56
MSc thesis Geology PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS FROM THE LANDSORT DEEP Raisa Alatarvas 2017 UNIVERSITY OF HELSINKI FACULTY OF SCIENCE DEPARTMENT OF GEOSCIENCES AND GEOGRAPHY GEOLOGY

MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

MSc thesis Geology

PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS FROM THE LANDSORT DEEP

Raisa Alatarvas

2017

UNIVERSITY OF HELSINKI FACULTY OF SCIENCE

DEPARTMENT OF GEOSCIENCES AND GEOGRAPHY GEOLOGY

Page 2: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

Tiedekunta/Osasto Fakultet/Sektion – Faculty

Faculty of Science Laitos/Institution– Department

Department of Geosciences and Geography Tekijä/Författare – Author

Raisa Alatarvas Työn nimi / Arbetets titel – Title

Physical properties of glacial sediments from the Landsort Deep Oppiaine /Läroämne – Subject

Geology Työn laji/Arbetets art – Level

MSc thesis Aika/Datum – Month and year

December 2017 Sivumäärä/ Sidoantal – Number of pages

55 Tiivistelmä/Referat – Abstracthttp://www.helsinki.fi/webmail/

The Landsort Deep is the deepest part of the Baltic Sea. It is located just south from the postulated

margins of the early Weichselian glacial, and the deep displays a high-resolution sediment

sequence from late Weichselian and Holocene. The physical properties and characterisation of

late and postglacial sediments from the Landsort Deep enabled the interpretation of the sediment

responses to late Weichselian and Holocene glacial settings. The sediment stratigraphy of the

Landsort Deep reflects variations of salinity in the Baltic Sea Basin, and it enables the

identification of four major stages in the history of the Baltic Sea: Baltic Ice Lake, Yoldia Sea,

Ancylus Lake and Littorina Sea.

The samples studied are from cores from Hole M0063C. The cores were recovered during the

Integrated Ocean Drilling Program (IODP) Expedition 347, “Baltic Sea Paleoenvironment”. Site

M0063 is divided into seven lithostratigraphic units, and the samples are from Unit VI (54–93

ambsf) and Unit V (48–54 ambsf). Sediment analyses and the interpretation of the sedimentary

environment of Units VI and V are based on grain-size analysis by laser diffractometry, loss on

ignition (LOI) determination, and on the IODP Expedition 347 physical properties dataset. The

uniquely long varved glacial clay sequence of Units VI and V gives indications of the

Scandinavian Ice Sheet and the sedimentation of the Baltic Sea Basin responded to climate

fluctuations. The lowermost part of the thick sediment column comprising Units VI and V,

displays ice-proximal settings, and the upper part represents ice-distal settings. The correlation of

the sedimentary features to well-known geological events in the Baltic Sea region enabled the

relative age determination of the Units VI and V. The sediments are deposited during the Baltic

Ice Lake stage and the transition into the Yoldia Sea stage, roughly estimated between 13.5 and

11.5 ka. Comparison to different late and postglacial settings e.g. marine and terrestrial, illustrated

conformities in various glacial sedimentary settings.

Avainsanat – Nyckelord – Keywords

Landsort Deep, Baltic Sea Basin, Scandinavian Ice Sheet, glacial sediments, grain size, clay

varves Säilytyspaikka – Förvaringställe – Where deposited

Helda Muita tietoja – Övriga uppgifter – Additional information

This study is a part of the CISU project funded by Academy of Finland and Russian Foundation

for Basic Research (RFBR).

Page 3: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

Tiedekunta/Osasto Fakultet/Sektion – Faculty

Matemaattis-luonnontieteellinen tiedekunta Laitos/Institution– Department

Geotieteiden ja maantieteen laitos Tekijä/Författare – Author

Raisa Alatarvas Työn nimi / Arbetets titel – Title

Physical properties of glacial sediments from the Landsort Deep Oppiaine /Läroämne – Subject

Geologia Työn laji/Arbetets art – Level

Pro gradu -tutkielma Aika/Datum – Month and year

Joulukuu 2017 Sivumäärä/ Sidoantal – Number of pages

55 Tiivistelmä/Referat – Abstracthttp://www.helsinki.fi/webmai

Itämeren syvin kohta, Landsortin syvänne sijaitsee Varhais-Veikselin-kauden oletetun

jäätikkörajan eteläpuolella. Syvänteen sedimentit edustavat korkean resoluution

sedimenttisekvenssejä Myöhäis-Veikselin ja holoseenin kauden ajalta. Landsortin syvänteen

glasiaalisedimenttien ominaisuuksien tutkiminen ja luonnehdinta mahdollistaa tulkitsemaan

Myöhäis-Veikselin ja holoseenin aikaisen glasiaaliolosuhteiden vaikutusta niihin. Landsortin

syvänteen sedimenttistratigrafia heijastaa Itämeren suolapitoisuuden vaihteluja, joka

mahdollistaa neljän suuren vaiheen tunnistamisessa Itämeren historiassa: Baltian jääjärvi,

Yoldiameri, Ancylusjärvi ja Litorinameri.

Tutkitut näytteet ovat kairareiästä M0063C ja ne kairattiin Itämeren syväkairaushankkeen

(IODP:n Expedition 347, ”Baltic Sea Paleoenvironment”) aikana. Kohde M0063 on jaettu

seitsemään litostratigrafiseen yksikköön ja näytteet ovat yksiköistä VI (54–93 ambsf) ja V (48–

54 ambsf). Sedimenttianalyysi ja sedimentaatioympäristön tulkinta perustuvat raekokoanalyysiin,

hehkutushäviön määritykseen ja IODP Expedition 347:n raportin fysikaalisten ominaisuuksien

aineistoon. Yksiköiden VI ja V muodostama ainutlaatuisen pitkä glasiaalisavilusto-sekvenssi

indikoi Skandinaavian jäätikön ja Itämeren altaan sedimentaation ja ilmastonvaihteluiden

vuorovaikutusta. Sekvenssin alin osa edustaa jäätikön proksimaaliosaa ja ylempi osa jäätikön

distaaliosaa. Yksiköiden sedimenttien korrelaatio tunnettuihin geologisiin tapahtumiin Itämeren

alueella mahdollisti niiden suhteellisen iän määrityksen. Sedimentit ovat kerrostuneet Baltian

jääjärven aikana, sekä sitä seuraavan Yoldiameri-vaiheeseen siirtymisen alkuaikoina, karkeasti

arvioiden 13.5 ja 11.5 ka välillä. Sedimenttien vertaaminen erilaisiin glasiaaliympäristöihin,

kuten mariinisiin ja terrestriaalisiin, vahvisti erilaisten glasiaalisedimenttien

kerrostumisympäristöjen yhdenmukaisuuksia.

Avainsanat – Nyckelord – Keywords

Landsort Deep, Baltic Sea Basin, Scandinavian Ice Sheet, glacial sediments, grain size, clay

varves Säilytyspaikka – Förvaringställe – Where deposited

Helda Muita tietoja – Övriga uppgifter – Additional information

Tämä tutkielma on osa Suomen Akatemian ja Venäjän perustutkimusrahaston (RFBR)

rahoittamaa CISU- projektia

Page 4: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

CONTENTS

1. INTRODUCTION ...................................................................................................4

2. GEOLOGICAL SETTING AND DEVELOPMENT OF THE STUDY AREA ...5

2.1. The geography and hydrology of the Baltic Sea ........................................................................... 5

2.2. Sedimentation of the Baltic Sea Basin ........................................................................................... 6

2.3. The development of the late and postglacial Baltic Sea Basin ...................................................... 6

2.3.1 The Baltic Ice Lake .................................................................................................................... 8

2.3.2. The Yoldia Sea ....................................................................................................................... 10

2.3.3. The Ancylus Lake and the Littorina Sea .................................................................................. 11

2.4. The Landsort Deep ...................................................................................................................... 11

3. MATERIALS AND METHODS ........................................................................... 13

3.1 Site M0063 .................................................................................................................................... 13

3.2. Physical properties measurements .............................................................................................. 14

3.3. Age determination, depth scale adjustment and correlation ...................................................... 14

3.4. Grain size analysis ....................................................................................................................... 15

3.5. LOI determination ...................................................................................................................... 17

4. RESULTS .............................................................................................................. 18

4.1. Lithostratigraphy of Units VI and V ........................................................................................... 18

4.2. Physical properties of Units VI and V ......................................................................................... 25

4.3. Grain size analysis ....................................................................................................................... 30

4.3.1. Clay, silt and sand content ...................................................................................................... 30

4.3.2. Distribution and d-values ....................................................................................................... 33

4.4. Carbon and LOI values, and water content ................................................................................ 39

5. DISCUSSION ........................................................................................................ 41

5.1. Sources of errors.......................................................................................................................... 41

5.2. Physico-stratigraphic zonation of Units VI and V ...................................................................... 41

5.2.1. Subunit VIa 87.5–93.3 ambsf .................................................................................................. 43

5.2.2. Subunit VIb 67.7–87.5 ambsf .................................................................................................. 44

5.2.3. Subunit VIc 54–67.7 ambsf ..................................................................................................... 44

5.2.4. Subunit Va 48–54 ambsf ......................................................................................................... 45

5.3. Reconstruction of the sedimentation environment of Units VI and V ........................................ 46

5.3.1. Relative age determination ..................................................................................................... 46

5.3.2. Sediments, processes and deposition environment ................................................................... 47

5.3.3. Correlation ............................................................................................................................ 49

6. CONCLUSIONS .................................................................................................... 51

REFERENCES .......................................................................................................... 53

Page 5: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

4

1. INTRODUCTION

The location of the Baltic Sea Basin (BSB) in a glaciation-sensitive latitude high in the

northern hemisphere, has led to a very dynamic development during the seas geological

history (Andrén et al. 2011.) The BSB sediments have been affected by the waxing and

waning of the Scandinavian Ice Sheet (SIS) (Andrén et al. 2015a). The recurring

Quaternary glaciations covered repeatedly the whole basin or parts of it, and the

interglacial/glacial cycles and subsequent deglaciations led to widely differential impacts

and uplift in the Baltic Sea region and in its drainage area (Andrén et al. 2011).

The deepest part of the BSB is the Landsort Deep, reaching to the depth of 459 m. It is

located just south from the postulated margins of the early Weichselian glacial, and the

deep displays a high-resolution sediment sequence from late Weichselian and Holocene

(Andrén et al. 2015b). The sediments of the Landsort Deep reflects variations of salinity

in the BSB and it enables the identification of four major stages in its history: Baltic Ice

Lake (BIL), Yoldia Sea (YS), Ancylus Lake (AL) and Littorina Sea (Böttcher and

Lepland 2000).

The Site M0063 in the Landsort Deep was recovered during the IODP Expedition 347.

The Site M0063 is divided into seven lithostratigraphic units (Andrén et al. 2015b), and

the samples are from Hole M0063C from Unit VI (54–93 ambsf) and Unit V (48–54

ambsf). Sediment analyses include grain-size analysis by laser diffractometry, loss on

ignition (LOI) determination, and the interpretation of the IODP Expedition 347 physical

properties dataset. The relative age determination for analysed cores is based on

knowledge of certain geological events in the BSB history.

The focus of this study is on the physical properties and characterisation of late and

postglacial sediments from the Landsort Deep and interpretation of the sediment

responses to late Weichselian and Holocene glacial settings. The interpretation is

complemented with correlation to different late and postglacial sedimentation settings

e.g. marine and terrestrial.

