15
Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14 C yr BP Tine L. Rasmussen a, , Erik Thomsen b , Marta A. Ślubowska a,c , Simon Jessen c , Anders Solheim d , Nalân Koç e a Department of Geology, University of Tromsø, Dramsveien 201, N-9037 Tromsø, Norway b Department of Earth Sciences, University of Aarhus, DK-8000 Århus C, Denmark c The University Centre in Svalbard (UNIS), PO Box 156, N-9171 Longyearbyen, Norway d Norwegian Geotechnical Institute (NGI), PO Box 3930, Ullevaal Stadion, N-0806 Oslo, Norway e Norwegian Polar Institute, Polar Environmental Centre, N-9296 Tromsø, Norway Received 2 December 2005 Available online 23 August 2006 Abstract Two cores from the southwestern shelf and slope of Storfjorden, Svalbard, taken at 389 m and 1485 m water depth have been analyzed for benthic and planktic foraminifera, oxygen isotopes, and ice-rafted debris. The results show that over the last 20,000 yr, Atlantic water has been continuously present on the southwestern Svalbard shelf. However, from 15,000 to 10,000 14 C yr BP, comprising the Heinrich event H1 interval, the BøllingAllerød interstades and the Younger Dryas stade, it flowed as a subsurface water mass below a layer of polar surface water. In the benthic environment, the shift to interglacial conditions occurred at 10,000 14 C yr BP. Due to the presence of a thin upper layer of polar water, surface conditions remained cold until ca. 9000 14 C yr BP, when the warm Atlantic water finally appeared at the surface. Neither extensive sea ice cover nor large inputs of meltwater stopped the inflow of Atlantic water. Its warm core was merely submerged below the cold polar surface water. © 2006 University of Washington. All rights reserved. Keywords: Paleoceanography; Atlantic water; polar water; Svalbard shelf and slope; Foraminifera; Oxygen isotopes; IRD Introduction Warm surface water from the Atlantic Ocean is the major source of heat and salt to high northern latitudes and important for the generation of deep water in the Nordic seas (Aagaard et al., 1985). The western Svalbard shelf and slope is situated at the northernmost reach of the warmer Atlantic water before it enters the Arctic Ocean and continues as the warm Atlantic layer below the cold polar surface waters (Fig. 1a). A branch of the warmer water flows into Storfjorden, where it meets with the cold, low-salinity polar water derived from the polar East Spitsbergen Current, which enters the Barents Sea from the Arctic Ocean (Loeng, 1991; Pfirman et al., 1994; Skogseth et al., 2005)(Fig. 1b). The boundary between the warm Atlantic water and the cold polar water in Storfjorden is termed the Arctic Front (Loeng, 1991; Saloranta and Svendsen, 2001)(Fig. 1b). The Atlantic water loses heat at the surface on its way towards the Arctic and around 78°N the warmest part of the current is found at depths between 100 and 500 m (Aagaard and Carmack, 1989; Boyd and D'Asaro, 1994; Rudels et al., 1999; Saloranta and Haugan, 2004). The flow of Atlantic water into high latitudes is dependent on the general climatic conditions and the distribution of po- lar surface water. In periods of colder climate than at present, the polar surface water has a wider distribution and the Atlantic water will cool more rapidly, causing the warm core to submerge at lower latitudes. A reduction, or even ter- mination, of the flow of Atlantic water into the Nordic seas caused by increased supply of meltwater and/or spreading of polar water have been postulated for several periods of the last glaciation (e.g., Ganopolski and Rahmstorf, 2001; Rahmstorf, 2002). In the present paper, we reconstruct the distribution of the Atlantic water and polar water on the southwestern margin of Quaternary Research 67 (2007) 100 114 www.elsevier.com/locate/yqres Corresponding author. Fax: +47 7764 5600. E-mail address: [email protected] (T.L. Rasmussen). 0033-5894/$ - see front matter © 2006 University of Washington. All rights reserved. doi:10.1016/j.yqres.2006.07.002

Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Embed Size (px)

Citation preview

Page 1: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

(2007) 100–114www.elsevier.com/locate/yqres

Quaternary Research 67

Paleoceanographic evolution of the SW Svalbard margin (76°N)since 20,000 14C yr BP

Tine L. Rasmussen a,⁎, Erik Thomsen b, Marta A. Ślubowska a,c, Simon Jessen c,Anders Solheim d, Nalân Koç e

a Department of Geology, University of Tromsø, Dramsveien 201, N-9037 Tromsø, Norwayb Department of Earth Sciences, University of Aarhus, DK-8000 Århus C, Denmark

c The University Centre in Svalbard (UNIS), PO Box 156, N-9171 Longyearbyen, Norwayd Norwegian Geotechnical Institute (NGI), PO Box 3930, Ullevaal Stadion, N-0806 Oslo, Norway

e Norwegian Polar Institute, Polar Environmental Centre, N-9296 Tromsø, Norway

Received 2 December 2005Available online 23 August 2006

Abstract

Two cores from the southwestern shelf and slope of Storfjorden, Svalbard, taken at 389 m and 1485 m water depth have been analyzed forbenthic and planktic foraminifera, oxygen isotopes, and ice-rafted debris. The results show that over the last 20,000 yr, Atlantic water has beencontinuously present on the southwestern Svalbard shelf. However, from 15,000 to 10,000 14C yr BP, comprising the Heinrich event H1 interval,the Bølling–Allerød interstades and the Younger Dryas stade, it flowed as a subsurface water mass below a layer of polar surface water. In thebenthic environment, the shift to interglacial conditions occurred at 10,000 14C yr BP. Due to the presence of a thin upper layer of polar water,surface conditions remained cold until ca. 9000 14C yr BP, when the warm Atlantic water finally appeared at the surface. Neither extensive sea icecover nor large inputs of meltwater stopped the inflow of Atlantic water. Its warm core was merely submerged below the cold polar surface water.© 2006 University of Washington. All rights reserved.

Keywords: Paleoceanography; Atlantic water; polar water; Svalbard shelf and slope; Foraminifera; Oxygen isotopes; IRD

Introduction

Warm surface water from the Atlantic Ocean is the majorsource of heat and salt to high northern latitudes and importantfor the generation of deep water in the Nordic seas (Aagaard etal., 1985). The western Svalbard shelf and slope is situated atthe northernmost reach of the warmer Atlantic water before itenters the Arctic Ocean and continues as the warm Atlanticlayer below the cold polar surface waters (Fig. 1a). A branch ofthe warmer water flows into Storfjorden, where it meets withthe cold, low-salinity polar water derived from the polar EastSpitsbergen Current, which enters the Barents Sea from theArctic Ocean (Loeng, 1991; Pfirman et al., 1994; Skogsethet al., 2005) (Fig. 1b). The boundary between the warmAtlanticwater and the cold polar water in Storfjorden is termed the Arctic

⁎ Corresponding author. Fax: +47 7764 5600.E-mail address: [email protected] (T.L. Rasmussen).

0033-5894/$ - see front matter © 2006 University of Washington. All rights reservdoi:10.1016/j.yqres.2006.07.002

Front (Loeng, 1991; Saloranta and Svendsen, 2001) (Fig. 1b).The Atlantic water loses heat at the surface on its way towardsthe Arctic and around 78°N the warmest part of the current isfound at depths between 100 and 500 m (Aagaard and Carmack,1989; Boyd and D'Asaro, 1994; Rudels et al., 1999; Salorantaand Haugan, 2004).

The flow of Atlantic water into high latitudes is dependenton the general climatic conditions and the distribution of po-lar surface water. In periods of colder climate than at present,the polar surface water has a wider distribution and theAtlantic water will cool more rapidly, causing the warm coreto submerge at lower latitudes. A reduction, or even ter-mination, of the flow of Atlantic water into the Nordic seascaused by increased supply of meltwater and/or spreadingof polar water have been postulated for several periods ofthe last glaciation (e.g., Ganopolski and Rahmstorf, 2001;Rahmstorf, 2002).

In the present paper, we reconstruct the distribution of theAtlantic water and polar water on the southwestern margin of

ed.

Page 2: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Figure 1. (a) Map of the Nordic seas and the Barents Sea showing major surface currents systems and location of cores. (b) Detail map of the Storfjorden area showingmean position of the Arctic Front (after Loeng, 1991) and location of cores. (c) Depth profile (A–Á) along the centreline of Storfjorden and the Storfjorden Fan.Position of brine basin is indicated. Location of profile is shown in panel b.

101T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Svalbard during the last ca. 20,000 yr. The purpose of the studyis to improve our understanding of the behavior and interplay ofthe two water masses under widely different climatic condi-tions. The study area is situated close to the boundary betweenthe warm Atlantic water and the cold, icy polar water in an areawhere even small changes in their distributional pattern shouldbe recorded (Fig. 1).

The study period comprises (1) the last glacial maximum(LGM); (2) the Heinrich event H1 interval, when numerousicebergs were delivered to the North Atlantic (e.g., Heinrich,1988; Bond et al., 1993); (3) the warmer Bølling and Allerødinterstades; (4) the cold Younger Dryas stade; and (5) theHolocene, including the early Holocene climate optimum,when the climate was warmer than today (e.g., Salvigsen et al.,1992).

Geological and oceanographic setting

Storfjorden is located between the island of Spitsbergen tothe west and the islands of Barentsøya, Edgeøya, and theStorfjordenbanken to the east (Fig. 1b). The inner part isshallow, deepening into an elongated basin with maximumdepth ca. 190 m and a sill of ca. 120 m water depth at the

southern end of the fjord. A broad trough, the StorfjordenTrough, extends from the mouth of the fjord to the shelf edge.The trough was generated by glacial erosion by former icestreams of the Svalbard-Barents Sea ice sheet (Laberg andVorren, 1995; Dowdeswell et al., 1996). A large fan, theStorfjorden Fan, is situated in front of the trough. It consists ofnumerous glacial debris flows separated by interglacialhemipelagic sedimentation (Hjelstuen et al., 1996; Laberg andVorren, 1996).

