordance with the latest landbased reconstruction of the SBIS, where early advances of the mid-Weichselian SBIS are envisioned
ŽFigs. 7 and 9; Mangerud et al., 1998..
Due to the limited data base for glaciation of
Spitsbergen, some advances registered in the marine
record during upper MIS 5 might be missing in the
onshore record Žcf. Mangerud et al., 1998.. The
cores investigated by Elverhøi et al. Ž1995. do not
reach back into MIS 5. Hence, this is the first
indication of an Event I-like development in an older
record. IRD records of Lloyd et al. Ž1996b. and
Hebbeln and Wefer Ž1997. also show multiple IRD
peaks in MIS 5Žb?.. Investigations of cores N and
NE of Spitsbergen gave evidence of earlier advances
of the northern rim of the SBIS, which coincide with
a strong northward influx of Atlantic water into the
Arctic Ocean by the WSC during MIS 6 ŽVogt,
1997; Knies and Stein, 1998.. Fig. 3a shows at least
two such events in MIS 6 ŽPS2212-3: increased
calcite content.. However, during MIS 5 there is no
indication of an intensive advance at the northern
rim of the SBIS according to IRD-data ŽKnies et al.,
1999, 2001.. This might be due to blocking of
icebergs by a more or less closed sea-ice cover.
At the isotopic stage 5r4 boundary and during
early stage 4, a marked increase of IRD, higher
quartz contents, CrN-ratios between 10 and 15, and
HI values - 50 mg HCrg C in the near coastal core
PS2123-2 indicate dominantly terrigenous sedimentation and the readvance of the Spitsbergen glaciers
ŽFigs. 4–8.. Further offshore, only stage 4 sediments
show the increase in IRD while stage 5r4 boundary
sediments contain several indicators of increased
sea-ice cover Žlow sedimentation rate, mainly fine
fraction sedimentation, low carbonate content being
mainly dolomite, a marked decrease in d13 C-values..
Only the fairly and solely high smectite contents in
PS2122-1 and PS2212-3 could point to some melting
of sea-ice during summer. This could also point to a
continuous sea-ice transport of the Transpolar Drift
with probably Laptev Sea sources Žcf. Vogt, 1997..
Latest reconstructions of the Kara Sea Ice Sheet
assume a strong middle Weichselian glaciation Žcf.
Astakhov et al., 1999.. Hence, sea-ice during MIS 4
could only be built east of the Kara Sea. While in the
central Arctic Ocean a continuos ice-cover is assumed ŽDarby et al., 1997; Nørgaard-Pedersen et al.,
1998 and references therein., a calcite peak in early
MIS 4 records of PS2212-3 ŽFig. 4. and other cores
near the Svalbard continental margin delineate some
influx of Atlantic Water even to the Northeast of
Svalbard ŽDokken and Hald, 1996; Knies and Stein,
1998..
5.7. Deglaciation at the MIS 4 r 3 boundary
Upper MIS 4 sediments of PS2212-3 from the
northeastern Yermak Plateau suggest input of very
fine-grained material with extremely low IRD and
sand fraction contents. No coccoliths were observed
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
in PS2212-3 ŽNowacyzk et al., 1994.. Thus, a perennial sea ice-cover can be assumed during MIS 4 at
the Yermak Plateau in accordance with published
reconstructions Žcf. Pagels, 1991; Kohler,
1992; Ku¨
bisch, 1992; Hebbeln and Wefer, 1997.. Planktonic
foraminifera occurrences are extremely low on the
Yermak Plateau, and in the Nansen Basin Žcf. Pagels,
1991. the calcite contents are reduced ŽFig. 4.. Increased dolomite contents in MIS 4 sediments point
to Svalbard sources of the sediment ŽFig. 4.. The
cores NW of Spitsbergen still receive coarse fraction
material being highest in the near coast core PS21232 ŽFigs. 4, 6 and 7.. The isotope data show a few
light values indicating several melting events west of
Spitsbergen ŽFig. 6.. Hence, at least some seasonally
open water conditions were present. Increased quartz
contents, and high KfsrPlg- and clay fraction Qzr
Fsp-ratios as well as increased smectite and kaolinite
contents point to melting of sea-ice with probably
Siberian origin.