Page 6: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

5

2. GEOLOGICAL SETTING AND DEVELOPMENT OF THE STUDY AREA

2.1. The geography and hydrology of the Baltic Sea

The Baltic Sea is a shallow inland sea and it consists of several sub-basins which all have

different physical and topographic features. Baltic Seas bedrock comprises of

Paleoproterozoic crystalline basement rocks that appear locally on the seafloor.

Sedimentary rocks are present in the southern areas, in the Gulf of Finland, and in tectonic

depressions in the Bothnian Bay and central Bothnian Sea. The bedrock is divided into

blocks by several fracture zones and ancient tectonic lineaments. The most notable

features are probably the pre-glacial originated depressions and troughs (Kaskela et al.

2012.).

Geographical position is between approximately 53° and 66°N, and the sea lies between

temperate and subarctic climate. The Baltic Sea is connected to the North Atlantic through

narrow straits between Denmark and Sweden, which leads to a restricted water exchange.

The Baltic Sea is one of the world's largest epicontinental and brackish water seas

(Kaskela et al. 2012), and it has an area of 377 000km². Björck (1995) reviewed that the

annual fresh-water input is 660 km³, annual discharge of brackish-water is 940 km³, and

the estimated annual inflow of saline bottom-waters is ca. 475 km³. The water depths of

the Baltic Sea are generally 25–75 m in the south, 100–200 m in the central part, and 50–

100 in the north. The salinity of the surface waters in the Baltic Sea basin varies between

ca.1‰ in the northern part and ca. 10‰ in the south (Andrén et al. 2000a). According to

Myrberg et al. (2006) the salinity of the Baltic Sea is one-fifth of average oceans salinity.

The drainage area is 1.7·10⁶ km² and it expands between 49–69°N, and the water

exchange time is about 50 years. One of the Baltic Seas special features is the absence of

tides (Björck 1995).

Page 7: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

6

2.2. Sedimentation of the Baltic Sea Basin

Four significant factors influence the stratigraphic characteristics and properties of the

BSB sediments. The source material and variations in the salinity contribute to grain size

mineralogical composition and organic components of the deposits. The bedrock of the

sea floor and the land areas surrounding it affects the sediments and the resulting

morphology of the sea floor. Sea floor topography and bathymetry in relation to

prevailing air pressure and wind fetch conditions, along with depth to wave base and near

bottom currents govern sedimentation, non-deposition and erosion balance. Biological

factors, such as bioturbation, have also influenced the BSB sediments. (Winterhalter

1992.)

The strong variability in bedrock lithology and the distribution of deformable Quaternary

sediments made the Baltic region favourable for the development of a thin and mobile ice

sheet (Lambeck et al. 1998). Various internal and external factors have affected the

climate in the Baltic Sea region during the Holocene. Changes in the Earth's orbit and

variations in incoming seasonal solar radiation, changes in the land surface, volcanic

activity and the concentration of atmospheric greenhouse gases were the main external

climate drivers during the deglaciation (Borzenkova et al. 2015). Intensive water-table

fluctuations characterized the different stages of the BSB. The subdivision of the BSB

into four different stages is related to water exchange between the Baltic Sea and the

world ocean (Lepland and Stevens 1998). River discharge and meltwater caused the

balance between water-table and glacio-isostatic uplift to rise, which led either to a

temporary connection between the BSB and the world ocean, or to isolation of the BSB

(Lampeck et al.1998).

2.3. The development of the late and postglacial Baltic Sea Basin

The Baltic Sea region had an interstadial sequence with occasional glaciation and boreo-

arctic proglacial fjords between ca. 40.0 and 30.0 ka (Houmark-Nielsen and Kjær 2003).

Page 8: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

7

This was followed by a Last Glacial Maximum (LGM) sequence at ca. 30.0–20.0 ka, with

the closure of subsequent ice streams and fjords (Houmark-Nielsen and Kjær 2003). After

25.0 ka the expansion of the ice sheet occurred mostly along the southern and eastern

terrestrial margins (Hughes et al. 2015). The estimated extent of the ice sheet between

25.0 and 18.0 ka is illustrated in Figure 1, based on the reconstructions by Hughes et al.

(2015). Uncertainty estimates of the extent of the ice sheet area are presented by

maximum, minimum and most-credible lines.

Figure 1. The extent of the ice sheet in the Baltic region between 25.0 and 18.0 ka. Three lines represent

the uncertainty in the data. Dashed line: maximum ice sheet area, dotted line: minimum ice sheet area, solid

white line and shaded white area: most-credible ice sheet area. Figures modified from Hughes et al. (2015).

Page 9: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

8

According to Hughes et al. (2015), a peak in the total ice sheet area and volume occurred

at 21.0–20.0 ka. Three ice-margin fluctuations of the south-eastern margin of the SIS

were identified between 25.0 ka and 12.0 ka (Rinterknecht et al. 2006). Deglacial

millennial-scale climate variability, along with its effect on surface mass balance,

influenced strongly to the rates of SIS margin retreat (Cuzzone et al. 2016). Lambeck et

al. (2010) summarised that the ice sheet advanced and retreated several times and the

maximum advance was not simultaneous in Scandinavia at the time of the LGM. The

advance of Baltic ice streams, the retreat of the Swedish ice, and the transgression of

arctic waters lead to interstadial conditions, which are indicated by a deglaciation

sequence between ca. 20.0 and 15.0 ka (Houmark-Nielsen and Kjær 2003). The glaciers

advanced and reached the largest extent in the Baltic region ca. 18.0 ka ago (Houmark-

Nielsen and Kjær 2003).

2.3.1 The Baltic Ice Lake

The first embryo of the BIL formed at ca. 16.0 ka (Houmark-Nielsen and Kjær 2003).

The BIL was most likely at sea level during its initial stage, and the updamming of the

lake possibly started due to uplift of the Öresund threshold area, which lifted the lake area

above sea level at ca. 14.0 ka (Andrén et al. 2011). The BIL captured meltwater from the

retreating ice sheet and the drainage of most river systems from Europe and western

Russia (Patton 2017). Andrén et al. (2011) concluded that the first drainage of the BIL is

estimated to have taken place 13.0 ka ago, due to the altitudinal difference between the

gradually risen sea and the level of the BIL. The drainage was possibly caused by the ice

recession at Mt. Billingen and at the northern part of the south Swedish highlands (Andrén

et al. 2011). Figure 2 demonstrates the deglaciation of the Baltic region and different

stages of the retreating ice sheet between 14.0 and 11.0 ka. At the time of the Younger

Dryas (YD) cooling at ca. 12.8 ka, the ice sheet re-advanced, blocking the northern

drainage of the BIL which led to transgression (Andrén et al. 2011).

Page 10: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

9

Figure 2. The retreat of the ice sheet and deglaciation of the Baltic region between 14.0 and 11.0 ka. Three

lines represent the uncertainty in the data. Dashed line: maximum ice sheet area, dotted line: minimum ice

sheet area, solid white line and shaded white area: most-credible ice sheet area. Figures modified from

Hughes et al. (2015).

Figure 3 illustrates the BIL at ca. 11.7 ka prior to the maximum extension and final

drainage, which took place when a new passage opened in south central Sweden

(Bergsten and Kjell 1992). The water level of the BIL was abruptly lowered – ca. 25 m

over a period of 1–2 years – down to sea level (Andrén et al. 2011). The final drainage of

the BIL is estimated to occur at ~11.7 ka (Andrén et al. 2015a). It had a great impact on

the BIL environment e.g. changes in fluvial systems, reworking of sediments, and

exposure of large coastal areas (e.g. Andrén et al. 2011, Hyttinen et al. 2011).

Page 11: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

10

Figure 3. Paleogeographic map of the Baltic Ice Lake (light grey area) at ca. 11.7 ka, prior to the maximum

extension and final drainage. Figure modified from Andrén et al. (2011).

2.3.2. The Yoldia Sea

According to Andrén et al. (2011) the onset of the YS stage coincides with the beginning

of the Holocene epoch at ~11.7 ka, and the rapid warming related with that. The lowering

of the BIL down to the YS level possibly was as rapid as the warming, and 25 m

displacement of the shores led to emerged zones rich with clay and silt, and most likely

the YS deposits consist of reworked BIL clays (Björck 1995). Björck (1995, according to

Brunnberg 1998) concluded that a distinct change from a brown clay, to a rather thick-

varved dark grey clay might indicate abrupt increase of deglaciation and sedimentation

rates, which correlates to the transition from the BIL to the YS.

The brackish phase of the YS is documented by the varve lithology, along with

geochemistry and marine/brackish fossils. An occasional show up of sulphide banding

implies a halocline, which gives an estimation of 350 years as maximum duration of the

brackish phase (e.g. Heinsalu et al. 2000, Andrén et al. 2011). The high uplift rate caused

Page 12: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

11

rapid shallowing of the strait area, and this prevented the saline water inflow to the Baltic,

turning the YS into a freshwater basin. (Andrén et al. 2011.) The end of the YS stage is

marked by the shallowing of the outlets in the west of Lake Vänern, caused by isostatic

rebound around the lake, which led to large outflow from the Baltic (Andrén et al. 2011).

2.3.3. The Ancylus Lake and the Littorina Sea

The AL began after the water level raised due to shallowing of the outlets in the west of

Lake Vänern. The melt water input from the final deglaciation of the ice sheet, and the

quite immaculate soils of the deglaciated drainage area, consequenced to organic-poor

sediment deposits in the large freshwater lake. This led to an aquatic environment, which

had low nutrient input and productivity. The AL stage ended when the lake was at

approximate level with the sea in the west and saltwater possibly entering the lake.

(Andrén et al. 2011.)

The first signs of saline water entering the basin, and a minor increase of brackish-

freshwater diatom taxa is recorded at ca. 9950–9750 ka (Andrén et al. 2000a). This period

with low saline influence is called the Initial Littorina Sea (Andrén et al. 2000b). The

Littorina Sea stage is characterized by lithological changes, apparent increase in organic

content, and abundance of brackish marine diatoms (Andrén et al. 2011).

2.4. The Landsort Deep

The Landsort Deep is the deepest part of the Baltic Sea and it is located at the western

part of the Gotland basin, between the southeast coast of Sweden and the island of

Gotland (Figure 4). The Landsort Deep is a crescent-shaped crack with very steep walls

(Myrberg et al. 2006). It is a comparatively narrow trench, with a width of 3.2 km and

length of 20 km, and the location of the deep is along a major fault line, which is deepened

Page 13: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

12

by glacial erosion (Lepland and Stevens 1998). The geometry of the Landsort Deep

enables protection to sediments from subsequent glacial erosion (Andrén et al. 2015b).

Figure 4. Map showing the location of the Landsort Deep in the central Baltic Sea.

The salinity of the water at the western part of the Gotland basin is 6.3–7.7‰ at the

surface, 8.7–10.3‰ beneath the halocline in the depth of 100 m, and 10–11.5‰ near the

seabed at Landsort Deep (Myrberg et al. 2006). The intense stratification of the basin and

great depth cause oxygen problems in the water near the seafloor (Myrberg et al. 2006).

The halocline of the water-column is at about 50 to 80 m, separating surface waters with

a salinity of 7–8‰, from the bottom waters with a salinity varying between 8‰ and 13‰

(e.g. Lepland and Stevens 1998, Böttcher and Lepland 2000,). The limited inflow of

oceanic water and the stratification of the water column leads to depletion of oxygen in

the bottom waters (Böttcher and Lepland 2000). Landsort Deeps water column is

stratified into three different zones: oxic, suboxic and anoxic (Berndmeyer et al. 2014).

