Вы находитесь на странице: 1из 303

Comparative study of co-seismic sedimentation in two

tectonically active areas : the sea of Marmara and the


gulf of Corinth : methodological developments,
implication for seismic hazards assessment
Corina Campos Serrano

To cite this version:


Corina Campos Serrano. Comparative study of co-seismic sedimentation in two tectonically ac-
tive areas : the sea of Marmara and the gulf of Corinth : methodological developments, implication
for seismic hazards assessment. Oceanography. Université de Grenoble, 2014. English. <NNT :
2014GRENU057>. <tel-01557531>

HAL Id: tel-01557531


https://tel.archives-ouvertes.fr/tel-01557531
Submitted on 6 Jul 2017

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
THÈSE
Pour obtenir le grade de
DOCTEUR DE L’UNIVERSITÉ DE GRENOBLE
Spécialité: Sciences de la Terre, de l’Univers et de l’Environnement
Arrêté ministériel : 7 août 2006

Présentée par
Corina CAMPOS SERRANO

Thèse dirigée par Christian BECK et


Codirigée par Christian CROUZET

préparée au sein de L’Institut des Sciences de la Terre (ISTerre),


Site de l’Université de Savoie
Unité Mixte de Recherche 5275 du C.N.R.S.

Etude comparative de la sédimentation co-


sismique sur deux sites tectoniques actifs:
la Mer de Marmara et le Golfe de Corinthe.
Développements méthodologiques, apports à
l'estimation de l'aléa sismique régional.

Comparative study of co-seismic


sedimentation in two tectonically active
areas: the Sea of Marmara and the Gulf of
Corinth.
Methodological developments, implication for
seismic hazards assessment.

Thèse soutenue publiquement le 23 Juin 2014


devant le jury composé de :

Charles Aubourg
Professeur, Université de PAU, Rapporteur
Jean-Nöel Proust
Directeur de Recherche, Université de Rennes 1, Rapporteur
Namik Ça#atay
Professeur, Istanbul Technical University, Examinateur
Miguel-Angel Rodriguez-Pascua
Checheur, Instituto Geologico y Minero de Espana, Examinateur
Christian Beck
Professeur, Université de Savoie, Directeur de Thèse
Christian Crouzet
Maître de conférences, Université de Savoie, Co-encadrant
Acknowledgments

The realization of this thesis required long hours of work, it would have been
impossible to carry out without the help and support of many people and institutions; here I
would like to thank some of them.

Firstly, I want to thank my two advisors: Christian Beck and Christian Crouzet for the
help that they have given me throughout these four long years. Christian Beck thanked for
having proposed me this thesis, for having confidence in me and for listening from the first
day. Christian Crouzet thank for his full availability, for personal lecture in rock magnetism
and for helping me considerably to improve my ideas.

I wish to give an especial thankful to Pierre Sabatier to guide me in the geochemical


analyses, to François Demory for you help in the rock magnetic analyses. I am grateful to
Annie Millerie, for her collaboration in the LOI measures. I also thank to Eduardo Carrillo,
for our precise and productive discussion.

I want to thank the members of the jury, Charles Aubourg, Jean-Nöel Proust, Namik
Ça÷atay, Miguel-Angel Rodriguez-Pascua for accepting to review my work.

The available cores for the present research were taken during the MARMACORE
Cruise, as a part of the large GEOSCIENCES MD 123 program, and during the
GEOSCIENCES-II program. The investigations was funded through different grants: i)
University of Savoie's AAP-2012-16 AGRASM Grant, dedicated to AMS sedimentological
application, ii) Agence Nationale pour la Recherche ANR's SISCOR Project dedicated to the
Gulf of Corinth seismic hazards, iii) CNRS-INSU funding through Institut des Sciences de la
Terre (ISTerre) Laboratory.

I would like to thank to Universidad Simon Bolivar and specifically the department of
earth sciences, for giving me permission to conduct this thesis. I would also to thank the
Fundacion Gran Mariscal de Ayacucho and the Ambassade of France in Venezuela for the
grant 20093262 assign me during four year for my stay in ISTerre-University of Savoie. I

iii
wish to thank the ISTerre laboratory for the scholarship assigned to me to finish this
dissertation.

A mis amigos venezolanos Carlos, Sirel, Enzo, Ili, Javier y Christian Beck, por
hacerme sentir tan cerca de casa. À Fabien, Pascale, Cecile, Eva, Silvi et Ileana pour leur
amitié. Je les aime. A Monique et Réne Veillas, pour être ma famille adoptive.

A mi hijo Leonardo por enseñarme lo que es realmente importante en la vida y a mi


compañero de vida y gran amigo Oswaldo, por siempre estar ami lado, por amarme y ser
paciente. Los amo a los dos.

iv
Abstract

The Eastern Mediterranean is one of most seismically active and rapidly deforming
regions in Europe. The presence in this region of numerous lakes and silled marine basins
makes this region an ideal area to study the record of the seismo-tectonic activity and the
main climatic changes in sub-aquatic environments. In this thesis we are interested in
studying the Late Quaternary sub-aquatic paleoseismic record of two basins: The Sea of
Marmara and The Gulf of Corinth. Both are silled deep marine basins, crossed by major faults
(e.g. the North Anatolian Fault). We focused on identifying the traces left by
paleoearthquakes in the sedimentary record, particularly, the layers named here:
“homogenites+turbidites” (HmTu), which represent individual seismically-induced
sedimentary events. Their identification was performed through classical sedimentological
analyzes and magnetic properties, in particular the Anisotropy of Magnetic Susceptibility.
Analyzes were conducted on sediment cores recovered by long Calypso piston core. These
sediments represent the last 17 cal kyr BP of sedimentation, and have recorded the last non-
marine (lacustrine) to marine transition. In the Sea of Marmara, this passage was identified
around 12.8 cal kyr BP, whereas in the Gulf of Corinth was around 11.7 cal kyr BP. In both
basins, the record of the earthquake-induced instantaneous events allowed the estimation of
the average earthquake recurrence interval. In the Çinarcik Basin of the Sea of Marmara, the
minimum average recurrence interval varies between a155 and 365 yr, while for the eastern
part of the Gulf of Corinth the minimum average recurrence interval varies between a400 and
a500 yr. Finally, in the Central Basin of the Sea of Marmara, the study of correlated co-seimic
sedimentary events at opposite sides of a segment of the North Anatolian Fault allows the
estimation of the co-seismic vertical offset for a continuous period of 2 kyr. Significant values
of vertical offset were observed (up to 144 cm), implying a dominant vertical (normal) throw
for this fault segment.

v
Résumé

La partie orientale de la Méditerranée, est une des régions sismiques les plus actives
en Europe. La présence dans cette région de nombreux lacs et bassins marins isolé fait de
cette région un endroit idéal pour étudier l’enregistrement de l'activité sismo- tectonique et
les principaux changements climatiques dans des environnements subaquatiques. Dans
cette thèse, nous nous intéressons à l'étude de l’enregistrement paléosismique subaquatique
du Quaternaire tardive dans deux bassins: la Mer de Marmara et le Golfe de Corinthe. Ces
derniers sont des bassins marins isolés traversés par des failles majeures (par exemple, la
faille Nord Anatolienne). Nous nous sommes concentrés sur l'identification des traces
laissées par des paléoséismes dans l'enregistrement sédimentaire, en particulier, les
couches nommées ici: "homogenites + turbidites" (HmTu), lequel représentent des
événements sédimentaires quasi instantanés induits par des secousses sismiques. Leur
identification a été réalisée à partir des analyses sédimentologiques classiques et des
propriétés magnétiques, en particulier, l'anisotropie de la susceptibilité magnétique. Les
analyses ont été effectuées sur les sédiments carottés à l’aide du système géant à piston
CALYPSO. Ces sédiments représentent les dernières 17 cal ka BP de sédimentation, et ont
enregistré le dernier passage de conditions non marine à des conditions marines. Dans la
mer de Marmara, ce passage a été identifié autour de 12,8 cal ka BP, tandis que, dans le
golfe de Corinthe était d'environ 11,7 cal ka BP. Dans les deux bassins, l’enregistrement
des événements instantanés considérés comme induits par des séismes a permis
l’estimation d’un intervalle moyen de récurrence. Dans le bassin de Çõnarcõk de la mer de
Marmara, l'intervalle de récurrence moyenne minimale varie entre ~155 et ~365 ans, tandis
que pour la partie orientale du golfe de Corinthe, l'intervalle de récurrence moyenne
minimale varie entre ~400 et ~500 ans. Enfin, dans le bassin central de la mer de Marmara,
l'étude des événements HmTu corrélées sur les côtés opposés d'un segment de la faille
nord-anatolienne permet d'estimer le décalage vertical co-sismique pour une période
continue de 2 ka. Des valeurs significatives observées étaient des décalées verticalement
(jusqu'à 144 cm), ce qui implique un déplacement verticalement dominant (normal) pour
ce segment.

vi
List of content

Acknowledgements ……………………………………………………………………...iii
Abstract …………………………………………………………………………………..v
Résumé …………………………………………………………………………………..vi
List of contents ………………………………………………………………………..…vii

CHAPTER 1. General Introduction …………………………………………………..12


1.1. Purpose of the study ……………………………………………………………14
1.2. Overview of record of the seismicity in sediments……………………………15
1.3 Paleoseismological records in sub-aqueous sediments ………………………16
1.3.1 Direct in situ disturbances……………………………………………....16
1.3.2 Indirect reworking processes………………………………………...…19
1.3. Thesis outline …………………………………………………………………..20

CHAPTER 2. Data acquisitions and analytical tools…………………………..……...22


2.1. Introduction ……………………………………………………………………..22
2.2. Coring …………………………………………………………………………...22
2.3. Macroscopic observations……………………………………………………....24
2.3.1. Split cores observations ………………………...……………………….24
2.3.2. X-ray images ………………………………………………………….…24
2.4.Laboratory analysis ……………………………………………………………..25
2.4.1 Sediment composition…………………………………………………...25
2.4.1-a Microscopic observations (smear slides)………………………………..25
2.4.1-b Loss on ignition (LOI): Organic and carbonate contents ………………25
2.4.1-c Geochemical analysis…………………………………………………...26
2.4.1-d Bulk (or volume) Magnetic Susceptibility (MS) *……………………..26
2.4.1-e Remanent magnetizations (Isothermal and Anhysteretic remanent
magnetization)…………………………………………………………..27
2.4.2 Sediment textures analyses.……………………………………………..27
2.4.2-a Laser microgranulometry……………………………………………….27
2.4.2-b Grain shape analysis………………………………………………….....27
2.4.2-c Anisotropy of Magnetic Susceptibility (AMS)…………………………28

vii
14
2.4.3 C dating ……………………………………...………………………...28

(*) Article: Deciphering hemipelagites from homogenites through Magnetic


Susceptibility Anisotropy. Paleoseismic implications (Sea of Marmara and Gulf of
Corinth) (Sedimentary Geology 292, p. 1-14)……………….………………….……30

CHAPTER 3. The Late Pleistocene/Holocene sedimentary record of the central and


eastern Sea of Marmara (Orta and Çinarcik basins)………................43
3.1. Introduction …………………………………………………………………….43
3.2. Geodynamic and geologic overview of the Sea of Marmara region................44
3.2.1. The “Anatolia Microplate” and the North Anatolian Fault (NAF)……44
3.2.2. Tectonic and geology of Turkey …………………………………………48
3.2.3. Geology of the Marmara region ………………………………………....51
3.2.4. The North Anatolian Fault within the Sea of Marmara …………….....52
3.3. The Sea of Marmara: overview of morpho-bathymetry, active tectonic
structures, and sedimentary fill. ……………………………………………..54
3.3.1. Morpho-bathymetric and tectonic features of the Sea of Marmara.......54
3.3.1-a The Çinarcik Basin………………………….…………………………..55
3.3.1-b The Central (Orta) Basin……………………..…………………………58
3.3.2. Previous data on paleoseismic archives within the Sea of Marmara’s
sediments…………………………………………………………………60
3.3.3. The Sea of Marmara a gateway between the Aegean and the Black
Seas ……………………………………………………………………… 61
3.3.4. Present-day physical oceanography and drainage ……………………..65
3.4. The Late Quaternary sedimentary infill of the central and eastern Sea of
Marmara………………………………………….……………………………...68
3.4.1. Introduction ……………………………………………………………….68
3.4.2. Description of facies ………………………………………………………70
3.4.3. Stratigraphy of the MD01-2425 core ……………………………………74
3.4.3-a Upper unit (marine deposits) ……………………………..……….……74
3.4.3-b Lower unit (non-marine deposits) ………………………………………76
3.4.3-c The contact between the upper and lower units ……………..………….76
3.4.4 Radiocarbon dating and age-depth curve of the MD01-2425 core ….....79
3.4.4-a. 14C samples …………………………………………………………....79

viii
3.4.4-b. Correlation between cores ……………………………………………..81
3.4.4-c. Calibration of radiocarbon dating ……………………………………...86
3.4.4-d. Age-depth curve of the MD01-2425 core ……………………………..86
3.4.4-e. Main results ……………………………………………………………90
3.5. The late Quaternary record of the seismotectonic activity in the sediments
of the Sea of Marmara’s eastern basin (core MD01-2425)….…….............…93
3.5.1. Introduction ……………………………………………………………….93
3.5.2. Textural analysis of the MD01-2425 core …………………………….....93
3.5.2-a Grain Size ………………………………………………………………93
3.5.2-b Base_to-top binary diagrams: Skewness vs Sorting and CM Passega’s
diagrams…………………………………………………………………99
- Upper unit (marine) ………...……………………………………….99
- Lower unit (non-marine) ……………………………………………100
3.5.2-c Particles shape of silty-sandy layers……………………………………104
3.5.2-d Magnetic fabric………………………………………………………....106
3.5.2-e Magnetic content……………………………………………………….110
3.6.Estimation of organic matter and carbonate content in the MD01-2425 core
(Loss on ignition) …..............................................................................................112
3.7. Data interpretation and conclusion of depositional processes in the MD01-
2425 ………………………………………………............................................114
3.7.1. Depositional processes……………………………………………………114
- Facies Mh. and Ls ……………………………………………………....114
- Facies Tuc ……………………………………………………………....115
- Facies HmTu …………………………………………………………...115
3.7.2. Depositional model for the HmTu events ……………………………....117
3.7.3. Possible triggering mechanism of sedimentary HmTu events………...120
3.7.4 Time distribution of the sedimentary HmTu events in the MD01-2425
core ………………………………………………………………………123
3.8. Correlation between the Çinarcik and Central basins through the MD01-
2425, MD01-2429 and MD01-2431 cores ………………………………….…128
3.9. Identification and measurement of the co-seismic fault offset along the
North Anatolian Fault in the Central Basin through the co-seismic
sedimentary episodes ………………………………………............................130

ix
(*) Article: Estimation of successive co-seismic vertical offsets using coeval
sedimentary events. Application to the Sea of Marmara’s Central Basin
(North Anatolian Fault). Natural Hazards and Earths System Science……..........131
3.10. Conclusions of chapter ………………………………………………….........157

CHAPTER 4. The Late Pleistocene/Holocene sedimentary record of the Gulf of


Corinth ………………………………………………………………..............................160
4.1. Introduction …………………………………………………………….……....160
4.2. Geological and tectonic setting ………………………………………………..161
4.3. The Gulf of Corinth ……………………………………………………………167
4.3.1. Seismic activity in the Gulf of Corinth …………………………………173
4.3.2. Oceanography and Quaternary sedimentation in the Gulf of Corinth 174
4.4. The Late Quaternary sedimentary infill of the Gulf of Corinth ……………177
4.4.1. Lithologic description of the MD01-2477, MD01-2478 and MD01
2481 cores …………………………………………………………….…177
4.4.1-a Muddy facies …………………………………………………………..178
4.4.1-b Coarser facies ………………………………………………………….180
4.5. Stratigraphy of the MD01-2477 and MD01-2481 cores ………….………….183
4.6. Chronostratigraphic framework ……………….……………………………..185
4.6.1. Analytical results ………………………………………………………...185
4.6.2. Age-depth curve ………………………………………………………….187
4.6.3. Mains results ……………………………………………………………..188
4.7. Magnetic properties of sediments of the MD01-2477, MD01-2478 and
MD01-2481 cores ……………………………………………………………….192
4.7.1. Magnetic Susceptibility (MS) …………………………………………...192
4.7.2. Anisotropy of Magnetic Susceptibility (AMS) …………………………193
4.7.2-a Anisotropy of Magnetic Susceptibility of the MD01-2477 core ……...194
- Results ……………………………………………………………….…194
- Interpretations ………………………………………………………….196
4.7.2-b Anisotropy of Magnetic Susceptibility of the MD01-2481 core ……...196
- Results ………………………………………………………………….196
- Interpretations ………………………………………………………….198
4.7.3. Magnetic content and their possible influence on magnetic fabric in the
MD01-2477 core …..………………………………………………….....200

x
4.8. High-resolution mineral geochemistry of the MD01-2477core……………...203
4.8.1. High-resolution geochemical analyses of the marine to non-marine
transition in the MD01-2477core ………………………………………..203
4.8.2. High-resolution mineral geochemistry of instantaneous events of the
MD01-2477 core …………………………………………………………..206
4.9. Estimation of organic matter and carbonate content in the MD01-2477 core
(Loss on ignition)………………………………………………………………...210
4.10. Interpretation of depositional processes in the MD01-2477 core through
sedimentary texture analysis ………………………………………….……….213
4.10.1. Results ……………...…………………………………………………....213
4.10.1-a Grain size ………………………………………………………….….213
4.10.1-b Skewness vs sorting and CM Passega’s diagrams ……………….…...216
4.10.1-c Particles shape ………………………………………………….……..221
4.10.2. Interpretation and conclusion of depositional processes in the MD01-
2477 core ……………………………………………………………...….223
4.11. (*) Article: Late Quaternary paleoseismic sedimentary archive from deep
central Gulf of Corinth: time distribution of inferred earthquake-induced layers.
ANNALS OF GEOPHYSICS, 56, 6, 2013, S0670; doi:10.4401/ag-6226……………………..…226

4.12. Conclusions of chapter ………………………………………………….….…242

CHAPTER 5. Conclusions …………………………………………………..……….…245

References …………………………………………………………………………….…250

Annex A……………………………………………………………………………….….284
Annex B……………………………………………………………………………….….285
Annex C……………………………………………………………………………….….290
Annex D……………………………………………………………………………….….294
Annex E……………………………………………………………………………….….297
Annex F…………………………………………………………………………………..299

xi
CHAPTER 1

General Introduction

The Eastern Mediterranean region, including the surrounding areas as western Turkey
and Greece, is one of the most seismically active and rapidly deforming regions in Europe,
with an active and rapid deformation within continents. This region has been intensively
studied, because it is a key area for understanding the main tectonic processes such as:
continental collision (e.g. Bitlis-Zagros, Caucasus), subduction of oceanic lithosphere and
back-arc extension (e.g. Cyprus arc, Hellenic arc, Aegean Sea), continental extension (e.g.
western Turkey, Sea of Marmara, Gulf of Corinth, Saronic Gulf), and intracontinental
transform faults (e.g. North and East Anatolian Faults, Dead Sea Fault). It is a key area for
understanding the causes and distribution of seismic events, and their impact on social life and
civilization (Dewey and ùengör, 1979; Dixon and Robertson, 1984; Robertson, 1998; Aksu
and Yaltrak, 2005; Taymaz et al., 2007).

The Eastern Mediterranean region is the product of diachronic collisions since the Late
Cretaceous. These collisions are the results of the partial consumption of the Neotethyan
oceanic crust through a complex pattern of subduction zones and collisions, leading to the
development of the Alpine Mountains across Southeastern Europe, and the evolution of the
small, elongated land-locked basin: the Eastern Mediterranean Sea (Aksu and Yaltrak, 2005).

The present-day tectonic framework of the Eastern Mediterranean is controlled by the


collision between the African and Eurasian plates, and the displacements of the Arabian and
Aegean – Anatolian Microplates (figure I.1). The final collision between the Arabian
Microplate and the Eurasian Plate in the Midle Miocene (25 Ma ago) produced the westward
tectonic escape of the Aegean–Anatolian Microplate. This escape occurs along two intra-
continental transform faults: the North Anatolian and East Anatolian Transform Faults
(Dewey et al., 1986, Aksu and Yaltrak, 2005; Taymaz et al., 2007) (figures I.1). The North
Anatolian Transform Fault has a right lateral strike slip displacements (24 r 2 mm/yr), while
the East Anatolian Transform Fault shows a complementary left lateral motion (9 r 2 mm/yr)
(Reilinger et al., 1997). The relative westward displacement (30 r 2 mm/yr) of the Aegean–

12
Anatolian Microplate and its counterclockwise rotation is driven by the north–northwest
pushing of the Arabian Microplate, as well as by the pulling of the subducting African Plate
beneath the Hellenic Arcs (Reilinger et al., 1997; McClusky et al., 2000). The boundary
between the African Plate and the Aegean–Anatolian Microplate corresponds to the Hellenic
and Cyprus Arc, and the Cephalonia and Paphos Transform faults.

Figure I.1: Active tectonic map of the Eastern Mediterranean region (adapted from Okay et al.,
2000). Lines with filled triangles show active subduction zones, lines with open triangles are
active thrust faults at continental collision zones, and lines with tick marks are normal faults. The
large solid arrows indicate the sense of motion of the lithospheric plates. EAF: East Anatolian
Fault. The studied areas are framed in a black box for the Sea of Marmara, and red box for the
Gulf of Corinth.