Over the slope at the Storfjorden Fan, CTD data identifywarm Atlantic water with a thickness of about 500 m and atemperature of 4–7°C overlying the cold intermediate water(T=−0.9 C) generated by convection in the Nordic seas(Fig. 2). The Storfjorden shelf and slope are generally ice-freeexcept for drift ice carried by the East Spitsbergen Current.Storfjorden is important, because wind-driven brine formsduring the winter in the innermost, shallow part of the fjord (e.g.,Quadfasel et al., 1988; Schauer, 1995; Schauer and Fahrbach,1999; Haarpainter et al., 2001a,b; Skogseth et al., 2005). Thebrine accumulates in the deeper basin behind the sill (Fig. 1c).Flow of brines over the sill does not occur every year and mayvary in duration and in water mass properties, but usually it takesplace from March to August (Schauer, 1995; Schauer and

Page 3: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Figure 2. Conductivity–temperature–depth (CTD) data of temperature and salinity from the sites of cores (a) JM02-460GC/JM02-460PC and (b) JM03-373PC2 takenon May 9, 2002 and May 8, 2003, respectively.

102 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Fahrbach, 1999; Haarpainter et al., 2001b). The rest of the year,the floor of the Storfjorden Trough is influenced by Atlanticwater with a bottom water temperature up to about +3.5°C(Schauer and Fahrbach, 1999).

Materials and methods

Two cores, gravity core JM02-460GC (length 387 cm) andpiston core JM02-460PC (length 714 cm), were taken in theStorfjorden Trough at the same position from a water depth of389 m. The core site is situated beneath warm Atlantic waterclose to the Arctic Front (Fig. 1b). Cold, dense brine fromStorfjorden occasionally overflows the core site on its passageinto the deeper areas (Fig. 2). The current from the brine flow isrelatively weak (Schauer and Fahrbach, 1999; Skogseth et al.,2004, 2005) and is not expected to cause reworking at the coresites. A typical brine flow measured in May 2002 had atemperature of −1.8°C and a salinity of 35.2‰ (Fig. 2). Asecond piston core, JM03-373PC2 (length 847 cm) was taken inthe Storfjorden Fan from a water depth of 1485 m beneath thewarmer Atlantic surface water and within the cold intermediatewater (Figs. 1b and 2).

Before opening, the cores were analyzed for magneticsusceptibility and density using a GEOTEK Multisensor CoreLogger. After logging, the cores were split, X-rayed, andsampled at 5-cm intervals with 1-cm-thick samples. Selectedintervals were sampled at 2-cm intervals. All samples weresieved over mesh sizes of 63 and 100 μm. Benthic and plankticforaminifera were counted in the fraction larger than 100 μm. Atotal number of >250 benthic and >250 planktic specimenswere picked and identified. The number of foraminifera pergram dry bulk sediment was calculated for all samples. Theforaminifera accumulation rates (in number/cm2/1000 yr) werecalculated using the formulas of Ehrmann and Thiede (1985). Incore JM03-373PC2, ice-rafted debris (IRD) was counted in the>150-μm fraction. In cores JM02-460GC and -PC, IRD was

counted in the >1-mm fraction. The preservation of theforaminifera was good throughout the shelf cores, whereascore JM03-373PC2 suffered from intervals with poor preserva-tion of the planktic foraminifera. To quantify the degree ofdissolution (see e.g., references in Conan et al., 2002), thenumber of fragments was counted and the number of fragmentsper gram dry weight was calculated. The percent fragmentationwas calculated relative to the total numbers of benthic andplanktic foraminifera and the total number of fragments in asample.

Stable isotope analyses were carried out at Woods HoleOceanographic Institution, USA, at the WHOI-MAT 252Laboratory using standard procedures (Ostermann and Curry,2000). For all cores, the δ18O values were measured on theplanktic foraminifera Neogloboquadrina pachyderma (s). Forcore JM03-373PC2, the benthic isotopes were measured onCibicidoides wuellerstorfi and Cassidulina neoteretis. A fewsamples were measured on Melonis barleeanum. For coreJM02-460GC and -PC, the δ18O values were measured on M.barleeanum. A few samples were measured on Cibicideslobatulus and Elphidium excavatum with an overlap withmeasurements of M. barleeanum. The values obtained on M.barleeanum were corrected for disequilibrium with seawater by+0.4‰ and C. lobatulus and C. wuellerstorfi by +0.64‰(Duplessy et al., 1980). The few measurements on E. excavatumwere adjusted to the corrected values of M. barleeanum bycorrection of +0.7‰. C. neoteretis was corrected by +0.16‰(Poole, 1994). The isotopic values were corrected for icevolume changes presuming that ten meter sea level changeequals 0.11‰ change in δ18O (Fairbanks, 1989).

Radiocarbon dates were obtained using accelerator massspectrometry (AMS). The dated material consisted of mono-specific samples of N. pachyderma (s), various bivalve shellsand/or mixed benthic foraminifera assemblages. The dates werecorrected for marine reservoir effect by −440 yr (Mangerud andGulliksen, 1975).

Page 4: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

103T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Results and interpretation

Dating and correlation of cores

A total of 23 samples were radiocarbon dated (Fig. 3 andTable 1). The uppermost part of core JM02-460PC is slightlydisturbed and it was decided to combine the piston and thegravity core into a single record. Sediments younger than12,640 14C yr BP are from JM02-460GC, and the oldersediments are from JM02-460PC. The splicing could be madewith great precision based on the AMS 14C dates, lithofacies,magnetic susceptibility records, faunistic data, and isotope andIRD records (Figs. 3, 4, 5, and 6). The combined record isreferred to as JM02-460 (Fig. 3).

Age models for the cores were constructed based on theassumption of constant sedimentation rates between the dated

Figure 3. Lithological log based on visual inspection and X-ray images, magnetic susc460GC.

levels (Fig. 4). All cores contained a dark grey, laminatedinterval of fine clay with a few dropstones. Dates from two ofthe cores gave an age of 12,450 14C yr BP for the top of theinterval and 12,640 14C yr BP for the bottom (Fig. 3). Thislaminated interval was used as an additional tie point in thecorrelation (Figs. 3 and 4).

Modern foraminiferal faunas

The modern planktic foraminiferal faunas off westernSvalbard are dominated by N. pachyderma (s) and (d), Globi-gerina bulloides,Globigerina glutinata,Globigerinita uvula, andTurborotalita quinqueloba (e.g., Johannessen et al., 1994;Carstens et al., 1997; Simstich et al., 2003). The modern benthicfaunas in the Barents Sea and on the shelf and upper slope ofwestern Svalbard are characterized by Cassidulina reniforme, C.

eptibility, and AMS 14C dates for cores JM03-373PC2, JM02-460PC, and JM02-

Page 5: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Table 1Radiocarbon dates

Core Depth, cm Specimens Laboratorycode

Age (1-sigma; uncorrected),14C yr BP

Corrected age,14C yr BP a

JM03-373PC2 25 N. pachyderma (s)/T. quinqueloba AAR-8925 1593±42 115555 N. pachyderma (s)/T. quinqueloba AAR-8926 2503±41 2065146 N. pachyderma (s) AAR-8927 7311±46 6870155 Thyasira sp.? AAR-8768 7790±60 7350185 N. pachyderma (s) AAR-8928 8505±60 8065220 N. pachyderma (s)/T. quinqueloba AAR-8929 9275±65 8835350 Thyasira sp.? AAR-8769 13310±90 12430 ⁎

510 N. pachyderma (s) AAR-8770 14370±100 13930560 N. pachyderma (s) AAR-8771 16920±120 16480625 N. pachyderma (s) AAR-8772 18690±120 18250650 N. pachyderma (s) AAR-8773 19310±140 18870

JM02-460GC 20 Nucula sp./Thyasira sp. AAR-8915 880±55 44090 N. labradorica/M. barleeanum/I. norcrossi/G. auricula AAR-8916 2570±60 2130160 Nucula sp./N. labradorica Tua-3975 4645±50 4305210 N. labradorica AAR-8917 7780±65 7340280 Thyasira sp./Nucula sp./ Tua-3976 10655±140 10215

N. labradorica340–345 Nucula sp./Thyasira ssp./ AAR-8918 12890±110 12450

Cylichna occulta?375 Nuculana sp./Thyasira sp. Tua-3977 13180±140 12640

JM02-460PC 193 N. labradorica/M. barleeanum AAR-8914 9130±100 8690406 Nuculana sp. AAR-8762 13450±90 13010413 N. pachyderma (s) AAR-9448 15250±130 14810451.5 N. pachyderma (s) AAR-8763 15940±110 15500463 N. pachyderma (s) AAR-8764 16750±110 16310

a Corrected for reservoir effect -440 years (Mangerud and Gulliksen, 1975).⁎ Glacial benthic date corrected for reservoir effect 880 years.

104 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

neoteretis,M. barleeanum, C. lobatulus, and E. excavatum (Haldand Steinsund, 1996; Wollenburg and Mackensen, 1998;Lubinski et al., 2001; Sejrup et al., 2004). On the slope at waterdepth below ca. 1200 m, C. wuellerstorfi, Epistominella exigua,Eponides tumidulus, and Oridorsalis umbonatus become com-mon (Belanger and Streeter, 1980; Sejrup et al., 1981;Mackensenet al., 1985). Table 2 summarizes the general environmentalpreferences of the most common species in the cores.

Figure 4. Age-depth plot for cores JM03-373PC2, JM02-460PC, and JM02-460GC.

Reconstruction of surface and bottom water conditions 18,87014C yr BP to present

Based on the age model and the results of the distributionpattern of IRD, stable oxygen isotopes of benthic andplanktonic foraminifera species, and relative abundance offoraminifera (Figs. 5 and 6), the records have been subdividedinto the following time intervals: 18,870–15,500 14C yr BP(LGM); 15,500–15,000 14C yr BP (LGM-H1 transition);15,000–13,000 14C yr BP (Heinrich event H1 time interval);13,000–11,000 14C yr BP (Bølling–Allerød interstades);11,000–10,000 14C yr BP (Younger Dryas stade); 10,000–9000 14C yr BP (Younger Dryas–Holocene transition); and9000 14C yr BP to present (Holocene).