Similar to the way the advance sequence of the
SBIS between upper MIS 5 and MIS 4 resembles the
last glaciation at the MIS 3r2 boundary, the
deglaciation record at the MIS 4r3 boundary resembles Termination I data ŽFigs. 4–9.. On the northeastern Yermak Plateau, a PS2212-3 core interval with full carbonate dissolution is overlain by
sediments with increased coarse fraction and IRD
contents ŽFig. 4.. The mineralogical data indicates
eastern sources. Increased kaolinite and smectite contents and increased KfsrPlg-ratios and QzrFsp-ratios
of the clay fraction are indicative of Franz Josef
Land and Siberian shelves east of the Archipelago
ŽVogt, 1997.. IRD of Nansen Basin sediments during
the early MIS 3 deglaciation have been attributed to
Siberian origin ŽKubisch, 1992; Vogt, 1997,
Nørgaard-Pedersen et al., 1998.. West of Spitsbergen, Svalbard and Fennoscandian sources were dominant Žcf. Hebbeln and Wefer, 1997.. Hence, the
Polar Front between warmer Atlantic and cold Polar
Waters has been in a similar position as today.
41
clay fraction can be related to Late Quaternary
glacialrinterglacial changes. The use of single-source
minerals as provenance indicators allows to decipher
different transport and sedimentation processes in the
Arctic Ocean during glacial and interglacial times.
Other geochemical and sedimentological parameters
support these changes Ži.e. isotope record, carbonate
and organic carbon content and composition, grainsize distribution and IRD content..
A very distinct signal combination including
high TOC content and CrN-ratios, high kaolinite
content, low carbonate and randomly ordered smectite contents, and the OLEM occurrence is related to
the 23–19 ka initial advance of the SvalbardrBarents
Sea Ice Sheet. We could add additional organic
geochemistry and mineralogical information to this
Event I especially with regard to its occurrence on
the north eastern Yermak Plateau.
Furthermore, we were able to recognize a Event
I-like signal during late MIS 5 for the first time. This
Svalbard ice sheet advance might not be properly
recorded in onshore data.
Using a combination of organic geochemical
and mineralogical parameters, intensified warm Atlantic water influx as far north as the Yermak Plateau
could be stated for late MIS 4rearly MIS 3, late
MIS 3rearly MIS 2, early Termination I and the
Holocene. Termination I and the MIS boundary 4r3
exhibit a very similar pattern of strong Atlantic water
influx into the Arctic Ocean, leading to a northward
migration of the summer ice edge. Early MIS 3 can
be compared to Termination I.
Paleoceanographic and ice sheet developments
of the Svalbard region can be recognized by means
of mineralogical and grain size parameter being independent from productivity and stable isotope data.
This enables us to use sediments from deeper water
cores for paleoceanographic reconstructions, although they have been influenced by corrosive bottom water conditions and carbonate dissolution.
v
v
v
v
Acknowledgements
6. Conclusion
Downcore variations of mineral assemblages
and distributions within the bulk sediment and the
v
The authors thank Antje Volker
and Trond Dokken
¨
for critical comments and thoughtful reviews which
improved and strengthened the manuscript. We thank
the captain and the crew of the RV Polarstern for
42
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
excellent cooperation during the 1991 expeditions.
For technical assistance and data discussion we thank
M. Wahsner, D. Nurnberg,
N. Nørgaard-Pedersen,
¨
C.J. Schubert, R. Frohlking,
M. Siebold, and H.
¨
Grobe and M. Seebeck for grain size data and clay
sample preparation. B. Diekmann and D.-K. Futterer
¨
critically read the manuscript in an early stage which
is much appreciated. The English was improved by
Emma Eades. This work was partially funded by
DFG ŽDeutsche Forschungsgemeinschaft. under the
contracts Fi-443r1,2 and 3. We benefited from the
European Science Foundation project APolar North
Atlantic Margins ŽPONAM.B including early data
discussions on PONAM workshops and the QUEEN
program. Data of this paper are stored in the PANGAEA database at http:rrwww.pangaea.de.
References
Aagaard, K., Carmack, E.C., 1994. The Arctic Ocean and climate:
a perspective. In: Johannessen, O.M., Muench, R.D., Overland, J.E. ŽEds.., The Polar Oceans and their Role in Shaping
the Global Environment: The Nansen Centennial Volume.