The complexity of the plate interactions and associated crustal deformation in the
Eastern Mediterranean region is reflected in many destructive earthquakes that have occurred
throughout its recorded history. This high seismic activity, as well as the presence of
numerous basins, makes the Eastern Mediterranean an ideal area to study the effects of
seismicity in the sub-aqueous sedimentation. Therefore, we will focus the present work in the
study of the sedimentary paleoseismic record in two basins located in this region (figure I.1):
 The Sea of Marmara (figure I.1), a marine deep basin located in north-western
Turkey, which is crossed by the North Anatolian Fault, one of the most active and

13
important structures of the Eastern Mediterranean region (e.g. Le Pichon et al., 2001;
Armijo et al., 2002; Fraser et al., 2010). The geometry of this fault in the Marmara
region is an important matter of debate (e.g. Parke et al., 1999; Okay et al., 2000; Le
Pichon et al., 2001; Imren et al., 2001; Armijo et al., 2002), because it has a direct
influence in the estimation of the seismic hazard in the region.
 The Gulf of Corinth, (figure I.1), a semi-enclosed deep marine basin, situated in
central Greece, within one of the most actively extending and highly seismic regions
in the world (e.g. Armijo et al., 1996; Briole et al., 2000). It is an active intra-
continental rift bounded by E-W striking, en échelons faults located onshore and
offshore (Brooks and Ferentinos, 1984; Armijo et al., 1996; Sakellariou et al., 2001).

Both regions underwent a significant amount of destructive large earthquakes during


the last 2.5 kyr, as revealed by the historical sources (e.g. Galanopoulos., 1953;
Papadopoulos, 1998; Ambraseys and Jackson, 1990; Papazachos and Papazachou, 1997;
Ambraseys, 2002). This data has been completed by paleoseismological studies conducted in
numerous trenches and submarines researches, which provide information about historic and
prehistoric record and show the potential of the region for paleoseismic recording (e.g. Collier
et al., 1998; Koukouvelas et al., 2001; Rockwell et al., 2001; Klinger et al., 2003; Pantosti et
al 2004, McHugh et al., 2006; Sari and Ça÷atay, 2006, Ça÷atay., et al 2012; Avúar, 2013). In
addition, both basins have similar morphologic characteristics (size, bathymetry and
watershed), and are undergoing the same climatic regimes. They also recorded the last Late
Quaternary climatic cycles. Thus the repective paleoseismological archive can be compared.

1.1 Purpose of the study

In the present work we attempt to decipher the Late Quaternary paleoseismic records
of the Sea of Marmara and Gulf of Corinth, through the analysis of sediments from long
piston cores. We are interested in identifying the traces left by paleoearthquakes in the
sedimentary record through classical sedimentological analyzes, as well as by the application
of the Anisotropy of Magnetic Susceptibility. Our interest is the reconstruction of the “seismic
signal” of the past time; resulting in the deunitination of the earthquake recurrence interval, as
well as in the estimation of the co-seismic offsets of the subaqueous active faults.
Additionally, as both basins suffered environmental changes during the last high-stand phase
associated to the Holocene glacio-eustasy, we are interested in the study of the last change

14
from non-marine to marine environment and its impact on the paleoseismic record. For these
reasons, a brief overview of record of the seismicity in sediments is discussed in the
following.

1.2 Overview of record of the seismicity in sediments

Paleoseismology is the search of preserved geological traces of significant/major


earthquakes, especially for remote periods lacking human testimony and archives.
Archeological traces may represent precious additional records. Paleoseismology aims to
precise the location and the age of paleo-earthquakes, and also attempts to estimate paleo-
intensities or paléo-magnitudes. The paleoseismologists interpret geologic evidence created
during individual paleoearthquakes, focusing in the instantaneous deformation of landforms
and sediments during earthquakes (McCalpin, 2009). Paleoseismological analysis relies on
the premise that earthquakes produce permanent and recognizable effects on the natural and
human environments, being only the paleoseismologists capable to study the earthquakes that
produce recognizable deformations (McCaplin, 2009, Michetti and Hancock, 1997).
Fingerprints of paleoearthquakes can be found in sedimentary archives (on land and offshore),
archeological relics, cave speleothem records, and dendrological records (McCaplin, 2009;
Avúar, 2013); being the sedimentary archives the most widely studied.

On land, the most commonly applied and systematically developed technique is on-
fault trenching. (McCaplin, 2009). This technique investigates structural, geomorphological
and sedimentological evidence of surface fault rupturing (Avúar, 2013). The basic motivation
behind this technique is to detect the stratigraphical event-horizons within fault-related
sedimentary sequence, and to estimate their approximate timing by using Quaternary dating
methods. If the characteristics of an event-horizon unambiguously indicate an earthquake
origin, then it can be further interpreted as an earthquake horizon (McCaplin, 2009). Trenches
across faults are typically sited to collect data on either paleoearthquake displacement or
paleoearthquake recurrence (Sieh, 1981). According to McCaplin (2009) the best places for
measuring displacement are those where all displacement is concentrated on a single, narrow
fault strand, and subsidiary faulting and folding are negligible. While, the best places for
measuring recurrence are local fault-zone depressions filled with fine-grained and/or organic
interfaulting sediments. Trench location is also deunitined by the number of paleoearthquakes
that the investigator wishes to observe. Trenches across faults on very young Quaternary

15
surfaces may expose only one or two paleoearthquake displacements. Whereas, trenches on
progressively older surfaces probably expose the cumulative deformation from many
paleoearthquakes.

One of the main objectives of the present work is to reveal the effects of seismicity in
the sub-aquatic sedimentation. Thus, the following paragraphs will be dedicated to summarize
some of the different sedimentary perturbation induced by earthquakes.

1.3 Paleoseismological records in sub-aqueous sediments

A variety of sedimentary perturbations may be induced by earthquakes. According to


Chapron et al (1999) they can be classified into two groups: in situ direct post-depositional
disturbances (seismites s.s.) and the indirect effects through gravity reworking (seismites s.l.).
The unit seismites s.s. was originally used by Seilacher (1969) to describe the stratigraphic
units containing sedimentary structures produced by shaking.

1.3.1 Direct or in situ disturbances.


They result from the action of seismic waves on in situ deposits, and involve soft and
brittle deformations which varies depending on the rheology of the materials affected
(Chapron et al., 1999). The soft deformations are classically the structures due to
liquefaction (e.g. sand volcanoes, diapirs, boudinage, convolute, ball-and pillow structures).
While, the brittle deformations can be observed in units of microfaults (e.g. Seilacher, 1969,
Monecke et al., 2004; Beck, 2009).

 The liquefaction is the transformation of a granular material from a solid state into a
liquefied state as a consequence of increased pore-water pressure (Youd, 1973;
Obermeier et al., 2001). Worldwide data on historical earthquakes show that the
minimum earthquake magnitude to form liquefaction features in most field settings is
about Mw: 5, but a magnitude of about Mw: 5.5 to 6 is the lower limit at which
liquefaction effects become relatively common (Ambraseys, 1988, Obermeier et al.,
2001, Audemard and De Santis, 1991). Obermeier (2009) indicates that the sediment
vented to the ground surface provides the most conspicuous evidence of liquefaction at
depth (figure I.2). These values concern an onshore situation with a free phreatic
surface. A water-sediment mixture typically erupts suddenly and violently to the

16
surface, through pre-existing holes or through fractures opened in the capping material
in response to liquefaction. They flow to the surface as sand volcanoes, which have
form of cones, often called sand blows or sand boils (figure I.2). The cones of sand
can be as much as a meter in height and tens of meters in width. Other structure
associated with liquefaction processes are the clastic dikes. They are a primary source
of the data used for paleoseismic interpretations. They consists typically of sand-filled
linear fractures in cross-section that can range from tens of centimeters to several
meters in vertical extent (Audemard and De Santis, 1991; Obermeier et al., 1993,
Rodriguez - Pazcua, 2000). The sand-filled fractures are commonly rooted in a sand
bed (figure I.2) and cut confining horizons of distinct lithologies (Rodriguez - Pascua,
2000).

Figure I.2: Schematic vertical section showing dikes cutting through overbank silt and clay strata
and the overlying sand-blow deposits (from Obermeier, 2009). (A) Stratigraphic details of
sediment vented to the surface. (B) Dikes that pinch together as they ascend. (C) Characteristics
of dikes in fractured zone of weathering, in highly plastic clays.

Others types of earquake-induced soft-sediment deformations in totally subaqueous


settings - as pseudonodules or ball-and-pillow, diapirs, loop bedding, etc. - are discussed by
Montenat et al. (2007) and Rodriguez - Pascua et al. (2000), among others. The last authors
study the soft-sediment deformations interpreted as seismites in the lacustrine sediments of
the Prebetic Zone, Spain; they propose empirical estimates of the magnitudes of the

17
earthquakes that may have caused these deformations, based on their geometry and
thicknesses. In the figure I.3 are represented some of these seismites recognized by
Rodriguez - Pascua et al. (2000) ordered according to a rank of earthquake magnitudes.

Figure I.3: Synthesis of seismites identified by Rodriguez-Pascua et al, (2000) in the lacustrine
sediments of the Prebetic Zone with indication of corresponding magnitude (MAG). The authors
separate shallow and deep lacustrine sedimentation due to their differences in response to seismic
shaking.

18
1.3.2 Indirect reworking processes.
Seismic shocks may trigger different gravity driven phenomena which include
external material input (e.g. land slides or mass transport deposits / MTD, rock falls,
slumps, debris and mud flows; mass flows, grain flows, turbidites, etc.), as well as re-
suspension processes.

Seismic shaking may trigger instabilities on steep slopes of marine or lacustrine


basin. These instabilities results in the remobilization of large masses of unconsolidated
sediments at proximal sites (e.g. slide, slumps, debris flows). They can evolve into density
and turbidite currents during they basinward travel, being observed their deposits in the
distal part of the basin (seismoturbidites and/or homogenites). An example of this
phenomenon has been identified and studied in Eastern Canada, specifically in the St
Pierre island slope. In this region, the 18 November 1962 an earthquake of Ms 7.2
triggered widespread seabed slumping, which were transformed into turbidity currents that
broke transatlantic telegraph cables for up to 13 hours after the earthquake (Heezen and
Ewing, 1952; Piper et al 1999). Under stable water conditions the turbidites deposits can
show the classical fining-upward and the different units defined by Bouma (1962) or Mutti
and Ricci Lucchi (1978). Although in special conditions (restricted basins, such as lakes or
isolated marine basins), the seismic shaking induce water oscillation (reflected tsunami /
seiche effects). These water oscillation may maintain the fine grained fraction of re-
suspended sediments in suspension into the water column over days to months. It results in
a complex deposit constituted by: a coarse graded lower part in sharp contact with a highly
homogeneous fine grained upper part, named “homogenite” (refering to the concept
proposed by Kastens and Cita, 1981). A long-lasting “cloud” of fine grained particles has
been observed immediately after recent major earthquakes in Cariaco Trough (Thunell et
al., 1999) and north-west of Haiti (McHugh et al., 2011). It reinforce the earthquake-
induced interpretation proposed for some homogeneous deposit defined as “homogenite-
type” layers (e.g. Chapron et al., 1999; Beck et al., 2007; Bertrand et al., 2008).

The subaquatic gravity driven deposits have proven to be a common way to record
earthquakes, especially following the work of Adams (1990). However not all gravity
driven sediment flows necessarily have seismic origin, for example overloading and over-
steepening subaqueous slopes, storm waves, extreme flooding (hyperpycnal flows), sea
level changes, and other factors can trigger gravity driven flows. Due to this wide range of

19
triggering mechanism of gravity flows, the paleoseismologist are interested in: 1)
differentiating between earthquake and non-earthquake trigger gravity driven deposits, and
2) the identification of the responsible active structure(s). In order to address the first issue
different method can be used:
- sedimentological and mineralogical examination. Textural and compositional
analysis has allowed the identification of: 1) shock-induced stacked multipulsed
turbidites (e.g. Nakajima and Kanai, 2000; Shiki et al, 2000), with different
mineralogies (Goldfinger et al., 2007); this turbidites are caused by multiple
slump events with different mineralogic sources, evidencing synchronous
triggering of multiple parts of the basin; 2) The “homogenite-type” layers
associated to shock-induced seiche effects (e.g. Sturm et al., 1995; Chapron et
al., 1999; Beck et al; 2007).
- Identification of multiple synchronouslytriggered mass movement deposits. If
these deposits can be correlated among widely spaced sites, synchronous
deposition can be established or inferred and an earthquake triggering
mechanism is interpreted. This approach has been used for deep structures as
subduction zones (e.g. Gracia et al., 2010, Moernaut, 2011; Pouderoux et al.,
2012) or major transform faults (e.g. Goldfinger et al, 2007).
- Correlation of the seismically-induced subaquatic gravity driven deposits with
historical earthquake records and land paleoseismic data.

Finally, in order to address the second issue, different authors have recently studied
the relation between active faults at the water/sediment interface (sea or lake bottom) and
the earthquake-induced sedimentary events (e.g. Carrillo et al., 2006, 2008; Bull et al.,
2006; Barnes and Pondard, 2010; Beck et al., 2012). These studies have achieved to
identify fault offsets and the slip rates (Beck et al., 2012) through a combination of high-
resolution seismic imagery and core analyses.

1.4. Thesis outline

This study is composed of five chapters. After this introductory Chapter 1, we


present the methods applied in this study for data acquisition and interpretative concepts.
Within Chapter 2, a new methodology to identify possible earthquake-induced deposits

20
type: “Homogenites + Turbidites” is presented as an article published in Sedimentary
Geology.

Chapter 3 presents the Late Pleistocene-Holocene paleoseismic record of the Sea of


Marmara, in the Çinarcik Basin. Also, we attempt to measure the co-seismic fault offset of
the North Anatolian Fault in the Central Basin through the co-seismic sedimentary
episodes, and the main results are presented in form of an article submitted to Natural
Hazards and Earth Science Systems. In Chapter 4 we present a tentative paleoseismic
record for the Central part of the Gulf of Corinth, concerning the Late Pleistocene-
Holocene interval. Parts of the obtained results are presented in an article published in
Annals of Geophysics. Finally, Chapter 5 provides the conclusions of the thesis.

21
CHAPTER 2

Data acquisitions and analytical tools

2.1. Introduction

The studied cores were sampled during the MARMACORE Cruise onboard R/V
MARION- DUFRESNES. This coring campaign was associated to site surveying with
very high resolution seismic reflection imaging (3.5 kHz). This cruise dedicated to the Sea
of Marmara was part of the (larger) GEOSCIENCES MD 123 program. The Gulf of
Corinth cores were recovered during the GEOSCIENCES-II program. Both survey used
the same devices and non destructive onboard processing, providing adequate conditions
for comparisons between the different investigated sites.

2.2 Coring

The two surveys provided the CALYPSO giant piston-coring device. It is a unique
system, with its 7 to 10 tons weight, a 40 to 60 m iron lance (5 inches ½ diameter) and an
internal high-pressure PVC liner (12/10 cm diameter). Although based on the classical
Küllenberg piston-coring design, it is characterized by its size and a triggering of the final
free fall of the lance added by the monitoring of the cable tension. When the counterweight
linked to the release mechanism touch the bottom, it activates the mechanism and allowed
the lead weight to fall by gravity. This pushes the steel tube into the sediments and samples
a core (Figure II.1). During the extraction, inside the nose of the core catcher, a trap
("peau d'orange") keeps the sediments into the core catcher and so the tube. Then the core
and the metallic tube are brought to the surface with the crane. This system allows
retrieving up to 55 m length cores.
https://perso-sdt.univ-brest.fr/~jacdev/uf08/calypso.html).

22
Figure II.1. Progress of a CALYPSO coring. 1) Preparation of the coring system (main tube and
weight linked to the triggering system). 2) Final retrieving of the core.

For the Sea of Marmara, cores were taken at depth between 1250 and 400 m, and
their lengths range from 21.8 to 37.3 m (in Beck et al., 2007). For the Gulf of Corinth,
cores were taken at depth between 837 and 867 m, and their lengths range from 2.97 to
20.08 m (Dannielou, B., 2002). The cores retrieved in both areas were divided into 1.5-m
long sections; non-destructive on board measurement profiles were performed using a
GEOTEK multi-sensors core logger (magnetic susceptibility, sonic velocity, gamma-ray

23
densimetry). Further laboratory samplings and measurements were done on half cores
horizontally stored at 4 °C in the core depository of the “Muséum National d’Histoire
Naturelle” in Paris.
The present work is dedicated to six long piston cores among the two sets:
- three of them for the Sea of Marmara, labeled MD01-2425, MD01-2429 and
MD01-2431 core. These cores were recovered in the deeper part of the Central
(Orta) and Eastern (Çinarcik) basins;
- three others for the central part of the Gulf of Corinth, labeled MD-2477, MD-2478
and MD-2481 core.

The studied cores represent approximately the same time interval and have
comparable lengths, which imply that they underwent similar diagenetic evolution; thus
good conditions for comparing the textures are fulfilled. Moreover previous investigations
of sedimentary archives in the two sites have shown their potential for paleoseismic
recording (Moretti et al., 2004; McHugh et al., 2006; Sarõ and Ça÷atay, 2006; Beck et al.,
2007; Drab et al., 2012).

2.3 Macroscopic observations


2.3.1 Split cores observations
The core was photographed with a digital camera and pictures and compiled with
Adobe Photoshop CS4 Version 11.0 software. The cores were described in units of colors
(as deunitined using Munsell Rock-Color Chart), underlining estimated grain size,
sedimentary structures (dynamic features and bioturbations), biogenic material, and
recognizable mineral fragments. Based on these descriptions different lithofacies were
identified, and a depositational framework under which the different lithofacies were
deposited, was developed. These visual descriptions were also used to deunitine the
optimum locations for creating smear slides of the sediment to best characterize the
different units that are present.

2.3.2 X-ray imaging


X-ray radiograph is a non-destructive method that allows revealing internal
physical or biological structures of the sediments. The X-ray images were obtained by
exposing the core to an X-ray source (160 kV). After X-rays generated by the source pass
through the sample, the a brightness amplifier improves the contrast and a CCD camera,

24
with a resolution of 756x581 pixels, detects the X-rays and supplies images to the
computer.

In this thesis, X-ray images were acquired with the SCOPIX system of Bordeaux
University (Migeon et al., 1999). These images were obtained for six thin slices of
sediment (1 cm thick, 1.5 m length and 10 cm width) from the MD01-2425 core. The slices
were selected in order to have high resolution images of the supposed shock-induced
sedimentary structure and of the transitional environmental zones.

2.4 Laboratory analyses


2.4.1 Sediment composition
2.4.1-a Microscopic observations (smear slides)
In order to identify and quantify the relative percentage of biogenic, detrital and
authigenic components in the coarse and fine-grained sediments of the cores, microscopic
observation of the smear slide were made. Sediment for smear slide was collected directly
from the core for each lithofacies. The sediment was mixed in distilled water on a glass
slide until the sediment was dispersed. The slide was dried, and a cover slip was attached
with a synthetic resin. A total of 161 smear slides (81 for Sea of Marmara and 80 for Gulf
of Corinth) were analyzed on a NIKONTM petrographic microscope.
2.4.1-b Loss on ignition (LOI): Organic and carbonate contents
Sequential loss on ignition (LOI) is a common and widely used method to estimate
the organic and carbonate content of sediments (e.g. Dean, 1974; Bengtsson and Enell,
1986). In this thesis, LOI was made following the procedure proposed by Heiri et al.
(2001). First, the sediment was dried in the oven during 12-24 h at a temperature around
105 °C. Second, the organic matter was combusted in a first step to ash and carbon dioxide
at a temperature between 500 and 550 °C during 4 h. The LOI was then calculated using
the following equation:
DW105  DW550
LOI 550 *100 (1)
DW105
- where, LOI550 represents LOI at 550 °C (as a percentage), DW105 represents the dry
weight of the sample before combustion and DW550 the dry weight of the sample
after heating to 550 °C (both in g). The weight loss should then be proportional to
the amount of organic carbon contained in the sample;

25
- finally, in a second reaction, carbon dioxide was evolved from carbonate at a
temperature between 900 and 1000 °C during 2 h, leaving oxide and LOI was
calculated as:
DW550  DW950
LOI 950 *100 (2)
DW105

- where LOI950 is the LOI at 950 °C (as a percentage), and DW950 represents the dry
weight of the after heating to 950 °C. The weight loss should then be proportional
to the amount of carbonate contained in the sample.
A total of 52 samples were made along the MD01-2425 core of the Sea of Marmara.
These samples were taken with a sampling distance of ~60 cm, and allowed to evaluate the
vertical variations of the carbon organic and carbonate along the core.
2.4.1-c Geochemical analysis
For selected portions of cores, along-core elementary relative abundance were
performed using an X-ray fluorescence (XRF) core scanner AVAATEC. In this method,
under the influence of an external X-ray radiation an electron is ejected from an inner shell
of an atom. The resulting vacancy is subsequently filled by an electron falling back from
an outer level, and the energy difference between both levels is emitted as an
electromagnetic radiation. The wavelength(s) of emitted radiation are characteristic for
each element, and the amplitudes of peaks in the XRF spectrum are proportional to the
concentration of corresponding elements in the analyzed sample (Richter et al., 2006).