Time interval 18,870–15,500 14C yr BP (last glacial maximum,LGM)

Both of the long piston cores JM03-373PC2 and JM02-460PC terminated in coarse, unsorted deposits composed ofclayey silt and numerous sharp, gravel-sized clasts. Thedeposits gave nearly constant magnetic susceptibility valuestypical for sediments mixed by debris flow processes or iceactivity (Kuijpers et al., 2001) (Fig. 3). In the slope core JM03-373PC2, the coarse unsorted layer is probably a debris flowdeposit. The age exceeds 18,870 14C yr BP (Table 1; Fig. 3).Debris flows are typical for the LGM deposits in the fan systemsin the Barents Sea (e.g., Laberg and Vorren, 1995, 1996;

Page 6: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Figure 5. Records for core JM03-373PC2 plotted versus 14C yr BP. (a) Planktic and benthic δ18O values corrected for ice volume changes measured on N.pachyderma (s), C. wuellerstorfi, C. neoteretis, and M. barleeanum, (b–d) relative abundance of planktic foraminifera species in percentages of total planktic fauna,(b) T. quinqueloba and N. pachyderma (s) (inverse scale), (c) cumulative % of the subpolar species N. pachyderma (d), G. bulloides, and G. glutinata, (d) G. uvula,(e–j) relative abundance of benthic foraminifera species in percentages of total benthic fauna, (e) C. wuellerstorfi, (f) O. umbonatus, (g) C. neoteretis, (h) the‘Atlantic species’ group (see text for explanation), (i) C. reniforme, (j) cumulative % of E. exigua and E. tumidulus, (k) percent fragmentation calculated relative tothe total numbers of benthic and planktic foraminifera including fragments, (l) flux of planktic foraminifera, (m) flux of benthic foraminifera, and (n) number of IRD>150 μm per gram dry weight sediment. Abbreviations: YD, Younger Dryas; B-A, Bølling–Allerød; H1, Heinrich event H1; LGM, last glacial maximum.

105T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Andersen et al., 1996; Vorren and Laberg, 1997). The coarsedeposit in the shelf core JM02-460PC is older than 16,310 14Cyr BP and is most likely a till.

The δ18O values and the distribution of benthic and plankticforaminifera in the marine sediments are very similar in the twocores (Figs. 5 and 6). However, on the shelf only the last part ofthe LGM from ca. 16,800 to ca. 15,250 14C yr BP is representedby hemipelagic sediments (Fig. 6). The planktic foraminiferaare dominated by N. pachyderma (s). T. quinqueloba constitutesup to 40% of the faunas. On the slope, the preservation offoraminifera is very good except for short intervals with highfragmentation indicating severe dissolution (Fig. 5).

The benthic faunas at both sites are strongly dominated by C.neoteretis. Similar high abundances characterize the H1 interval(Figs. 5 and 6). We are not aware of comparable high abundanceof C. neoteretis in any modern faunas, and so the situationseems unique. However, in Greenland fjords, the species reachup to 60% in pre-modern sediments overlain by AtlanticIntermediate Water (Jennings and Weiner, 1996; Jennings et al.,2006). The close coupling between the surface and bottomproductivity and the similarity of the benthic shelf and slopefaunas (Figs. 5 and 6) suggest that the water column washomogenous down to a depth of about 1500 m and occupied bythe same water mass (Table 2). This conclusion is supported bythe fact that the planktic and benthic δ18O values were muchmore similar during the LGM than during the Holocene and

today (Figs. 5 and 6). It is important that this tendency is seen inboth cores, as the values in JM03-373 were measured ondifferent species and the correction for the isotopic disequili-brium of C. neoteretis is somewhat uncertain.

Time interval 15,500–15,000 14C yr BP (LGM-H1 transition)Towards the end of the LGM and close to 15,250 14C yr BP,

the relative abundance of N. pachyderma (s) increased to >90%on both the shelf and slope indicating a cooling. On the slope,the fragmentation of planktic foraminifera is high (Fig. 5).Concomitant with the cooling, the concentration of coarsegrains increased on the shelf to very high values (Fig. 6), and weinterpret the grains as IRD (Fig. 3). The short spikes of lowδ18O values seen in the planktic records may reflect localcontributions of meltwater (Figs. 5 and 6). E. excavatum, whichshows fluctuating abundances in this interval with peaks ofmore than 50%, indicates low salinity and cold polar conditions(Table 2).

Time interval 15,000–13,000 14C yr BP (Heinrich event H1interval)

The interval is time-equivalent with the Heinrich 1 event asdescribed in the Northeast Atlantic (e.g., Bond et al., 1993) andit has many of the same characteristic features. We will thereforeuse the name Heinrich event H1, even though our study area issituated far outside of the type area for the Heinrich events.

Page 7: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Figure 6. Records for core JM02-460 plotted versus 14C yr BP. (a) Planktic and benthic δ18O values corrected for ice volume changes measured on N. pachyderma (s),M. barleeanum, C. lobatulus, and E. excavatum, (b–d) relative abundance of planktic foraminifera species in percentages of total planktic fauna, (b) T. quinquelobaand N. pachyderma (s) (inverse scale), (c) cumulative % of subpolar species N. pachyderma (d), G. bulloides, and G. glutinata, (d) G. uvula, (e–j) relative abundanceof benthic foraminifera species in percentages of total benthic fauna (e) C. reniforme, (f) C. lobatulus, (g) M. barleeanum, (h) C. neoteretis, (i) the ‘Atlantic species’group (see text for explanation), (j) E. excavatum, (k) flux of planktic foraminifera, (l) flux of benthic foraminifera, (m) number of IRD >1 mm per gram dry weightsediment. Abbreviations: YD, Younger Dryas; B-A, Bølling–Allerød; H1, Heinrich event H1; LGM, last glacial maximum.

106 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

The low planktic δ18O values could indicate the presence ofmeltwater at the surface and stratification of the upper watercolumn (e.g., Bond et al., 1993) (Figs. 5 and 6). Theaccumulation rate is high for planktic and low for benthicforaminifera (Figs. 5 and 6), suggesting that the surface and thebottom productivity were decoupled and that nutrients weremainly recycled in the upper part of the water column.

The most conspicuous faunal change is the appearance of agroup of species, which Rasmussen et al. (1996a,b) termed‘Atlantic species’ (Figs. 5 and 6; Table 2). On the shelf, the‘Atlantic species’ group constitutes >25% of the assemblage.On the slope, the ‘Atlantic species’ group constitutes 4–7% ofthe benthic fauna (Figs. 5 and 6).

The appearance of the ‘Atlantic species’ group indicate thepresence of Atlantic water and a strong warming of the bottomwater compared to the LGM. The warming is most conspicuouson the shelf (Fig. 5). The faunal changes agree well with thechanges in the oxygen isotope values. Heinrich event H1 isaccompanied by a decrease in the benthic isotope values of1.8‰ on the shelf and of 0.5‰ on the slope (Figs. 5 and 6),corresponding to temperature rises of minima 7°C and 2°C,respectively (cf. Duplessy et al., 2001).

Time interval 13,00–11,000 14C yr BP (Bølling–Allerødinterstades)

The surface conditions were still dominated by cold polarwater as indicated by the strong dominance of N. pachyderma(s) (Figs. 5 and 6). The benthic and planktic δ18O values

increase, indicating reduction in surface meltwater (Figs. 5and 6). The sedimentation rate was extremely high in thelaminated sediments deposited from 12,640 to 12,450 14C yrBP (Figs. 3–6), whereas the accumulation rates of plankticforaminifera were low.

On the slope, the benthic foraminifera were dominated by C.reniforme with presence of C. neoteretis after 12,450 14C yr BP(Fig. 5). However, the conditions during the Bølling–Allerødwere different from today, as indicated by the absence of C.wuellerstorfi and O. umbonatus (Fig. 5; Table 2). The maincause could be high turbidity related to the high accumulationrates.

On the shelf, the benthic faunas were dominated by E.excavatum from 13,010 to 12,450 14C yr BP (Fig. 6). Weinterpret the increase of E. excavatum to indicate colder bottomwater, more variable salinities, high turbidity, and a reducedinfluence of Atlantic water as compared to Heinrich event H1(Table 2). From 12,450 14C BP, C. neoteretis increases togetherwith M. barleeanum and C. lobatulus, pointing to a relativelystrong influence of Atlantic water and increased food supply(Table 2). However, the bottom water temperatures wereprobably lower than during H1 as indicated by the increase inbenthic δ18O values (Fig. 6).

Time interval 11,000–10,000 14C yr BP (the Younger Dryasstade)

Ice rafting continues almost on the same level as during thepreceding interstades (Figs. 5 and 6). On the slope, the interval

Page 8: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Table 2Summary of the environmental preferences of the most common foraminifera species in the cores JM03-373 and JM02-460

Species Environmental conditions/water mass References

N. pachyderma (s) Polar conditions, Arctic and polar surface water Bé and Tolderlund (1971), Johannessen et al. (1994)N. pachyderma (d), G. bulloides,G. glutinata

Subpolar conditions, warm Atlantic surface water Bé and Tolderlund (1971), Johannessen et al. (1994),Carstens et al. (1997), Simstich et al. (2003)

T. quinqueloba Subpolar and Arctic conditions, productive Arctic Waterand Arctic and polar fronts

Bé and Tolderlund (1971), Johannessen et al. (1994),Carstens et al. (1997), Simstich et al. (2003)

G. uvula Polar front and upwelling areas, cold productive surfacewater, opportunist species

Boltovskoy et al. (1996), Kemle-von Mücke andHemleben (1999)

C. neoteretis Slope: chilled Atlantic Water (−1 to +2°C). Shelf:bottom temperatures >1°C. Occurs below Arctic surfacewater, in seasonal ice-free conditions and high seasonalproductivity

Mackensen and Hald (1988), Hald and Steinsund (1992,1996), Polyak and Mikhailov (1996), Wollenburg andMackensen (1998), Hald et al. (1999), Lubinski et al.(2001), Jennings et al. (2004), Wollenburg et al. (2004)

C. reniforme Arctic species in cooled Atlantic Water in faunas ofrelatively high diversity. Also in distal glaciomarineenvironments in faunas of low diversity.