Page 14: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

13

Landsort Deep has a high sedimentation rate of 100–500 cm kyr–1 (Andrén et al. 2015b),

and it has varied between 100 and 650 cm kyr–1 from 11.0 ka ago to present day (Obrochta

et al. 2017). Sediments of the deep display the transition from low-organic glacial clay

deposits to organic-rich Holocene clay deposits, and this transition indicates the

changeover from glacial to interglacial, and from freshwater to brackish marine

conditions (Andrén et al. 2015b).

3. MATERIALS AND METHODS

3.1 Site M0063

The cores studied here were recovered during Integrated Ocean Drilling Program (IODP)

Expedition 347, “Baltic Sea Paleoenvironment”, at the Baltic Sea. The drilling occurred

at Site M0063 (58°37.34ʹN, 18°15.25ʹE, water depth 437 m), in the Landsort Deep. Five

holes were drilled at the site: Hole M0063A, M0063B, M0063C, M0063D and M0063E.

The average site recovery was 97.9%. Materials used in this study are from cores

recovered from Hole M0063C, as it has the greatest continuous penetration (Andrén et al.

2015b).

Details of the operations at Hole M0063C are described in the Expedition report 347

(Andrén et al. 2015b), and summarised here. The recovery was 99.29% at Hole M0063C.

For gaining the maximum recovery of the expanding sediments, before firing the piston

corer system (PCS), it was pulled back 1.3 m from the base of each washed-out interval.

The bottom-hole assembly (BHA) was advanced 2 m after each PCS sample run. The

procedure provided the most complete sequence within the expanding sediment package.

The coring continued until diamicton was encountered at 93.3 adjusted meters below

seafloor (ambsf). This nonoverlapping depth scale for each hole was constructed by

Obrochta et al. (2017) by calculating a new depth scale for all cores. The ambsf depth

Page 15: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

14

scale took account of the expansion that the cores experienced due to degassing during

decompression. The hole was opened 3 m, and a hammer sample was taken. The sample

confirmed that the coring had encountered diamicton, and the hole was terminated.

Altogether 40 cores were successfully recovered from Hole M0063C, including one open-

hole section. The maximum depth reached was 96.40 ambsf

3.2. Physical properties measurements

The condensed method description is based on the Expedition 347 report (Andrén et al.

2015b). All physical parameters were measured in one run from core sections shortly after

recovery. During the whole analytical process, all sections were handled in the same way

to guarantee consistency in treatment. Measuring device was automated GEOTEK®

Multi-Sensor Core Logger (MSCL). Standard - MSCL track outfitted with four primary

sensors was used for measuring P-wave velocity, gamma density, magnetic susceptibility

(MS) and non-contact electrical resistivity. Fast Track – MSCL track outfitted with two

primary sensors were working offset measuring magnetic susceptibility. For a complete

lithological sample information, visual examination and written descriptions were

produced from all the cores and core catcher material, along with images of the core

archive halves.

3.3. Age determination, depth scale adjustment and correlation

Age determinations of the studied sediments were derived from different types of

methods, or compilation of those methods. The different stages of the BSB and phases of

the SIS were determined for example by geochronological data from varves (e.g. Andrén

et al. 2002, Stroeven et al. 2015a), radiocarbon, 14C, (e.g. Obrochta et al. 2017,

Rinterknecht et al. 2006), optically-stimulated luminescence, OSL (Houmark-Nielsen

Page 16: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

15

and Kjær 2003) and cosmogenic nuclide,10Be dating (e.g. Cuzzone et al. 2016,

Rinterknecht et al. 2006).

15 samples from Hole M0063C (43–86 ambsf) were radiocarbon dated at the University

of Tokyo for age model and determination (Obrochta et al. 2017). The calibrated

radiocarbon ages of 12 cores from Units VI and V (49.408–85.836 ambsf) vary between

21533 and 24820 cal BP and they are not consistent in relation to depth. The reworked

material of the units strongly affected the radiocarbon ages and the resulted radiocarbon

dates are older than the construed age of deglaciation at Landsort Deep and not in relation

with the known knowledge of the geological events in the Baltic region (Obrochta et al

2017).

Results from this study's analyses are compared with former analyses from the Baltic Sea

area and Holocene marine sedimentation. The relatively age determination for sediments

of Units VI and V between 48 and 93 ambsf are based on the existing knowledge of the

geological history of the BSB, interpretation of the lithostratigraphy of the units and

educated guesses.

3.4. Grain size analysis

The 76 samples (Figure 4) studied are from the depths of Units VI (54–93 ambsf) and V

(48–54 ambsf). Before the laser diffractometer analysis, the samples were oxidized to

reduce the amount of organic matter. The oxidation was carried out with aliquots of 30%

hydrogen peroxide (H2O2) and with an addition of hydrochloride (HCl). Before the laser

analysis, the samples were also chemically treated with 300 mg of sodium pyrophosphate

(Na4P2O7) to prevent particles from flocculating. The samples were also mechanically

treated with 10µm amplitude ultrasound (Figure 5) for 30 seconds before measurements.

Page 17: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

16

Figure 5. Two samples from the 76 samples from Hole M0063C, Units VI and V, and Soniprep 150 ultrasound

device at the Department of Geosciences and Geography at the University of Helsinki.

The samples were analysed using the Department of Geosciences and Geography

Malvern Mastersizer 2000G device (Figure 6). The necessary amount of material from

the samples was used to achieve a beam obscuration of 15–20% on the laser equipment.

Each sample was measured three times with 750rpm stirrer speed and 2000rpm pump

speed.

Figure 6. Malvern Mastersizer 2000G laser diffractometry device at the Department of Geosciences and

Geography at the University of Helsinki.

The analysis of grain size statistics from the laser diffractometry results is derived from a

computer program called GRADISTAT (Blott and Pye 2001). The program runs within

the Microsoft Excel spreadsheet package.

Page 18: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

17

3.5. LOI determination

The loss on ignition (LOI) is a method used in estimating the organic matter and water

content of sediments. The samples were heated in a muffle furnace to determine the

weight percent of organic material by means of LOI (Heiri et al. 2001). The technique is

commonly used in reconstructions of Holocene glacier variations, especially with

glaciolacustrine sediments.

29 samples were chosen to be studied with the LOI method. Heating the samples in the

hot air oven removed pore water from the samples. By igniting the samples in furnace,

the amounts of organic matter and crystallization water could be estimated. First the

empty ceramic crucibles were weighted, and then the crucibles and samples were

weighted. Samples were dried in the oven at 105°C overnight and then cooled in

desiccator. Dried samples and the crucibles were weighted and placed in a muffle furnace

and kept at 550° C for 2 hours. After that, the samples were cooled again in desiccator

and weighted. The weights of crucibles were subtracted to determine the mass loss, and

the result reported was LOI%. LOI was calculated using the following equation from

Heiri et al. (2001)

LOI550 = ((DW105–DW550)/DW105)*100

In the equation DW105 represents the dry weight of the sample before combustion, and

DW550 is the dry weight of the sample after heating to temperature of 550 °C In addition,

total carbon (TC), total organic carbon (TOC), total inorganic carbon (TIC) and sulphur

(S) measurements from Units VI and V (Expedition 347 Scientists 2014a) were used.

Page 19: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

18

4. RESULTS

4.1. Lithostratigraphy of Units VI and V

The description of the lithostratigraphy of Units VI and V is based on the report by

Andrén et al. (2015b), visual images are from Expedition 347 Scientists (2014b), and

these are supplemented by grain size concentration of 76 samples derived from

GRADISTAT.

Unit VI is more than 40 m thick and it is the thickest unit at Site M0063. The unit consists

of finely laminated varved clay, and is featured by a down-core increase of silt and sand

content, which is indicated by dispersed pebble content and thicker laminae. Well-sorted

sand laminae and dispersed clasts occur in the lower parts of the unit, and the percentage

of very coarse sand is 1–2 vol% at 72.58, 75.38–75.88, 76.88, and 86.75 ambsf, and 14

vol% at 70.33 ambsf. Coarse sand is present in a few samples between 68.83 and 82.45

ambsf, with a maximum concentration of 1 vol%.

The concentration of very coarse silt varies from 1 to 11 vol% at depths between 66.67

and 90.08 ambsf. Mostly brown and weak red coloured clay-silt-sand interlaminations

are present upward from ~85 ambsf in thicker units. Occasional internal mm-scale silt

laminae and slumping, packed to cm-scale, within 2–20 cm thick clay laminae is also a

common feature of Unit VI. At 82.45 ambsf medium and fine sand are present with a

percentage of 4–5 vol%. Very fine sand with a concentration of 1–3 vol% appears only

in few depths in Unit VI. Scattered cm-scale carbonate concretions within mm-scale silt

laminae can be identified upward from ~75 ambsf.

The percentage of very fine silt varies between 14 and 28 vol%, with the lowest values

concentrated below ~68 ambsf. The percentage of coarse silt varies between 1–19 vol%,

with the highest values (>10 vol%) concentrated below ~67 ambsf. There are several

distinct fining-upward couplets present with depth, as well as mm-scale sand laminae and

Page 20: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

19

pockets. The share of fine silt varies between 12–25 vol%, with values >20 vol%

appearing below ~62 ambsf. The concentration of medium silt varies between

13– 23 vol%, with the highest values (>10 vol%) appearing also below ~62 ambsf. The

uppermost part of the unit is presented by mm- to cm-scale dark grey-brown clay, and it

is interlaminated by mm-scale silt laminae. Very coarse silt with a concentration of

0– 2 vol% is present only at various depths in Unit VI. Very coarse silt with a

concentration of 18 vol%, along with very fine sand with a concentration of 9 vol% are

present at 58.43 ambsf.

Unit V is distorted, convolute bedded clay which varies from dark grey to greyish brown

in colour and is very well sorted. In parts the internal structures show a massive

appearance, although contorted silt laminated intervals with clear convolute bedding are

visible. At the lower parts of the unit cm-scale laminated clay is displaced by microscale

transverse faults. Very coarse silt with a concentration of 0–2 vol% and very fine sand

with a concentration of 1–3 vol% occur only in a few depths in Unit V.

The lithostratigraphy of Unit VI and V is illustrated in Figure 7. The lowest section

between 92.12 and 93.02 ambsf has a massive, diamictic and homogeneous structure, and

it is moderately or highly disturbed, and poorly sorted. The sections from 77.37 to 92.12

and between 54.5 and 77.37 ambsf have laminated structure, and they are slightly

disturbed. The lower section (77.37–92.12 ambsf) of this continuous unit is well or very

well sorted, and the upper section (54.5–77.37 ambsf) very well sorted. The section from

50.0 to 54.5 ambsf has also laminated and slump bedding structure, with an addition of

homogeneous structure. These sections are slightly or moderately disturbed and very well

sorted. The lithology of the units is mainly of clay and silt content, with a small amount

of sand content at the depth of 55–92 and 48–50 ambsf. The uppermost section between

48 and 50 ambsf is laminated, and has a convolute and slump bedding structure.

Page 21: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

20

Figure 7. Lithostratigraphy of Units VI and V at 48–93 ambsf. Data from Expedition 347 Scientists (2015).

Figure 8 represents visual examples of lithostratigraphy features from Figure 7 of Unit

VI and V. Core 39–1 (90.8–93.3 ambsf) shows massive, moderately or highly disturbed

diamict and sandstone (Figure 8a). Silt is more dominant in core 37–1 (84–87.5 ambsf)

and the sediment is laminated by grain size and colour and slightly disturbed (Figure 8b).