This method provides a rapid high-resolution (down to 1 mm) record of chemical


composition on split sediment cores. The measurements are non-destructive and require
very limited sample preparation (Richter et al., 2006). The core sections were measured
every 5 or 1 mm, with energies of fluorescence radiation of 10 keV and 30 keV to reach a
large spectrum of elements from Al to U. The elemental distributions are expressed in
counts per second (cps), which is proportional to the chemical concentration (Tjallingii et
al., 2007). The analyses were carried out at the EDITEM laboratory (Université de Savoie).
In the MD01-2425 core (Sea of Marmara) and in the MD01-2477 core (Gulf of Corinth)
the measures were performed on sections of 1.5 m and 4.5 m long respectively.
2.4.1-d Bulk (or volume) Magnetic Susceptibility (MS)
Volume (non oriented) bulk measurements of Magnetic Susceptibility were initially
performed onboard on non split core section using a BARTINGTONTM MS2 sensor

26
through a ring, with a 2 cm spacing. MS profiles presented and used in the following were
achieved with a contact sensor, permitting a 5 mm spacing for a better accuracy.
2.4.1-e Remanent magnetizations (Isothermal and Anhysteretic remanent
magnetization)
Isothermal Remanent Magnetization (IRM) and Anhysteretic Remanent Magntization
(ARM) were performed on selected portion of cores where we developed the use of
Anisotropy of Magnetic Susceptibility (AMS). As for the latter, we refer to a published
article (joint hereafter)* for detailed description of the procedure.

2.4.2 Sediment textures analyses


As our investigations were mainly directed towards specific ways of settlement in
relations with shaking effects and their consequences on sediments, special attention were
paid to textures (grain size distributions, grains shapes, grains array).
2.4.2-a - Laser microgranulometry
A MALVERNTM Mastersizer 2000 was used for two purposes: 1) to characterize
the distribution of the different types of deposits, 2) to investigate the base-to-top evolution
of specific layers supposed to represent individual depositional events. For the second
approach we constructed different binary diagrams involving different major indices and
parameters (Sorting index, Skewness index, median, mean size, percentile Q99, etc.).
Except for the coarse sandy layers, a 2 mm measuring interval could be applied respecting
the implied measurement conditions (approx. 10 % obscuration). To determine the
processes of transport and deposition prevailing during the deposition of some lithofacies,
we used binary diagrams: i) skewness vs sorting index, and ii) percentile C99 vs median
(Passega, 1964). On these diagrams, base-to-top paths appeared to be able to differentiate
between different subaqueous gravity flows as: “flood turbidite” and “slump turbidite”
(Lignier, 2001; Arnaud et al., 2002; Carrillo, 2006; van Welden 2007; Beck, 2009). In the
present study, we used them to distinguish the earthquake-induced deposits - homogenites
+turbidites type - from other type of gravity driven deposits or “normal” hemipelagitec
sediments.
2.4.2.b Grain shape analysis
In order to validate non direct measurements of grain array (especially AMS) -
especially to avoid effect of particle shape and density (controlled by composition) - we
performed automatic shape analysis on fine-grained sediments (clay, silt and very fine

27
sand, up to 100 µm). We used a SYSMEXTM FPIA-2100 apparatus. Examples of obtained
measurements and pictures will be given in Chapters 3 and 4.
2.4.2.c Anisotropy of Magnetic Susceptibility (AMS)
Since several decades, the use of Anisotropy of Magnetic Susceptibility has been
developed to identify or confirm, and to quantify, the possible orientation of grains or
crystals, either related to deformation, or to initial settling. More specifically, AMS appears
to correlate with other textural parameters in the study of re-suspended and re-settled
sediments. Thus we develop their use for inferred earthquake-induced layers following
preliminary previous examples (in Beck, 2009). We refer to a published article (jointed at
the end of this Chapter)* for detailed description of the procedure.

2.4.3 Chronology: 14C dating and correlations


Chronological control of the sedimentary sequences analyzed in this thesis was
14
obtained through C dating. Thus, a brief discussion about the general bases of this
method is provided in the following. Carbon has three naturally occurring isotopes. The
two most abundant are the stable isotopes, 12C(98.99 %) and 13
C (1.11 %). 14
C, with an
abundance of less than 10 -10
%, is unstable and undergoes ȕ-decay with a half life of 5730
± 40 years (Stuiver and Polach, 1977). This unstable isotope is produced by the interaction
14 14
of cosmic rays with nitrogen and oxygen, with the majority formed by the N (n,p) C
14 14
reaction in the atmosphere (Kamen, 1963). The C formed is rapidly oxidized to C O2
and enters the earth's plant and animal lifeways through photosynthesis and the food chain.
Thus, the level of 14C in plants and animals when they died approximately equals the levels
of 14C in the atmosphere at that time (Taylor and Lloyd, 1992). After death of an organism,
14
the C in its tissues is no longer in equilibrium with atmospheric CO2, and undergoes
14
radioactive decay back to C. If the plants or animals tissues remain intact and isolated
from exchange, the decrease in its 14C content (expressed as a ratio of 14C /12C) from that in
living organism, may be used to indicate the time since the death of the organism
(Trumbore, 2000). Assuming that the death of this organism was contemporaneous to the
deposit that contains it, an age for this deposit can be estimated (Libby, 1955; Taylor and
Lloyd, 1992). Due to the short half life of the 14C and the available technology to measure
14
the concentration of C remain, the range of applicability of 14C dating is fixed between
>300 years and <55000 years (Trumbore, 2000).

28
In this work, two set of 14C measurements are used: 1) Published preliminary dating
performed by Moretti et al. (2004), Beck et al. (2007), Lykousis et al. (2007) and van
Welden (2007) in BETA ANALYTIC Laboratory and at Woods Hole Oceanographic
Institution (National Ocean Sciences Accelerator Mass Spectrometry facility NOSAMS),
and at Poznan Radiocarbon Laboratory; 2) New measurements series performed through
CNRS-INSU ARTEMIS facility program at Saclay CEA Laboratory. The samples concern
plant debris, wood fragments, particulate organic matter, shell fragments and foraminifers. As
the samples were taken in sub-aquatic environments (marines and non-marine)
unprocessed values were calibrated according to the Marine09 calibration curve (Reimer et
al., 2009), using the OxCal program v4.1. (Ramsey, 2001), and a local marine reservoir
correction (ǻR) was applied. The ages (cal. year BP) are reported as time intervals that
represent 95% of probability (2ı).

(*) Article: Deciphering hemipelagites from homogenites through Magnetic Susceptibility


Anisotropy. Paleoseismic implications (Sea of Marmara and Gulf of Corinth), Sedimentary
Geology, 292:1-14, doi: 10.1016/j.sedgeo.2013.03.015)

29
Sedimentary Geology 292 (2013) 1–14

Contents lists available at SciVerse ScienceDirect

Sedimentary Geology
journal homepage: www.elsevier.com/locate/sedgeo

Deciphering hemipelagites from homogenites through anisotropy of magnetic


susceptibility. Paleoseismic implications (Sea of Marmara and Gulf of Corinth)
Corina Campos a, b, Christian Beck a,⁎, Christian Crouzet a, François Demory c,
Aurélien Van Welden a, d, Kadir Eris a, e
a
Laboratoire ISTerre, UMR CNRS 5275, Université de Savoie, Grenoble University Group, 73 376 Le Bourget du Lac, France
b
Departamento de Ciencias de la Tierra, Universidad Simón Bolívar, Sartenejas, Baruta, Venezuela
c
Aix-Marseille Université, CNRS, IRD, CEREGE UM34, 13545 Aix-en-Provence Cedex 04, France
d
CGGVeritas, OH Bangs. vei 70, NH 1323 HØVIK, Norway
e
Faculty of Engineering, Geology Department, Fırat University, 23100, Elazığ, Turkey

a r t i c l e i n f o a b s t r a c t

Article history: In closed marine basins and large lakes, gravity re-depositional processes often result in specific turbidites with
Received 22 November 2012 two abruptly separated layers: a coarse graded lower term and an upper homogenous fine-grained term. An ad-
Received in revised form 14 March 2013 ditional mixed term generally occurs in between, indicating to and from particle displacements. The later ones
Accepted 17 March 2013
are related to oscillating bottom currents responsible for a high increase of the fine fraction segregation, within
Available online 28 March 2013
the reworked wasted mass. The whole sedimentary event is the association “homogenite + turbidite” (HmTu),
Editor: B. Jones which specific a settling condition area here characterized through Anisotropy of Magnetic Susceptibility (AMS).
Homogenites's magnetic foliation appears anomalously high with respect to their expected state of compaction,
Keywords: and strikingly higher than hemipelagites's values. We applied this approach to Late Pleistocene/Holocene
Homogenite sediments from the Sea of Marmara and from the Gulf of Corinth. Grain-size and other magnetic parameters
Turbidite related to mineralogy are added to better assess the granular array influence on AMS. As HmTu is considered
Magnetic fabric as often related to earthquake-triggering and tsunami/seiche effects, AMS appears as a useful tool for subaqueous
Paleoseismicity paleoseismic investigations. First, it may evidence the signature of paleo-earthquakes; second, it permits to
Marmara
decipher hemipelagic intervals which are inferred to represent the time elapsed between two successive
Corinth
reworking events.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction and Rimoldi, 1997) was triggered by a giant volcanic crater collapse,
most of the further surveyed and analyzed homogenites have been
Following the discovery of Mediterranean fine-grained “homogenites” related to major seismic events, by mean of tsunami/seiche effect. Ex-
and their interpretation as tsunami-induced reworking and re-deposition amples of turbidite/homogenite associations have been described in
(Kastens and Cita, 1981; Cita and Rimoldi, 1997), similar layers have been deep pull-apart basins where they have been attributed to co-seismic
reported in various settings and frequently associated to subaqueous faulting (Beck et al., 2007). A 15 m-thick sedimentary “event” is men-
earthquake effects. Mentionned thicknesses range from few centimeters tioned; comprising mud-clast breccia, coarse sand, and a 8 m-thick
to ten meters. Homogenites often appear associated to an underlying homogenite. In situ observations done immediately after recent major
coarser layer which may vary from a thin silty laminae to thick, sandy/ earthquakes have confirmed the importance of a thick and long-
gravelly and normally graded layers, similar to the lower terms of a turbi- lasting “cloud” of fine particles: in Cariaco Trough (Thunell et al., 1999),
dite. The word “turbidite” refers to the definitions and models by Bouma north-west of Haïti (McHugh et al., 2011).
(1962) and Mutti and Ricchi-Lucchi (1978). The systematic occurrence of Focusing on the paleoseismic sedimentary record, we have to en-
a sharp limit between the fine-grained “homogenites” and the underlying visage two major problems: i) the demonstration of earthquake ori-
turbidites has been underlined by Sturm et al. (1995), who proposed a gin of a sedimentary “event”, ii) the distinction of homogenites from
conceptual model for “homogenites” in large lakes, focusing on their overlying hemipelagites. Almost all the different features of a sub-
possible association with earthquake-triggered mass wasting; this aqueous earthquake imprint are gathered on Fig. 1 example: lique-
concept was later developed by Chapron et al. (1999). Although the faction, microfracturation, and HmTu-type gravity reworking deposit;
above-mentioned “Minoan” homogenite (Kastens and Cita, 1981; Cita here the re-suspended fine-grained part probably resulted in a two-
phased homogenite.
⁎ Corresponding author. As hemipelagites represent continuous slow (“normal”) sedi-
E-mail address: beck@univ-savoie.fr (C. Beck). mentation, they are used to estimate time elapsed between two

0037-0738/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.sedgeo.2013.03.015
2 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

Core MD01-2431 - Section IX - 1380/1410 cm

Fig. 1. Illustration of major features of a re-depositional event in a deep basin; descriptions (left) and inferred responsible processes (right). The shown example (supposed to be
earthquake-related; Central Basin of the Sea of Marmara) displays an abrupt separation of a coarse lower part (lower term of a turbidite sensu Bouma) from a fine-grained homogenous
upper part (called “homogenite”). Event n + 1 is responsible for microfracturing (green arrows) and a (almost synchronous) new turbidity current. An hemipelagic interval separates the
two successive events. Direct core observation, X-ray picture, and/or vertical grain-size evolution, indicate an episode of to-and-fro bottom water displacement, between the two major
layers. The subdivision of the homogenite is not often present, depending of initial grain-size spectrum and on composition of the reworked deposit.

events (n/n + 1, Fig. 1 as an example); this approach has been pro- 2. The investigated sites and their settings
posed by Adams (1990) and further developed in different tectonically
active areas (Gorsline et al., 2000; Huh et al., 2004; Sarı and Çağatay, Both studied areas are situated in the Eastern Mediterranean re-
2006; Beck et al., 2007; Goldfinger et al., 2007; Noda et al., 2008; gion (Fig. 2a). The deep basins of the Sea of Marmara are associated
McCalpin, 2009 and references therein; Barnes and Pondard, 2010; to an east–west trending right lateral transform boundary, well
Grácia et al., 2010; Beck et al., 2012). The need to distinguish the known as the North Anatoila Fault. The Gulf of Corinth corresponds
fine-grained part of reworking events from hemipelagites has also to north–south extension above the hellenic subduction. In addition
been underlined for ancient sediments (Hesse, 1977). A similar ap- to their tectonic-geodynamic settings, the following reasons led us
proach is also used in large lacustrine basins (Marco and Agnon, 1995; to elect these two sites:
Rodriguez Pascua et al., 2003).
As homogenites should basically be characterized by a specific granu- - they are undergoing similar climatic regimes and underwent same
lar array, related to a specific settling and independently from their com- paleoenvironmental changes during Late Quaternary;
position, their textural parameters appear fundamental discriminating - their size, bathymetry, and watershed are comparable, and they
criteria. Grain-size, undrained shear stress, and magnetic fabric, have both were disconnected from open sea during Late Glacial Maxi-
been combined in an attempt to characterize a well-constrained historical mum (MIS 2);
homogenite (Chapron et al., 1999; Carrillo et al., 2008; Beck, 2009). AMS - the two available cores represent approximately the same time in-
(especially foliation) appeared as a promising tool for this purpose. Its terval and have comparable lengths, which imply they underwent
previously-developed use for soft sediments was mainly related to similar diagenetic evolution, especially compaction;
paleomagnetism investigations as different processes may alter the paleo- - their recent sedimentary fills were sampled with the same giant
magnetic field recording and its measurement: sediments settling, piston coring device, and core processing and storage were identi-
compaction, bioturbation, diagenetic mineralogical changes, early soft de- cal; thus good conditions for comparing the textures are fulfilled.
formation, and coring and sampling (LØvlie et al., 1971; Kent and Lowrie, Previous investigations of sedimentary archives in the two sites
1975; Schneider and Kent, 1988; Hounslow et al., 1990; Rochette et al., have shown their potential for paleoseismic recording (Moretti
1992; Collombat, 1993; Demory et al., 2005; etc.). Beside, AMS has been et al., 2004; McHugh et al., 2006; Sarı and Çağatay, 2006; Beck et
proposed to characterize primary settling (Rees, 1965; Hamilton and al., 2007; Drab et al., 2012).
Rees, 1970, Joseph et al., 1998). Post-depositional disturbances inferred
as earthquake-induced have also been analyzed through AMS (Levi et 2.1. The Çinarcik Basin, Sea of Marmara
al., 2006; Mörner and Sun, 2008; Campos et al., 2011). In the here-
presented investigation, we focus on the use of AMS to characterize The Sea of Marmara (Northwestern Turkey) is an East–west elon-
re-depositional processes in two deep basins directly related to active gated 200 km-long pull-apart basin developed along the North Anato-
faulting and undergoing frequent strong earthquakes: the Sea of lian Fault (Hancock and Barka, 1981). It is composed of several
Marmara, and the Gulf of Corinth. Previous investigations in the Sea of aligned sub-basins (Tekirdağ, Orta/Central, Kumburgaz, and Çinarcic/
Marmara have shown that homogenites may represent a significant Eastern) up to 1280 m-deep (Fig. 2b). It is presently connected to the
part of the total sediments accumulation (Beck et al., 2007). Black Sea by the Istanbul Strait (Bosphorus), and to the Aegean Sea by
C. Campos et al. / Sedimentary Geology 292 (2013) 1–14 3

22 00’ 22 30’ 23 00’


38 30’
20 30
c
Apu

Eurasian Plate
North Anatolian Fault 20 Km
Itea
lian

40 25 mm/yr
Pla

-200
-400
W-ANT F.
te

Anatolian Plate -600


c Aegean MD01-2477
Sea GULF
OF CORINTH
M
ed

Arabian Plate
Akrata
ite

lt
He 35 mm/yr Xyloca
stro
rra

ea Fau
l le F. Co
ni rinth
ne

35 c F.
-800
a

Trench
n

Xylocastro
R
id

Dead S
38 00’
g

10 mm/yr
e

2500 m Faults

-100
Major towns
Water Depth (m)
Corinth
a African Plate 0m Core location
SARONICOS GULF

27 00’ 28 00’ 29 00’

b
41 00’
-100
-250
Istambul 25 Km

-1000 -1000 -750


Orta Basin 500 -
Tekirdag -1000
Basin inarcik Basin
MD01-2425
Izmit
-750 -750 -1250
-500
-100
ault
an F
An atoli -100
-100
SEA OF MARMARA -250

40 30 N orth
-100

2500 m Thrust faults Major towns


-100
Normal faults Water Depth (m)
Strike slip faults Core location
0m

Fig. 2. Simplified geodynamic setting of the Sea of Marmara and the Gulf of Corinth. a) plate kinematics (simplified from McClusky et al., 2003; Barrier et al., 2004); b) the Sea of Marmara:
geomorphological setting and North-Anatolian Fault (from Armijo et al., 2002); c) the Gulf of Corinth: geomorphological setting and major active faults (from Lykousis et al., 2007;
Sakellariou et al., 2007); relief from NASA SRTM.

the Çanakkale Strait (Dardanelles). These two straits have respective sill Ferentinos, 1984; Papatheodorou and Ferentinos, 1993; Armijo et al.,
depths of approximately 35–45 m and 55–75 m, (Londeix et al., 2009). 1996; Stefatos et al., 2002; McNeill et al., 2005; Lykousis et al., 2007;
The North Anatolian Fault (NAF) is a 1200-km-long dextral strike-slip Sakellariou et al., 2007; Bell et al., 2009). This basin reaches at its centre
fault (Sengör et al., 2004) and is considered as a major active boundary a maximum depth of 900 m, and it is characterized by steep north and
between Anatolia and Eurasia plates. (Armijo et al., 1999, 2002; south dipping slopes (Stefatos et al., 2002). The focal mechanism studies
McClusky et al., 2003; Flerit et al., 2003). It is responsible for recent indicate a main N–S direction of extension (Bernard et al., 1997, and ref-
high magnitude earthquakes, part of them responsible for deep sub- erences therein). The extension rate is at least 3–4 mm year−1 across
aqueous scarps (Armijo et al., 2005). the most active part of the rift system (Armijo et al., 1996) and it has
The northern branch of the NAF crosses the different deep sub-basins, up to 1.5 mm year−1 uplift rate in the southern margin since the late
where giant pistons cores (from 27 to 37.7 m) have been retrieved, with Middle Pleistocene (Keraudren and Sorel, 1987; Armijo et al., 1996; De
location based on high resolution seismic reflection imaging (Beck et al., Martini et al., 2004; Palyvos et al., 2005; Lykousis et al., 2007; Ford et
2007). The cored sediments are Holocene-Late Pleistocene in age and al., 2013). Recent to Present Day faulting activity is surveyed through
they represent marine and non marine environments; this change is detailed seismological and GPS (kinematics) surveys (Jackson et al.,
considered to have occurred around 12 cal. kyr BP (Londeix et al., 2009). 1982; Briole et al., 2000; Bernard et al., 2006). Possible coastal, onshore,
and offshore catastrophic consequences of major earthquakes have also
been mentioned (Papadopoulos and Chalkis, 1984; Perissoratis et al.,
2.2. The Gulf of Corinth 1984; Papadopoulos and Plessa, 2000; Papadopoulos, 2003).