Belanger and Streeter (1980), Sejrup et al. (1981);Mackensen et al. (1985), Hald and Korsun (1997),Korsun and Hald (2000)

E. excavatum Eurytopic species tolerant of variable environmentalconditions with low salinity and temperatures and highturbidity. Dominant in front of glaciers and abundant inbrine influenced areas

Hald and Steinsund (1992, 1996), Lubinski et al. (1996),Hald and Korsun (1997), Korsun and Hald (2000), Sejrupet al. (2004)

M. barleeanum High sedimentation rates and steady supply of food Hald and Steinsund (1992, 1996), Korsun and Hald (1998)C. lobatulus Bottom current influenced areas with high food supply Sejrup et al. (1981), Mackensen et al. (1985), Hald and

Steinsund (1992, 1996)C. wuellerstorfi Interglacial conditions, epifaunal suspension feeder and

indicator of bottom current activityBelanger and Streeter (1980), Sejrup et al. (1981),Mackensen et al. (1985), Haake and Pflaumann (1989),Linke and Lutze (1993), Struck (1997), Bauch et al.(2001)

O. umbonatus Interglacial conditions, deep food starved basins in theNordic seas

Belanger and Streeter (1980), Sejrup et al. (1981),Mackensen et al. (1985), Haake and Pflaumann (1989),Struck (1997), Bauch et al. (2001)

E. exigua, E. tumidulus Opportunistic phytodetritus feeders, unpredictable foodproductivity

Thomas et al. (1995), Wollenburg and Mackensen (1998),Gooday (2003)

‘Atlantic species group’, shelf: Epistominelladecorata/Alabaminella weddellensis,Discospirina italica, Spirophthalmidiumacutimargo

The group has a southern affinity and lives in warmerAtlantic Water. E. decorata/A. weddellensis occurs on thenorthern Norwegian shelf in Atlantic Water withtemperature >4.5°C. It is a phytodetritus species livingin areas of pulsed, unpredictable food supply.

E.g., Hald and Steinsund (1992, 1996), Thomas et al.(1995), Rasmussen et al. (1996a,b), Wollenburg andMackensen (1998), Lassen et al. (2002), Gooday (2003),Wollenburg et al. (2004), Rasmussen (2005)

Slope: Bulimina costata, Gyroidinoidesumbonata, Sigmoilopsis schlumbergeri

107T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

is marked by a small maximum in the relative abundance of N.pachyderma (s), indicating cold conditions and the presence ofpolar surface water. The higher abundance of C. neoteretisindicates influence of Atlantic water at depth (Fig. 5; Table 2).

On the shelf, the Younger Dryas is marked by a peak in therelative abundance of E. excavatum. The start of the period ischaracterized by a decrease in the δ18O values at both thesurface and bottom, which we interpret to be a result of reducedsalinity (Fig. 6). Together with the near absence of plankticforaminifera it points to the presence of polar surface water (Fig.6). Decreases in the relative abundance of M. barleeanum, C.lobatulus, and C. neoteretis suggest that the influence ofAtlantic water on the shelf was reduced (Table 2). The increasein the relative abundance of E. excavatum indicates lowersalinity and colder bottom water as compared to the precedingBølling–Allerød interstades.

Time interval 10,000–9000 14C yr BP (the Younger Dryas–Holocene transition)

On the slope, the boundary between the Younger Dryasand the Holocene is marked by the first conspicuous surface

warming as indicated by a maximum in the relative abun-dance of subpolar planktic species at 9500 14C yr BP (Fig. 5;Table 2). The warming is immediately followed by a cooling,but around 9000 14C yr BP the surface conditions becomewarm once again. This new warming is shown by the nearcomplete disappearance of N. pachyderma (s) and a signi-ficant increase in benthic and planktic foraminifera produc-tivity (Fig. 5). The first warming corresponds to appearanceand a rapid increase in the relative abundance of C.wuellerstorfi and O. umbonatus (Fig. 5). It signifies condi-tions similar to the modern environment on the deeper slopesof the Nordic seas (Table 2).

On the shelf, the transition is marked by an increase inthe relative abundance of C. lobatulus and M. barleeanumand decrease in C. reniforme (Fig. 6) indicating a strongerinflow of Atlantic water and ameliorated conditions withincreased supply of food (Table 2). Planktic foraminifera arenearly absent until ca. 9000 14C yr BP (Fig. 6). A high pointin the relative abundance of C. lobatulus probably indicatesthe Holocene maximum in the inflow of Atlantic water(Fig. 6).

Page 9: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

108 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Time interval 9000 14C yr BP to present (the Holocene)In both cores, the planktic fauna is dominated by T.

quinqueloba (60–80%) and N. pachyderma (s). The abundanceof subpolar species is relatively low and they decrease graduallythroughout the Holocene from a small maximum in the earlyHolocene (Figs. 5 and 6). The planktic and benthic δ18O valuesincrease gradually from the mid-Holocene to the present (Figs.5 and 6). On the slope, an early sea surface temperatureoptimum from 9000 to 8000 14C yr BP is indicated by lowrelative abundance of N. pachyderma (s) (Fig. 5). This earlyoptimum is less conspicuous on the shelf, which appears to beunder influence of the Arctic Front with T. quinquelobaconstituting about 60% of the fauna (Fig. 6; Table 2).

The increase in the relative abundance of N. pachyderma (s)from 6800 to 2500 14C yr BP in JM03-373 is probably due tosevere dissolution of T. quinqueloba, and the environmentalchanges were probably less abrupt and the temperaturessomewhat higher than indicated by the percentages of the twospecies (Fig. 5). In fact, the decreasing δ18O values stronglyindicate that the fall in the SST during the last 9000 14C yr hasbeen gradual (Figs. 5 and 6). This interpretation agrees wellwith the slow increase in the small planktic species G. uvuladuring the last 4000 14C yr observed in both records. It indicatesa cooling and an increase in food supply probably inconjunction with an advance of the Arctic Front and a reductionin the flow of Atlantic water (Table 2). An increase in theamount of IRD from ca. 4000 14C yr BP also points to anincrease in meltwater and iceberg supply.

The decrease in the abundances of the benthic foraminiferaC. wuellerstorfi and O. umbonatus, and the increase in E.exigua, and E. tumidulus (Fig. 5), is in good agreement with thechanges in the planktic foraminifera, which, as described above,indicate a gradual reduction in the inflow of Atlantic water andconsequently a change to more unpredictable seasonal produc-tivity (Table 2). On the shelf, C. reniforme, C. lobatulus, andM. barleeanum decrease, whereas E. excavatum increases(Fig. 6). This shift is indicative of a spread of polar water on theshelf.

Discussion

The interplay of Atlantic water, polar water, and meltwaterduring the last 20,000 14C yr BP

The presence of normal marine hemipelagic sediments inthe Storfjorden Trough show that the ice sheet had retreatedfrom the shelf edge before 16,310 14C yr BP. The benthicfaunas, the oxygen isotopes, and the abundant IRD indicate anaccelerated retreat between 15,500 and 14,810 14C yr BP.Studies from the western Svalbard shelf at 78°N haveindicated that the early deglaciation here began before14,800 14C yr BP with a retreat of the shelf ice from theouter shelf area (Mangerud et al., 1992; Elverhøi et al., 1995;Svendsen et al., 1996; Landvik et al., 1998). Marine recordsfrom the northern Barents Sea margin have dated the initialbreak-up of the northern parts of the Barents Sea ice sheet to15,400 14C yr BP (Knies et al., 1999). It appears that the more

southern Storfjorden Trough deglaciated earlier than at thesenorthern sites.

Previous studies from the Svalbard margin and the FramStrait have demonstrated that the LGM was a period with highproductivity and a high concentration of planktic foraminifera(Hebbeln et al., 1994; Dokken and Hald, 1996; Lloyd et al.,1996; Hald et al., 2001; Nørgaard-Pedersen et al., 2003).Reconstructions by transfer functions give SST for the Svalbardmargin of 3–4°C (Weinelt et al., 1996). This is slightly belowpresent temperatures in the area (4–7°C). Large parts of theNordic seas including the Fram Strait experienced seasonallyopen waters with inflow of cooled Atlantic surface water andabundant icebergs (Hebbeln et al., 1994; Dokken and Hald,1996; Lloyd et al., 1996; Weinelt et al., 1996; Knies et al., 1999;Vogt et al., 2001; Hald et al., 2001; Pflaumann et al., 2003;Sarnthein et al., 2003a; Nørgaard-Pedersen et al., 2003;Wollenburg et al., 2004).

Our data from the Storfjorden slope and trough support thisinterpretation. The high benthic and planktic productivityindicates the presence of extensive areas of open water duringsummer. The surface conditions were probably fairly similar tothe conditions in the central Nordic seas today, where the SSTvaries from 3 to 5°C. However, icebergs and drift ice must havebeen much more abundant than at present (Figs. 5 and 6). Ourdata also indicate that Atlantic water occupied the entire watercolumn on the Storfjorden slope down to a depth of at least1500 m. Studies from the Arctic Ocean and the Fram Straitshow that during the LGM Atlantic water, as a subsurfacecurrent, even penetrated into the Arctic Ocean (Vogt et al., 2001;Nørgaard-Pedersen et al., 2003; Wollenburg et al., 2004).

Heinrich event H1 is interpreted as being very cold in NorthAtlantic records with high concentrations of IRD and lowplanktic δ18O values (Bond et al., 1993). Our records andseveral earlier studies from Svalbard and the Fram Strait show asimilar pattern, indicating that the surface conditions also herewere very cold with abundant icebergs and meltwater at thesurface (Hald et al., 1996; Dokken and Hald, 1996; Andersen etal., 1996; Svendsen et al., 1996; Lloyd et al., 1996; Knies et al.,1999; Nørgaard-Pedersen et al., 2003; Birgel and Hass, 2004).