Cores 33–2 (71–74.3 ambsf), 32–2 (67.7–71 ambsf) and 30–2 (61.1–64.4 ambsf) are silt

dominated, and, also laminated by grain size and colour (8c, d, e). There is also a clear

carbonate concretion in core 33–2, and a massive clay interbed in core 32–2. Core 27–2

(52–54.5 ambsf) shows a clayey and silty homogenous structure, and core 26–2

(50– 52 ambsf) slightly or moderately disturbed lamination and slump bedding. Figure

8h illustrates convolute bedding and slumping in moderately disturbed and deformed

sediment in core 25–2 (48–50 ambsf).

Page 22: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

21

a) 39–1 b) 37–1 c) 33–2 d) 32–2 e) 30–2 f) 27–2 g) 26–2 h) 25–2

Figure 8. Examples of diamict and sandstone (a), clay and fine sand-silt laminated by color and grain size

(b, c, d, e), carbonate concretion (c), massive clay interbed (d), homogeneous grey clay and silty clay (f),

slumping, deformation and disturbance (g and h). Images modified from Expedition 347 Scientists (2014b).

The recovered cores were described in detail and photographed by the Expedition 347

Scientists. The description of the cores is based on Expedition 347 Scientists (2015) report

and the core images are modified from Expedition 347 Scientists (2014b). The lowermost

part of Unit VI, Core 39 (90.8–93.3 ambsf) is brown, poorly sorted, and clast poor sandy

diamict and sandstone with pebbles (Figure 8a and 9a). There is also slightly disturbed

Page 23: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

22

interlamination by grey/brown silt and clay, with few dispersed pebbles. Core 38 (87.5–

90.8 ambsf) is well sorted grey clay and silty clay, with fine sand-silt laminations

throughout the core at varying intensities (Figure 9b). Figures 8b and 9c show that core

37 (84.2–87.5 ambsf) is grey clay with mm-scale fine sand-silt and color laminations

throughout the core at varying intensities and mm-cm-scale spacings.

a) b) c)

Figure 9. Core images from Hole M0063C: a) 39/90.8–93.3 ambsf, b) 38/87.5–90.8 ambsf, c) 37/84.2–87.5

ambsf. Images modified from Expedition 347 Scientists (2014b).

Core 36 (80.9–84.2 ambsf) is greyish brown clay, which is interlaminated by silt laminae,

and they are occasionally bundled (Figure 10a). Grey clay with mm-scale fine sand-silt

and color laminations at varying intensities and spacings is the dominant feature

throughout core 35 (77.6–80.9 ambsf) (Figure 10b). Core 34 (74.3–77.6 ambsf) is greyish

brown clay, which is interlaminated by silt laminae (Figure 10c). Grey clay and silty clay

varies in core 33 (71–74.3 ambsf). It is well sorted with mm-scale color and fine silt-sand

laminations occurring in cm-scale in unequal distributions (Figure 8c and 10d). There is

also a carbonate concretion at 120 cm in section 2.

Page 24: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

23

a) b) c) d)

Figure 10. Core images from Hole M0063C: a) 36/80.9–84.2, b) 35/77.6–80.9 ambsf c) 34/74.3–77.6 ambsf,

d) 33/71–74.3 ambsf. Images modified from Expedition 347 Scientists (2014b).

The content of core 32 (67.7–71 ambsf) is mainly dark grey clay, which is planar, parallel

laminated by color (Figure 8d and 11a). There is also a massive clay interbed in section

1 at 45–55 cm, and in section 2 at 65–85 cm. Section 4 (CC) is highly disturbed by coring.

All though core 31 (64.4–67.7 ambsf) is dark grey clay, which is mostly laminated by

color and grain size, there are also homogeneous parts in section 2 and 3 (Figure 11b).

Planar, parallel laminated dark grey clay is the main feature of core 30 (61.1– 64.4 ambsf)

but there are silt laminae with uneven thickness and distribution throughout the sections,

and section 3 also has a massive homogeneous dark grey clay interbed (Figure 11c). Core

29 (57.8–61.1 ambsf) is dark grey clay with lamination by color and grain size (Figure

11d).

Page 25: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

24

a) b) c) d)

Figure 11. Core images from Hole M0063C: a) 32/67.7–71 ambsf, b) 31/64.4–67.7 ambsf, c) 30/61.1–64.4

ambsf, d) 29/57.8–61.1 ambsf. Images modified from Expedition 347 Scientists (2014b).

The main feature of core 28 (54.5–57.8 ambsf) is dark grey clay with faint lamination by

color (Figure 12a). Section 3 has visible layers, and there are also dispersed sand grains

in sections 2 and 3. Core 27 (52–54.5 ambsf) is dark grey clay with lamination or faint

lamination by color (Figure 8f and 12b). There are dispersed intraclasts and slumping

features, and sections 1 and 2 are slightly deformed and disturbed. Interlaminated dark

grey and dark greyish brown clay is the main feature of core 26 (50–52 ambsf) (Figure

8g and 12c). Bedding is strongly contorted in sections 1 and 2, and there are also micro-

faults and steep inclining and disturbance by slumping. Core 25 (48–50 ambsf) is dark

grey clay with intraclasts (Figure 8h and 12d). There occurs soft-sediment deformation

and convolute bedding or slumping. The sediment is moderately or highly disturbed.

Page 26: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

25

a) b) c) d)

Figure 12. Core images from Hole M0063C: a) 28/54.5–57.8 ambsf, b) 27/52–54.5 ambsf, c) 26/50–52

ambsf, d) 25/48–50 ambsf. Images modified from Expedition 347 Scientists (2014b).

4.2. Physical properties of Units VI and V

Natural gamma ray (NGR) data shows that, regardless of the expansion of the sediment

and coring disturbances, the sediment record recovery was nearly continuous (Andrén et

al. 2015b). The NGR values remain relatively high (~20cps) and constant through Units

VI and V, and there are also meter-scale peaks of lower values throughout the units

(Figure 13) The distortion between 45–53 ambsf is caused by expansion of the sediment

(Andrén et al. 2015b). Discrete P–wave velocity in Unit VI is typically ~1500 m s–1, and

~1250 m s–1 in Unit V. Zero values are due to distortion at the top and bottom of the core

(Andrén et al. 2015b). Dry density increases downhole and through Unit VI it is generally

>1 g cm³. There is a general declining trend of density from bottom of Unit VI to the top

of Unit V, which is positively related to a decrease in sample wet density

(Andrén et al. 2015b). Large excursions in magnetic susceptibility (MS) are limited to the

lower portion of Unit VI.

Page 27: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

26

Figure 13. Natural gamma radiation, P-wave velocity, density and magnetic susceptibility values of Units VI

(54–93 ambsf) and V (48–54 ambsf). Dashed lines marking the changes in physical properties trends. Data

from Expedition 347 Scientists (2014c).

48

53

58

63

68

73

78

83

88

93

10 15 20 25

DE

PT

H (

am

bsf)

NATURAL GAMMA RADIATION (cps)

48

53

58

63

68

73

78

83

88

93

0 400 800 1200 1600 2000

DE

PT

H (

am

bsf)

P-WAVE VELOCITY (m s–1)

48

53

58

63

68

73

78

83

88

93

0,8 1,2 1,6 2 2,4

DE

PT

H (

am

bsf)

GAMMA DENSITY (g cm³)

48

53

58

63

68

73

78

83

88

93

0 50 100 150 200 250 300

DE

PT

H (

am

bsf)

MAGNETIC SUSCEPTIBILITY (IUs)

Page 28: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

27

The MS, density, P-wave velocity and NGR values of Units VI and V are divided into

sections of certain depths and are illustrated in Figures 14–17. There are larger excursions

in MS at the lower parts of Unit VI between 81 and 93.3 ambsf, and the values vary from

40 to 160 10–5 instrument unit (IUs). The average value between 68 and 81 ambsf is 60

IUs. The values are quite constant from 48 to 68 ambsf, settling between 30 and 40 IUs,

with the addition of a few peaks of higher values.

Figure 14. Magnetic susceptibility of cores from Unit VI and V from depths between 48 and 93 ambsf. Data

from Expedition 347 Scientists (2014c).

87

88

89

90

91

92

93

60 160 260

am

bsf

IUs

81

82

83

84

85

86

87

40 80 120 160

am

bsf

IUs

75

76

77

78

79

80

81

40 60 80 100

am

bsf

IUs

69

70

71

72

73

74

75

20 40 60 80

am

bsf

IUs

63

64

65

66

67

68

69

20 40 60 80

am

bsf

IUs

58

59

60

61

62

63

20 40 60 80

am

bsf

IUs

53

54

55

56

57

58

20 40 60

am

bsf

IUs

48

49

50

51

52

53

0 60 120 180

am

bsf

IUs

Page 29: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

28

Figure 15 shows that density decreases upward in the units. Density is ~2 g m3 from 87

to 93 ambsf, and the highest value (>2 g m3) appears at 92–93 ambsf. Peaks of lower

values also occur at 90.5 and 93 ambsf. At 75–87 ambsf the values vary between 1.8 and

2 g m3, and at 50–75 ambsf between 1.6 and 1.8 g m3. Upward from 49 ambsf the density

decreases to 1.4 g m3.

Figure 15. Density of cores from Unit VI and V from depths between 48 and 93 ambsf. Data from Expedition

347 Scientists (2014c).

P-wave velocity varies approximately between 1000 and 1500 m s-1 throughout Units VI

and V, with an exception of few peaks of lower values (Figure 16). A relatively constant

value of 1500 m s-1 is present at 81–83, 78.5–80.5, 61–63 and 59–60.5 ambsf. There are

frequent variations of velocity in relation to depth between 48 and 53 ambsf.

87

88

89

90

91

92

93

0,8 1,4 2 2,6

am

bsf

g m3

81

82

83

84

85

86

87

1,6 1,8 2 2,2

am

bsf

g m3

75

76

77

78

79

80

81

1,6 1,8 2 2,2

am

bsf

g m3

69

70

71

72

73

74

75

1,4 1,8 2,2

am

bsf

g m3

63

64

65

66

67

68

69

1,6 1,8 2

am

bsf

g m3

58

59

60

61

62

63

1,4 1,6 1,8 2 2,2

am

bsf

g m3

53

54

55

56

57

58

1,4 1,6 1,8

am

bsf

g m3

48

49

50

51

52

53

1,2 1,4 1,6 1,8

am

bsf

g m3

Page 30: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

29

Figure 16. P-wave velocity of cores from Unit VI and V from depths between 48 and 93 ambsf. Data from

Expedition 347 Scientists (2014c).

Figure 17 illustrates that the NGR values vary between 15 and 20 cps at 87–93 ambsf,

and from 69 to 87 ambsf the values are ~20 cps, with a few peaks of lower values. From

61 to 69 ambsf, there is a slight increase of values to >20 cps, and between 48 and 61 the

values settle again at ~20 cps. The lowest value (10 cps) occurs at 52 ambsf.

87

88

89

90

91

92

93

500 1500

am

bsf

m s-1

81

82

83

84

85

86

87

500 1000 1500

am

bsf

m s-1

75

76

77

78

79

80

81

750 100012501500

am

bsf

m s-1

69

70

71

72

73

74

75

0 500 10001500

am

bsf

m s-1

63

64

65

66

67

68

69

250 750 1250

am

bsf

m s-1

58

59

60

61

62

63

750 1500

am

bsf

m s-1

53

54

55

56

57

58

500 1000 1500

am

bsf

m s-1

48

49

50

51

52

53

500 1000 1500

am

bsf

m s-1

Page 31: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

30

Figure 17. Natural gamma ray of cores from Unit VI and V from depths between 48 and 93 ambsf. Data from

Expedition 347 Scientists (2014c).