The Gulf of Corinth is a semi-enclosed marine basin, located in Cen-


tral Greece (Fig. 2c); it is 115 km long and approximately 30 km wide at 3. Analytical tools
maximum. This N°100E elongated basin separates the Peloponnese to
the south from continental Greece to the north, and is developed per- The two analyzed cores were retrieved during R/V MARION-
pendicular to the Alpine Hellenides mountain chains (Lykousis et al., DUFRESNES cruises using the CALYPSO giant-piston coring device. Core
2007).To the west the gulf is connected to open marine Ionian Sea MD01-2425 (water depth: 1215 m, length: 31.29 m) was taken during
through the Rion-Antirion strait. The latter is 2 km wide and 65 m the MARMACORE Cruise (August-September 2001), as part of a larger
deep. To the east, it is linked with the Saronikos Gulf and the western GEOSCIENCES MD 123 program. Core MD01-2477 (water depth:
Aegean Sea by an artificial channel (The Corinth Canal), 6.3 km long, 867 m, length: 20.08 m) was retrieved as part of GEOSCIENCES-II pro-
21 m wide and 8 m deep. gram, carried out in October 2001. Shipboard processing included
The Gulf of Corinth is an active half-graben (Fig. 2c) bounded by E–W GEOTEK core logger profiles and splitting; further laboratory samplings
striking, en échelon faults located onshore and offshore (Brooks and and measurements were done on half cores horizontally stored at 4 °C.
4 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

MD01-2425
0
MD01-2477
Gulf
0
Tu
of 200
Sea of Marmara
Tu
Tu Corinth Eastern Basin
Tu
200 HmTu?
Tu
400
HmTu

HmTu
3064 - 2734 cal yr BP
Tu
Tu 600

400
HmTu
Tu
D.S
390 cm= 3280 - 2913 cal yr BP 5625 - 5321 cal yr BP
Tu
HmTu ? 800
HmTu 516 cm= 6052 - 5706 cal yr BP

HmTu

600 HmTu 1000


A HmTu
HmTu
HmTu
650 cm= 13990 - 13493 cal yr BP

HmTu ? Legend
HmTu
HmTu
Tu
1200 Laminated facies
800 HmTu
Tu
HmTu
HmTu
Homogeneous mud
Tu
Depth (cm)

Depth (cm)
HmTu
HmTu 1400 Laminate mud
Tu
HmTu
Tu
1000 D.S.= Deformational structures
HmTu
(ball and pillow, slumps and dikes)
B HmTu
1600
1107 cm= 9239 - 8779cal yr BP Tu= Turbidite
Non-marine to marine (fining upward deposit)
Tu
marine Tu
Tu
transition
Tu
HmTu = Turbidite + Homogenite
HmTu
1200 Tu
1800 (Two associated layers. (1) Coarse fining upward
HmTu
HmTu ? base, overlain by (2) Homogenous silty-claily layer)
HmTu
Cal. AMS 14C age
HmTu
HmTu ? HmTu ?

non marine D.S


Non-marine to marine
HmTm
HmTu Non-marine to marine 2000 transition. (approx. 12,000 yr BP)
Tu
Tu
1400 HmTu transition
HmTu? HmTu ? A
Sections s tudied by Anisotropy o f Mag-
HmTu
netic Susceptibility
HmTu 2200 HmTu
HmTu
HmTu ?
HmTu ?
Tu HmTu
HmTu ?
Tu
1600 HmTu
HmTu
HmTu
Tu
14038 - 13618 cal yr BP 2400
HmTu ?
HmTu ?
D.S
HmTu ?
Tu
HmTu ? D.S

Tu Tu
HmTu
Tu 1766 cm= 16928 - 16218 cal yr BP Tu
HmTu ?
Tu
1800 2600 HmTu
HmTu
A HmTu
HmTu
HmTu
HT ? HmTu
HmTu
HmTu
14945 - 14034 cal yr BP
2800 HmTu
HmTu ?
HmTu ? 2868 cm= 15650 - 14655 cal yr BP
2000
HmTu

3000 B HmTu
HmTu
0,1 1 10 100 1000 HmTu
-5
MS (x10 S.I.)
1 10 100 1000
Magnetic Susceptibility (x10-5 S.I.)

Fig. 3. Synthetics logs of Cores MD01-2477 (Gulf of Corinth) and MD01-2425 (Sea of Marmara, Cynarcik Basin). In both sites, the Late Glacial/Holocene interval was recovered, with
open sea connection during the Holocene; gravity reworking with coarser deposits are more frequent during the non marine part. Rectangles: intervals with detailed continuous
AMS measurements.

Focusing on particular sedimentary events, selected portions of A total of 377 discrete samples were analysed (172 for Marmara and
the two cores were investigated (red rectangles on Fig. 3). 205 for Corinth). Prior to these sampling, a Magnetic Susceptibility
curve was established using a BARTINGTON™ MS2 contact sensor on
3.1. Grain size split cores, at a 5 mm spatial resolution.
The AMS measurements were carried out in the CEREGE Laboratory
Grain-size analyses were conducted with a laser diffraction micro- of Magnetism (Aix-Marseille University), using AGICO MFK1-FA
granulometer MALVERN™ Mastersizer 2000 at the ISTerre laboratory; Kappabridge (spinning specimen method). During measurement, the
the displayed size-range – 0.1 μm to 2 mm – being particularly adapted specimen slowly rotates subsequently about three perpendicular axes
to the investigated sediments. Measurements were performed each and the susceptibility differences are measured during specimen spin-
2 cm for the whole cores and each 2 to 5 mm for the selected sections. ning (64 measurements per spin). The result is a sensitive determina-
Among the different classical statistic parameters, we choose Median/ tion of the anisotropic component of the susceptibility tensor. For
D50 (expressed in μm) to represent the grain-size evolution within the dif- each sample the measurements allowed to reconstruct the AMS tensor,
ferent layers. defined by three eigenvectors (Kmax, Kint and Kmin) (Hrouda, 1982;
Tarling and Hrouda, 1993). The magnetic susceptibility tensor may be
3.2. Magnetic properties represented geometrically by a tri-axial ellipsoid, whose axes are paral-
lel to the eigenvectors of the AMS tensor. This susceptibility ellipsoid
For each core section, two U-channels were prepared, one for con- represents the combined result of the susceptibility anisotropy pro-
tinuous measurements, one for discrete sampling and measurements. duced by individual grain shape and/or crystallography in a sample
The samples were collected into 8 cm3 non-magnetic plastic boxes. In (Stoner et al., 1996). For sedimentary deposits, the maximum suscepti-
order to limit sampling disturbance of the sediment, continuous cubic bility axis, Kmax, is generally parallel to the mean long axis of the in-
samples were cut from an additional U-channel (the core diameter al- dividual particles. The magnetic fabric is usually comparable to the
lows to collect two U-channels along the axis of each section). This sediment fabric (Borradaile, 1988; Rochette et al., 1992; Tarling and
avoids disturbances related to direct introduction of successive boxes. Hrouda, 1993).
C. Campos et al. / Sedimentary Geology 292 (2013) 1–14 5

The AMS ellipsoid shape is illustrated by several parameters de- 3.3. Chronology
fined by Jelinek (1981) as:
Two sets of Accelerator Mass Spectrometry 14C measurements are
- magnetic foliation: F = Kint/Kmin, used: 1) published preliminary dating performed on the two cores
- magnetic lineation: L = Kmax/Kmin, (Moretti et al., 2004; Beck et al., 2007; Lykousis et al., 2007) in BETA AN-
- shape parameter: T = 2 ln(Kint/Kmin)/ln(Kmax/Kmin)-1. ALYTIC Laboratory, at Woods Hole Oceanographic Institution (National
Ocean Sciences Accelerator Mass Spectrometry facility NOSAMS), and
Additionally, the magnetic susceptibility (MS) and the mean in Poznan Radiocarbon Laboratory; 2) new measurements series
(volume) magnetic susceptibility (Km) were measured for each performed through CNRS-INSU ARTEMIS facility program at Saclay CEA
specimen. Laboratory. All ages concern Organic Matter (plant debris and particular
In order to identify magnetism-carrying particles, different mea- matter from fine-grained deposits). As the samples were taken in sub-
surements have been performed, in addition to grain-size distribution. aquatic environments (marines and non-marine) unprocessed values
In order to distinguish different mineralogical and granulometric frac- were calibrated according to the Marine09 calibration curve (Reimer et
tions within the magnetic components, two types of laboratory rema- al., 2009), using the OxCal program v4.1. (Bronk Ramsey, 2001). An aver-
nent magnetisations were artificially imparted: Isothermal Remanent age value of local marine reservoir correction (ΔR) proposed by Reimer
Magnetization (IRM) and Anhysteretic Remanent Magnetization (ARM). and McCormac (2002) for Aegean Sea is applied to the Gulf of Corinth
For IRM and ARM measurements, U-channels subsampling was used. 5 (ΔR = 89 ± 58), and by Siani et al. (2000) for the Marmara Sea
U-channels from Marmara and 3 from Corinth have been analyzed in (ΔR = 81 ± 56). We applied these ΔR corrections to the whole sedi-
CEREGE laboratory; their lengths range from 40 to 92 cm. IRM and ARM mentary sequence (marine and non-marine conditions; Holocene and
were measured at intervals of 2-cm. ARM was produced in-line Late Pleistocene) both for the Gulf of Corinth to the Marmara Sea. This
along the u-channel axis, using a 100 mT alternating field with a assumption is probably simplistic, but the lack of data does not allow
superimposed 0.05 mT steady field. For ARM demagnetization was us to make more precise corrections. Values represented in the core
done following steps of 10, 20, 30 and 40 mT. IRM was obtained by logs (Fig. 3) correspond to highest confidence intervals.
passing the u-channels through two different Halbach cylinders
(permanent rare earth magnets) that develop fields of 1 T and 4. Results
0.3 T, respectively (Rochette et al., 2001). Additionally, the S-ratio
(= 0.5*(1 − (IRM− 0.3T / SIRM1T)) from Bloemendal et al. (1992) 4.1. Overview of cores content
and HRM (=0.5*(IRMIT − IRM-0.3T)) from Robinson (1986) were calcu-
lated. Remanent magnetization was measured using a superconducting Fig. 3 summarizes lithology and layering of the two investigated
quantum interference device (SQUID) pass-through magnetometer (2G giant piston cores (location on Fig. 2b and c). The dominant sediment
760R), located in a shielded room. is an olive grey to dark yellowish brown calcareous mud interbedded

0 40 80 120 160 1 1.04 1.08 1.12 1.16 -1 0 1


Median µm Foliation T
2630

HmTu
2640

2650
h.d.

2660
HmTu
Depth cm

2670 h.d.

2680
HmTu

2690

2700 100 200 300 400 500 1 1.02 1.04 1.06 1.08 0 30 60 90
Magnetic susceptibility Lineation Inc Kmin
(10-6 S.I.)
Marmara Section-A
h.d.: Hemipelagic deposits HT: re-depositional event Turbidite Homogenite
(turbidite+homogenite)

Fig. 4. AMS profiles of selected portions A and B from Core MD01-2425 (Sea of Marmara); 2 cm sampling intervals. Volume magnetic susceptibility (Km) and a grain-size parameter
(5 mm sampling interval for both) are added. (stratigraphic position on Fig. 3).
6 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

0 10 20 30 40 50 60 1.04 1.06 1.08 1.1 1.12 1.14 -1 0 1


2960
Median µm Foliation T

HmTu
2970

2980
h.d.

Tu
Depth cm

2990 h.d.

HmTu
3000

3010 h.d.

HmTu
3020

100 150 200 250 300 350 400 1 1.02 1.04 1.06 1.08 0 30 60 90
Magnetic susceptibility Lineation Inc Kmin
(10-6 S.I.)
Marmara Section-B
h.d.: Hemipelagic deposits HT: re-depositional event Turbidite Homogenite
(turbidite+homogenite)

Fig. 4 (continued).

with millimetric to decimetric layers of yellowish brown sand, and silty between 12 kyr BP (Perissoratis et al., 1993, 2000; Collier et al.,
sand. Additionally, in the non-marine sections, a few centimetres-thick 2000) and 13.2 kyr BP (Moretti et al., 2004). Both occurrences of
series of thin laminated muds (olive grey, brownish olive grey, greyish reworking events (Tu, HmTu) and detailed bulk volume MS curves
yellow and brownish black) are present; for the Gulf of Corinth. They display clear differences: i) between the two sites, ii) between the
have been interpreted as a seasonal imprint by Moretti et al. (2004) marine and the non marine successions in the sames core. Typical
and Lykousis et al. (2007). Microscopic observation, both on coarse HmTu are thinner and less frequent in the Gulf of Corinth core, but
and fine fractions, point out a high variety of terrigenous components they are present in both non marine and marine episodes. MS profiles
(siliciclastics and carbonates). Calcareous nannoplancton, diatoms, and also underline a difference between the two episodes in both cores;
bio-induced micrite represent the biogenic and bio-induced fraction of the non marine-to-marine changes is marked by a small increase
the silty/clayey fraction, depending on the marine vs. non marine (Corinth) or a small decrease (Marmara) of mean MS. In the Gulf of
situation. Corinth the occurrence of reworking events up to Present is due to
Concerning sedimentary “events” linked to density currents and the continuous uplift (normal faulting) of the bounding relieves and
reworking processes, we have distinguished: associated terrigenous feeding; for the Sea of Marmara, the difference
between non marine (Late Glacial) and marine (Holocene) is mainly
- turbidites (Tu on Fig. 3), when observing the classical succession with
due to the change of erosion conditions on a less changing relief
a transition from coarse, normally-graded lower part to fine-grained
(strike slip faulting).
upper part,
Detailed magnetic analyses were conducted on selected section
- from the association “turbidite + homogenite” (HmTu on Fig. 3),
indicated on Fig. 3 (red rectangles); for Core MD01-2477, a unique
when a sharp contact separates these two parts and when we have ar-
thick event could be investigated.
guments for to-and-fro particle displacements at the top of the coarse
lower part (cf. Introduction, Fig. 1; and in Beck et al., 2007; Beck,
4.2. Magnetic fabric
2009).
The second criteria may be: either directly deduced from flaser bed- For each of the four selected portion of sedimentary succession,
ding and microprogradation visible on X-ray pictures, or evidenced four parameters represent the magnetic fabric: magnetic foliation,
using a high-resolution microgranulometric profile (each 2 mm). For magnetic lineation, T parameter, inclination of Kmin (Fig. 4a and b,
several cases (noted “HmTu ?” on Fig. 3), the two parts are separated Fig. 5a and b). Mean volume magnetic susceptibility (Km) is added
by a neat contact, but we don’t have arguments to infer a temporary os- in parallel with Median (Q50) as an indicator of grain size.
cillatory movement of bottom water.
According to 14C ages, the two cores represent the same period, 4.2.1. Core MD01-2425-Marmara Sea-sections (Fig. 4)
approximately the last 18 kyr. The non marine-to-marine change All layers (HmTu, h.d., and coarse turbidites bases) have oblate mag-
(connexion with Mediterranean Sea) is considered to have occurred netic ellipsoid shapes (0.4 b T b 1) with a minimum susceptibility axis
C. Campos et al. / Sedimentary Geology 292 (2013) 1–14 7

very close to the vertical (75° b Inc Kmin b90°) and a sub-horizontal 1.083), whereas the basal turbititic layers show the lowest values; a
maximum axis (0° b Inc Kmax b 20°) The Km values average in the two sharp boundary and a high contrast are also present.
sections are 195 ± 82 and 193 ± 74 × 10−6 S.I., the highest values
being observed in the basal turbiditic unit (293 ± 106 and 250 ±
50 × 10 −6 S.I.). Only in one hemipelagic interval of Section-A (Fig. 4a), 4.3. Grain size, magnetic content, and their possible influence on magnetic
the minimum susceptibility axis is not subvertical (Kmin ~60°). This low fabric
value suggests a normal settling fabric subsequently disturbed after depo-
sition, possibly by bioturbation effects. The magnetic lineation is very low 4.3.1. Core MD01-2425-Marmara Sea-sections (Figs. 6a and 7)
and relatively constant, being almost invariable along all sections, with The median grain size profiles (Fig. 4) clearly discriminate
low values between 1.001 and 1.018. homogenites and hemipelagites, on one hand, from coarse graded bases
The magnetic foliation is variable in the two sections. The upper ho- of turbidites, on the other hand. In “Homogenite + Turbidite” (HmTu),
mogeneous layers present the highest values (1.100 to 1.148), while the the basal units (Tu) are 5 mm to 6 cm-thick. They show one or two fining
basal turbiditic unit shows the lowest (1.02 to 1.07). The contrast be- upward sequences, with a median grain size ranging from 160 to 10 μm
tween homogenites and hemipelagites is high and coincides with a (fine sand to fine silt). The upper units (Hm) are 4 to 12 cm-thick and
sharp boundary. have a relatively constant median grain size from 6 to 10 μm (very fine
silt to fine silt), and a slight variable sorting index (2.2 to 2.5). The
hemipelagic deposits (h.d.) have median grain size ranging from 1.9 to
4.2.2. Core MD01-2477-Gulf of Corinth-sections (Fig. 5) 8.0 μm and a moderate variable sorting index (1.7 to 2.6). Thus, no
As in the Sea of Marmara site, the “Homogenite + Turbidites” and significant difference could be noticed between homogenites and
hemipelagites deposits fabrics are prominently oblate (0.3 b T b 1) hemipelagites, excepted for the clay content (b4 μm) in homogenites.
with a minimum susceptibility axis sub-vertical (75° b Inc Kmin b 90°) Regarding the relationships between grain size and Km, coarse
and a horizontal maximum axis (0° b Inc Kmax b 12°). The magnetic lin- turbiditic bases generally show highest Km values; nevertheless, seve-
eation is very low (~1.011) and almost invariable. The highest values of ral Km peaks are observed in hemipelagites or at top of homogenites.
Km are present in the upper homogeneous units (140 ± 16 and Beside, fine-grained layers show a rather high Km (approx. 120 to
217 ± 24 × 10−6 S.I.), while the basal turbiditic unit display the lowest 150 × 10−6 S.I.), related to the non biogenic content (mainly clay
values (102 ± 22 and 124 ± 4 × 10−6 S.I.). minerals).
Here also, we have to underline that the highest values of magnet- To discuss relationships between the magnetic foliation and the mag-
ic foliation characterize the upper homogeneous unit (1.060 and netic content, first indications are given by the evolution of Km. On

0 40 80 120 1.02 1.04 1.06 1.08 1.1 -1 0 1


610
Median µm Foliation T
620

HmTu
630

640 h.d.

650 HmTu
Depth cm

h.d.
660

670
HmTu

680

690 h.d.
Tu
h.d.
HmTu

700

710 60 80 100 120 140 160 180 1 1.02 1.04 1.06 1.08 0 30 60 90
Magnetic susceptibility Lineation Inc Kmin
(10-6 S.I.)
Corinth Section-A
h.d.: Hemipelagic deposits HT.: re-depositional event Turbidite Homogenite
(turbidite+homogenite)

Fig. 5. AMS profiles of selected portions A and B from Core MD01-2477 (Gulf of Corinth); 2 cm sampling intervals. Volume magnetic susceptibility (Km) and a grain-size parameter
(5 mm sampling interval for both) are added. (stratigraphic position on Fig. 3). Interval B represents a unique thick event.
8 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

0 100 200 300 400 500 1 1.02 1.04 1.06 1.08 1.1 -1 0 1
1030 Median µm Foliation T
h.d.
1040

1050

1060
Depth cm

1070

HmTu
1080

1090

1100

1110

1120
50 100 150 200 250 300 1 1.02 1.04 1.06 1.08 0 30 60 90
Magnetic susceptibility Lineation Inc Kmin
(10-6 S.I.)
Corinth Section-B
h.d.: Hemipelagic deposits HT: re-depositional event Turbidite Homogenite
(turbidite, homogenite)

Fig. 5 (continued).

Fig. 4, the comparison of Km curves and magnetic foliation curves points Marmara. The basal turbiditic parts of homogenites + turbidites
out: opposite tendencies in coarse turbiditic layers, and independence be- (HmTu) have thicknesses between 1 and 26 cm, they show one or
tween Km and foliation in fine-grained successions (h.d. and Hm). Fig. 6A two fining upward sequences. Their median grain size ranges from
figures out these variations and the clear separation homogenites/ 400 to 10 μm (medium sand to fine silt), and the sorting index dis-
hemipelagites; a great variability of both parameters characterizes the plays high variability (1.5 to 5.5). The upper layers (homogenites)
coarse turbiditic samples. A more detailed approach, based on ARM and are 2 to 44 cm-thick. They have a relatively constant median grain
IRM, is used to discuss the influence of magnetic content and grain size size from 5.5 to 4.5 μm (very fine silt), and a sorting index ranging
on magnetic foliation (Fig. 7a and b). The highest values of JARM are from 1.95 to 2.30. The hemipelagic deposits have median grain size
generally present in the basal turbiditic units (1.43 × 10 − 2 ± ranging from 5.2 to 4.0 μm (very fine silt) and a moderate variably
0.004 and 1.14 × 10 − 2 ± 0.002 Am − 1). The HIRM values and the of sorting index from 2.1 to 2.5.
ratio Km/JIRM1T are variable, being observed the highest Km/JIRM1T Relationships between Km and grain size are significantly different
values in hemipelagites and homogenites (144 ± 30 and 118 ± from what we observe on the Sea of Marmara’s core: in all “events”
28 S.I. × 10–6 Am−1). The S-ratio is very constant, with values very MS is low in coarse turbiditic bases and high in fine-grained layers
close to 1 (0.96 ± 0,01), suggesting a predominance of low-coercivity (hemipelagites and homogenites). Nevertheless, in the unique thick
minerals (magnetite). The highest HIRM values in a basal turbiditic HmTu shown on Fig. 5b, significant variations occur in fine-grained
unit (Fig. 7a), and in hemipelagites (Fig. 7b), indicate the presence, in part (Hm) and in overlying hemipelagite.
minor proportion, of high-coercivity minerals (hematite, goethite). High Differences between the two sites are noticeable regarding MS vs
ARM intensity in the turbiditic layers, and its gradual decrease in foliation (Fig. 6a and b): a general increase of magnetic foliation ap-
homogenites, indicate that the variations of anhysteretic remanent mag- pears with increasing MS in hemipelagites and homogenites). This
netization are associated, either to grain size changes, or to magnetic min- could indicate a higher relative influence of magnetic content on
erals concentration changes. Beside, the variability of Km/JIRM1T ratio magnetic fabric especially in section B homogenite (blue squares);
does not seem to be related to changes of the magnetic particles size (he- nevertheless, for section A (blue circles) it is not the case.
matite or goethite); it rather appears related to their concentration. With respect to IRM and ARM (Fig. 8a and b), the evolution and
the relationships between fabric parameters and magnetic mineralo-
4.3.2. Core MD01-2477-Gulf of Corinth-sections (Figs. 6b and 8) gy markers is quite similar for both cores. The values of ARM, IRM 1 T
Median grain size values and variations (Fig. 5a and b) are quite and S-ratios along of the two sections show similar pattern. The
identical to what we observe on Core MD01-2425 from the Sea of highest values of these parameters are generally present in the
C. Campos et al. / Sedimentary Geology 292 (2013) 1–14 9

A 1.16

1.14

1.12

1.1 Ho
Foliation

He + T Marmara MD01-2425
1.08 Hemipelagic deposits
Section-A
Hemipelagic deposits
Section-B
1.06 Homogenites-Section-A
Homogenites-Section-B
Turbidites-Section-A
1.04
Turbidites-Section-B
Ho: Homogenites
He: Hemipelagites
1.02
T: Turbidites

1
100 200 300 400 500
Magnetic susceptibility
(10-6 S.I.)