Geophysical Monograph, vol. 85. American Geophysical
Union, Washington, DC, pp. 5–20.
Alley, R.B., 1995. Resolved: the arctic controls global climate
change. In: Smith Jr., W.O., Grebmeier, J.M. ŽEds.., Arctic
Oceanography: Marginal Ice Zones and Continental shelves.
Coastal and Estuarine Studies, vol. 49. American Geophysical
Union, Washington, DC, pp. 263–284.
Andersen, E.S., Dokken, T.M., Elverhøi, A., Solheim, A., Fossen,
I., 1996. Late Quaternary sedimentation and glacial history of
the western Svalbard margin. Mar. Geol. 133, 123–156.
Anderson, L.G., Bjork,
¨ G., Holby, O., Jones, E.P., Kattner, G.,
Koltermann, K.P., Liljeblad, B., Lindegren, R., Rudels, B.,
Swift, J., 1994. Water masses and circulation in the eurasian
basin: results from the Oden 91 North Pole expedition. J.
Geophys. Res. 99 ŽC2., 3273–3283.
Astakhov, V.I., Matiouchkov, A., Svendsen, J.I., Mangerud, J.,
Maslenikova, O., Tveranger, J., 1999. Marginal formations of
the last Kara and Barents ice sheets in northern European
Russia. Boreas 28 Ž1., 23–45.
Bauch, H.A., Erlenkeuser, H., Fahl, K., Spielhagen, R.A., Weinelt,
M.S., Andruleit, H., Henrich, R., 1999. Evidence for steeper
Eemian than Holocene sea surface temperature gradient between Arctic and sub-Arctic regions. Palaeogeogr., Palaeoclimatol., Palaeoecol. 145 Ž1–3., 95–117.
Behrends, M., 1999. Reconstruction of sea-ice drift and terrigenous sediment supply in the late quaternary: heavy-mineral
associations in sediments of the Laptev-Sea continental margin
and the central Arctic Ocean. Reports on Polar Research, vol.
310. Alfred Wegener Institut Bremerhaven, Bremerhaven, 167
pp.
Bergmann, U., 1996. Interpretation of digital Parasound
echosounder records of the eastern Arctic Ocean on the basis
of sediment physical properties. Rep. Pol. Res., vol. 183.
Alfred Wegener Institute, Bremerhaven, 164 pp.
Berner, H., Wefer, G., 1994. Clay–mineral flux in the Fram Strait
and Norwegian Sea. Mar. Geol. 116, 327–345.
Birks, H.H., Paus, A., Alm, T., Mangerud, J., Landvik, J.Y., 1994.
Late Weichselian environmental change in Norway including
Svalbard. J. Quat. Sci. 9 Ž2., 133–145.
Boulton, G.S., 1990. Sedimentary and sea level changes during
glacial cycles and their control on glacimarine facies architecture. In: Dowdeswell, J.A., Scourse, J.D. ŽEds.., Glacimarine
Environments: Processes and Sediments. Geol. Soc. Spec.
Publ. vol. 53. The Geological Society, London, pp. 15–52.
Darby, D.A., Bischof, J.F., Jones, G.A., 1997. Radiocarbon
chronology of depositional regimes in the western Arctic
Ocean. Deep-Sea Res. II 44 Ž8., 1745–1757.
Dokken, T.M., 1995. Poleoceanographic changes during the last
Interglacial–Glacial cycle from the Svalbard–Barents Sea
margin: Implications for ice sheet growth and decay. Dr.
Scient. Thesis Žunpublished., Inst. Biol. Geol., Univ. of
Tromsø, 175 pp.
Dokken, T.M., Hald, M., 1996. Rapid climatic shifts during
isotope stages 2–4 in the Polar North Atlantic. Geology 24
Ž7., 599–602.
Elverhøi, A., Andersen, E.S., Dokken, T., Hebbeln, D., Spielhagen, R.F., Svendsen, J.I., Sørflaten, M., Rørnes, A., Hald, M.,
Forsberg, C.F., 1995. The growth and decay of the Late
Wechselian Ice Sheet in western Svalbard and adjacent areas
based on provenance studies of marine sediments. Quat. Res.