4.3. Grain size analysis

4.3.1. Clay, silt and sand content

The mean grain size of the 76 samples from Unit VI and V from depths between 49

and 90 ambsf is described by using a scale in which the grain size of <2 µm is classified

as clay, 2–63 µm is silt, and >63 µm is sand. Figure 18 shows the change of the clay-silt

ratio in relation to depth. The silt content decreases and clay content increases upward,

87

88

89

90

91

92

93

10 15 20 25

am

bsf

cps

81

82

83

84

85

86

87

15 20 25

am

bsf

cps

75

76

77

78

79

80

81

10 15 20 25

am

bsf

cps

69

70

71

72

73

74

75

10 15 20 25

am

bsf

cps

63

64

65

66

67

68

69

15 20 25

am

bsf

cps

58

59

60

61

62

63

15 20 25

am

bsf

cps

53

54

55

56

57

58

15 20 25

am

bsf

cps

48

49

50

51

52

53

10 15 20 25

am

bsf

cps

Page 32: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

31

and between 66 and 90 ambsf the clay content is 20–30 vol% and silt content 70–80 vol%.

The sand content in the lower part has a maximum percentage of 5 vol%, and notable

peaks approximately at 58, 70 and 82 ambsf. Between 49 and 66 ambsf the clay content

is 40–60 vol%, silt 40–60 vol%, and sand 0–5 vol%. The percentage of clay and silt are

relatively equal in the upper part.

Figure 18. Clay, silt and sand content (vol%) of the 76 samples. The samples are from Units VI and V, from

different depths between 49 and 90 ambsf. Arrow pointing the boundary between Unit VI and V.

The grain size scale was divided into certain size fractions of silt and sand. This enabled

the construction of the volume percentage in each size fraction. The scale of sand grain

size fraction is divided between the ranges of 63–200 µm, 200–600 µm and 600–2000 µm

(Figure 19). Between 57 and 63 ambsf the content is fine/very fine and medium sand.

Below 68 ambsf the dominant size fraction of sand content varies from medium to

coarse/very coarse sand. The highest coarse content is approximately at 75 ambsf. A

notable factor is the absence of sand content at depths of 50–57 and 64–67 ambsf. The

silt grain size fraction is divided into 2–4 µm, 4– 16 µm and 16–63 µm (Figure 19). There

0 % 10 % 20 % 30 % 40 % 50 % 60 % 70 % 80 % 90 % 100 %

49

53

55

57

59

61

63

65

67

69

71

73

75

76

78

81

82

84

86

Depth

(am

bsf)

Content (vol%)

Clay < 2µm Silt 2-63 µm Sand 63-2000 µm

Page 33: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

32

is a decrease of medium/coarse silt content and an increase of very fine silt/clay content

upward the units. A peak of coarser material appears at 58 ambsf.

Figure 19. The percentage of grains falling into each size fraction of sand and silt. The samples are from

different depths between 49 and 90 ambsf from Units VI and V.

0 % 10 % 20 % 30 % 40 % 50 % 60 % 70 % 80 % 90 % 100 %

49

53

55

57

59

61

63

65

67

69

71

73

75

76

78

81

82

84

86

Depth

(am

bsf)

Sand content (vol%)

63-200 µm 200-600 µm 600-2000 µm

0 % 10 % 20 % 30 % 40 % 50 % 60 % 70 % 80 % 90 % 100 %

49

53

55

57

59

61

63

65

67

69

71

73

75

76

78

81

82

84

86

Depth

(am

bsf)

Silt content (vol%)

2-4 µm 4-16 µm 16-63 µm

Page 34: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

33

4.3.2. Distribution and d-values

Figure 20 illustrates the grain size distribution of the 76 samples. The type and textural

group of the samples is unimodal and poorly sorted mud, with an exception of three

samples, which are bimodal and very poorly sorted sandy mud. Sediment varies from

very fine, fine, and medium silt to medium sandy fine silt, very coarse sandy fine silt,

very fine sandy coarse silt, and mud.

Figure 20. Grain size distribution of the 76 samples from Unit VI and V.

Figures 21–24 provide visual images of 14 samples, as well as their grain size distribution.

Figure 21 shows samples from the lower part of Unit VI from depths between 87.5 and

93.3 ambsf. The sediment name of sample 38–2 is very fine silt and sample 37–2 is fine

silt. The content of these samples is 96–100 vol% mud and 0–4 vol% sand.

0

1

2

3

4

5

0,2 2 20 200 2000

Vol%

Grain size (µm)

Page 35: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

34

Figure 21. Grain size distribution and cm-scale images of samples from Unit VI. Images modified from

Expedition 347 Scientists (2014b).

Samples from depths 67.7–87.5 ambsf are illustrated in Figure 22. Sample 36–2 is

medium sandy fine silt, sample 32–2 is very coarse sandy fine silt. These samples are

13– 16 vol% sand and 84–87 vol% mud. Sample 35–1 is very fine silt, and samples 34– 2

and 33–1 are fine silt. These samples are 96–100 vol% mud and 0–4 vol% sand.

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

38–2/108–110cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

37–2/105–107cm

Page 36: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

35

Figure 22. Grain size distribution and cm scale images of samples from Unit VI. Images modified from

Expedition 347 Scientists (2014b).

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

36–2/5–7cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

35–1/82–84cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

34–2/8–10cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

33–1/57–59cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

32–2/113–115cm

Page 37: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

36

Samples from Unit VI between 54 and 67.7 ambsf are presented in Figure 23, and samples

from Unit V between 48 and 54 ambsf are illustrated in Figure 24. The samples from

31– 1 to 25–2 are mud, with the exception of sample 30–2 beeing very fine silt. These

sample are almost all 100% mud, and the main grain size is clay with distribution between

31 and 57 vol%.

Figure 23. Grain size distribution and cm scale images of samples from Unit VI. Images modified from

Expedition 347 Scientists (2014b).

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

31–1/17–19cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

30–2/47–49cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

29–1/114–116cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

28–1/70–72cm

Page 38: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

37

Figure 24. Grain size distribution and cm scale images of samples from Unit V. Images modified from

Expedition 347 Scientists (2014b).

The d-values were used to illustrate volume distribution of the 76 samples from depths

between 49–90 ambsf. The d-values are the intercepts for 10 vol%, 50 vol% and 90 vol%

of the cumulative mass. Figure 26 shows an increase of coarser material downward from

~67 ambsf, which is in line with Figures 17–19.

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

27–1/18–20cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

26–2/78–80cm

0

2

4

6

0,01 0,1 1 10 100 1000 10000

Vol%

µm

25–2/64–66cm

Page 39: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

38

Figure 25. The d(10), d(50) and d(90) values based on a volume distribution from all 76 measured samples

from Unit VI and V.

A table of d(10), d(50) and d(90) values from four samples from Units VI and V displays

examples of the volume distribution of different sediment types (Table 1). Sample

36– 2/5–7cm (82.45–82.47 ambsf) is medium sandy fine silt, and sample 33–1/57–59cm

(71.57–71.59 ambsf) is fine silt. The latter represents the most common volume

distribution of the samples. Sample 32–2/113–115cm (69.93–69.95 ambsf) is very coarse

sandy fine silt, and the high d(90) value of the sample and the grain size variation is also

visible in the cm scale image of the sample in Figure 23. Sample

27– 1/18– 20cm (52.18– 52.20 ambsf) values represents the volume distribution of mud.

Table 1. The d(10), d(50) and d(90) values and depths from samples from Unit VI and V.

Sample Depth (ambsf) d (10) d (50) d (90)

36–2/5–7cm 82.45–82.47 1.202 6.535 124.3

33–1/57–59cm 71.57–71.59 1.270 6.167 23.3

32–2/113–115cm 69.93–69.95 1.363 8.084 1402.4

27–1/18–20cm 52.18–52.20 0.595 1.992 7.8

48

53

58

63

68

73

78

83

88

93

0 1 10 100 1000

Depth

(am

bsf)

d (10) d (50) d (90)

Page 40: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

39

4.4. Carbon and LOI values, and water content

The sedimentary total carbon (TC), total organic carbon (TOC), total inorganic carbon

(TIC), and sulphur values determined by Expedition 347 Scientists (2014a) are illustrated

in Figure 27. TC values are relatively low and vary from 0.2 to 0.5 wt% in Units VI and

V. The sediments of Units VI and V are lean in TOC, with values of <5 wt%. TIC values

are also relatively low, varying around 0.1 wt% throughout the units. The TOC and TIC

values follow the TC trend, except for the TIC values at 50–58 ambsf. Generally, there

are low values of TC reaching only up to 5 wt% only in the uppermost part of Unit V.

The sulphur content is relatively flat-rated throughout the units, with the value settling

around 0.1 wt%.

Figure 26. Sulphur (S), total carbon (TC), total organic carbon (TOC) and total inorganic carbon (TIC) values

of Unit VI and V, between 48 and 91 ambsf. Data from Expedition 347 Scientists (2014a).

48

53

58

63

68

73

78

83

88

93

0 0,1 0,2 0,3 0,4 0,5 0,6

Depth

(am

bsf)

wt%

TC S TOC TIC

Unit V

Unit VI

Page 41: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

40

The LOI and water content determination from 29 samples show that the values generally

increase from bottom to top throughout the sediment column (Figure 27). In general, the

LOI values measured range between 2.1wt% and 5.5wt% and the water content is

between 21wt% and 40.5%. In the lower part of Unit VI there is an increase of the values

around 85 ambsf. There is a decrease of the values at 79 ambsf, which is followed by a

short increase, and then a small decrease again at 75 ambsf. Between 74 and 69 ambsf the

values are relatively steady, and from there the values increase upwards. Both values

decrease at the bottom of Unit V at 53 ambsf, and then increase at the upper part of the

unit. The water content and LOI values follow the same trend throughout the column with

an exception at 63 ambsf.

Figure 27. Water content and LOI values of 29 samples from Units VI and V between 49 and 90 ambsf. Data

from this study's LOI and water content determination from 29 samples from Units VI and V.

2 2,5 3 3,5 4 4,5 5 5,5

49

54

59

64

69

74

79

84

89

20 25 30 35 40 45

wt%

Depth

(am

bsf)

wt%

water content LOI

Unit V

Unit VI

Page 42: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

41

5. DISCUSSION

The physico-stratigraphic zonation of Units VI and V presented here is based on the

lithostratigraphy, physical properties and results from the grain size analysis from the

units. The reconstruction of the sedimentation environment is concluded from the

physico-stratigraphic zonation and previous knowledge of the Holocene glacial

environment of the BSB. Included are also relative age determination of the units and

correlation to different Holocene glacial sedimentation environments.

5.1. Sources of errors

In their reports, Andrén et al. (2015b) and Obrochta et al. (2017) discussed the difficulties

and challenges in the coring processes at Site M0063. Methane gas expansion caused

physical core disturbance, disturbing the between-sample accuracy of the paleomagnetic

record of the sediments recovered from the site. Coring disturbances and large nonlinear

sediment expansion were the cause of the hole-to-hole correlation remaining at a relative

approximate level of 0.5–1.0 m. At the top of each core, the physical property data had a

strong noise, which had to be cleaned. The lithologic information and seismic profiles

suggest the possibility of some type of gravity flows occurring in the area, which caused

the deforming of the sediment. There were many challenges in the correlation and there

are visible larger scale trends in all physical properties.