B 1.1

1.08

Ho
1.06
He + T
Foliation

Corinth MD01-2477

Hemipelagic deposits
1.04 Section-A
Hemipelagic deposits
Section-B
Homogenites-Section-A
Homogenite-Section-B

1.02 Turbidites-Section-A
Turbidite-Section-B
Ho: Homogenites
He: Hemipelagites
T: Turbidites
1
50 100 150 200 250 300
Magnetic susceptibility
(10-6 S.I.)

Fig. 6. Magnetic fabric versus volume magnetic susceptibility for selected portions of Core MD01-2425 (6a, Sea of Marmara) and MD01-2477 (6b, Gulf of Corinth).

upper homogeneous units (ARM averages = 3.09 10 − 2 ± 0.039 predominance of low coercitivity minerals. Nevertheless, some
and 7 10 − 2 ± 0.029 Am- 1; and HIRM values = 1.4 ± 0.02 and high-coercitivity components (hematite, goethite) are present, in
0.18 ± 0.02 Am- 1). The Km/JRM1T ratio shows slight variations minor amount, as evidenced by a HIRM increase. The variation of
with relatively higher values in basal turbidititic terms (126 ± 18 the intensity of ARM can be associated with changes in the grain
and 86 ± 29 S.I. × 10–6 Am − 1). As for the Sea of Marmara, the size or concentration of magnetic mineral, whereas, the variability
S-ratio is quite constant and close to 1 (0.93 ± 0.01), indicating a in the Km/J IRM1T relation is not relate to the variability of the
10 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

A 80 160
Km/J IRM 1T (S.I.* 10-6 *m/A)
240 0 4E-2 8E-2
H IRM (A*m-1)
0.12
2630

HmTu
2640

2650

HmTu
2660
Depth cm

2670

2680

HmTu
2690

2700

8E-3 1.6E-2 2.4E-2 0.9 1 1 1.04 1.08 1.12 1.16


JARM (A*m-1) S-Ratio F

B
29 60 40 80 120 160 200 0 8E-2 0.16 0.24
Km/J IRM 1T (S.I.*10-6 *m/A) H IRM (A*m-1)

2970

HmTu
2980

Tu
Depth cm

2990
HmTu

3000

3010
HmTu

3020

6E-3 8E-3 1E-2 1.2E-2 1.4E-2 0.9 1 1.04 1.06 1.08 1.1 1.12 1.14
JARM (A*m-1) S-Ratio F

Fig. 7. ARM, Km/JRM1T, S-ratio and HIRM profiles of selected portions A and B from Core MD01-2425 (Sea of Marmara); 2 cm sampling intervals (stratigraphic position on Fig. 3).
Magnetic foliation is added to check its relationships with magnetic content.

grain magnetic size (hematite or gohetite). It seems related with the retrieved in northern Indian Ocean (MD90-943, in Collombat, 1993).
concentration of the magnetic particles. We plotted the value obtained for the different sections from Gulf
of Corinth and Sea of Marmara, at the depth they were found, assuming
5. Interpretation and discussion similar compaction (Fig. 9). The lowest magnetic foliation values we
obtained from hemipelagites, are slightly higher than for MD90-943
5.1. Comparison of investigated layers with published examples profile; thus, the hemipelagites from Cores MD01-2477 and MD01-2425
do not appear as displaying particularly low magnetic foliation.
To better assess the specificity of magnetic fabric of homogenites with We also compare our results with a first attempt done by Chapron et
respect to hemipelagites from the same deposition area, we searched for al. (1999 and in Beck, 2009) for a historical lacustrine homogenite (an
published similar studies. To compare with an almost entirely hemiplegic earthquake which occurred in 1822). These authors considered the “nor-
succession, we used the results from a similar deep giant piston core mal” continuous lacustrine sedimentation similar to marine hemipelagite.
C. Campos et al. / Sedimentary Geology 292 (2013) 1–14 11

A 40 80 120 160 200 0.4 0.8 1.2 1.6 2


610 Km/J IRM 1T (S.I.* 10-6 *m/A) H IRM (A*m-1)

620

HmTu
630

640

HmTu
650
Depth cm

660

HmTu
670

680

690

HmTu Tu
700

710
2E-2 3E-2 4E-2 0.9 1 1.02 1.04 1.06 1.08 1.1
JARM (A*m-1) S- Ratio F

B
0 40 80 120 160 200 8E-2 0.16 0.24
1030 Km/J IRM (S.I.*10-6*m/A) H IRM (A*m-1)

1040

1050

1060
Depth cm

1070

HmTu
1080

1090

1100

1110

1120
0 2E-2 4E-2 6E-2 8E-2 0.9 1 1 1.02 1.04 1.06 1.08 1.1
JARM (A*m-1) S - Ratio F

Fig. 8. ARM, Km/JRM1T, S-ratio and HIRM profiles of selected portions A and B from Core MD01-2477 (Gulf of Corinth); 2 cm sampling intervals (stratigraphic position on Fig. 3).
Magnetic foliation is added to check its relationships with magnetic content.

In this example, they measured, at 40 cm below the sediment/water in- one displayed by hemipelagites, both for marine hemipelagites and
terface, a magnetic foliation ranging from 1.015 to 1.020 in a calcareous for similar lacustrine fine-grained sediments (“hemipelagites” in a
and clayey mud. A homogenite (15 cm thick) intercalated within this broad sense). This specificity is observed for homogenites representing
hemipelagite, displays values between 1.035 and 1.040. They are lower the upper part of composite sedimentary events (HmTu), overlying a
than the here-obtained values, but they concern a very recent layer coarser turbiditic episode; but also when the latter is poorly developed.
(1822 AD). These authors also underlined a sharp contrast coinciding Furthermore, the along-core variations correspond to abrupt changes of
with a sharp boundary and considered the high values as anomalous magnetic foliation with contrasted values. This contrast is especially neat
with respect to expected diagenetic state. at the top of homogenites, where the limit homogenite/hemipelagite is
generally neither eye-visible nor underlined by other parameters (grain
5.2. Discussion size, composition). Measurement of isolated samples of hemipelagites
(sampling intervals ~60 cm) throughout MD01-2477 core and in the
As for the two above-mentioned examples, our results show that non marine section of MD01-2425 core do not show any increase of foli-
the magnetic foliation of homogenites is clearly distinct from the ation values with deep. Moreover, the regular oblate deposits fabrics and
12 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

Fig. 9. Comparison of magnetic fabric of the two investigated cores with an almost entirely hemipelagic succession (Core MD90-943 from Indian Ocean). This “reference” core were
retrieved with exactly the same device (corer, liner) as for the here-analyzed ones. Samples from Cores MD01-2425 and MD01-2477 (this work) have been plot at an equivalent
depth, assuming similar states of compaction. Hemipelagites from the three sites display low and very close values of magnetic foliation; homogenites display highly anomalous
foliation with respect to the expected compaction at the equivalent depth.

the position of Kmax parallel and Kmin perpendicular to the bedding plane re-suspension of fine fraction, “ponding” in deep basins subdivisions. As
suggest that these magnetic parameters are diagnostic of primary sedi- this process has been related to earthquake-triggering in isolated basins
mentary fabric, being not affected the described units by deformation or (see Introduction), the use of AMS for subaqueous paleoseismology
bioturbation. Therefore, a specific array of magnetic and clay particles appears reinforced. Furthermore, according to the vertical resolution of
may explain the strong foliation observed in all homogenites. In the two our analyses, AMS permits to precisely locate the limit between a
investigated cores, the control of a possible influence of composition homogenite and the overlying hemiplegic layer, being this limit often
and grain size on AMS, confirms a general independence between the not clearly visible and/or not identifiable with another proxy. Consider-
two groups of parameters and reinforces the proposed interpretation. ing that a HmTu may represent a co-seismic event, the thickness of an
We consider that the particular magnetic foliation of homogenites is hemipelagic interval separating two events may be precisely estimated.
mainly driven by the grain array, and, by mean, by the way of particles Thus, recurrence time interval measurement may also benefit from this
settling. technique.
As underlined by different authors (see Introduction), improving the
interpretation of magnetic fabric in terms of depositional/redepositional
6. Conclusions and perspectives process is necessary and has to be carried out through a complete control
of the magnetism carriers (terrigenous ?, early diagenetic ?, post-coring
On the two analyzed cores, AMS, measured with high resolution on modifications ?, size and shape of magnetic particles ? etc.). Shipboard
all intervals of sedimentary accumulation, displays strong variations measurements on fresh cores should be preferred. Further analyses
with layering at 10 to 30 m depth-below-seafloor (10 to 16 kyr BP de- will be conducted on the phyllosilicate contents and the shapes of silt
posits). Based on complementary measurements and observations, we and clay particles.
consider that – despite the differences between the two cores – the
dominant factor for the observed magnetic fabric is the grain array.
For fine-grained layers, the magnetic foliation values appear clearly Acknowledgements
divided into two groups: 1) low magnetic foliations, which can be
considered as “normal” with respect to burial depth, characterizing pri- The here-presented investigations were funded through different
mary hemipelagic deposits; 2) anomalously high magnetic foliations, grants: i) University of Savoie's AAP-2012-16 AGRASM Grant, dedi-
which are much higher than expected with a primary settling, signing cated to AMS sedimentological application, ii) Agence Nationale
homogenites. For both cases, the magnetic lineation is close to 1 and pour la Recherche ANR's SISCOR Project dedicated to the Gulf of
Kmin axis is sub-vertical. Corinth seismic hazards, iii) CNRS-INSU funding through ISTerre Lab-
The deduced strong planar horizontal array of homogenites fits with oratory. Our investigations also benefit from CNRS-INSU ARTEMIS facility
the hypothesis of a specific settling (Chapron et al., 1999; Beck, 2009 and for 14C dating. Corina Campos's PhD thesis and stay in ISTerre Laboratory
references therein): strong initial segregation from gravity-reworked is funded through Venezuela's FUNDAYACUCHO Grant 20093262.
soft sediments under oscillatory conditions, long lasting to almost stable Authors are grateful to these different institutes and governments.
C. Campos et al. / Sedimentary Geology 292 (2013) 1–14 13

They acknowledge two anonymous reviewers for improving their Chapron, E., Beck, C., Pourchet, M., Deconinck, J.F., 1999. 1822 earthquake-triggered
homogenite in Lake Le Bourget (NW Alps). Terra Nova 11, 86–92.
initial manuscript. Cita, M.B., Rimoldi, B., 1997. Geological and geophysical evidence for a Holocene tsu-
nami deposit in the eastern Mediterranean deep-sea record. In: Hancock, P.L.,
References Michetti, A.M., et al. (Eds.), Paleoseismology; understanding past earthquakes
using Quaternary geology. : Journal of Geodynamics, 24. 1–4, pp. 293–304.
Adams, J., 1990. Paleoseismicity of the Cascadian subduction zone: evidence from tur- Collier, R.E.L., Leeder, M.R., Trout, M., Ferentinos, G., Lyberis, E., Papatheodorou, G.,
bidites off the Oregon–Washington margin. Tectonics 9 (4), 569–583. 2000. High sediment yields and cool wet winters: test of last glacial paleoclimates
Armijo, R., Meyer, B., King, G.C.P., Rigo, A., Papanastassiou, D., 1996. Quarternary evolu- in the northern Mediterranean. Geology 28, 999–1002.
tion of the Corinth rift and its implications for the Late Cenozoic evolution of the Collombat, H., 1993. Etude des propriétés magnétiques des sédiments non consolidés:
Aegean. Geophysical Journal International 126, 11–53. anisotropie et erreurs d’inclinaison paléomagnétique. PhD Thesis, J. Fourier Uni-
Armijo, R., Meyer, B., Hubert, A., Barka, A., 1999. Westward propagation of the North versity Grenoble, 214 pp.
Anatolian Fault into the northern Aegean: timing and kinematics. Geology 27 De Martini, P.M., Pantosti, D., Palyvos, N., Lemeille, F., McNeill, L., Collier, R.E.L., 2004.
(3), 267–270. Slip rates of the Aigion and Eliki Faults from uplifted marine terraces, Corinth
Armijo, R., Meyer, B., Navarro, S., King, G., Barka, A., 2002. Asymmetric slip partitioning Gulf, Greece. Comptes-Rendus Geosciences 336, 325–334.
in the Sea of Marmara pull-apart: a clue to propagation processes of the Anatolian Demory, F., Oberhänsli, H., Nowaczyk, N.R., Gottschalk, M., Wirth, R., Naumann, R.,
Fault. Terra Nova 14, 80–86. 2005. Detrital input and early diagenesis in sediments from Lake Baikal revealed
Armijo, R., Pondard, N., Meyer, B., Uçarkus, G., Mercier de Lépinay, B., Malavieille, J., by rock magnetism. Global and Planetary Change 46 (1–4), 145–166.
Dominguez, S., Gutscher, M.-A., Schmidt, S., Beck, C., Çagatay, N., Çakir, Z., Imren, C., Drab, L., Hubert Ferrari, A., Schmidt, S., Martinez, P., 2012. The earthquake sedimentary
Eris, K., Natalin, B., Özalaybey, S., Tolun, L., Lefèvre, I., Seeber, L., Gasperini, L., Rangin, record in the western part of the Sea of Marmara, Turkey. Natural Hazards and
C., Emre, O., Sarikavak, K., 2005. Submarine fault scarps in the Sea of Marmara pull- Earth System Sciences 1235–1254. http://dx.doi.org/10.5194/nhess-12-2012
apart (North Anatolian Fault): implications for seismic hazard in Istanbul. Geochemis- (Special Issue “Subaqueous Paleoseismology” (D. Pantosti Edt.)).
try, Geophysics, Geosystems 6 (6), Q06009. http://dx.doi.org/10.1029/2004GC000896. Flerit, F., Armijo, R., King, G.C.P., Meyer, B., Barka, E., 2003. Slip partitioning in the Sea of
Barnes, P.M., Pondard, N., 2010. Derivation of direct on-fault submarine Marmara pull-apart determined from GPS velocity vectors. Geophysical Journal Inter-
paleoearthquake records from high-resolution seismic reflection profiles: national 154, 1–7.
Wairau Fault, New Zealand. Geochemistry, Geophysics, Geosystems 11, Ford, M., Rohais, S., Williams, E.A., Bourlange, S., Jousselin, D., Backert, N., Malartre, F.,
Q11013. http://dx.doi.org/10.1029/2010GC003254. 2013. Tectono-sedimentary evolution of the western Corinth rift (Central Greece).
Barrier, E., Chamot-Rooke, N., Giordano, G., 2004. Geodynamic map of the Mediterra- Basin Research 25 (1), 3–25.
nean. Sheet 1, Tectonics and Kinematics, CGMW-UNESCO. Goldfinger, C., Morey, A.E., Nelson, C.H., Gutierrez-Pastor, J., Johnson, J.E., Karabanov, E.,
Beck, C., 2009. Late Quaternary lacustrine paleo-seismic archives in north-western Chaytor, J., Eriksson, A., Shipboard Scientific Party, 2007. Rupture lengths and tem-
Alps: Examples of earthquake-origin assessment of sedimentary disturbances. poral history of significant earthquakes on the offshore and north coast segments
Earth-Science Reviews 96, 327–344. of the Northern San Andreas Fault based on turbidite stratigraphy. Earth and Plan-
Beck, C., Mercier de Lépinay, B., Schneider, J.-L., Cremer, M., Çağatay, N., Wendenbaum, etary Science Letters 254, 9–27.
E., Boutareaud, S., Ménot, G., Schmidt, S., Weber, O., Eris, K., Armijo, R., Meyer, B., Gorsline, D.S., De Diego, T., Nava-Sanchez, E.H., 2000. Seismically triggered turbidites in
Pondard, N., Gutscher, M.-A., and the MARMACORE Cruise Party, Turon, J.-L., small margin basins: Alfonso Basin, Western Gulf of California and Santa Monica
Labeyrie, L., Cortijo, E., Gallet, Y., Bouquerel, H., Gorur, N., Gervais, A., Castera, M.- Basin, California Borderland. Sedimentary Geology 135, 21–35.
H., Londeix, L., de Rességuier, A., Jaouen, A., 2007. Late Quaternary co-seismic sed- Grácia, E., Vizcaino, A., Escutia, C., Asioli, A., Rodes, A., Pallas, R., Garcia-Orellana, J.,
imentation in the Sea of Marmara's deep basins, in: Bourrouilh-Le Jan, F., Beck, C., Lebreiro, S., Goldfinger, C., 2010. Holocene earhquake record offshore Portugal
Gorsline, D. (Eds.), Sedimentary Records of Catastrophic Events. Spec. Iss. Sedi- (SW Iberia): testing turbidite paleoseismology in a slow-convergence margin.
mentary Geology, 199:65–89. Quaternary Science Reviews 29, 1156–1172.
Beck, C., Reyss, J.-L., Leclerc, F., Moreno, E., Feuillet, N. and GWADASEIS Cruise Scientific Hamilton, N., Rees, A.I., 1970. The use of magnetic fabric in paleocurrent estimation.
Party: Barrier, L., Beauducel, F., Boudon, G., Clément, V., Deplus, C., Gallou, N., Lebrun, Paleogeophysics 445 (S.K. Runcorn Edtr.).
J.-F., Le Friant, A., Nercessian, A., Paterne, M., Saurel, J.-M., Pichot, T., Vidal, C., 2012. Hancock, P.L., Barka, A.A., 1981. Opposed shear senses inferred from neotectonic
Identification of deep subaqueous co-seismic scarps through specific coeval sedimen- mesofractures systems in the North Anatolian fault zone. Journal of Structural
tation in Lesser Antilles: implication for seismic hazard. Natural Hazards and Earth Geology 3, 383–392.
System Sciences, Special Issue “Subaqueous Paleoseismology” (D. Pantosti Edt.), Hesse, R., 1977. Turbiditic and non-turbiditic mudstone of Cretaceous flysch sections of
http://dx.doi.org/10.5194/nhess-12-1-2012, 1755–1767. the East Alps and other basins. Sedimentology 22, 387–416.
Bell, R., McNeill, L., Bull, J.M., Henstock, T.J., Collier, R.E.L., Leeder, M.R., 2009. Fault ar- Hounslow, M., Bootes, P.A., Whyman, G., 1990. Remanent magnetization of sediments
chitecture, basin structure and evolution of the Gulf of Corinth Rift, central Greece. undergoing deformation in the Barbados accretionary prism: ODP Leg 110. In:
Basin Research 21, 824–835. http://dx.doi.org/10.1111/j.1365-2117.2009.00401.x. Moore, J.C., Mascles, A., et al. (Eds.), Proceedings of the Ocean Drilling Program, Sci-
Bernard, P., Briole, P., Meyer, B., Lyon-Caen, H., Gomez, J.-M., Tiberi, C., Berge, C., Catin, R., entific Results 110, 371–391.
Hatzfeld, D., Lachet, C., Lebrun, B., Deschamps, A., Courboulex, F., Laroque, C., Rigo, A., Hrouda, F., 1982. Magnetic anisotropy of rocks and its application in geology and geo-
Massonet, D., Papadimitriou, P., Kassaras, J., Diagourtas, D., Makropoulos, K., Veis, G., physics. Geophysical Surveys 5, 37–82.
Papazisi, E., Mitsakaki, C., Karakostas, V., Papadimitriou, P., Papanastassiou, D., Huh, C.A., Su, C.C., Liang, W.T., Ling, C.Y., 2004. Linkage between turbidites in thesouthern
Chouliaras, G., Stavrakakis, G., 1997. The Ms = 6.2, June 15, 1995 Aigion Earthquake Okinawa Trough and submarine earthquakes. Geophysical Research Letters 31,
(Greece): evidence for low angle normal faulting in the Corinth Rift. Journal of Seismol- L12304. http://dx.doi.org/10.1029/2004GL019731.
ogy 1, 131–150. Jackson, J.A., Gagnepain, J., Houseman, G., King, G.C.P., Papadimitriou, P., Soufleris, C.,
Bernard, P., Lyon-Caen, H., Briole, P., Deschamps, A., Boudin, F., Makropoulos, K., Virieux, J., 1982. Seismicity, normal faulting and the geomorphological develop-
Papadimitriou, P., Lemeille, F., Patau, G., Billiris, H., Paradissis, D., Castarède, H., ment of the Gulf of Corinth (Greece): the Corinth earthquakes of February and
Charade, O., Nercessian, A., Avallone, D., Pachiani, F., Zahradnik, J., Sacks, S., Linde, March 1981. Earth and Planetary Sciences Letters 57, 377–397.
A., 2006. Seismicity, deformation and hazard in the western rift of Corinth: new in- Jelinek, V., 1981. Characterization of the magnetic fabric of rocks. Tectonophysics 79,
sights from the Corinth Rift Laboratory (CRL). Tectonophysics 426, 7–30. 63–67.
Bloemendal, J., King, J.W., Hall, F.R., Doh, S.-J., 1992. Rock magnetism of Late Neogene Joseph, L.H., Rea, D.K., Pluijl, B.A.V., 1998. Use of grain size and magnetic fabric analyses to
and Pleistocene deep-sea sediments relationship to sediment source, diagenetic distinguish among depositional environments. Paleoceanography 13, 491–501.
processes, and sediment lithology. Journal of Geophysical Research 97, 4361–4375. Kastens, K., Cita, M.B., 1981. Tsunami-induced sediment transport in the abyssal
Borradaile, G., 1988. Magnetic susceptibility, petrofabrics and strain. Tectonophysics 156, Mediterrranean Sea. Geological Society of America Bulletin 92, 845–857.
1–20. Kent, D.V., Lowrie, W., 1975. On the magnetic susceptibility anisotropy of deep-sea sed-
Bouma, A.H., 1962. Sedimentology of Some Flysch Deposits. Elsevier Publishing (168 pp.). iments. Earth and Planetary Sciences Letters 28 (1), 1–12.
Briole, P., Rigo, A., Lyon-Caen, H., Ruegg, J., Papazissi, K., Mistakaki, C., Balodimou, A., Keraudren, B., Sorel, D., 1987. The terraces of Corinth (Greece) — a detailed record of
Veis, G., Hatzfeld, D., Deschamps, A., 2000. Active deformation of the Gulf of eustatic sea-level variations during the last 500 000 years. Marine Geology 77,
Korinthos, Greece: results from repeated GPS surveys between 1990 and 1995. 99–107.
Journal of Geophysical Research 105, 25605–25625. Levi, T., Weinberger, R., Aïfa, T., Eyal, Y., Marco, S., 2006. Earthquake-induced clastic
Bronk Ramsey, C., 2001. Development of the Radiocarbon calibration program OxCal, dikes detected by anisotropy of magnetic susceptibility. Geology 34, 69–72.
Radiocarbon. Proceedings of 17th International 14C Conference, 43, pp. 355–363. Londeix, L., Herreyre, Y., Turon, J.-L., Fletcher, W., 2009. Last Glacial to Holocene
Brooks, N., Ferentinos, G., 1984. Tectonics and sedimentology in the Gulf of Corinth and hydrology of the Marmara Sea inferred from a dinoflagellate cyst record. Re-
Zakynthos and Kefallinia channels, western Greece. Tectonophysics 101, 25–54. view of Palaeobotany and Palynology, 158(1–2). Elsevier B.V., pp. 52–71.
Campos, C., Beck, C., Crouzet, C., Carrillo, E., 2011. Characterization of Late Pleistocene- http://dx.doi.org/10.1016/j.revpalbo.2009.07.004.
Holocene earthquake-induced “homogenites” in the Sea of Marmara through mag- LØvlie, R., Lowrie, W., Jacobs, M., 1971. Magnetic properties and mineralogy of four
netic fabric. Implications for co-seismic offsets detection and measurements. Pro- deep-sea cores. Earth and Planetary Sciences Letters 15, 157–168.
ceedings 2nd INQUA-IGCP-567 International Workshop on Active Tectonics, Lykousis, V., Sakellariou, D., Moretti, I., Kaberi, H., 2007. Late Quaternary basin evolu-
Earthquake Geology, Archaeology and Engineering, Corinth, Greece, pp. 19–21. tion of the Gulf of Corinth: Sequence stratigraphy, sedimentation, fault–slip and
Carrillo, E., Beck, C., Audemard, F.A., Moreno, E., Ollarves, R., 2008. Disentangling Late Qua- subsidence rates. Tectonophysics 440, 29–51.
ternary climatic and seismo-tectonic controls on Lake Mucubají sedimentation Marco, S., Agnon, A., 1995. Prehistoric earthquake deformations near Masada, Dead Sea
(Mérida Andes, Venezuela). In: De Batist, M., Chapron, E. (Eds.), Lake systems: sedi- graben. Geology 23, 695–698.
mentary archives of climate change and tectonicPalaeogeography, Palaeoclimatology, McCalpin, J.M., 2009. Paleoseismology. International Geophysics Series, vol. 95. Aca-
Palaeoecology 259, 284–300. demic Press978-0-12-373576-8 (798 pp.).
14 C. Campos et al. / Sedimentary Geology 292 (2013) 1–14