44, 303–316.
Emmermann, R., Lauterjung, J., 1990. Double X-ray analysis of
cuttings and rock flour: a powerful tool for rapid and reliable
determination of borehole lithostratigraphy. Sci. Drill. 1, 269–
282.
Forman, S.L., Lubinski, D., Miller, G.H., Matishov, G., Korsun,
S., Snyder, J., Herlihy, F., Weihe, R., Myslivets, V., 1996.
Postglacial emergence of Western Franz Josef Land, Russia
and retreat of the Barents Sea Ice Sheet. Quat. Sci. Rev. 15,
77–90.
Futterer,
D.K. ŽEd.., 1992. ARCTIC ’91: The Expedition ARK¨
VIIIr3 of RV APolarsternB in 1991. Rep. Pol. Res., vol. 107.
Alfred Wegener Institute, Bremerhaven, 267 pp.
Grobe, H., 1987. A simple method for determination of ice rafted
debris in sediment cores. Polarforschung 57 Ž3., 123–126.
Haake, F.W., Plaumann, U., 1989. Late Pleistocene foraminiferal
stratigraphy on the Vøring Plateau, Norwegian Sea. Boreas 18
Ž4., 343–356.
Hebbeln, D., Wefer, G., 1997. Late Quaternary paleoceanography
in the Fram Strait. Paleoceanography 12 Ž1., 65–78.
Hebbeln, D., Dokken, T., Andersen, E.S., Hald, M., Elverhøi, A.,
1994. Moisture supply for northern ice-sheet growth during
the Last Glacial Maximum. Nature 370, 357–359.
Kassens, H., Bauch, H.A., Dmitrenko, I., Eicken, H., Hubberten,
H.-W., Melles, M., Thiede, J., Timokhov, L. ŽEds.., 1999.
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
Land–Ocean Systems in the Siberian Arctic: Dynamics and
History. Springer-Verlag, Berlin, Heidelberg, 711 pp.
Kleiber, H.P., 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 Ž1–4., 25–44.
Knies, J., 1994. Spatquatare
¨
¨ Sedimentation am Kontinentalhang
nordwestlich Spitzbergens. Der letzte GlazialrInterglazialZyklus Žunpublished Dipl. thesis. Justus-Liebig-Universitat,
¨
Gießen, 95 pp.
Knies, J., Stein, R., 1998. New aspects of organic carbon deposition and its paleoceanographic implications along the northern
Barents Sea margin during the last 30,000 years. Paleoceanography 13 Ž4., 384–394.
Knies, J., Vogt, C., Stein, R., 1999. Late Quaternary growth and
decay of the SvalbardrBarents sea ice sheet and paleoceanographic evolution in the adjacent Arctic ocean. Geo-Mar. Lett.
18, 195–202.
Knies, J., Nowaczyk, N., Muller,
C., Vogt, C., Stein, R., 2000. A
¨
multiproxy approach to reconstruct the environmental changes
along the Eurasian continental margin over the last 150,000
years. Mar. Geol. 163 Ž1–4., 317–344.
Knies, J., Kleiber, H.-P., Nowaczyk, N., Mathiessen, J., Muller,
¨
C., Niessen, F., Stein, R., Weiel, D., 2001. Marine ice-rafted
debris records constrain maximum extent of Saalian and Weichselian ice-sheets along the northern Eurasian Margin. Global
and Planetary Change 31, 45–64.
Kohler,
S.E.I., 1992. Spatquartare
En¨
¨
¨ palao-ozeanographische
¨
twicklung des Nordpolarmeeres und Europaischen
Nord¨
meeres anhand von Sauerstoff-und Kohlenstoffisotopenverhaltnissen
der planktischen Foraminifere Neogloboquadrina
¨
pachyderma Žsin... GEOMAR Rep., vol. 13. GEOMAR, Kiel,
104 pp.
Kubisch, M., 1992. Die Eisdrift im Arktischen Ozean wahrend
der
¨
letzten 250,000 Jahre. GEOMAR Rep., vol. 16. GEOMAR,
Kiel, 100 pp.
Leirdal, G., 1997. Senkværtær Utvikling Av Kontinentalmarginen
Nord for Svalbard. unpubl. Ms. Thesis, University of Oslo,
Oslo, 141 pp.