5.2. Physico-stratigraphic zonation of Units VI and V

The general trend in the clay-silt-sand ratio of Units VI and V is the upward increase of

clay content and decrease of silt content (Figure 17). In their study, Andrén et al. (2002)

stated that in the relation of grain size and colour, the only general factor is that the silt

content in brownish clay is higher than in grey clay. The upward decreasing silt and sand

Page 43: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

42

content of Units VI and V is concordant to the colour change from brownish to grey.

According to Andrén et al. (2002), the change in colour can be caused by the change of

the sediment source area, the change in the input of coarser material, or the change of the

chemical and/or physical composition of the sediment. The upward fining of sediments

into weakly colour laminated clay can indicate increasing distal conditions and

continuous retreat of the ice sheet (Streuff et al. 2017a).

The sediments of Unit VI vary by colour from brownish/reddish grey clay at the

lowermost parts of the unit to brownish/greyish clay at the upper parts of the unit. This

colour change is also visible in the core images presented in Figures 8–12 and 21–24. The

volume distributions of the 76 samples are consistent with the variation of the grain size.

Silt is dominant at depths between 67 and 90 ambsf, and finer clay is dominant upward

from ~67 ambsf. The d-values from the 76 samples also illustrated coarsening of the

material downward from ~67 ambsf (Figure 25).

Andrén et al. (2015b) state that peaks in NGR indicate variation and reduction of grain

size and the occurrence of meter-scale peaks throughout Unit VI indicate a

correspondence to intervals of reduced grain size. Figure 13 showed the upward

increasing NGR values which are concordant with the upward decreasing silt and sand

content illustrated in Figure 18. Density decreases upward in Units VI and V in relation

to the fining of the sediment (Figure 13 and 18). According to Harff et al. (2011) changes

in the depositional regime are reflected in the MS values. There are notable changes in

the MS values at certain depths throughout the sediment column.

Figure 27 illustrated that the water content and LOI values of the 29 samples from Units

VI and V follow the general trend of upward increasing values. Correlative shifts in LOI

values and the variation of grain size are related to sedimentary change. The LOI values

are relatively low throughout the sediment column and, according to MacLeod et al.

(2006), low LOI values represent the expected low levels of productivity in recently

deglaciated basin. The TOC and TIC values are also relatively low and in line with the

TC values throughout Units VI and V (Figure 26). Changes in ice sheet extent are thought

Page 44: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

43

to be illustrated by the alternations between organic and non-organic sedimentation

regimes. (Smith 2003.)

Based on these differences in grain size variations and physical properties, the sediment

column of Units VI and V is divided into four subunits between certain depths. This

division is also illustrated in Figure 13.

5.2.1. Subunit VIa 87.5–93.3 ambsf

The frequent sand laminations and increased grain size, with pebble and sand grains,

between 87.5 and 93.9 ambsf indicate a proximal glaciolacustrine environment (Andrén

et al. 2015b). The massive brownish/reddish sandy diamict and pebbly sandstone between

90.8 and 93.3 ambsf (Figure 8a and 9a) are an indication of high-energy mechanism such

as ice-rafting. These lowermost parts of the subunit display a deposition environment of

an ice-proximal setting. Slightly disturbed interlaminations of grey and brown silt and

clay are also present at lowermost part of the subunit.

Unconsolidated sediments usually have lower density than lithified fragments, which is

why an increase of fragments and clasts could increase the density (MacLeod et al. 2006).

The density increases in the diamict and sandy sediment at the lowermost part between

92 and 93.3 ambsf (Figures 13 and 15). Because wet bulk density is sensitive to changes,

increased input of terrigenous sandy material can also cause an increase in density (Harff

et al. 2011). The increase in density is concordant with the increasing values of MS

illustrated in Figure 13. There occur larger excursions of MS at the lower part of the unit

(Figure 14), and according to MacLeod et al. (2006), the correlative shifts in MS values,

along with grain size variation, are related to sedimentary change. There also occurs slight

variation in velocity and NGR values. The TOC value increases and TIC value decreases

at 87 ambsf.

Page 45: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

44

5.2.2. Subunit VIb 67.7–87.5 ambsf

The sediments are silt dominated and laminated by grain size and colour at 84–87.5,

71.0– 74.3 and 67.7–71 ambsf (Figures 9c, 10d and 11a). These sediments are most likely

formed by low-energy mechanism. The colour of the sediments varies from greyish

brown clay at the lower part between 74 and 87.5 ambsf to darker grey clay at 67.7– 74

ambsf. The intensified melting of the ice sheet leading to increasing sedimentation rate

can also be the cause for the colour change in the sediment (Andrén et al. 2002).

Sequences of darker clayey and lighter silty sediment couplets indicate winter and

summer seasons (e.g. Figure 8e and 10). Massive clay interbeds appear at 69.85–70.05

and 68.15–68.25 ambsf (Figure 8d and 11a). Andrén et al. (2011) write that varved clay

forms at ice-proximal areas and homogeneous clay deposits at ice-distal areas.

According to MacLeod et al. (2006) the decrease in MS is due to the increase of organic

and water content. A decrease in MS value is in relation to an increase in water content

and LOI, TIC, and TC values at approximately 85 ambsf (Figures 13, 26 and 27). A peak

in LOI values at 85 ambsf is possibly due to the formation of different precipitates.

Several alterations of water content and LOI values occur throughout the sediment

column. TOC values show an increase upward from 73 ambsf, and, according to Andrén

et al. (2002) increased organic production or enhanced preservation of organic carbon in

anoxic conditions can be the cause of increase in the content of organic carbon. The NGR

values are relatively constant, except for a few lower values. The velocity varies from

constant to frequent oscillations of the values.

5.2.3. Subunit VIc 54–67.7 ambsf

Common features are fine laminaes of varved clay, as well as mm-scale silt laminae and

slumping within thicker clay laminae. The variations in the concentration of the sediments

throughout the subunit might be related to changes in clay availability from the source

areas. There is also an increase of water content and LOI and carbon values upward from

67.7 ambsf. Homogeneous parts are present between 66 and 67 ambsf (Figure 11b) and

Page 46: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

45

massive interbeds of homogeneous grey clay are present at 64.1–64.4 ambsf (Figure 11c).

The thickness of the homogenous sediments varies which indicates nonconformities in

sedimentation. The sound velocity drops in homogeneous deposits, but the density

increases (Harff et al. 2011). There are notable declines in density between 64 and 67

ambsf and from 50 to 57 ambsf (Figure 13), which is related to the absence of coarser

sandy material (Figures 18 and 19). The MS values are relatively constant and low, but a

slight increase of NGR values occur at lower part, and velocity is constant in certain parts

of the subunit. The water content and LOI and carbon values show variations in the upper

part of the subunit.

5.2.4. Subunit Va 48–54 ambsf

The laminated clay between 48 and 54 ambsf is an indication of a possible depositional

environment (Andrén et al. 2015b). The internal structures show, in part, massive,

contorted silt laminated intervals with convolute bedding (Figures 12c and d) and

slumping features (Figures 8g and h). There are a few steeply inclined soft-sedimentary

microfaults at ~50 ambsf and bedding is strongly contorted and microfaulted at ~51

ambsf. The sediment also illustrates soft-sediment deformation. Due to compaction,

density decreases gradually in soft clay and silt (Harff et al. 2011). The decrease in density

between 48 and 50 ambsf (Figures 13 and 15) is related to the soft-sediment deformation

at the uppermost part of the unit. The very well sorted and distorted, convolute bedded

clay varies by colour from greyish brown to dark grey. The colour change from brownish

to grey can also be an indication of oxygen deficiency caused by an increase in the

sedimentation rate (Andrén et al. 2002).

Velocity decreases upward the subunit and more frequent variations of the values also

take place (Figure 16). The physical properties data from Expedition 347 Scientists

(2014b) suggests that the velocity drop is possibly due to the change of sediment type and

increase of organic material and water content. An increase of water content and LOI

values and, also the highest values of TC, TIC and TOC appear between 48 and 50 ambsf

(Figures 26 and 27). The MS and NGR values are relatively constant, except for the peak

of a lower NGR value at 52 ambsf.

Page 47: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

46

5.3. Reconstruction of the sedimentation environment of Units VI and V

5.3.1. Relative age determination

The ice recession and the first drainage of the BIL are estimated to have taken place at

13.5– 13.0 ka (e.g. Andrén et al. 2011, Björck 1995). Proximal sediments of the

lowermost part of Unit VI are possibly deposited at the time of the first drainage of the

BIL and ice recession. The inferred age of the deglaciation of the Landsort Deep is ~13.5

ka (Obrochta et al. 2017). The reconstruction of the BIL by Vassiljev and Saarse (2013)

indicates that the Landsort Deep was deglaciated at 12.2 ka, and the water level at ice

margin was up to 140–150 m a.s.l.

Distinct pollen types at ~66 ambsf might imply a pre-Holocene age for that spectrum, and

it could also be related to the Bølling/Allerød interstadials (Andrén et al. 2015b). The

change in the sediment content in the lower part of subunit VIc (~66–67.7 ambsf)

substantiate warmer climate conditions. Previously dated pollen spectrum typical at 11.0–

10.5 ka was found at ~56.8 ambsf, suggesting a correlative age for the sediments (Andrén

et al. 2015b). According to Andrén et al. (1999), an ice-rafted debris (IRD) sequence

preceded the final drainage of the BIL. Dispersed intraclasts and sand grains in various

depths at 52.5–58 ambsf might reflect this IRD sequence of debris settling from melting

ice sheet and/or icebergs.

Sediments of the upper part of Unit V most likely represent the final drainage of the BIL

~11.7 ka ago (Andrén et al. 2015a), and the onset of the transition into the YS stage. The

silt content shows a slight increase upward Unit V which might have resulted from

intermixing of finer and coarser material and indicate reworking of sediments. Sharp

contacts of such sediments are interpreted as mass-transport deposits (Streuff et al.

2017a). The structure of Unit V also corresponds to an abrupt event such as the drainage

and lowering of the water level.

Page 48: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

47

The Hole M0063C was analysed for siliceous microfossils for locating the 150-year

brackish phase of the YS stage by combining diatom stratigraphy with its stratigraphic

position, and a typical diatom assemblage of the YS stage was found at five levels in Hole

M0063C between 41 and 43 ambsf (Andrén et al. 2015b). According to Obrochta et al.

(2017) the brackish phase of YS is represented by the laminated grey clays with iron

sulphides that are overlying the thick sequence of varved glacial clays between 48 and 93

ambsf. Variations of the geochemistry in pore water from the Site M0063 reflects the

transition from the underlying glaciolacustrine sediments to Holocene brackish marine

deposits (Andrén et al. 2015b).

5.3.2. Sediments, processes and deposition environment

The stratification of the glacial lake controls the resulting processes of sedimentation that

may occur through processes such as the deposition from meltwater flows, direct

deposition from the glacier front, settling from suspension, re-sedimentation by gravity

flows, current reworking and rain-out from icebergs (Bennett and Glasser 2009). The

depth and extent of ice-marginal lake, such as the BIL and its sedimentation are controlled

by different factors e.g. location of the ice sheet margin, volume of sediment supply and

differential isostatic rebound (Teller 2013).