McClusky, C.M.G., Reilinger, R., Mahmoud, S., Ben Sari, D., Tealeb, A., 2003. GPS con- Rees, A., 1965. The use of anisotropy of magnetic susceptibility in the estimation of
straints on Africa (Nubia) and Arabia plate motions. Geophysical Journal Interna- sedimentary fabric. Sedimentology 4, 257–271.
tional 155, 126–138. Reimer, P.J., McCormac, F.G., 2002. Marine radiocarbon reservoir corrections for the
McHugh, C.M.G., Seeber, L., Cormier, M.-H., Dutton, J., Cagatay, N., Polonia, A., Ryan, W.B.F., Mediterranean and Aegean Seas. Radiocarbon 44, 159–166.
Gorur, N., 2006. Submarine earthquake geology along the North Anatolia fault in the Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Bronk Ramsey,
Marmara Sea, Turkey: a model for transform basin sedimentation. Earth and Plane- C., Buck, C.E., Burr, G.S., Edwards, R.L., Friedrich, M., Grootes, P.M., Guilderson, T.P.,
tary Sciences 248, 661–684. http://dx.doi.org/10.1016/j.epsl.2006.05.038. Hajdas, I., Heaton, T.J., Hogg, A.G., Hughen, K.A., Kaiser, K.F., Kromer, B., McCormac,
McHugh, C., Seeber, L., Braudy, N., Cormier, M.-H., Davis, M.B., Diebold, J.B., Dieudonne, F.G., Manning, S.W., Reimer, R.W., Richards, D.A., Southon, J.R., Talamo, S., Turney,
N., Douilly, R., Gulick, S.P.S., Hornbach, M.J., Johnson III, H.E., Ryan Miskin, K., C.S.M., van der Plicht, J., Weyhenmeyer, C.E., 2009. IntCal09 and Marine09 radio-
Sorlien, C., Steckler, M., Symithe, S.J., Templeton, J., 2011. Offshore sedimentary ef- carbon age calibration curves, 0–50,000 years cal BP. Radiocarbon 51 (4),
fects of the 12 January 2010 Haiti earthquake. Geology 39 (8), 723–726. http:// 1111–1150.
dx.doi.org/10.1130/G31815.1. Robinson, S.G., 1986. The late Pleistocene palaeoclimatic record of North Atlantic deep-
McNeill, L.C., Cotterill, C.J., Henstock, T.J., Bull, J.M., Stefatos, A., Collier, R.E.L., Papatheodorou, sea sediments revealed by mineral magnetic measurements. Physics of the Earth
G., Ferentinos, G., Hicks, S.E., 2005. Active faulting within the offshore western Gulf of and Planetary Interiors 42, 22–47.
Corinth, Greece: implications for model of continental rift deformation. Geology 33, Rochette, P., Jackson, M., Aubourg, C., 1992. Rock magnetism and the interpretation of
241–244. anisotropy of magnetic susceptibility. Reviews of Geophysics 30 (3), 209–226.
Moretti, I., Lykousis, V., Sakellariou, D., Reynaud, J.-Y., Benziane, B., Prinzhoffer, A., 2004. Rochette, P., Vadeboin, F., Clochard, L., 2001. Rock magnetic applications of Halbach
Sedimentation and subsidence rate in the Gulf of Corinth: what we learn from the cylinders. Physics of the Earth and Planetary Interiors 126, 109–117.
Marion-Dufresne’s long-piston coring. Comptes-Rendus Géosciences 336 (4/5), Rodriguez Pascua, M.A., De Vicente, G., Calvo, J.P., Perez-Lopez, R., 2003. Similarities be-
291–299. tween recent seismic activity and paleoseismites during the late miocene in the
Mörner, N.A., Sun, G., 2008. Paleoearthquake deformations recorded by magnetic vari- external Betic Chain (Spain): relationship by the b value and the fractal dimension.
ables. Earth and Planetary Sciences Letters 267, 495–502. Journal of Structural Geology 25, 749.
Mutti, E., Ricchi-Lucchi, F., 1978. Turbidites of the northerne Apennines. Introduction to Sakellariou, D., Lykousis, V., Alexandri, S., Kaberi, H., Rousakis, G., Nomikou, P., Georgiou, P.,
facies analysis. International Geological Review 20, 25–166. Ballas, D., 2007. Faulting, seismic-stratigraphic architecture and Late Quaternary evo-
Noda, A., Tuzino, T., Kanai, Y., Furukawa, R., Uchida, J., 2008. Paleoseismicity along the lution of the Gulf of Alkyonides Basin, East Gulf of Corinth, Central Greece. Basin Re-
southern Kuril Trench deduced from submarine-fan turbidites. Marine Geology search 19, 273–295. http://dx.doi.org/10.1111/j.1365-2117.2007.00322.x.
254, 73–90. Sarı, E., Çağatay, M.N., 2006. Turbidities and their association with past earthquakes in
Palyvos, N., Pantosti, D., De Martini, P.M., Lemeille, F., Sorel, D., Pavlopoulos, K., 2005. the deep Çınarcık Basin of the Marmara Sea. Geo-Marine Letters 26, 69–76.
The Aigion-Neos Erineos coastal normal fault system (west Corinth Gulf Rift, Schneider, D.A., Kent, D.V., 1988. Inclination anomalies from Indian Ocean sediments
Greece): geomorphological signature, recent earthquake history and evolution. and the possibility of a standing nondipole field. Journal of Geophysical Research
Journal of Geophysical Research, B 110, B09302. 93 (B10, 11), 621–630.
Papadopoulos, G.A., 2003. Tsunami hazard in the Eastern Mediterranean: strong earth- Sengör, A.M.C., Tuysuz, O., Imren, C., Sakinc, M., Eyidogan, H., Gorur, N., Le Pichon, X.,
quakes and tsunamis in the Corinth Gulf, Central Greece. Natural Hazards 29 (3), Rangin, C., 2004. The North Anatolian Fault. A new look. Annual Review of Earth
437–464. and Planetary Sciences 33, 1–75.
Papadopoulos, G.A., Chalkis, B.J., 1984. Tsunamis observed in Greece and the surround- Siani, G., Paterne, M., Arnold, M., Bard, E., Metivier, B., Tisnerat, N., Bassinot, F., 2000.
ing area from Antiquity up to the present times. Marine Geology 56, 309–317. Radiocarbon reservoir ages in the Mediterranean Sea and Black Sea. Radiocarbon
Papadopoulos, G.A., Plessa, A., 2000. Magnitude–distance relations for earthquake- 42 (2), 271–280.
induced landslides in Greece. Engineering Geology 58, 377-38. Stefatos, A., Papatheodorou, G., Ferentinos, G., Leeder, M., Collier, R., 2002. Seismic re-
Papatheodorou, G., Ferentinos, G., 1993. Sedimentation processes and basin-filling de- flection imaging of active offshore faults in the gulf of Corinth, their seismotectonic
positional architecture in an active asymmetric graben: Strava graben, Gulf of Cor- significance. Basin Research 14 (4), 487–502.
inth, Greece. Basin Research 5, 235–253. Stoner, J.S., Channell, J.E.T., Hillaire-Marcel, C., 1996. The magnetic signature of rapidly
Perissoratis, C., Mitropoulos, D., Angelopoulos, I., 1984. The role of earthquakes in in- deposited detrital layers from the deep Labrador Sea: relationships to North Atlan-
ducing sediment mass movements in the eastern Corinthiakos Gulf: an example tic Heinrich layers. Paleoceanography 40, 276–286.
from the February 24–March 4 activity. Marine Geology 55, 35–45. Sturm, M., Siegenthaler, C., Pickrill, R.A., 1995. Turbidites and “homogenites”: a concep-
Perissoratis, C., Piper, D.J.W., Lykousis, V., 1993. Late Quaternary sedimentation in the tual model of flood and slide deposits. Publ. International Association of Sedimen-
Gulf of Corinth: the effects of marine-lake fluctuations driven by eustatic sea tologists, 16th Regional Meeting, Paris, 22, p. 40.
level changes. Special Publication Dedicated to Prof. A. Panagos.Nat. Tech. Univ. Tarling, D.H., Hrouda, F., 1993. The Magnetic Anisotropy of Rocks. Chapman & Hall,
of Athens 693–744. London (218 pp.).
Perissoratis, C., Piper, D.J.W., Lykousis, V., 2000. Alternating marine and lacustrine sed- Thunell, R., Tappa, E., Valera, R., Llano, M., Astor, Y., Muller-Karger, F., Bohrer, R., 1999.
imentation during late Quaternary in the Gulf of Corinth rift basin, central Greece. Increased marine sediment suspension and fluxes following an earthquake. Nature
Marine Geology 167, 391–411. 398, 233–236.
CHAPTER 3

The Late Pleistocene/Holocene sedimentary record of the


central and eastern Sea of Marmara (Orta and Çinarcik basins)

3.1. Introduction

The Sea of Marmara and their surrounded areas is a seismically active region
located in north-western Turkey. It is crossed by the North Anatolian Fault (NAF), which
is one of the most active and important structures of the eastern Mediterranean region (e.g.
Le Pichon et al., 2001; Armijo et al., 2002; Fraser et al., 2010). This region has suffered a
significant amount of destructive large earthquakes, as is revealed by the historical sources.
These sources have been completed by the paleoseismological studies conducted in
numerous trenches and submarines researches, providing information about historic and
prehistoric record (e.g. Ambraseys and Finkel, 1991; Rockwell et al., 2001; Klinger et al.,
2003; McHugh et al., 2006; Sari and Ça÷atay, 2006).

This work presents a contribution to subaqueous paleoseismological study of the


Sea of Marmara along long piston cores. It tries to identify the traces left by
paleoearthquakes in the sedimentary record, as previously achieved by different authors
(Hiscott and Aksu, 2002; McHugh et al., 2006; Sari and Ça÷atay, 2006; Beck et al., 2007;
Drab et al., 2012; Ça÷atay et al., 2000; 2003; 2009; 2012; Eriú et al., 2012). These previous
studies were based on short gravity cores analyses, high-resolution bathymetry and seismic
profiling. The longer cores here studied permitted to investigate a long lasting archive.

Beck et al. (2007) proposed a preliminary study of giant piston cores obtained
during the MARMACORE Cruise, reaching a higher temporal span (at least 20 kyr). They
tentatively identified earthquake-induced events and their distribution in time, in the
Central Basin. Thus, the present work will be focus: 1) on the identification of the
earthquake-related traces recorded in the MD01-2425 core collected in the Çinarcik Basin
(Eastern Basin), 2) apply the same detailed approaches to the Central Basin. We will apply

43
the methodology proposed in Chapter 2: to identify Homogenites+Turbidites (HmTu)
layers and to distinguish them from the “normal” hemipelagic-type sedimentation through
Anisotropy of Magnetic Susceptibility (AMS). Following the identification of these shock-
induced events we attempt to deunitine the earthquake recurrence interval for this basin. In
addition, we try to understand how and when was the last transition between the non-
marine to marine environment in the Sea of Marmara, and its impact over the transport and
deposition dynamics of the basin. A comparison between the paleoseismological record of
the Central Basin (Beck et al., 2007) and that proposed for the Çinarcik Basin (present
work) will be performed. Finally; we attempt to measure the vertical fault offset of the
North Anatolian Fault (NAF) in the Central Basin through the co-seismic sedimentary
episodes, applying the methodology proposed by Beck et al. (2012) in the Lesser Antilles.
This method allows decipher the co-seismic vertical fault displacements from those whose
origin is aseismic (fault creep). The deunitination of the existence or not of a significant
vertical component in the movement of the NAF in the Central Basin will contribute to
clarify the existing debate about the geometry of this active fault in the Sea of Marmara
and their level of segmentation; which control the magnitude of future events.

3.2. Geodynamic and geologic overview of the Sea of Marmara region

3.2.1. The “Anatolia Microplate” and the North Anatolian Fault (NAF)

Present-day kinematics (figure III-1) and actives tectonic structures led several
authors to distinguish an “Anatolia” microplate. The convergence of Arabia and Eurasia
caused – among other major Circum-Mediterranean structures – the Aegean subduction
and the extrusion of the Anatolian Plateau to the west. This extruding, and partly rotating,
block is bounded to the north by the right-lateral North Anatolian Fault system (NAF) and
to the southeast by the left-lateral East Anatolian Fault system (figure III.2). The NAF is
thus a major dextral transform fault (McKenzie, 1972; ùengör, 1979), along which relative
motion between the Anatolian and Pontide blocks belonging to the Eurasian plate is
accommodated (Sengör, 1979).

44
Figure III.1: Synthesis of Present kinematics based on. From Noquet (2012).

Figure III.2: Tectonic setting of continental extrusion in eastern Mediterranean. Anatolia-Aegean block
escapes westward from Arabia-Eurasia collision zone, toward Hellenic subduction zone (from Armijo et
al., 1999). In Aegean, two deformation regimes are superimposed, widespread, slow extension starting
earlier (orange stripes, white diverging arrows), and more localized, fast transtension associated with
later, westward propagation of North Anatolian fault (NAF). EAF—East Anatolian fault, K—Karliova
triple junction, DSF—Dead Sea fault, NAT—North Aegean Trough, CR—Corinth Rift. Box outlines
Marmara pull-apart region.

Hubert-Ferrari et al. (2003) proposed a timing of the propagation of the NAF from
east to west at a rate of about 20 cm/yr (figure III.3). In their model, the propagation of the

45
NAF started in the east during the uplift of the Anatolian Plateau and after the formation of
the East Anatolian Fault at least about 13 Ma. The movement in the central part of the
NAF postdates 8.5 Ma, and in its western part, around the Sea of Marmara, movement
started about 5 Ma; while further to the west, the Gulf of Corinth was reactivated at 1 Ma
(Armijo et al., 1999).

The slip rate estimates in the NAF may have a slight variation depending on how
they were measured and the time period over which the rate is estimated (Kozacõ et al.,
2009). Geological studies indicate that the slip rate on the NAF over the past 1 to 10 kyr
was about 10 to 22 mm/yr (Hubert et al., 1997; Polonia et al 2004; Kozaci et al., 2009;
2011), whereas geodetic studies reveal somewhat present-day faster slip rates of ~24 to 29
mm/yr (McClusky et al., 2000; Reilinger et al., 2006). The NAF shows a total right-lateral
offset of ~85 km (Sengor et al. 1985, Hubert-Ferrari et al., 2003).

Figure III.3: Propagation of the North Anatolian Fault (NAF)(from Hubert-Ferrari et al., 2003).
Initiated in the east during the uplift of the Anatolian Plateau and after the formation of the East
Anatolian Fault (1, 2). The NAF propagated across north Anatolia (3) and reached the Sea of
Marmara 5 Ma ago (4) and the Gulf of Corinth 1 Ma ago (5). When the NAF entered the Aegean
region, it interacted with the existing active extensional structures. Black lines represent present
plate-boundary geometry as introduced in the modelling. The present (gray) and past (dashed)
(Armijo et al. 1996), coastlines of Greece and Western Turkey are indicated.

Geometrically, the NAF extends for ѝ1500 km from the Karliova triple junction in
eastern Anatolia to the Aegean Sea, paralleling roughly the southern Black Sea shores,
with an upward-convex trajectory (Sengör et al., 2004) (figures III.2). The easternmost
46
part of the NAF runs from Karlõova to Erbaa with a strike of N110°E, mostly as a
continuous trace, but is disrupted by releasing step-overs which leads to the creation of
various Neogene to Pleistocene extensional basins (Suzanne et al., 1990). In its central
section the NAF is characterized by a change in its main direction, from N110° to N75°. It
runs N75°E in a fairly linear trend all the way to Bolu where the NAF divides into two
major fault strands known as northern and southern NAF (Suzanne et al., 1990) (figure
III.2).

The major geometric features of the NAF coincide with three different seismo-
tectonics behaviors identified by Fraser et al. (2010) in the NAF: the eastern
transpressional, the central translational, and the western transtentional sections. In the
transpressional section (from Karliova to Erbaa), the earthquake recurrence rate is longer,
with a variable fault rupture patterns over time and long fault ruptures lengths
(compressional influence caused by the collision of Arabia and Anatolian). The
translational section (from Erbaa to Bolu) have little compression or extensional influence.
It shows sequences of rupture in long segments, successive and fast. The transtensional
section (from Bolu to the Aegean Sea) is characterized by short rupture lengths with
recurrence rate relatively fast. It is caused by the interaction of the NAF with the west
Anatolian extensional province, which results in relatively low shear stress failure
thresholds.