Mid-depth records of H1 from the North Atlantic Ocean arealso characterized by low benthic δ18O values (Van Kreveld etal., 2000; Rasmussen et al., 2002, 2003a,b; Rasmussen andThomsen, 2004). The low values are always accompanied byincrease in the abundance of the warm water ‘Atlantic species’group (Rasmussen et al., 1996b, 2002, 2003a,b; Lassen et al.,2002; Rasmussen and Thomsen, 2004). Between 15,400 and13,000 14C yr BP, they migrated into the SE Nordic seas (Jansenand Erlenkeuser, 1985; Rasmussen et al., 1996a,b, 2002;Rasmussen and Thomsen, 2004) and from 13,600 to 13,100 14Cyr BP they were present on the northern shelf of Iceland(Eiríksson et al., 2000). According to Rasmussen and Thomsen(2004), the spreading of the ‘Atlantic species’ group into theNordic seas during H1 reflects the northward advection ofAtlantic water as a subsurface water mass below a polarmeltwater layer at the surface. The presence of the Atlanticspecies group at 76°N indicate that bottom waters on thesouthwest Svalbard shelf warmed at least 7°C as compared to

Page 10: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

109T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

the temperatures of the preceding LGM time interval. DuringH1 the bottom water temperatures increased to the highestregistered in the whole record.

Haugan (1999) discussed the present heat loss of the WestSpitsbergen Current, suggesting that heavy sea-ice cover andmeltwater layer during winter and spring act as an insulationcover preventing heat loss from the warm Atlantic waterbeneath. This would allow transport of more heat northwardinto the Arctic Ocean. Similar insulation effects may havecontributed to the presence of warm water on the southwesternSvalbard shelf during H1.

From the start of the Bølling interstade at 13,000 14C yr BP,the Svalbard-Barents Sea ice sheet rapidly retreated (Vorren andKristoffersen, 1986; Mangerud et al., 1992; Elverhøi et al.,1995; Polyak et al., 1995; Polyak and Mikhailov, 1996; Landviket al., 1998; Lubinski et al., 2001). The Bølling and Allerød aredescribed as warm interstades in the northwest European landrecords (Koç and Jansen, 1992 and references therein) and inthe Greenland ice cores (Johnsen et al., 1992, 2001). West andnorth of Svalbard, the Bølling interstade is typically representedby high sedimentation rates from turbid meltwater plumes fromthe Svalbard-Barents Sea ice sheet (Elverhøi et al., 1995;Andersen et al., 1996; Svendsen et al., 1996; Birgel and Hass,2004). In Storfjorden, the sedimentation rate reaches amaximum from ∼12,640 to 12,430 14C yr BP resulting in thedeposition of laminated fine clays (Figs. 3–6).

According to diatom records, warm Atlantic water appearedat the surface in the southern Norwegian Sea between 13,400and 13,000 14C yr BP (Koç et al., 1993). From the same time inthe Norwegian Channel, Klitgaard Kristensen et al. (2001)recorded up to 50% of the subpolar planktic foraminifera N.pachyderma (d) replacing the polar N. pachyderma (s). In theStorfjorden area, the surface conditions remained cold (Figs. 5and 6), whereas on the shelf, the bottom water temperaturesincreased from 12,450 14C yr BP indicating increasinginfluence of Atlantic water. However, the temperaturesremained lower than during H1. We attribute the colderconditions of the Atlantic water to loss of heat to the atmosphereand outpouring of cold meltwater. The water column wasprobably not permanently stratified as during H1. Thisinterpretation agrees well with studies from the northernBarents Sea margin and the Fram Strait, where Atlantic waterwas continuously present after 13,000 14C yr BP (Polyak andSolheim, 1994; Lubinski et al., 1996; Koç et al., 2002;Wollenburg et al., 2004; Ślubowska et al., 2005).

During the Younger Dryas, the Barents Sea ice sheet retreatedfrom the Barents Sea shelf and into the fjords (Elverhøi et al.,1995; Bondevik et al., 1995; Polyak et al., 1995; Landvik et al.,1998; Svendsen et al., 1996, 2004). The Younger Dryas was apronounced cold event in records from the North Atlantic andNordic seas (e.g., Koç et al., 1993, 1996). At high latitudes, theperiod is not very conspicuous (e.g., Birgel and Hass, 2004),because the surface conditions were cold both before and after(Figs. 5 and 6). The initiation of the East Spitsbergen Currentafter the melting of the Barents Sea ice sheet may havecontributed to the polar conditions in the Storfjorden area.Atlantic water was present below the polar surface water, but it

was colder and probably of lower salinity than during theBølling–Allerød interstades. In the Storfjorden Trough, Atlanticwater occurs at the surface from about 9000 14C yr BP, markingthe transition to the Holocene. Overall, the shift to interglacialconditions occurs earlier in the benthic than in the plankticenvironment (Figs. 5 and 6). Over the slope, the transition wasstepwise with at least one significant setback (Fig. 5). Similarunstable transitions are seen in many records from the Nordicseas region and often referred to as the Pre-Boreal Oscillation(e.g., Koç and Jansen, 1992; Björck et al., 1996; Hald andHagen, 1998; Sarnthein et al., 2003b; Ślubowska et al., 2005). Inboth of the Svalbard records, the pronounced appearance ofAtlantic water at the surface is at 9000 BP. In core 23258-2 from75°N south of Svalbard, the surface warming also occurredaround 9000 14C yr BP (Sarnthein et al., 2003b).

The benthic foraminiferal records from core JM02-460 in theStorfjorden Trough with decreasing C. reniforme and increasingE. excavatum (Fig. 6) resemble most other Holocene recordsretrieved from the Svalbard margin and the Barents Sea in areaswith Atlantic water at the bottom (see e.g., Polyak andMikhailov, 1996; Lubinski et al., 1996; Hald et al., 1999,2004; Duplessy et al., 2001; Ślubowska et al., 2005). Theserecords indicate an early Holocene temperature optimumfollowed by a gradual cooling through the mid-late Holocenetime interval. Similar trends for the Holocene have beenrecorded in numerous cores from the Nordic seas (e.g., Koç andJansen, 1992, 2002; Hald and Aspeli, 1997; Sarnthein et al.,2003b; Andersen et al., 2004; Ślubowska et al., 2005).

The distribution pattern of planktic foraminifera in coreJM03-373 from the Storfjorden slope and in core 23258-2 from75°N (water depth 1768 m; Sarnthein et al., 2003b) (Fig. 1)shows some differences from the general pattern describedabove. According to the calculations of Sarnthein et al. (2003b),the SST during the Holocene temperature optimum were 4°Chigher than today. They found that the temperature optimumwasshort and occurred early in the Holocene interval between 9000and 8000 14C yr BP. Our records from Storfjorden support thisinterpretation (Fig. 5). However, in both cores JM03-373 and23258-2, the mid-Holocene interval from about 7000 14C yr BPis characterized by an abrupt increase in the relative abundanceof N. pachyderma (s) at the expense of T. quinqueloba. The shiftindicates a rapid and lasting temperature fall (Fig. 5). In coreJM03-373, this interval is marked by severe fragmentation anddissolution, and we interpret that the increase in N. pachyderma(s) partly is the result of differential dissolution, affecting thesmall, thin shelled T. quinqueloba more than the larger andthicker shelled N. pachyderma (s) (e.g., Berger, 1973;Boltovskoy and Wright, 1976) (Fig. 5). Thus, the abrupt andapparently severe cooling is probably exaggerated by dissolu-tion. Bauch et al. (2001) described a similar situation withdissolution and over-representation of N. pachyderma (s) fromthe same time interval in the western Fram Strait.

Brine formation in Storfjorden

In the St. Anna and Franz Victoria troughs on the northernmargin of the Barents Sea, the difference between surface and

Page 11: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

110 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

benthic δ18O values increased during the late Holocene,showing increased stratification between the surface and thebottom water (Duplessy et al., 2001). They attribute theincreased stratification to cooling of the bottom water causedby increased formation of brine during wintertime.

Brine formation occurs today in Storfjorden during thewinter (Fig. 2). For the Holocene interval, we have comparedthe δ18O values measured on M. barleeanum in core JM02-460 from the Storfjorden shelf with values measured on M.barleeanum in core NP94-51 from the northern shelf ofSvalbard at a comparable depth (400 m) (Ślubowska et al.,2005) (Fig. 7). The values from the Storfjorden Trough aregenerally higher than the values from the northern Svalbardshelf, indicating that the bottom water in Storfjorden wascolder than the bottom water on the northern shelf (Fig. 7).However, during the early Holocene temperature optimumfrom 9000 to 8000 14C yr BP there was almost no differencebetween the two sites. The differentiation began after 800014C yr BP and has been increasing during the last 5000 14C yr

Figure 7. δ18O records corrected for ice volume changes plotted versus 14C yrBP for the last 10,500 yr from core JM02-460 (this study) and core NP94-51(north Svalbard shelf, Ślubowska et al., 2005). All values are measured on M.barleeanum.

BP. Today, the difference in δ18O values between the twosites is 0.34–0.4‰, corresponding to a temperature differenceof 1.3-1.6°C. Given the southern position of the StorfjordenTrough it is likely that the lower temperatures in Storfjordenand the increasing temperature difference between Storfjor-den and the northern shelf is due to increasing brine formationin the inner part of Storfjorden and a higher overflow ofbrines.

Summary and conclusions

Over the last 20,000 14C yr BP, Atlantic water has beencontinuously present on the southwestern Svalbard shelf.However, from 15,000 to 10,000 14C yr BP, comprisingHeinrich event H1, the Bølling–Allerød interstades, and theYounger Dryas stade, it flowed as a subsurface water massbelow a layer of polar surface water. In the benthic environment,the shift to interglacial conditions occurred at 10,000 14C yr BP.Due to persistent presence of a thin upper layer of polar water,surface conditions remained cold until ca. 1000 yr later, whenthe warm Atlantic water finally broke to the surface. The studyshows that neither extensive sea ice cover nor large inputs ofmeltwater stopped the inflow of Atlantic water. It simplycontinued as a submerged water mass below the cold cover ofpolar surface water.

During the LGM, cooled Atlantic water was present on theStorfjorden margin from the surface to a water depth of at least1485 m (Fig. 8a). The water column was homogenous and themaximum SST was only slightly lower than today. However,the winter ice coverage was undoubtedly larger and the drift icedenser.