The poorly sorted massive sandy diamict of subunit VIa were presumably deposited

directly at the ice sheet margin. The diamict with silt interlaminas appearing in an upward-

fining succession suggests depositional environment with initially higher energy but

increasingly lower depositional energy. According to Björck and Möller (1987), this kind

of subglacial melt-out sediment is deposited by slow melt-out release of basal debris from

stagnant ice. Thick sequences of proximal sediments possibly depict very slow retreat of

the ice margin (Streuff et al. 2017b). The ice-distal sediments overlaying the ice-proximal

sediments of subunit VIa possibly indicate that the retreat of the ice sheet was relatively

continuous. The interlaminas of more sorted sediment are synsedimentary and melt-out

deposits, and they are deposited in cavities at the ice sheet sole/bed boundary or in

englacial cavities produced by running water (Björck and Möller 1987).

Page 49: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

48

Varved glacial clay is present between 71 and 90 ambsf with varying thickness

(Expedition 347 Scientists 2015). Varves form during the melting of the ice sheet, runoff

and proglacial sedimentation and demonstrate the pace of deglaciation, and they generally

reflect regular, most likely seasonal variations in sediment and meltwater supply. The

varves are very compact and disturbed in the lowermost part of Unit VI (Expedition 347

Scientists 2015). Differences in varve thickness are generally caused by changes in

sediment input by meltwater, slight oscillations of the ice margin and local iceberg

activity. The varve thickness between 81 and 87 ambsf is 10–20 mm, and some layers are

a little darker (Expedition 347 Scientists 2015). At 77.5–81 ambsf, the varve thickness is

1–5 mm, and there are no changes between darker and lighter layers (Expedition 347

Scientists 2015) which reflects seasonal invariability.

The input of water and sediment is highly seasonal, and thin finer clayey layers form in

the winter, while thicker and coarser silty and sandy layers form during the summer. The

varve thickness is 10 mm and grain size does not change between lighter and darker layers

between 71 and 77.5 ambsf (Expedition 347 Scientists 2015). At ~74 ambsf, the content

is generally all clay and there is no change in grain size between winter and summer layers

(Expedition 347 Scientists 2015). This might be due to generally cooler conditions and

glacier re-advance which can lead to decrease of meltwater and sediment availability and

be related to a lack of turbidites. The formation of varves is related to the forming of

freshwater basin at the changing ice sheet margin (Stroeven et al. 2015a). According to

Andrén et al. (2015b), varved internal grading of rhythmically banded clays indicate lake

deposits. Continuous varved clay that are overlain by silt and sand layers were formed

during the regression of the BIL (Björck and Möller 1987).

Sediment changes to dark grey clay with sandy and silty interlaminas upward from 71

ambsf (Expedition 347 Scientists 2015). The exposed till after deglaciation offers an

ample source of clay (MacLeod et al. 2006). The upward decreasing grain size indicates

a low-energy and depositional environment, and sediments were most likely deposited by

suspension rainout from meltwater or from the water column in ice-distal setting. Thicker

sand laminae occurring at ~69 ambsf might be due to larger-scale event, such as ice sheet

outwash. Increased sediment supply can result in rapid sediment accumulation which

Page 50: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

49

increases the clay content (Andrén et al. 2002). This change in sediment content in subunit

VIb possibly indicates enhanced meltwater and material input caused by warmer climate

and deglaciation. Massive interbed at ~68 ambsf and homogeneous parts between 66 and

67.7 ambsf indicate rapid deposition possibly through meltwater flows. A rapid warming

event occurred at 11.5 ka (Borzenkova et al. 2015). The fast retreat of the SIS and the

relative sea level changes caused a rapid regression in the recently deglaciated regions,

which led to increasing sedimentation rate (Andrén et al. 2011). The sedimentation rate

was ~0.4–0.55 mm year–1 at 11.0 ka and decreased to ~0.4 mm year–1 at 10.0 ka (Obrochta

et al. 2017).

According to Harff et al. (2011), homogeneous sediments replaced by laminated

sequences is an indication of the transition from freshwater to brackish-marine

environment. According to Andrén et al. (2015b), convolute bedding indicates that the

clay deposits were possibly remobilized in a slump event. The accumulation of contorted

layers can relate to gravitational slump events which reworked and redeposited the

sediments down-slope (e.g. Forsberg et al. 1999, Streuff et al. 2017b). Folded and faulted

clays and silts might be a result of icebergs settling into the soft bottom mud during the

declining of the lake level, and icebergs can supply detritus of all size and add it to the

sediments on the lake floor, deforming the pre-existing laminaes (Teller 2013). The

sediment between 48 and 50 ambsf is grey clay with intraclasts, which can be formed by

debris flows.

5.3.3. Correlation

In their climate models Rinterknecht et al. (2006) suggest that changes in the North

Atlantic climate affected the mass balance of the SIS, due to its location downwind of the

North Atlantic Ocean. Temperature fluctuations of only ±1°C, and increases in total

precipitation can lead to significant changes in ice sheet evolution (Patton et al. 2016).

Between 14.5 and 10.0 ka, the SIS melt-out contributed ~8m to global sea-level rise

(Cuzzone et al. 2016). The final drainage of the BIL led to rapid release of fresh water

into the North Atlantic, which possibly caused a minor temporary disturbance in its

circulation (Andrén et al. 2002). The geographical location of the BSB serves as a link to

Page 51: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

50

North Atlantic marine records and Greenland ice cores and, also between the north-

western European and the Eurasian terrestrial records (Expedition 347 Scientists 2014d).

The comparison of the climatic information from the Greenland ice core to the retreat

rates of the ice sheet display a general correspondence between climatic forcing and

retreat rate (Stroeven et al. 2015b). The variability between iceberg-dominated and

meltwater-dominated glacimarine sedimentation might not only be related to glaciology

and climate, but the sedimentation can also be dependent on local factors such as seafloor

topography, glacier size and distance to the ice margin (e.g. Ó Cofaigh et al. 2001, Streuff

et al. 2017b). Glacimarine sediments of fjords deposited during the deglaciation of the

Svalbard-Barents Sea Ice Sheet (Streuff et al. 2017a) show similar principal sedimentary

processes as during the deglaciation of the SIS. Sandy turbidites formed by gravitational

mass flows in ice-proximal setting and weakly stratified mud deposited in ice-distal areas

by suspension settling from meltwater (Streuff et al. 2017a).

Varved clays deposited during the deglaciation in the BIL were connected to the existing

chronology of ice recession and depositional environment of southern Sweden. The

coastal area and valley below the highest shoreline were dominated by varved clays,

littoral sand and glaciofluvial gravel. Above the highest shoreline, areas of transverse and

hummocky moraine were formed from Late Weichselian glaciofluvial sediments. During

the retreat of the ice sheet, the areas above the highest shoreline were drained by large

glacial river systems. The river systems carried clay and silt into the BIL, and these annual

deposits of varved clay draped the bedrock, till and subaqueous glaciofluvial deposits

(Wohlfarth et al. 1994.).

Sedimentary sequences from the Sokli basin In Finnish Lapland represents the variations

of the interglacial/glacial cycle (Helmens et al. 2000). In comparison to Units VI and V,

the sediment sequences from the Sokli basin show also the upward fining of the sediments

overlaying diamict deposits. These finer sediment sequences from the Sokli basin are

interpreted as deglaciation sequences (Helmens et al. 2000).

Page 52: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

51

According to Ignatius et al. (1981), the ice sheet had retreated far north at the end of the

YS stage, and today's BSB was mostly deglaciated, except the Bothnian Bay. Therefore,

varved glacial clay deposit are found in the Bothnian Bay, and postglacial sedimentation

occurs in the southern and central parts of the basin (Ignatius et al. 1981).

Postglacial deposits from Otokomi Lake in Montana have similar features like the

deposits from the Landsort Deep. The trend of LOI values might be due to coarser

sediment, and the parallel trends of LOI and water content could indicate initial

postglacial warming, which was interrupted by cooling and followed by warming shortly

after (MacLeod et al. 2006). A relatively shallow thermocline might had been a significant

distribution mechanism pending long-term periods of ice cover (MacLeod et al. 2006).

6. CONCLUSIONS

Grain size analysis enables the interpretation of depositional conditions, sediment

provenance and transport history, and sediment grain size is a common proxy of climatic

changes in studies. The determination of the stratigraphic characteristics of the sediments

from Units VI and V from the Landsort Deep enabled the interpretation of the late and

postglacial stage of the BSB. The relative age determination of the sediments connects

them to the BIL stage and the onset of the transition into the brackish YS stage, roughly

approximated between 13.5 and 11.5 ka.

The sedimentation of the Landsort Deep is characterized by low- and high-energy

transport and distribution mechanisms. High-energy mechanism, such as ice rafting and

mass movements, has led to coarser sand and clast deposits. Clay and silt deposits are

likely formed by low-energy mechanisms. The lower ice-rafted part of Unit V

demonstrates a deposition environment of an ice-proximal setting, and the upper part

displays ice-distal facies types. The coarser ice-proximal varved sediments are changed

at ice-distal areas into thinner varved clays and eventually into clayey and silty sediments.

Page 53: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

52

The sediments from the upper part of Unit V have probably deposited during the final

drainage of the BIL at ~11.7 ka and the beginning of the transition into the YS stage.

Timing and chronology of the SIS retreat and the different stages of the BSB vary and are

debated due to different dating methods. Different glaciological and climatic settings

affect the glacimarine sedimentary processes, and their variability is also caused by local

controls e.g. the topography of the seafloor and the distance to the ice margin. The wide

range of the deposited sediments of the BSB is due to variations in the water exchange

between the BSB and the world ocean, the prevailing climatic conditions and the

proximity of the retreating ice sheet.

Page 54: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

53

REFERENCES

Andrén, T., Björck, J., and Johnsen, S. 1999. Correlation of Swedish glacial varves with the

Greenland (GRIP) oxygen isotope record. Journal of Quaternary Science, 14, 4, 361–371.

Andrén, E., Andrén, T., and Kunzendorf, H. 2000a. Holocene history of the Baltic Sea as a background for assessing records of human impact in the sediments of the Gotland Basin. The

Holocene, 10, 687–702.

Andrén, E., Andrén, T., and Sohlenius, G. 2000b. The Holocene history of the southwestern Baltic Sea as reflected in a sediment core from the Bornholm Basin. Boreas, 29, 233–250.

Andrén, T., Lindeberg, G., and Andrén, E. 2002. Evidence of the final drainage of the Baltic Ice

Lake and the Brackish phase of the Yoldia Sea in glacial varves from the Baltic Sea. Boreas, Vol. 31, 226–238.

Andrén, T., Björck, S., Andrén, E., Conley, D., Zillén, L., and Anjar, J. 2011. The Development

of the Baltic Sea During the Last 130 ka. In: Harff, J., Björck, S., Hoth, P. (eds.). The Baltic

Sea Basin. Springer, Berlin, 75–97. Andrén, T., Jørgensen, B.B., Cotterill, C., Green, S., and the IODP expedition 347 scientific party.

2015a. IODP expedition 347: Baltic Sea basin paleoenvironment and biosphere. Scientific

Drilling, 20, 1–12. Andrén, T., Jørgensen, B.B., Cotterill, C., Green, S., Andrén, E., Ash, J., Bauersachs, T., Cragg,

B., Fanget, A.-S., Fehr, A., Granoszewski, W., Groeneveld, J., Hardisty, D., Herrero-Bervera,

E., Hyttinen, O., Jensen, J.B., Johnson, S., Kenzler, M., Kotilainen, A., Kotthoff, U., Marshall,

I.P.G., Martin, E., Obrochta, S., Passchier, S., Quintana Krupinski, N., Riedinger, N., Slomp, C., Snowball, I., Stepanova, A., Strano, S., Torti, A., Warnock, J., Xiao, N. and R. Zhang.