Earthquakes along the NAF have caused destruction in the region throughout
history (see subchapter of historical seismicity). The seismicity along the active segments
of the NAF is characterized by frequent moderate to large earthquakes (M>7), most of
them accompanied by surface ruptures (Barka and Kandiski-Cade, 1988), with focal
mechanisms that show especially right-lateral strike-slip solutions and a few normal
faulting solutions (Sengör et al., 2004). Most of the NAF ruptured during the 20th century
show a spatio-temporal progressive failure pattern, generally westward-propagating. The
last sequence comprises the 1939 (Ms=7.9), 1942 (Ms=7.1), 1943 (Ms=7.6), 1944
(Ms=7.3), 1957 (Ms=7.0), 1967 (Ms=7.1), 1992 (Ms=6.8) and 1999 (Ms=7.4 and Ms=7.2)
earthquakes (Ambraseys, 1970; Ambraseys and Finkel, 1995; Stein et al., 1997; Barka,
1999; Kozaci et al., 2011). Taking into account the 1912 Gulf of Saros earthquake to the
west, then, the succession of earthquakes from east to west, ending in the 1999 Izmit event,
a large seismic gap can be defined in the Sea of Marmara (figure III.4). This fact has

47
promoted the interest of the scientific community for a better understanding of the
formation and tectonic evolution of the Sea of Marmara together with identification of
major seismogenic faults in the sea, in order to evaluate the seismic hazard of the region
(Hubert-Ferrari et al., 2000; Parsons et al., 2000, 2004; Pondard et al., 2007).

Figure III.4: Western part of the 1939-to-1999 earthquakes “sequence” along the NAF, displaying
the “Marmara seismic gap”.Simplified from Uçarkus (2010)

In order to better interpret the sedimentation, it is also necessary to take into


account the geological setting of the Sea of Marmara and the long unit regional evolution,
as they directly concern sedimentation through terrigenous supply. The following will
summarize the main geological and tectonic aspects of this area.

3.2.2 Tectonic and geology of Turkey

The tectonic and geologic history of Turkey is intimately related with the evolution
of the Tethys Ocean. This ocean was located between the two large continents of
Gondwana in the south and Laurussia in the north, at least since 350 Ma. It was not a
single wide ocean. Rather it was formed by several relatively narrow oceanic seaways
separated by island arcs or continental slices, called terranes (Sengör and Yilmaz, 1981,

48
Okay, 2008). The final closure of the Tethyan oceans and the collision between the
different terranes resulted in the Alpine orogeny and the creation of Anatolia as a single
landmass during the Oligocene (Okay, 2008).

Turkey is constituted by different terrains: (1) The Pontides terrain to the north, (2)
the Anatolides - Taurides terrain in the center, (3) and the Arabian Platform (Ketin, 1966;
Okay, 2008) (figure III.5). These terrains are now separated by sutures, which mark the
tectonic lines or zones along which the oceans have disappeared. The sutures located
between and within the Pontides and Anatolides-Taurides terrains (Intra-pontide Suture,
Izmir-Ankara-Erzincan Suture, Inner-Tauride suture, indicated with a dotted red line in the
figure III.5) were originated from different stages of subduction of Tethys Ocean. (Dewey
et al., 1973). The age of these sutures are controversial, and can be diachronous, starting at
Late Cretaceous and continues into the Eocene (Okay, 1989). The south Assyrian-Zagros
Suture represents the closure of the Blitis Ocean due to collision between the Arabian and
Eurasian Plates, during the Middle Miocene (indicated with a dotted blue line in the figure
III.5) (Sengör and Yilmaz, 1981, Okay, 2008).

The Pontides terrain exhibits Laurasia affinities, and was located north of the
northern branch of the Neo-Tethys. It consists of three tectonics units which different
geological evolutions: The Strandja, østanbul and Sakarya units (figure III.5). These units
will be explained in detail in the Geology of Marmara subchapter. The Anatolides-
Taurides terrain show Gondwana affinities but were separated from the main mass of
Gondwana by the southern branch of Neo-Tethys (Okay, 2008).

The collision between the Arabian and Eurasian Plates had very profound effects on
the overall tectonics of Turkey. It is responsible of a north-south shortening and important
thickening across eastern Turkey (Sengör and Yilmaz, 1981). In response to these
conditions an important part of Anatolia wedged out westward along two new plate
boundaries, the North Anatolian and East Anatolian transform faults (NAF and EAF),
giving birth to the “Anatolian block” (Sengör and Yilmaz, 1981). During the Neogene the
westward escape of the Anatolian Plate along its boundary structures (NAF and EAF); as
well as, the regional extension associated with the southward migration of the Hellenic
trench, which result in the opening of the Aegean Sea, are the elements that constitute the
present day neotectonic framework of Turkey (Okay, 2008).

49
Figure III.5: Distribution of different terrains, accretionary complexes and ophiolites in Turkey (modified from Okay et al. 2006). The Pontides terrain
is represented by a yellow area, the Anatolides – Taurides by a blue area, and the Arabian Platform by a pink area. The Izmir-Ankara-Erzincan Suture
50

by a dotted red line and the Assyrian-Zagros Suture by a dotted blue line. A green area within the Sakarya zone, in the southern margin of the Sea of
Marmara, indicates the location of a portion of the Karakaya Complex and the blueschist and eclogites of the Bandirma region.
3.2.3 Geology of the Marmara region.

The Sea of Marmara and surrounding area have a complex basement, which was
subsequently covered by lower Eocene to Recent sediments and associated magmatic rocks
(Siyako et al., 1989; Wong et al., 1995). The basement sedimentary cover is more susceptible
to be transported to the different sub-basins of the Sea of Marmara (Wong et al., 1995). The
lithology of the different outcropping units may be summarized as follow:
The Strandja zone consists of Paleozoic quartzo-feldespathic gneisses intruded by
granitoids (Sunal et al., 2006; Okay, 2008), unconformably overlain by Triassic - Jurassic
continental to shallow marine sediments (Okay, 1989, 2008). This sequence is unconformably
overlain by Cretaceous volcanic rocks (ùengör and Yõlmaz, 1981). This zone is separated
from the Istanbul zone by a small area of undeformed Eocene sediments (Okay, 1989).
The Istanbul zone show a late Precambrian crystalline basement made of gneiss,
amphibolite, metavolcanic rocks, metaophiolite and granitoids (Chen et al., 2002; Yi÷itbaú et
al., 2004). It is unconformably overlain by continuous and well-developed Ordovician to
Carboniferous sedimentary succession, and by Permian - Triassic continental red beds and
limestones; which are unconformably overlain by Cretaceous - Paleocene clastic, carbonate
and andesitic rocks (Okay, 1989, 2008).
The Sakarya zone is characterized by a variable metamorphosed and strongly
deformed Permian and Triassic basement called Karakaya Complex overlained with a major
unconformity by Jurassic conglomerates and sandstones, which passes up to Jurassic –
Cretaceous limestones and Upper Cretaceous flysch (Okay, 1989). The Karakaya Complex is
generally metamorphosed in the low greenschist facies. Although, in some regions
metamorphism reaches epidote - amphibolite, blueschist and eclogite facies (Okay and
Göncüo÷lu, 2004). The blueschist and eclogite-facies rocks form tectonic slices, and has been
described in the southern margin of the Sea of Marmara (Okay and Monié, 1997) (figure
III.5).
The Thrace Basin is a large Tertiary depression located between the Strandja, Sakarya
and Istanbul zones. Its infill comprises sedimentary rocks of Eocene – Oligocene age (Görür
and Okay, 1996; Okay, 2008). Their cumulative thickness has been estimated at more than 9
km (Turgut et al., 1991). The basin is flanked predominantly by Triassic to Tertiary rocks,
overlaying the Triassic to Jurassic metamorphic rocks of the Strandja massif. The sedimentary
rocks are composed predominantly of an upward shallowing sequence of sandstone and shale.

51
The sequence starts with deep marine turbidites of Eocene age and ends with continental
Oligocene sandstone, shale and lignite (Okay, 2008).

3.2.4. The North Anatolian Fault within the Sea of Marmara

The Sea of Marmara is located in the western part of the 1600-km long North
Anatolian Fault Zone and forms a 270-km long and 80-km wide intracontinental marine basin
between the Mediterranean Sea and the Black Sea (Halbach et al., 2002). In the Sea of
Marmara, the NAF splits into three strands that locally trend east-west (Yaltõrak, 2002,
Armijo et al., 2002; Muller and Aydin, 2005). These branches unitinate in the northern
Aegean Sea (Yaltõrak, 2002) (figure III.2 and III.3). The northern strand of the North
Anatolian Fault extends from Bolu to Izmit and Gulf of Saros. The middle branch is situated
along the Iznik Lake and the Gemlik Bay (Yaltõrak and Alpar, 2002), while the southern
branch crosses Bursa (an industrial heartland and densely populated city of Turkey) (Yaltõrak
and Alpar, 2002). GPS measurements indicate that the northern zone is the most active with
an approximate slip rate of 10-15 mm/yr, the middle branch shows 5-10 mm/yr of distributed
deformation, whereas the southern fault strand accommodates 2-4 mm/yr of dextral slip
(Straub et al., 1997).

There are different views concerning the fault configuration of the NAF within the Sea
of Marmara, with direct influence in the estimation of the seismic hazard in the region. Early
models, based on a limited number of 2-D seismic reflection profiles, proposed a series of
northeast-southwest trending strike-slip faults and northwest-southeast trending normal faults
that bound the individual basins within the Marmara trough (Barka, 1992; Ergün and Özel,
1995; Wong et al., 1995). Recent models can be classified in three groups (Yaltõrak, 2002).
They are summarized and illustrated in figure III.6 from Muller and Aydin, (2005):
(1) models incorporating en échelon fault segments (Parke et al., 1999; Okay et al., 2000).
These authors proposed models with northern and southern fault strands bounding the
northern Marmara trough with different fault patterns between the two bounding
structures (Muller and Aydin, 2005) (figure III.6a).
(2) models with a single strike-slip fault cutting longitudinally through the entire Marmara
deep part (Le Pichon et al., 2001; Imren et al., 2001) (figure III.6b).
(3) a pull-apart model (Armijo et al., 2002). These authors concluded, from the same data
sets studied by Imren et al. (2001) and Le Pichon et al. (2001), that no evidence for a

52
single, continuous, purely strike-slip fault exists and that fault segments bounding pull-
apart features remain active (Muller and Aydin, 2005) (figure III.6c).

Figure III.6: Active fault patterns proposed for the Sea of Marmara (from Muller and Aydin, 2005).
a) Fault interpretation from Okay et al. (2000) (solid lines) and Parke et al. (1999) (dashed). b)
Fault interpretation from Le Pichon et al. (2001) (solid lines) and Imren et al. (2001) (dashed). c)
Fault interpretation from Armijo et al. (2002). Topography DEM derived from Gtopo30 global data
set. High resolution bathymetry (Rangin et al., 2000). Depth scale is only represented in c. In a, TB:
Tekirda÷ basin; CMB: central Marmara basin; KB: Kumburgaz basin; CB: Çinarcõk basin; IB:
imrali basin.

53
The location, nature, level of segmentation, and slip partitioning of active fault strands
in the Sea of Marmara will govern the magnitude of the future events. The earthquake
produced in a single continuous strike-slip fault (Le Pichon et al., 2001) will be larger than
that produced in a segmented fault models (Parke et al., 1999; Okay et al., 2000; Armijo et al.,
2002). For this reason, some authors have attempted to identify which model best produces
the observed deformation pattern within the Sea of Marmara (e.g. Muller and Aydin, 2005;
Carton et al., 2007; Laigle et al., 2008). These studies suggest that the single continuous
strike-slip fault model (Le Pichon et al., 2001) do not appear to sustain the new observations.

3.3. The Sea of Marmara: overview of morpho-bathymetry, active tectonic


structures, and sedimentary fill.

In this section we present the general features driving the sedimentation in the Sea of
Marmara, and especially the morpho-bathymetry and deeper structures of the Central and
Çinarcik sub-basins. Additionally, we will make a brief overview of previously published data
about sedimentary paleoseismic archives and about the Late Quaternary to present day
predominant oceanographic conditions.

3.3.1. Morpho-bathymetric and tectonic features of the Sea of Marmara

The Sea of Marmara is an East-West elongated 270 km-long basin developed along
the North Anatolian Fault (Hancock and Barka, 1981). It comprises three main basins up to
1276 m-deep (figure III.6). From west to east the basins are: the Tekirda÷ or Western Wasin
(maximum depth 1152 m), Central (1265 m depth) and Çinarcik or Eastern Basin (1276 m
depth). They are separated by two major structural highs at about 700 m depth, the Western
and Central Highs (Wong et al., 1995; Armijo et al., 2002; Yaltõrak, 2002a, Zitter el al.,
2012). The Sea of Marmara can be considered as a gateway between the Mediterranean Sea
and the Black Sea. It is connected to the Black Sea through the Straits of Bosphorus, at a 40
m depth, and to the Aegean Sea through the Strait of Dardanelles at a 70 m depth. Figure
III.7 shows a striking difference between the northern (15 km wide) and southern (50 km
wide) shelves. The shelf break is located at a depth of about -110 m (Ça÷atay et al., 2000;
McHugh et al., 2008). The slopes of the northern margin are generally steeper and less
covered with sediments than the southern ones. The steepest slopes are located on the north-
western side of Tekirda÷ Basin, and on the northern and southeastern sides of Çinarcik Basin
54
(gradient of a 20°). The Central Basin exhibits slope gradients of about 10°, and the southern
and northern sides of Tekirda÷ Basin have slope gradients of a 8° (Zitter et al., 2012).

Figure III.7: Bathymetric map and active faults in the North Marmara Basin. The data was
collected during the French–Turkish cruise carried out aboard R.V. Suroît in September-October
2000, and published by Armijo et al. (2002). It shows the location of the basin, become smaller
towards the west. Fnw, fn, fi, fls: fault nomenclature according to Carton et al. (2007), explained in
the text.

3.3.1-a The Çinarcik Basin


It is the largest sub-basin of the Sea of Marmara. It is a wedge-shaped basin oriented
N110°E, about 50 km long and up to 15–18 km wide. The maximum seafloor depth is of 1276
m. It is bounded on its north and south sides by large topographic escarpments and to the west
by the topographic Central high which isolates it from the Central Basin (Carton et al., 2007).
Numerous canyons and mass wasting cut the steepest slopes of the basin (Zitter et al., 2009;
see subchapter 3.7.3).

Based on seismic profiles and high resolution bathymetric data, Carton et al. (2007)
interpret the present-day tectonic activity of the Çinarcik Basin as governed by transtension
between two strike-slip segments: the Izmit fault in the east (fi), which enters through the
Gulf of Izmit, and the linear strike-slip transfer fault connecting the Çinarcik Basin to the
Central Basin in the west (fnw) (figure III.7). Along its northeastern and southern margin, the
basin is bounded by faults having a significant extensional component of slip. These faults are
named ‘‘fn’’ and ‘‘fs1,’’ respectively (figure III.7).

Deep seismic reflection data analyzed by Laigle et al. (2008) imaged a thick infill of
the Çinarcik Basin, its basement, and the whole crust over 30 km depth; a clear basement
reflector deepening northeastwards under the sedimentary infill is also revealed. A
55
geometrical fanning of the sedimentary infill above this dipping basement attests to a strong
component of extension in normal faulting at the NE Çinarcik Fault (Laigle et al., 2008)
(figure III.8b). According Okay et al. (2000) and Carton et al. (2007) the sedimentary infill in
this basin is at least 4 km thick, and consist of Pliocene-Quaternary sediments. The figure
III.8b reveals local north-dipping small normal-faults at the southern Çinarcik rim, which are
responsible for the basement outcrop at sea-bottom in its footwall and deformation of recent
sediments. These faults are conjugate to the major fault at the northern rim. Laigle et al.
(2008) conclude that these faults might be viewed as petals of a large scale negative flower-
structure that spreads over a width of 30 km at surface and roots into the lithosphere.

These faults that were mapped at the eastern half of the Çinarcik Basin from deep
profiles can also be observed at shallow depths, as is shows in the figure III.8c (profile s30),
where the seismic sparker reflection data have a shallow penetration of a 400 m (Rangin et
al., 2001), and in the figure III.8d (profile c14), where a high resolution seismic reflection is
obtained with the 3.5 kHz sediment penetrator, which images, with good accuracy, the
structure at a50 m depth below the seafloor (Armijo et al., 2005). Both figures (III.8c, d)
display the same northeastwards deepening of the sedimentary basin. They are characterized
by several “en echelon” faults offsetting its southern rim and a dominant northern boundary
fault, that creates the asymmetric sedimentary deposition, due to it strong normal faulting
component. Toward the western half of the Çinarcik Basin this pattern slightly changes, being
observed a smaller increase in sediment thickness toward the north. This can be observed in
the high resolution 3.5 kHz seismic profile c10b+p01 (figure III.8e). This phenomenon has

also been described by Carton et al. (2007) and Uηarkuú (2010) indicating a deepening of the

basin to the northeast.

56
Figure III.8: Seismic profiles across The Çõnarcõk Basin. a) Shaded elevation topography,
bathymetry and active faults of the Çõnarcõk basin, with location of seismic profiles and MD01-2427
and MD01-2425 cores. b) Depth seismic profile SM36 (from Laigle et al., 2008). c) Sparker
reflection data, profile s30 (from Rangin et al., 2001). d and e) 3.5 kHz high resolution profiles
(profile c14 and profile c10b+p01) acquired during the MARMARASCARPS cruise 2001.

57
3.3.1-b The Central (Orta) Basin
The Central Basin is the second largest basin along the Northern Marmara fault
system. According to Armijo et al. (2002) this basin is a pull-apart structure, where an internal
pull-apart has been formed (figure III.9). En échelons fault scarps enclose a pull-apart with a
distinctive rhomb shape. The individual scarps striking NW–SE show clear normal fault
morphology across the most recent sediment. They average 1–2 km long and the flat floor of
the pull-apart is 20–60 m deeper than the surrounding seafloor (Armijo et al., 2002). Clear
young fault scarps with comparable strike also run at the base of the larger escarpments
located at the edges of the Central Basin. Both the inner and outer faults of the basin have
normal components of slip and seem to spread from linear right-lateral fault segments striking
ENE–WSW, which are located east and west of the pull-apart structure (figure III.9). The
extensional step-over between these two main strike-slip segments is about 4 km (Armijo et
al., 2002).

Deep seismic reflection data interpreted by Laigle et al. (2008) and Bécel et al. (2010)
over the Central Basin reveal a deep sedimentary basin of about 16 km long and 6 km thick,
indicating strong subsidence between boundary faults. Four main faults (F1 to F4) (figure
III.9a) are clearly identified through this depth section down to 4–6 km depth (Bécel et al.,
2010). F1 and F2 are basin border faults and F3 and F4 are cross basin faults. The faults have
a very high dip (e.g. 75° to 85°) and clear vertical slip component in the plane of section.
They may be transtensional faults, with both normal and strike-slip components (Bécel et al.,
2010).

The Central Basin and theirs respective faults were also studied at shallow depths. The
seismic sparker reflection data with a shallow penetration of a 400 m (Rangin et al., 2001) is
shows in the figure III.9c, whereas the figure III.9d illustrate the high resolution seismic
reflection obtained with the 3.5 kHz sediment penetrator, with a penetration of a50 m depth
below the seafloor (Armijo et al., 2005). In both figures (figures III.9c, d) can be observed
the transtensional faults with both normal and strike-slip components. In these sections can
also be observed the fault scarps; as the cumulative scarp of a50 m high, formed at the SW
boundary fault of the inner pull-apart, where the late Pleistocene-Holocene sediment pile is

58
Figure III.9: Seismic profiles across the Central Basin. a) Shaded elevation topography, bathymetry
and active faults (from Armijo et al., 2002), with location of seismic profiles and MD01-2429 and
MD01-2431 cores. b) Depth seismic profile SM46 from Bécel et al. (2010). c) Sparker reflection
data, profile s52 (from Rangin et al., 2001). d) 3.5 kHz high resolution profile (cb3) acquired during
the MARMARASCARPS cruise 2001.

59
cut. In the cb03 profile have been projected the MD01-2429 and MD01-2431 cores, which
represent at least the last 18 cal kyr BP of sedimentation (Beck et al., 2007). In this profile and
in the cores studied by Beck et al. (2007) a difference in the thickness of the sedimentary
sequence on both sides of the SW boundary fault of the inner pull-apart is observed, the
thickest accumulation characterizing the deepest central part. According to theses authors, this
difference in the thickness is not the result of a continuous process, with mean values of
sedimentation rates and fault displacements; but it is rather the product of a succession of co-
seismic incremental events and co-seismic faults scarps.

3.3.2. Previous data on paleoseismic archives within the Sea of Marmara’s


sediments.

The Marmara region underwent a significant amount of destructive large earthquakes


during the last 2 kyr, as is revealed by the historical sources (Ambraseys and Finkel, 1991;
Ambraseys and Jackson, 2000; Ambraseys, 2002b). This historical record has been completed
by the paleoseismological studies conducted in numerous trenches (Rockwell et al., 2001;
Klinger et al., 2003; Fraser et al., 2010) and also submarines researches, providing
information about historic and prehistoric record (Hiscott and Aksu, 2002; McHugh et al.,
2006; Sari and Ça÷atay, 2006; Beck et al., 2007; Drab et al., 2012; Ça÷atay et al., 2012; Eriú
et al., 2012).

The region has probably one of the most complete earthquakes historical records of the
earth because it is densely populated, being the centre for the trade routes between East and
West and the capital city of the Byzantine and Ottoman Empires (Ambraseys, 2002b). Annex
A table summarizes the principal earthquakes (a total of 43 with magnitude Ms• 6.8) occurred
along the North Anatolian Fault Zone in the Marmara region, after 400 AD. Through the
study of these historical and paleoseismological data, different authors have been proposed
different recurrence intervals (RI) for destructive earthquakes. These recurrences may be
variable according to the fault segment studied, but simplifying, a RI comprises between 200
and 300 yr as proposed Ambraseys (2002b) would represent an average.