Heinrich event H1 was characterized by a stratified watercolumn with a surface layer of cold polar meltwater withreduced salinity overlying a mass of warm Atlantic water (Fig.8b). The sea ice cover was probably extensive. The warmAtlantic water entered the deeper parts of the shelf, where thebottom water temperatures could be at least 7°C higher thanduring the LGM.

During the Bølling and Allerød interstades, the surfaceconditions were relatively cold due to plumes of turbidmeltwater, which mixed with and cooled the Atlantic water(Fig. 8c). The deeper Atlantic water on the outer part of the shelfand on the upper slope was warmer than today, although colderthan during H1.

The Younger Dryas period was cold due to the advectionof polar water from the Arctic Ocean and reduced inflow ofAtlantic water (Fig. 8d). Atlantic water was present on thedeeper shelf and upper slope, but it was colder and fresherthan during the Bølling–Allerød interstades. Polar surfaceconditions continued until 9000 14C yr BP, well into the earlyHolocene.

The surface conditions over the shelf and slope became verywarm at 9000 14C yr BP, as the polar water retreated and theAtlantic water appeared at the surface (Fig. 8e). The Holocenetemperature maximum lasted from ca. 9000 to 8000 14C yr BP,when a gradual cooling began. The cooling has continued untiltoday, slightly accelerating during the last 4000 yr (Fig. 8f).

Page 12: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

Figure 8. (Summer) paleotemperature profiles for the southeastern Svalbard margin reconstructed for (a) LGM, (b) Heinrich event H1, (c) Bølling–Allerød, (d)Younger Dryas, and (e) the Holocene temperature optimum. The reconstructions are based on the distribution of planktic and benthic foraminifera, δ18O values, andIRD and compared with the present temperature profile for the area (f).

111T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Some of the cooling in the benthic environment may be localdue to increased overflow of brines from the inner part ofStorfjorden.

Acknowledgments

The captain and crew of RV Jan Mayen are thanked for theirefficient work in the coring procedures and Steinar Iversen forhis invaluable assistance with the coring and seismic acquisi-tion. Edel Ellingsen performed the GEOTEK logging and tookthe X-ray images of the cores. Ragnheid Skogseth measured theCTD data in 2002. The stable isotope measurements weresupervised by Dorinda Ostermann, Woods Hole OceanographicInstitution. The manuscript benefited from highly constructivesuggestions by reviewers A. Jennings and J. Lloyd. The projectwas financed by UNIS and forms part of the ESF-EuroClimateProgram RESOLuTION (grant no. 04-ECLIM-FP33).

References

Aagaard, K., Carmack, E.C., 1989. The role of sea ice and other freshwater inthe Arctic circulation. Journal of Geophysical Research 94, 14,485–14,498.

Aagaard, K., Swift, J.H., Carmack, E.C., 1985. Thermohaline circulation in theArctic Mediterranean Seas. Journal of Geophysical Research 90 (C5),4833–4846.

Andersen, E.S., Dokken, T.M., Elverhøi, A., Solheim, A., Fossen, I., 1996. LateQuaternary sedimentation and gladial history of the western Svalbardcontinental margin. Marine Geology 133, 123–156.

Andersen, C., Koç, N., Jennings, A., Andrews, J.T., 2004. Nonuniform responseof the major surface currents in the Nordic seas to insolation forcing:implications for the Holocene climate variability. Paleoceanography 19,1–16, PA2003, doi:10.1029/2002PA000873.

Bauch, H.A., Erlenkeuser, H., Spielhagen, R.F., Struck, U., Matthiessen, J.,Thiede, J., Heinemeier, J., 2001. A multiproxy reconstruction of theevolution of deep and surface waters in the subarctic Nordic seas over thelast 30,000 yr. Quaternary Science Reviews 20, 659–678.

Bé, A.W.H., Tolderlund, D.S., 1971. Distribution and ecology of livingplanktonic foraminifera in surface waters of the Atlantic and Indian Oceans.

Page 13: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

112 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

In: Funnell, B.M., Riedel, W.R. (Eds.), The Micropaleontology of theOceans. Cambridge University Press, pp. 105–149.

Belanger, B.E., Streeter, S.S., 1980. Distribution and ecology of benthicforaminifera in the Norwegian-Greenland Sea. Marine Micropaleontology 5,401–428.

Berger, W.H., 1973. Deep-sea carbonates: Pleistocene dissolution cycles.Journal of Foraminiferal Research 3, 187–195.

Birgel, D., Hass, H.C., 2004. Oceanic and atmospheric variations during the lastdeglaciation in the Fram Strait (Arctic Ocean): a coupled high-resolutionorganic-geochemical and sedimentological study. Quaternary ScienceReviews 23, 29–47.

Björck, S., Kromer, B., Johnsen, S., Bennike, O., Hammarlund, D., Lemdahl,G., Possnert, G., Rasmussen, T.L., Wohlfarth, B., Hammer, C.U., Spurk, M.,1996. Synchronized terrestrial-atmospheric deglacial records around theNorth Atlantic. Science 274, 1155–1160.

Boltovskoy, E., Wright, R., 1976. Recent foraminifera. Dr. Junk. The Hague515.

Boltovskoy, E., Boltovskoy, D., Correa, N., Brandini, F., 1996. Plankticforaminifera from the southwestern Atlantic (30°–60°S): species-specificpatterns in the upper 50 m. Marine Micropaleontology 28, 53–72.

Bond, G., Broecker, W.S., Johnsen, S.J., McManus, J., Labeyrie, L., Jouzel, J.,Bonani, G., 1993. Correlations between climate records from North Atlanticsediments and Greenland ice. Nature 365, 143–147.

Bondevik, S., Mangerud, J., Ronnert, L., Salvigsen, O., 1995. Post-glacial sea-level history of Edgeøya and Barentsøya, eastern Svalbard. Polar Research14, 153–180.

Boyd, T.J., D'Asaro, E.A., 1994. Cooling of the West Spitsbergen Current:wintertime observations west of Svalbard. Journal of Geophysical Research99 (C11), 22,597–22,618.

Carstens, J., Hebbeln, D., Wefer, G., 1997. Distribution of planktic foraminiferaat the ice margin in the Arctic (Fram Strait). Marine Micropaleontology 29,257–269.

Conan, S.M.-H., Ivanova, E.M., Brummer, G.-J.A., 2002. Quantifyingcarbonate dissolution and calibration of foraminiferal dissolution indicesin the Somali basin. Marine Geology 182, 325–349.

Dokken, T.M., Hald, M., 1996. Rapid climatic shifts during isotope stages 2–4in the polar North Atlantic. Geology 24, 599–602.

Dowdeswell, J.A., Kenyon, N.H., Elverhøi, A., Laberg, J.S., Hollender, F.-J.,Mienert, J., Siegert, M.J., 1996. Large-scale sedimentation on the glacier-influenced Polar North Atlantic margins: long-range side-scan sonarevidence. Geophysical Research Letters 23, 3535–3538.

Duplessy, J.-C., Moyes, J., Pujol, C., 1980. Deep water formation in the NorthAtlantic Ocean during the last ice age. Nature 286, 479–482.

Duplessy, J.-C., Ivanova, E., Murdmaa, I., Paterne, M., Labeyrie, L., 2001.Holocene paleoceanography of the Northern Barents Sea and variations ofthe northward heat transport by the Atlantic Ocean. Boreas 30, 2–16.

Ehrmann, W.U., Thiede, J., 1985. History of mesozoic and cenozoic sedi-ment fluxes to the North Atlantic Ocean. Contributions to Sedimentology15, 1–109.

Eiríksson, J., Knudsen, K.L., Haflidason, H., Henriksen, P., 2000. Late-glacialand Holocene paleoceanography of the North Icelandic shelf. Journal ofQuaternary Science 15, 23–42.

Elverhøi, A., Andersen, E.S., Dokken, T., Hebbeln, D., Spielhagen, R.,Svendsen, J.I., Sørflaten, M., Romes, A., Hald, M., Forsberg, C.F., 1995.The growth and decay of the Late Weichselian ice sheet in western Svalbardand adjacent areas based on provenance studies of marine sediments.Quaternary Research 44, 303–316.

Fairbanks, R.G., 1989. A 17,000-years glacio-eustatic sea level record:influence of glacial melting rates on the Younger Dryas event and deep-ocean circulation. Nature 342, 637–742.

Ganopolski, A., Rahmstorf, S., 2001. Rapid changes of glacial climate simulatedin a coupled climate model. Nature 409, 153–158.

Gooday, A.J., 2003. Benthic Foraminifera (Protista) as tools in deep-waterpalaeoceanography: environmental influences on faunal characteristics. In:Southward, A.J., Tyler, P.A., Young, C.M., Fuiman, L.A. (Eds.), Advancesin Marine Biology, vol. 46. Academic Press, Amsterdam, pp. 1–90.

Haake, F.-W., Pflaumann, U., 1989. Late Pleistocene stratigraphy on the VøringPlateau, Norwegian Sea. Boreas 18, 343–356.

Haarpainter, J., Gascard, J.C., Haugan, P.M., 2001a. Ice production and brineformation in Storfjorden, Svalbard. Journal of Geophysical Research 106(C7), 14,001–14,013.

Haarpainter, J., O'Dwyer, J., Gascard, J.C., Haugan, P.M., Schauer, U.,Østerhus, S., 2001b. Seasonal transformation of water masses, circulationand brine formation observed in Storfjorden, Svalbard. Annals ofGlaciology 33, 437–443.

Hald, M., Aspeli, R., 1997. Rapid climatic shifts of the northern Norwegian Seaduring the last deglaciation and the Holocene. Boreas 26, 15–28.

Hald, M., Hagen, S., 1998. Early Preboreal cooling in the Nordic seas regiontriggered by meltwater. Geology 26, 615–618.

Hald, M., Korsun, S., 1997. Distribution of modern benthic foraminifera fromfjords of Svalbard, European Arctic. Journal of Foraminiferal Research 27,101–122.

Hald, M., Steinsund, P.I., 1992. Distribution of surface sediment benthicforaminifera in the southwestern Barents Sea. Journal of ForaminiferalResearch 22, 347–362.