Letzig, T., 1995. Sea ice-transported lithogenic fine fraction of
late quaternary deep-sea sediments of the central eastern Arctic Ocean and the Fram Strait. Reports on Polar Research, vol.
162. Alfred Wegener Institute, Bremerhaven, 98 pp.
Lloyd, J.M., Kroon, D., Laban, C., Boulton, G.S., 1996a.
Deglaciation history and paleoceanography of the western
Spitsbergen margin since the last glacial maximum. In: Andrews, J.T., Austin, W.E.N., Bergsten, H., Jennings, A.E.
ŽEds.., Late Quaternary Paleoceanography of the North Atlantic Margins, vol. 111. Geol. Soc. Spec. Publ, London, pp.
289–301.
Lloyd, J.M., Kroon, D., Boulton, G.S., Laban, C., Fallick, A.,
1996b. Ice rafting history from the Spitsbergen ice cap over
the last 200 kyr. Mar. Geol. 131, 103–121.
Lubinski, D.J., Korsun, S., Polyak, L., Forman, S.L., Lehman,
S.J., Herlihy, F.A., Miller, G.H., 1996. The last deglaciation of
the Franz Victoria Trough, northern Barents Sea. Boreas 25,
89–100.
43
Mangerud, J., Gulliksen, S., 1975. Apparent radiocarbon ages of
recent marine shells from Norway, Spitsbergen, and Arctic
Canada. Quat. Res. 5, 273–296.
Mangerud, J., Dokken, T., Hebbeln, D., Heggen, B., Ingolfsson,
´
Ó., Landvik, J.I., Mejdahl, V., Svendsen, J.I., Vorren, T.O.,
1998. Fluctuations of the Svalbard Barents Sea Ice Sheet
during the last 150,000 years. Quat. Sci. Rev. 17 Ž1–3.,
11–42.
Manley, T.O., Bourke, R.H., Hunkins, K.L., 1992. Near-surface
circulation over the Yermak Plateau in northern Fram Strait. J.
Mar. Syst. 3, 107–125.
Markussen, B., Zahn, R., Thiede, J., 1985. Late Quaternary
sedimentation in the eastern Arctic Basin: Stratigraphy and
depositional environment. Palaeogeogr., Palaeoclimatol.,
Palaeoecol. 50, 271–284.
Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C.,
Shackleton, N.J., 1987. Age dating and the orbital theory of
the ice ages: development of a high-resolution 0 to 300,000
years chronostratigraphy. Quat. Res. 27, 1–27.
Mathiessen, J., Knies, J., Nowaczyk, N.R., Stein, R., 2001. Late
Quaternary dinoflagellate cyst stratigraphy at the Eurasian
continental margin, Arctic Ocean: indications for Atlantic
water inflow in the past 150,000 years. Global and Planetary
Change 31, 65–86.
Muller,
G., 1967. Methods in sedimentary petrology. In: von
¨
Engelhardt, W., Fuchtbauer,
H., Muller,
G. ŽEds.., Sedimen¨
¨
tary Petrology, vol. 1. Schweizerbart, Stuttgart, 283 pp.
Nees, S., Struck, U., 1994. The biostratigraphy and paleoceanographic significance of Siphotextularia rolshauseni Phleger
and Parker in Norwegian–Greenland Sea sediments. J.
Foraminiferal Res. 24 Ž4., 233–240.
Nørgaard-Pedersen, N., Spielhagen, R.F., Thiede, J., Kassens, H.,
1998. Central Arctic surface ocean environment during the
past 80,000 years. Paleoceanography 13 Ž2., 193–204.
Nowaczyk, N.R., Fredrichs, T.W., Eisenhauer, A., Gard, G., 1994.
Magnetostratigraphic data from late Quaternary sediments from
the Yermak Plateau, Arctic Ocean: evidence for four geomagnetic polarity events within the last 170 Ka of the Brunshes
Chron. Geophys. J. Int. 117, 453–471.
Nurnberg,
D., Wollenburg, I., Dethleff, D., Eicken, H., Kassens,
¨
H., Letzig, T., Reimnitz, E., Thiede, J., 1994. Sediments in
Arctic sea ice: implications for entrainment, transport and
release. Mar. Geol. 119, 185–214.