2015b. Site M0063. In: Andrén, T., Jørgensen, B.B., Cotteril, C., Green, S., and the Expedition

347 Scientists. Proceedings of the Integrated Ocean Drilling Program, Volume 347. Bennett, M. R., and Glasser, N. F. 2009. Glacial geology, ice sheets and landforms. Wiley-

Blackwell, UK, p. 385.

Bergsten, H., and Nordberg, K. 1992. Late Weichselian marine stratigraphy of the southern Kattegat, Scandinavia: evidence for drainage of the Baltic Ice Lake between 12,700 and 10,300

years BP. Boreas, 21, 223–252.

Berndmeyer, C., Thiel, V., Schmale, O., Wasmund, N., and Blumenberg, M. 2014. Biomarkers

in the stratified water column of the Landsort Deep (Baltic Sea). Biogeosciences, 11, 7009– 7023.

Björck, S., and Möller, P. 1987. Late Weichselian Environmental History in Southeastern Sweden

during the Deglaciation of the Scandinavian Ice Sheet. Quaternary Research, 28, 1–37. Björck, S. 1995. A Review of the History of the Baltic Sea, 13.0–8.0 ka BP. Quaternary

International, 27, 19–40.

Blott, S.J. and Pye, K. 2001. GRADISTAT: a grain size distribution and statistics package for the analysis of unconsolidated sediments. Earth Surface Processes and Landforms, 26,

1237– 1248.

Borzenkova, I., Zorita, E., Borisova, O., Kalniņa, L., Kisielienė, D., Koff, T., Kuznetsov, D.,

Lemdahl, G., Sapelko, T., Stančikaite, M., and Subetto, D. 2015. Climate Change During the Holocene (Past 12,000 Years). In: The BACC II Author Team (eds.), Second Assessment of

Climate Change for the Baltic Sea Basin. Regional Climate Studies. Springer, Cham, p.

25– 49. Brunnberg, L. 1988. Diatoms in glacial clay in the Södertörn area, Stockholm region. In: Miller,

U. and Robertsson, A.-M. (eds.), Proceedings of Nordic Diatomist Meeting. University of

Stockholm, Department of Quaternary Research Report, 12, 7–9.

Böttcher, M.E., and Lepland, A. 2000. Biogeochemistry of sulphur in a sediment core from the west-central Baltic Sea: Evidence from stable isotopes and pyrite textures. Journal of Marine

Systems, 25, 299–312.

Cuzzone, J.K., Clark, P.U., Carlson, A.E., Ullman, D.J., Rinterknecht, V.R., Milne, G.A., Lunkka, J-P., Wohlfarth, B., Marcott, S.A., and Caffee, M. 2016. Final deglaciation of the Scandinavian

Page 55: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

54

Ice Sheet and implications for the Holocene global sea-level budget. Earth and Planetary

Science Letters 448, 34–41. Expedition 347 Scientists. 2014a. Carbon and sulphur content (raw data) of IODP Hole

347- M0063C. https://doi.org/10.1594/PANGAEA.839492.

Expedition 347 Scientists. 2014b. Visual core description and core images IODP Hole 347- M0063C. https://doi.org/10.1594/PANGAEA.837860.

Expedition 347 Scientists. 2014c. Physical properties of IODP Hole 347-M0063C.

https://doi.org/10.1594/PANGAEA.838479.

Expedition 347 Scientists. 2014d. Paleoenvironmental evolution of the Baltic Sea Basin through the last glacial cycle. Integrated Ocean Drilling Program.

https://doi:10.2204/iodp.pr.347.2014.

Expedition 347 Scientists. 2015. Description of lithologic section units of IODP Hole 347- M0063. https://doi.org/10.1594/PANGAEA.843039.

Forsberg, C. F., Solheim, A., Elverhøi, A., Jansen, E., Channell, J.E.T., and Andersen, E.S. 1999.

The Depositional environment of the western Svalbard margin during the Late Pliocene and the Pleistocene: Sedimentary facies changes at Site 986. In: Raymo, M.E., Jansen, E., Blum,

P., and Herbert, T.D. (Eds.). Proceedings of the Ocean Drilling Program, Scientific Results,

Vol. 162.

Harff, J., Endler, R., Emelyanov, E., Kotov, S., Leipe, T., Moros, M., Olea, R., Tomczak, M., and Witkowski, A. 2011. Late Quaternary Climate variations Reflected in Baltic Sea Sediments.

In: Harff, J., Björck, S., Hoth, P. (eds.). The Baltic Sea Basin. Springer, Berlin; 99–132.

Heinsalu, Atko., Veski, Siim., Vassiljev, J. 2000. Palaeoenvironment and shoreline displacement on Suursaari Island, the Gulf of Finland. Bulletin of the Geological Society of Finland. 72, 1–

2, 21–46.

Heiri, O., André, F.L., and Lemcke, G. 2001. Loss on ignition as a method for estimating organic and carbonate content in sediments: reproducibility and comparability of results. Journal of

Paleolimnology, 25, 101–110.

Helmens, K.F., Räsänen, M.E., Johansson, P.W., Junger, H., and Korjonen, K. 2000. The Last

Interglacial-Glacial cycle in NE Fennoscandia: a nearly continuous record from Sokli (Finnish Lapland). Quaternary Science Reviews, 19, 1605–1623.

Houmark-Nielsen, M., and Kjær, K.H. 2003. Southwest Scandinavia, 40–15 ka BP:

palaeogeography and environmental change. Journal of Quaternary Science, 18, 769–786. Hughes, A.L.C., Gyllencreutz, R., Lohne, Ø.S., Mangerud, J., and Svendsen, J.I. 2015. The last

Eurasian ice sheets – a chronological database and time-slice reconstruction, DATED-1.

Boreas, 45, 1–45.

Hyttinen, O., Kotilainen, A., Salonen, V-P. 2011. Acoustic evidence of a Baltic Ice Lake drainage debrite in the Northern Baltic Sea. Marine Geology, 284, 139–148.

Ignatius, H., Axberg, S., Niemistö, L., and Winterhalter, B. 1981. Quaternary geology of the

Baltic Sea. In: Voipio, A. (ed.). The Baltic Sea. Elsevier Scientific Publishing Company, the Netherlands, 418 p.

Kaskela, A.M., Kotilainen, A.T., Al-Hamdani, Z., Leth, J.O., Reker, J. 2012. Seabed geomorphic

features in a glaciated shelf of the Baltic Sea. Estuarine, Coastal and Shelf Science, 100, 150– 161.

Lambeck, K., Smither, C., Johnston, P. 1998. Sea-level change, glacial rebound and mantle

viscosity for northern Europe. Geophysical Journal International, 134, 102–144.

Lambeck, K., Purcell, A., Zhao, J., and Svensson, N-O. 2010. The Scandinavian Ice Sheet: from MIS 4 to the end of the Last Glacial Maximum. Boreas, 39, 410–435.

Lepland, A., and Stevens, R.L. 1998. Manganese authigenesis in the Landsort Deep, Baltic Sea.

Marine Geology, 151, 1–25. MacLeod, D.M., Osborn, G., and Spooner, I. 2006. A record of post-glacial moraine deposition

and tephra stratigraphy from Otokomi Lake, Rose Basin, Glacier National Park, Montana.

Canadian Journal of Earth Sciences, 43, 447–460. Myrberg, K., Leppäranta, M., and Kuosa, H. 2006. Itämeren fysiikka, tila ja tulevaisuus.

University Press, Helsinki, 202 p.

Page 56: MSc thesis PHYSICAL PROPERTIES OF GLACIAL SEDIMENTS …

55

Obrochta, S.P., Andrén, T., Fazekas, S.Z., Lougheed, B.C., Snowball, I., Yokoyama, Y., Miyairi,

Y., Kondo, R., Kotilainen, A.T., Hyttinen, O., and Fehr, A. 2017. The undatables: Quantifying uncertainty in a highly expanded Late Glacial-Holocene sediment sequence recovered from

the deepest Baltic Sea basin—IODP Site M0063. Geochemistry Geophysics Geosystems, 18,

3, 858–871. Ó Cofaigh, C., Dowdeswell, J.A., Grobe, H. 2001. Holocene glacimarine sedimentation, inner

Scoresby Sund, East Greenland: the influence of fast-flowing ice-sheet outlet glaciers. Marine

Geology, 175, 1–4, 103–129.

Patton, H., Hubbard, A., Andreassen, K., Winsborrow, M., Stroeven A.P. 2016. The build-up, configuration, and dynamical sensitivity of the Eurasian ice-sheet complex to Late

Weichselian climatic and oceanic forcing. Quaternary Science Reviews, 153, 97–121.

Patton, H., Hubbard, A., Andreassen, K., Auriac, A., Whitehouse, P.L., Stroeven, A.P., Shackleton, C., Winsborrow, M., Heyman, J., Hall, A.M. 2017. Deglaciation of the Eurasian

ice sheet complex. Quaternary Science Reviews 169, 148–172.

Rinterknecht, V.R., Clark, P.U., Raisbeck, G.M., Yiou, F., Bitinas, A., Brook, E.J., Marks, L., Zelčs, V., Lunkka, J-P., Pavlovskaya, I.E., Piotrowski, J.A., and Raukas, A. 2006. The Last

Deglaciation of the Southeastern Sector of the Scandinavian Ice Sheet. Science, 311,

1449– 1452.

Smith, J. 2003. Aspects of the loss-on-ignition (LOI) technique in the context of clay-rich, glaciolacustrine sediments. Geogr. Ann, 85 A (1), 91–97.

Streuff, K., Cofaigh, C.Ó., Noormets, R., and Lloyd, J.M. 2017a. Submarine landforms and

glacimarine sedimentary processes in Lomfjorden, East Spitsbergen. Marien Geology, 390, 51–71.

Streuff, K., Cofaigh, C.Ó., Hogan, K., Jennings, A., Lloyd, J.M., Noormets, R., Nielsen, T.,

Kuijpers, A., Dowdeswell, J.A., Weinrebe, W. 2017b. Seafloor geomorphology and glacimarine sedimentation associated with fast-flowing ice sheet outlet glaciers in Disko Bay,

West Greenland. Quaternary Science Reviews, 169, 206–230.

Stroeven, A.P., Heyman, J., Fabel, D., Björck, S., Caffee, M.W., Fredin, O., and Harbor, J.M.

2015a. A new Scandinavian reference 10Be production rate. Quaternary Geochronology 29, 104–115.

Stroeven, A.P., Hättestrand, C., Kleman, J., Heyman, J., Fabel, D., Fredin, O., Goodfellow, B.W.,

Harbor, J.M., Jansen, J.D., Olsen, L., Caffee, M.W., Fink, D., Lundqvist, J., Rosqvist, G.C., Strömberg, B., and Jansson, K.N. 2015b. Deglaciation of Fennoscandia. Quaternary Science

Reviews, 147, 91–121.

Teller, J.T. 2013. Subaquatic Landsystems: Large Proglacial Lakes. In Evans, D.J.A. (edit.)

Glacial Landsystems. Routledge, New York, 532 p. Vassiljev, J., and Saarse, L. 2013. Timing of the Baltic Ice Lake in the eastern Baltic. Bulletin of

the Geological Society of Finland, 85, 9–18.

Winterhalter, B. 1992. Late-Quaternary Stratigraphy of Baltic sea Basins – a review. Bull. Geol. Soc. Finland 64, Part 2, 189–194.

Wohlfarth, B., Björck, S., Holmqvist, B., Lemdahl, G., and Ising, J. 1994. Ice recession and

depositional environment in the Blekinge archipelago of the Baltic Ice Lake. GFF, 116:1, 3– 12.