Hald, M., Steinsund, P.I., 1996. Benthic foraminifera and carbonate dissolutionin the surface sediments of the Barents and Kara Seas. Berichte zurPolarforschung 212, 285–307.

Hald, M., Dokken, T., Hagen, S., 1996. Paleoceanography of the EuropeanArctic margin during the last glaciation. Geological Society of LondonSpecial Publication 111, 275–287.

Hald, M., Kolstad, V., Polyak, L., Forman, S.L., Herlihy, F.A., Ivanov, G.,Nescheretov, A., 1999. Late-glacial and Holocene paleoceanographyand sedimentary environments in the St. Anna Trough, Eurasian ArcticOcean margin. Palaeogeography, Palaeoclimatology, Palaeoecology 146,229–249.

Hald, M., Dokken, T., Mikalsen, G., 2001. Abrupt climatic change during thelast interglacial-glacial cycle in the polar North Atlantic. Marine Geology176, 121–137.

Hald, M., Ebbesen, H., Forwick, M., Godtliebsen, F., Khomenko, L., Korsun, S.,Olsen, L.R., Vorren, T.O., 2004. Holocene paleoceanography and glacialhistory of the West Spitsbergen area, Euro-Arctic margin. QuaternaryScience Reviews 23, 2075–2088.

Haugan, P.M., 1999. Structure and heat content of the West Spitsbergen Current.Polar Research 18, 183–188.

Hebbeln, D., Dokken, T., Andersen, E.S., Hald, M., Elverhøi, A., 1994.Moisture supply for northern ice-sheet growth during the last glacialmaximum. Nature 370, 357–360.

Heinrich, H., 1988. Origin and consequences of cyclic ice rafting in thenortheast Atlantic Ocean during the past 130,000 years. QuaternaryResearch 29, 142–152.

Hjelstuen, B.O., Elverhøi, A., Faleide, J.I., 1996. Cenozoic erosion andsediment yield in the drainage area of the Storfjorden Fan. Global andPlanetary Change 12, 95–117.

Jansen, E., Erlenkeuser, H., 1985. Ocean circulation in the Norwegian Seaduring the last deglaciation: isotopic evidence. Palaeogeography, Palaeocli-matology, Palaeoecology 49, 189–206.

Jennings, A.E., Weiner, N.J., 1996. Environmental change in eastern Greenlandduring the last 1300 years: evidence from foraminifera and lithofacies inNansen Fjord, 68 degrees N. The Holocene 6, 179–191.

Jennings, A.E., Weiner, N.J., Helgadottir, G., Andrews, J.T., 2004. ModernForaminiferal faunas of the southwestern to northern Iceland shelf:oceanographic and environmental controls. Journal of ForaminiferalResearch 34, 180–207.

Jennings, A.E., Hald, M., Smith, M., Andrews, J.T., 2006. Freshwater forcingfrom the Greenland Ice Sheet during the Younger Dryas: evidence fromsoutheastern Greenland shelf cores. Quaternary Science Reviews 25,282–298.

Johannessen, T., Jansen, E., Flatøy, A., Ravelo, A.C., 1994. The relationshipbetween surface water masses, oceanographic fronts and paleoclimaticproxies in surface sediments of the Greenland, Iceland, Norwegian Seas.NATO ASI Series 61–86.

Johnsen, S.J., Clausen, H.B., Dansgaard, W., Fuhrer, K., Gundestrup, N.,Hammer, C.U., Iversen, P., Jouzel, J., Stauffer, B., Steffensen, J.P., 1992.Irregular glacial interstadials recorded in a new Greenland ice core. Nature359, 311–313.

Page 14: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

113T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Johnsen, S.J., Dahl-Jensen, D., Gundestrup, N., Steffensen, J.P., Clausen, H.B.,Miller, H., Masson-Delmotte, V., Sveinbjörnsdóttir, A.E., White, J., 2001.Oxygen isotope and paleotemperature records from six Greenland ice-corestations: camp century, Dye-3, GRIP, GISP2, Renland and NorthGRIP.Journal of Quaternary Science 16, 299–307.

Kemle-von Mücke, S., Hemleben, C., 1999. Foraminifera. In: Boltovskoy, D.(Ed.), South Atlantic Zooplankton. Backhuys Publishers, Leiden, TheNetherlands, pp. 43–73.

Klitgaard Kristensen, D., Sejrup, H.P., Haflidason, H., 2001. The last 18 kyrfluctuations in Norwegian Sea surface conditions and implications for themagnitude of climatic change: evidence from the North Sea. Paleoceano-graphy 16, 455–467.

Knies, J., Vogt, C., Stein, R., 1999. Late Quaternary growth and decay of theSvalbard/Barents Sea ice sheet and paleoceanographic evolution in theadjacent Arctic Ocean. Geo-Marine Letters 18, 195–202.

Koç, N., Jansen, E., 1992. A high-resolution diatom record of the lastdeglaciation from the SE Norwegian Sea: documentation of rapid climatechanges. Paleoceanography 7, 499–520.

Koç, N., Jansen, E., 2002. Holocene climate evolution of the North AtlanticOcean and the Nordic seas-a synthesis of new results. In: Wefer, G., Berger,W., Behre, K.-E., Jansen, E. (Eds.), Climate and History in the NorthAtlantic Realm. Springer-Verlag, Berlin, pp. 165–173. Heidelberg.

Koç, N., Jansen, E., Haflidason, H., 1993. Paleoceanographic reconstructions ofsurface ocean conditions in the Greenland, Iceland and Norwegian Seasthrough the last 14 ka based on diatoms. Quaternary Science Reviews 12,115–140.

Koç, N., Jansen, E., Hald, M., Labeyrie, L., 1996. Late-glacial-Holocene seasurface temperatures and gradients between the North Atlantic and theNorwegian Sea: implications for the Nordic heat pump. Geological Societyof London Special Publication 111, 177–185.

Koç, N., Klitgaard Kristensen, D., Hasle, K., Forsberg, C.F., Solheim, A., 2002.Late glacial paleoceanography of Hinlopen Strait, northern Svalbard. PolarResearch 21, 307–314.

Korsun, S., Hald, M., 1998. Modern benthic foraminifera off Novaya Zemlyatidewater glaciers, Russian Arctic. Arctic and Alpine Research 30, 61–77.

Korsun, S., Hald, M., 2000. Seasonal dynamics of benthic foraminifera in aglacially fed fjord of Svalbard, European Arctic. Journal of ForaminiferalResearch 30, 251–271.

Kuijpers, A., Nielsen, T., Akhmetzhanov, A., de Haas, H., Kenyon, N.H., vanWeering, T.C.E., 2001. Late Quaternary slope instability on the Faeroemargin: mass flow features and timing of events. Geo-Marine Letters 20,149–159.

Laberg, J.S., Vorren, T.O., 1995. Late Weichselian debris flow deposits on theBear Island Trough Mouth Fan. Marine Geology 127, 45–72.

Laberg, J.S., Vorren, T.O., 1996. The glacier fed fan at the mouth of StorfjordenTrough, western Barents Sea: a comparative study. Geologisches Rundschau85, 338–349.

Landvik, J.Y., Bondevik, S., Elverhøi, A., Fjeldskaar, W., Mangerud, J.,Salvigsen, O., Siegert, M.J., Svendsen, J.I., Vorren, T.O., 1998. The lastglacial maximum of Svalbard and the Barents Sea area: ice sheet extent andconfiguration. Quaternary Science Reviews 17, 43–75.

Lassen, S.L., Kuijpers, A., Kunzendorf, H., Lindgren, H., Heinemeier, J.,Jansen, E., Knudsen, K.L., 2002. Intermediate water signal leads surfacewater response during northeast Atlantic deglaciation. Global and PlanetaryChange 32, 111–125.

Linke, P., Lutze, G.F., 1993. Microhabitat preferences of benthic foraminifera-astatic concept or a dynamic adaptation to optimize food acquisition? MarineMicropaleontology 20, 215–234.

Lloyd, J., Kroon, D., Laban, C., Boulton, G., 1996. Deglaciation history andpaleoceanography of the western Spitsbergen margin since the last glacialmaximum. Geological Society of London Special Publication 111, 289–301.

Loeng, H., 1991. Features of the physical oceanographic conditions of theBarents Sea. Polar Research 10, 5–18.

Lubinski, D.J., Korsun, S., Polyak, L., Forman, S.L., Lehman, S.J., Herlihy, F.,Miller, G.H., 1996. The last deglaciation of the Franz Victoria Trough,northern Barents Sea. Boreas 25, 89–100.

Lubinski, D.J., Polyak, L., Forman, S.L., 2001. Freshwater and Atlantic waterinflows to the deep northern Barents and Kara seas since ca. 13 14C ka:

foraminifera and stable isotopes. Quaternary Science Reviews 20,1851–1879.

Mackensen, A., Hald, M., 1988. Cassidulina teretis Tappan and C. laevigatad'Orbigny: their modern and Late Quaternary distribution in northern seas.Journal of Foraminiferal Research 18, 16–24.

Mackensen, A., Sejrup, H.P., Jansen, E., 1985. The distribution of living benthicforaminifera on the continental slope and rise off southwest Norway. MarineMicropaleontology 9, 275–306.

Mangerud, J., Gulliksen, S., 1975. Apparent radiocarbon ages of recent marineshells from Norway, Spitbergen, and Arctic Canada. Quaternary Research 5,263–273.

Mangerud, J., Bolstad, M., Elgersma, A., Helliksen, D., Landvik, J.Y., Lønne,L., Lycke, A.K., Salvigsen, O., Sandahl, T., Svendsen, J.I., 1992. The lastglacial maximum on Spitsbergen, Svalbard. Quaternary Research 38, 1–31.

Nørgaard-Pedersen, N., Spielhagen, R.F., Erlenkeuser, H., Grootes, P.M.,Heinemeier, J., Knies, J., 2003. Arctic Ocean during the last glacialmaximum: Atlantic and polar domains of surface water mass distributionand ice cover. Paleoceanography 18, 1063, doi:10.1029/2002PA000781.

Ostermann, D.R., Curry, W.D., 2000. Calibration of stable isotope data: anenriched δ18O standard used for source gas mixing detection and correction.Paleoceanography 15, 353–360.