Nurnberg,
D., Schubert, C.J., Stein, R., Vogt, C., 1995. Biogenic
¨
barium and opal in Arctic Ocean sediments - do they reflect
paleoproductivity? EOS Trans. AGU 76 Ž16rSpring Meet.
Suppl.., S172.
Pagels, U., 1991. Sedimentologische Untersuchungen und Bestimmungen der Karbonatlosung
in spatquartaren
¨
¨
¨ Sedimenten des
ostlichen
Arktischen Ozeans. GEOMAR Rep., vol. 10. GEO¨
MAR, Kiel, 106 pp.
Petschick, R., Kuhn, G., Gingele, F.X., 1996. Clay mineral distribution in surface sediments of the South Atlantic: sources,
transport, and relation to oceanography. Mar. Geol. 130, 203–
229.
Polyak, L., Forman, S.L., Herlihy, F.A., Ivanov, G., Krinitsky, P.,
44
C. Vogt et al.r Global and Planetary Change 31 (2001) 23–44
1997. Late Weichselian deglaciation history of the Svyataya
ŽSaint. Anna Trough, northern Kara Sea, Arctic Russia. Mar.
Geol. 143, 169–188.
Quadfasel, D., Rudels, B., Kurz, K., 1988. Outflow of dense water
from a Svalbard fjord into the Fram Strait. Deep-Sea Res. I 35
Ž7., 1143–1150.
Rachor, E., 1992. Scientific Report of RV APolarsternB Cruise
ARK-VIIIr2. Rep. Pol. Res., vol. 115. Alfred Wegener Institute, Bremerhaven, 150 pp.
Reynolds Jr., R.C., 1970. The nature of interlayering in mixedlayer illite–montmorillonites. Clays Clay Mineral. 18, 25–36.
Schauer, U., Muench, R.D., Rudels, B., Timokhov, L., 1997.
Impact of eastern Arctic shelf waters on the Nansen Basin
intermediate layer. J. Geophys. Res. 102 ŽC2., 3371–3382.
Schubert, C.J., Stein, R., 1996. Deposition of organic carbon in
late quaternary Arctic ocean: terrigenous supply vs. marine
productivity. Org. Geochem. 24 Ž4., 421–436.
Stein, R., Schubert, C., Vogt, C., Futterer,
D., 1994. Stable
¨
isotope stratigraphy, sedimentation rates, and salinity changes
in the Latest Pleistocene to Holocene eastern central Arctic
ocean. Mar. Geol. 119, 333–355.
Steinsund, P.I., Hald, M., 1994. Recent calcium carbonate dissolu-
tion in the Barents Sea: paleoceanographic applications. Mar.
Geol. 117, 303–316.
Voelker, A.H.L., Sarnthein, M., Grootes, P.M., Erlenkeuser, H.,
Lay, C., Mazaud, A., Nadeau, M.-J., Schleicher, M., 1998.
Correlation of marine 14 C ages from the Nordic Seas with the
GISP2 isotope record: implications for 14 C calibration beyond
25 ka BP. Radiocarbon 40 Ž1., 517–534.
Vogt, C., 1997. Regional and temporal variations of mineral
assemblages in Arctic Ocean sediments as climatic indicator
during glacialrinterglacial changes. Rep. Pol. Res., vol. 251.
Alfred Wegener Institute, Bremerhaven, 309 pp.
Wahsner, M., Muller,
C., Stein, R., Ivanov, G., Levitan, M.,
¨
Shelekhova, E., Tarasov, G., 1999. Clay–mineral distribution
in surface sediments of the Eurasian Arctic Ocean and continental margin as indicator for source areas and transport
pathways—a synthesis. Boreas 28 Ž1., 215–233.
Winsnes, T.S., 1988. Geological map 1:1000000. Bedrock map of
Svalbard and Jan Mayen. Nor. Polar. Temakart. 3, 12 pp.
Wollenburg, J.E., Kuhnt, W., Mackensen, A., 2001. Changes in
Arctic Ocean paleoproductivity and hydrography during the
last 145 kyr: the benthic foraminiferal record. Paleoceanography 16 Ž1., 65–77.