Pfirman, S.L., Bauch, D., Gammelsrød, T., 1994. The Northern Barents Sea: watermass distribution and modification. In: Johannessen, O.M., Muench, R.D.,Overland, J.E. (Eds.), The Polar Oceans and Their Role in Shaping the GlobalEnvironment. AGU Geophysical Monograph Series, vol. 85, pp. 77–94.

Pflaumann, U., Sarnthein, M., Chapman, M., d'Abreu, L., Funnell, B., Huels,M., Kiefer, T., Maslin, M., Schulz, H., Swallow, J., van Kreveld, S.,Vautravers, M., Vogelsang, E., Weinelt, M., 2003. Glacial North Atlantic:sea-surface conditions reconstructed by GLAMAP 2000. Paleoceanography18 (1065), 1–28, doi:10.1029/2002PA000774.

Polyak, L., Mikhailov, V., 1996. Post-glacial environments of the southeasternBarents Sea: foraminiferal evidence. Geological Society of London SpecialPublication 111, 323–337.

Polyak, L., Solheim, A., 1994. Late- and postglacial environments in thenorthern Barents Sea west of Franz Josef Land. Polar Research 13, 197–207.

Polyak, L., Lehman, S.J., Gataullin, V., Jull, A.J.T., 1995. Two-step deglaciationof the southeastern Barents Sea. Geology 23, 567–571.

Poole, D.A.R., 1994. Neogene and Quaternary paleoenvironments on thenorthern Norwegian shelf. Unpublished Phd Thesis, University of Tromsø,Norway, 1–82.

Quadfasel, D., Rudels, B., Kurtz, K., 1988. Outflow of dense water from aSvalbard fjord into the Fram Strait. Deep-Sea Research A35, 1143–1150.

Rahmstorf, S., 2002. Ocean circulation and climate during the past120,000 years. Nature 419, 207–214.

Rasmussen, T.L., 2005. Systematic paleontology and ecology of benthicforaminifera from the Plio-Pleistocene Kallithea Bay section, RhodosGreece. In: Thomsen, E. (Ed.), Lagoon to Deep Water Foraminifera andOstracods from the Plio-Pleistocene Kallithea Bay Section, Rhodes, Greece,39. Cushman Foundation Special Publication, pp. 53–157.

Rasmussen, T.L., Thomsen, E., 2004. The role of the North Atlantic Drift in themillennial timescale glacial climate fluctuations. Palaeogeography, Palaeo-climatology, Palaeoecology 210, 101–116.

Rasmussen, T.L., Thomsen, E., van Weering, T.C.E., Labeyrie, L., 1996a. Rapidchanges in surface and deep water conditions at the Faeroe Margin duringthe last 58,000 years. Paleoceanography 11, 757–771.

Rasmussen, T.L., Thomsen, E., Labeyrie, L., van Weering, T.C.E., 1996b.Circulation changes in the Faeroe-Shetland Channel correlating with coldevents during the last glacial period (58–10 ka). Geology 24, 937–940.

Rasmussen, T.L., Bäckström, D., Heinemeier, J., Klitgaard Kristensen, D.,Knutz, P.C., Kuijpers, A., Lassen, S., Thomsen, E., Troelstra, S.R., vanWeering, T.C.E., 2002. The Faeroe-Shetland Gateway: late Quaternarywater mass exchange between the Nordic seas and the northeastern Atlantic.Marine Geology 188, 165–192.

Rasmussen, T.L., Thomsen, E., Troelstra, S.R., Kuijpers, A., Prins, M., 2003a.Millennial-scale glacial variability versus Holocene stability: changes inplanktic and benthic foraminifera faunas and ocean circulation in the NorthAtlantic during the last 60,000 years. Marine Micropaleontology 47,143–176.

Page 15: Paleoceanographic evolution of the SW Svalbard margin (76°N) since 20,000 14C yr BP

114 T.L. Rasmussen et al. / Quaternary Research 67 (2007) 100–114

Rasmussen, T.L., Oppo, D., Thomsen, E., Lehman, S.J., 2003b. Deep-searecords from the SE Labrador Sea: timing of ocean circulation changes andice rafting events during the last 150,000 years. Paleoceanography 18, 1–15,doi:10.1029/2002PA000764.

Rudels, B., Friedrich, H.J., Quadfasel, D., 1999. The Arctic circumpolarboundary current. Deep-Sea Research II 46, 1023–1062.

Saloranta, T.M., Haugan, P.M., 2004. Northward cooling and freshening of thewarm core of the West Spitsbergen Current. Polar Research 23, 79–88.

Saloranta, T.M., Svendsen, H., 2001. Across the Arctic Front west ofSpitsbergen: high-resolution CTD sections from 1998–2000. Polar Research20, 177–184.

Salvigsen, O., Forman, S.L., Miller, G.H., 1992. Thermophilous molluscs onSvalbard during the Holocene and their paleoclimatic implications. PolarResearch 11, 1–10.

Sarnthein, M., Pflaumann, U., Weinelt, M., 2003a. Past extent of sea ice in thenorthern North Atlantic inferred from foraminiferal paleotemperatureestimates. Paleoceanography 18, 1047, doi:10.1029/2002PA000771.

Sarnthein, M., van Kreveld, S., Erlenkeuser, H., Grootes, P.M., Kucera, M.,Pflaumann, U., Schulz, M., 2003b. Centennial-to-millennial-scale periodi-cities of Holocene climate and sediment injections off the western Barentsshelf, 75°N. Boreas 32, 447–461.

Schauer, U., 1995. The release of brine-enriched shelf water from Storfjord intothe Norwegian Sea. Journal of Geophysical Research 100 (C8),16,015–16,028.

Schauer, U., Fahrbach, E., 1999. A dense bottom water plume in the westernBarents Sea: downstream modification and interannual variability. Deep-SeaResearch I 46, 2095–2108.

Sejrup, H.-P., Fjæran, T., Hald, M., Beck, L., Hagen, J., Miljeteig, I., Morvik, I.,Norvik, O., 1981. Benthonic foraminifera in surface samples from theNorwegian continental margin between 62°N and 65°N. Journal ofForaminiferal Research 11, 277–295.

Sejrup, H.P., Birks, H.J.B., Klitgaard Kristensen, D., Madsen, H., 2004.Benthonic foraminiferal distributions and quantitative transfer functions forthe northwest European continental margin. Marine Micropaleontology 53,197–226.

Simstich, J., Sarnthein, M., Erlenkeuser, H., 2003. Paired δ18O signals of N.pachyderma (s) and T. quinqueloba show thermal stratification structure inthe Nordic seas. Marine Micropaleontology 48, 107–125.

Skogseth, R., Haugan, P.M., Haarpaintner, J., 2004. Ice and brine production inStorfjorden from four winters of satellite and in situ observations andmodelling. Journal of Geophysical Research 109 (C10008), 1–15doi:10.1029/2004JC002384.

Skogseth, R., Haugan, P.M., Jakobsson, M., 2005. Watermass transformations inStorfjorden. Continental Shelf Research 25, 667–695.

Ślubowska, M.A., Koç, N., Rasmussen, T.L., Klitgaard Kristensen, D., 2005.Changes in the flow of Atlantic water into the Arctic Ocean since the lastdeglaciation: evidence from the northern Svalbard continental margin, 80°N.Paleoceanography 20, 1–15, doi:10.1029/2005PA001141.

Struck, U., 1997. Paleoecology of benthic foraminifera in the Norwegian-Greenland Sea during the past 500 ka. In: Hass, H.C., Kaminski, M.A.(Eds.), Contributions to the Micropaleontology and Paleoceanography of theNorthern North Atlantic, vol. 5. Grzybowski Foundation Special Publica-tion, pp. 51–82.

Svendsen, J.I., Elverhøi, A., Mangerud, J., 1996. The retreat of the Barents SeaIce Sheet on the western Svalbard margin. Boreas 25, 244–256.

Svendsen, J.I., Gataullin, V., Mangerud, J., Polyak, L., 2004. The glacial historyof the Barents and Kara Sea region. In: Ehlers, J., Gibbard, P.L. (Eds.),Quaternary Glaciations—Extent and Chronology, vol. 1. Elsevier, Amster-dam, pp. 369–378. Europe.

Thomas, E., Booth, L., Maslin, M., Shackleton, N.J., 1995. NortheasternAtlantic benthic foraminifera during the last 45,000 years: changes inproductivity seen from the bottom up. Paleoceanography 10, 545–562.

Van Kreveld, S., Sarnthein, M., Erlenkeuser, H., Grootes, P., Jung, S., Nadeau,M.J., Pflaumann, U., Voelker, A., 2000. Potential links between surging icesheets, circulation changes, and the Dansgaard–Oeschger cycles in theIrminger Sea, 60–18 kyr. Paleoceanography 15, 425–442.

Vogt, C., Knies, J., Spielhagen, R.F., Stein, R., 2001. Detailed mineralogicalevidence for two nearly identical glacial/interglacial cycles and Atlanticwater advection to the Arctic Ocean during the last 90,000 years. Global andPlanetary Change 31, 23–24.

Vorren, T.O., Kristoffersen, Y., 1986. Late Quaternary glaciation in thesouthwestern Barents Sea. Boreas 15, 51–60.

Vorren, T.O., Laberg, J.S., 1997. Trough Mouth Fans-paleoclimate and ice sheetmonitors. Quaternary Science Reviews 16, 865–881.

Weinelt, M., Sarnthein, M., Pflaumann, U., Schulz, H., Jung, S., Erlenkeuser,H., 1996. Ice-free Nordic Seas during the last glacial maximum? Potentialsites of deepwater formation. Paleoclimates 1, 283–309.

Wollenburg, J.E., Mackensen, A., 1998. Living benthic foraminifers from thecentral Arctic Ocean: faunal composition, standing stock and diversity.Marine Micropaleontology 34, 153–185.

Wollenburg, J.E., Knies, J., Mackensen, A., 2004. High-resolution paleopro-ductivity fluctuations during the last 24 kyr as indicated by benthicforaminifera in the marginal Arctic Ocean. Palaeogeography, Palaeoclima-tology, Palaeoecology 204, 209–238.