The paleoseismological studies can go beyond than the historical record. In the Sea of
Marmara’ Central Basin, data obtained by Beck et al. (2007) display a longer temporal
archive. Based on the study of giant piston cores and VHR seismic profiles obtained during

60
the MARMACORE Cruise, a 20 kyr-lasting sedimentary “chronicle” was depicted. In the
Central Basin, 71 inferred earthquake-induced events were identified along the entire
stratigraphic record (marine and non-marine sections). In the non-marine section, these events
are represented by the association Homogenite+Turbidite (HmTu), highlighting the presence
of a great HmTu event (8 to 15 m thick), which can be clearly identified in the seismic
profiles. The textural characteristics of this event were studied at very high resolution.
Moreover, in the marine section other types of earthquake-related features (figure III.10,
from Beck et al., 2007) have been identified. The different examples display:
- silty-sandy laminæs that exhibiting low angle symmetrical flaser bedding, like
structures indicatives of oscillating currents associated to reflected tsunamis possibly
leading to seiche effect (figure III.10a).
- liquefaction features (figure III.10b) as ball-and-pillow structures (figure III.10c),
favoured by textural contrasts between superimposed layers;
- boudinage of laminæted layers (figures III.10d, f) interprets as lateral spreading
effect.
- subvertical water-escape “vein structures” (figure III.10d), structures which have been
described in soft sediments submitted to frequent tectonic and seismo-tectonic activity.
Based on the identification of the shock-induced events and the estimation of their
time distribution, Beck et al. (2007) proposed for the Central Basin a mean recurrence time
interval which varies between 80 yr for the non-marine section and 480 yr for the marine one.

3.3.3. The Sea of Marmara, gateway between the Aegean and the Black seas.

The Sea of Marmara can be considered as a gateway between the Mediterranean Sea
and the Black Sea. The sediments in this gateway are a sensitive recorder of water mass
exchanges between the two Seas, which during the Late Quaternary was controlled by the
eustatic sea level variations (Vidal et al., 2010). During glacial periods, the water level of the
Aegean Sea and the Black Sea are below the depth of the straits and isolate the Marmara and
Black Seas that become in large brackish lakes (Ryan et al., 1997; Aksu et al., 1999;
Kaminsky et al., 2002). During interglacial periods, the sea level rises and the three seas can
be reconnected again establishing a system of two-layer flow in the basin of Marmara
(explained in the following subchapter, figure III.11) (Beúiktepe et al., 1994). The last
connection between the Aegean and Black seas occurred at about 9.4 kyr BP. (Ça÷atay et al.,
2000; Major et al., 2006; Vidal et al., 2010; Badertscher et al., 2011).

61
Core MD01-2429

Core MD01-

Figure III.10: Scopix X-ray imagery of possible earthquake-related disturbances identified in the
Holocene marine section of the MD01-2429 and MD01-2431cores (from Beck et al., 2007). Two
mechanisms can be deduced: effect of oscillating bottom currents (reflected tsunamis, seiches?) on
silt; and water escape with horizontal offsets (lateral shear by shaking?).

62
Different reconstructions for water mass exchange between the basins during the last
glacial to Holocene has been proposed based on lithology, geochemistry, and fauna, among
others. Ryan et al. (1997, 2003) based on data from the northern Black Sea shelf proposed that
the Black Sea was a giant freshwater lake during the Last Glacial Maximum (LGM), with the
water level standing at -150 m below present day sea level. During the post-glacial sea-level
rise at a7150 yr BP the Mediterranean Sea breached the Strait of Bosphorus, where conditions
for a catastrophic flooding were met, contributing to the Noah’s Flood myth (Mestel, 1997).

Other hypothesis is proposed by Aksu et al. (1999), Hiscott et al. (2007), and Londeix
et al. (2009); among others. They suggested - based on seismic profiles, and cores analyses -
that the Black Sea first overflowed into the Sea of Marmara through the Bosphorus.

Based on oxygen and strontium (Sr) isotopes from carbonate shells, major and trace
elements, as well as, specific organic biomarker measurements made over the MD01-2430
long piston core, Vidal et al. (2009) proposed a more complex alternative model for the
reconnection between the three seas. This complex hydrologic history of the Sea of Marmara
can be divided in six phases (figure III.11).
Phase 1: 23 to18 kyr BP
During the Last Glacial Maximum (LGM), the Sea of Marmara basin exhibited a
stable hydrologic state, receiving only water flowing from the Black Sea.
Phase 2: 18 to16 kyr BP
The Marmara and the Black seas continue connected, this fact was evidenced by the
presence of G18O-depleted freshwater in both basin.
Phase 3: 16 to12.8 kyr BP
- Between 16 to 14.7 kyr BP, (pre-Bølling-Allerød) the climate is characterized by
warm and humid conditions, potentially resulting in an increase in regional evaporation, but
with relatively high precipitation, which would have contributed to maintain the outflow from
the Black Sea into the Marmara basin.
- At 14.7 kyr BP, occurs the onset of the latest incursion of Mediterranean waters in
the Sea of Marmara.
- Between 14.7 to 12.8 kyr BP, the Sea of Marmara was influenced by both
Mediterranean and Black seawaters. Marine conditions were established gradually, taking
about 2 kyr for euryhaline Mediterranean organisms to colonize the Sea of Marmara, probably

63
due to the strong overflowing of brackish waters from the Black Sea, which could have
weakened the marine water inflow.
Phase 4: 12.8 to 11.5 kyr BP
This phase correspond to the Younger Dryas (YD) characterized by colder and drier
conditions (Valsecchi et al., 2012), an increase of organic material of terrigenous origin, as
well as a reduction of the evaporation rates. The connection between the Aegean and Sea of
Marmaras would have persisted until the end of the YD.
Phase 5: 11.5 to 9.4 kyr BP
The surface water salinity continued to increase until 9 kyr BP, confirming that the
Black Sea outflow was negligible at the onset of sapropel deposition in the Sea of Marmara.
Phase 6: 9.4 kyr BP to Present.
At 9.4 kyr BP is established the connection between the Aegean and the Black seas. A
sapropel deposition occurred during this period (between 11.5 and 7 kyr BP). This period was
characterized by a different origin for organic matter with terrigenous and marine inputs.

Figure III.11: Hypothetical representation of different hydrologic phases between 23 kyr and 9.4 kyr
BP (from Vidal et al., 2010). BS represents the Bosphorus sill depth, DS represents the Dardanelles
sill depth, and MS represents the Sea of Marmara. Arrows indicate water masses entering the Sea of
Marmara through the BS and/or the DS. Dashed arrows represent a smaller water flux than solid
arrows.

64
3.3.4. Present-day physical oceanography and drainage

Since the last connection between the Black Sea and the eastern Mediterranean Sea, a
permanent system of water flow of two layers is installed in the Sea of Marmara (figure
III.12). The cooler (5-15°C) and lower salinity (17-20‰) surface layer coming from the
Black Sea flows south-southwest across the straits at a 10-30 cm/s velocitiy, this water mass
form a 25 m-thick surface layer in the Sea of Marmara. Warmer (15-20°C) and high-salinity
(38-39‰) Mediterranean water mass plunges beneath the low-salinity surface layer in the
northeastern Aegean Sea and penetrates the Strait of Dardanelles flowing northeast at a 5-25
cm/s velocity. The Mediterranean water occupies the entire Marmara Basin below 20 to 30-
m-thick low-salinity surface layer. It travels farther northeast across the Strait of Bosphorus at
a 5-15 cm/s velocitiy, and penetrates the Black Sea where it constitutes the bottom water mass
below the 100-200-m-thick surface layer (Beúiktepe et al., 1994 ; Ça÷atay et al., 2000; Aksu
et al., 2002).

The drainage basin of the Sea of Marmara are strikingly asymmetrical, with a very
narrow northern drainage basin of 4.438 km2 (figure III.12) and only small streams without
signiÞcant discharge (Okay and Ergun, 2005). At the opposite, the drainage of the southern
margin extends over an area of 30.600 km2, and several medium-sized rivers deliver a 6.3 ×
106 tons/yr total sediment supply (Ergin et al., 1991). The Kocasu is the largest of these
streams and supplies most of the terrigenous sediment (2.0 x 106 tons/yr suspended sediment).

Okay and Ergun (2005) propose that the recent coarse sediment in the Marmara basins
must have been derived from the land area surrounding the Sea of Marmara. They indicate
that the geology of the northern and southern drainage basins of the Sea of Marmara is
different. The Sea of Marmara’s northern catchment is dominated by Oligocene and Miocene
sandstones, whereas the southern drainage basin is geologically more heterogeneous, and
comprises a variety of Paleozoic to Neogene metamorphic, granitic, and sedimentary rocks
(see subchapter 3.2 for detailed information). Based on a comparison of heavy minerals
recovered in the deep basin sediments and in the beach sands, Okay and Ergun (2005)
conclude that: (1) the Tekirda÷ Basin, with predominant chlorite and mica, is probably fed by
the friable soft Miocene sandstones of the northwestern margin; (2) the Central Basin sands -
dominant chlorite, mica, and amphibole+pyroxene – is mainly fed from the south; (3) in the
Çinarcik Basin, with dominant amphibole+pyroxene and epidote grains, terrigenous feeding is

65
probably coming from the East through the Gulf of Izmit, or from the South through the
Kocasu delta (figure III.13).

 Figure III.12: Surface- (a) and depth- (b) water circulation in the Sea of Marmara (from
 Beúiktepe et al., 1994).

Our study of sediment composition along the MD01-2425 core (Çinarcik Basin)
shows a significant content of amphibole, pyroxene and epidote (see subchapters 3.4.2).
These results are in agreement with the observations made by Okay and Ergun (2005), and
confirm a predominantly southern and/or eastern source.

66
Figure III.13: Map showing possible clastic source regions for the deep Marmara basins and mean
compositions of heavy minerals in the Marmara sand provinces and in the basins (from Okay and
Ergun 2005). HMP: heavy-mineral province.

67
3.4. The Late Quaternary sedimentary infill of the central and eastern Sea of
Marmara

3.4.1. Introduction

The Sea of Marmara’s sedimentary infill has been the subject of numerous studies.
Several investigations have been focused on the understanding of paleoclimatic and
palaeoceanographic conditions prevailing during the Quaternary. Especially, many authors
have made an effort to understand when and how was the last connection between the Sea of
Marmara, the Aegean Sea and the Black Sea (e.g. Aksu et al., 1999; Ça÷atay et al., 2000;
2009; Hiscott et al., 2007; Londeix et al., 2009; Vidal et al 2010). On the other hand, due to
the high seismic activity of the area, others authors have directed their research on
paleoseismic record (e.g. McHugh et al., 2006; Sarõ and Ça÷atay ; 2006; Beck et al., 2007;
Eriú et al., 2012; Ça÷atay et al.; 2012).

In the present work we focus on the study of three of the eight long piston core
recovered during the MARMACORE cruise; the MD01-2425, MD01-2429 and MD01-2431
cores. These cores were recovered in the deeper part of the Central (Orta) and Eastern
(Çõnarcõk) basins (figure III.14 and table III.1 for location). In a preliminary report, Beck et
al. (2007) identified specific layers interpreted as earthquake-induced deposits. Through the
detailed study of these three cores we would like to provide additional insights about: 1) the
effects of the seismicity on the mechanisms of transport and deposition of sediments in the
deep environment of the Sea of Marmara; attempting to reconstruct the “paleoseismic record”
and trying to determine a recurrence interval of recorded earthquakes; 2) how and when was
the last change from non-marine to marine environment and their impact on the transport and
deposition dynamics; and 3) the identification and estimation of the co-seismic fault offsets
associated to the NAF system in the Central Basin through their co-seismic sedimentary
events. Our work aimed to contribute to a better understanding of the fault configuration of
the NAF within the Sea of Marmara, and a better estimation of the seismic hazard in the
region.

In this study, only Core MD01-2425’s detailed characteristics will be presented here
(visual descriptions and microscopic observations) in order to avoid repetitions, due to very

68
high similarities between the three cores. Preliminary lithological analyses of Cores MD01-
2429 and MD01-2431 were made by Beck et al. (2007).

Figure III.14: Simplified geodynamic setting of the Sea of Marmara with the location of the
cores. Bigger stars represents the cores analized in the present work Geomorphological setting
and North Anatolian Fault are from Armijo et al., 2002. Plate kinematics is simplified from
McClusky et al. (2003) and Barrier et al. (2004). Relief from NASA SRTM.

Water depth Latitude Longitude Length


Core N° Location
(m) (N) (W) (m
MD01-2424 800 Kumburgaz 40°51.84 N 028°30.56E 36,44
MD01-2425 1215 Cinarcik 40°47.95N 028°59.55E 31,29
MD01-2427 1163 Cinarcik 40°49.43'N 028°58.81E 25,78
MD01-2428 400 Imrali 40°41.23'N 028°33.70E 21,77
MD01-2429 1230 Orta 40°50.30'N 028°00.52'E 37,3
MD01-2430 580 N. Marmara Is. 40°47.81'N 027°43.51'E 28,88
MD01-2431 1170 Orta 40°47.84'N 028°00.53'E 26,4
MD01-2432 1100 Tekirdag 40°48.71'N 027°32.49'E 26,09

Table III.1: Location of 8 long Calypso piston cores recovered by the R/V MARION-DUFRESNES,
during the MARMACORE Cruise, August-September 2001

The sedimentary infill of the MD01-2525 core is characterized by its heterogeneity;


although the homogeneous muds are the most common and widely distributed sedimentary
deposit. Through visual descriptions (color, grain-size, structure and contacts), X-ray images
and the microscopic observation of coarse sediments (binocular) and fine-grained part (81
smear slides), we achieve to identify six (6) different sedimentary facies in the MD01-2425
core: two muddy facies, and four coarser facies.
69
3.4.2. Description of facies

Muddy facies
Facies Mh (Homogeneous mud): Light olive gray to dark yellowish homogeneous or
poorly laminæted calcareous muds. It can form massive layers from few centimetres to meter
thick. The grains are angular to sub-rounded with variable sphericity (high and low) (see
subchapter 3.5.2-c; figure III.30). They are constituted by detritic carbonate crystals, bio-
induced or biogenic components, and terrigenous particles (figure III.15). The silty
terrigenous fraction is siliciclastic, generally it is well to moderately sorted. It is constituted
by quartz, feldspar, micas, and, in minor amounts, by green and blue amphiboles, framboidal
pyrite, and others opaque minerals. Calcareous fraction is represented by detrital carbonate
crystals, micrite and aragonite needles. The biogenic components are mainly composed by
nannoplankton calcareous fossils, few planktonic foraminifers, sponge spicules, traces of
radiolarians and freshwater diatoms frustules. Flakes of organic matter are present. These
muds can be slightly to moderately bioturbated. The textural characteristics of this facies
allow us to interpret that it was deposited by processes of fall-out in calm conditions. In this
study, sediments belonging to this facies are interpreted as hemipelagic deposits (h.d.) or
hemipelagites (massive deep marine muds settled out from suspension following Selley
(2000). In order to simplify, in the present work the non-marine deep muds will also be called
hemipelagic-type deposits. The Mh facies is the predominant facies, and represents ~60% of
the total sediments in the MD01-2425 core (figure III.15, annex B).

Facies Mso (Slightly laminated organic muds): in contrast with the Mh facies, this
facies is slightly more organic and can show beds with abrupt or progressive changes of color.
In general they are dark yellowish brown color, although this may gradually or abruptly
change to light olive gray. This facies may also display very thin laminae of olive black and
moderate brown color (figure III.15). Facies Mso has textural and mineralogical
characteristics similar to those described in the facies Mh. Although it may have highest
percentages of framboidal pyrite, organic particles and some laminæ may present a slightly
higher percentage of particles of terrigenous origin. This facies is found in the marine unit and
represents ~12% of the total sediments in the MD01-2425 core (figure III.15).

70
Coarser facies
Facies Ls (Silty-sandy laminae): it consists of millimetric silty-sandy calcareous
laminae (<3 mm thick), which can form thin layers of few centimetres (1 to 2.5 cm) thick.
They generally show visual parallel planar lamination (figure III.15). They are olive gray and
light brown color. A detailed observation of this facies using X-ray images shows that these
laminæ not only exhibit parallel planar laminætion, also millimetric structures as: boudinage,
ball and pillow, convolute and flaser laminætion; and overlapping in some cases with
apparently liquefied muds layers (figures III.10, III.25, III.27). The laminæ are constituted
mainly by terrigenous components and in less proportion by detrital carbonate grains and low
or absence of clay or micrite. The terrigenous components are similar to those described in the
Mh facies, despite this, the percentage of quartz decreases and the percentage of feldspar, blue
and green amphibole, pyroxenes and epidote increase (figure III.15). Traces of framboidal
pyrite, chromite, organic fragments, calcareous nannoplankton, mollusk shells fragments and
planktonic foraminifers are present. The grains are angular to sub-rounded with variable
sphericity; they are moderately to well sorted. This facies is found mainly in the marine unit
(figure III.15), showing an apparent episodic occurrence. It represents ~1.5 % of the total
sediments in the MD01-2425 core. Sediments similar to those described in this facies have
been identified and studied by Beck et al. (2007) and Drab (2012) and has been interpreted as
possible earthquake-related imprints.

Facies Tuc (Classical turbidite): this facies consist of coarse sand to coarse silt, which
can grade upward to very fine silts or muds. They are predominantly of moderate yellowish
brown, moderate brown and dark yellowish brown color; they show net or erosive bases, and
are massive or present rare current ripples, parallel lamination and slight normal gradation.
This facies form very thin to thin layer from 5 to 50 mm thick (figure III.15). The grains are
angular to sub-angular with variable sphericity (high and low). It shows few to moderate
sorting. Their components (terrigenous, biogenic and trace minerals) are similar to those
described in the facies Ls, but the proportion of the terrigenous and carbonate components can
vary from one turbidite to an other. In general the terrigenous components are more abundant,
but few turbidites show an abundance of detrital carbonate particles, which may indicate at
least two different sources. This facies can be interpreted as a classical turbidites as defined
by Bouma (1962), and Mutti and Ricci Lucci (1978). This Tuc facies occurs in the non-marine
unit (figure III.15), and represents ~2 % of the total sediments in the MD01-2425 core.

71
Facies HmTu (Homogenite-Turbidites): This facies is composed by two units,
generally with a sharp contact and oscillation sedimentary structures in between (see figure 1
in subchapter 2.4.2). The basal unit (Tu) is coarse grained. It has similar characteristics to
those described in the facies Tuc, but in general, the top of the unit show an inunitediate unit
with sedimentary structures indicative of to-and-fro particles displacement. The bottom of the
Tu units are poorly to moderately sorted, with considerable amounts of clay and micrite. The
top, and specially the inunitediate unit can be better sorted, with little or absent clay and
micrite. On the other hand, the upper unit (Hm), is a homogeneous calcareous mud of dark
yellowish brown color, it shows characteristics similar to those described in Mh facies. Due to
this similarity, it is not easy to differentiate it from the Mh facies. The association with the
basal Tu unit, the occurrence of oscillatory sedimentary structures in between, and in some
case, their characteristic color can help in their visual identification. A sharp base of the upper
“homogenite” interval and it extremely granulometric homogeneity represent main visual
criteria. Despite this, the best tool to differentiate the Hm unit from the facies Mh is the AMS,
specifically the magnetic foliation (see subchapter 2.4.2 and 3.5.2). This facies was
recognized within the non-marine unit (figure III.15), and represents ~22% of the total
sediments in the MD01-2425 core.

Facies Sd (Soft-sediment deformed units): This facies is composed by deformed muds


sands and silts. These sediments present textural and mineralogical characteristics similar to
those described in the facies Tuc and Mh, but the original layering and sedimentary structures
were subjected to modification, with deformational structures such as: dykes, sills and ball-
and-pillow (figure III.15, annex B, sections XVI and XX). These structures are intruded into
calcareous muds as those described in the facies Mh. The dykes and sills can reach 15 cm
length and have thicknesses between 25 and 2 mm. The balls-and-pillows can reach
thicknesses of 6 cm, and 10 cm length. We interpret that these sand dykes and sills as having
seismic origin (Audemard and De Santis, 1991; Obermeier, 1996; Rodriguez-Pascua et al.,
2007), although, other agents could also produce dykes within the soft-sediments as storm
waves and breaking waves (Bowman et al., 2004). In the Sea of Marmara, the depth at which
the sediments belonging to this facies were deposited (> 1000 m depth) allows us to discard
this possibility. Likewise, the ball-and-pillow structures can also have seismic origin;
although contrast in density between layers can produce such structures, earthquakes seem the
most probable cause (Rodriguez-Pascua et al., 2007). This facies was recognized within the

72
non-marine unit (figure III.15), and represents ~3 % of the total sediments in the MD01-2425
core.

a b

c d

Figure III.15: Illustration of different lithofacies defined in the MD01-2425 core. 1).
Macroscopic photographs of fresh split core (scale in cm). 2) Optical microscope observation of
smear slides of different facies, evidencing the presence of: calcareous nanoplankton
(microphoto a, facies Mh), crystals of calcite and quartz, as well as, diatoms (microphoto b,
facies Mh). Blue and green amphibole and quartz (microphoto c, facies Ls); gastropods and
algal fragments (microphoto d, facies Tuc).

73
3.4.3. Stratigraphy of the MD01-2425 core

The vertical distribution of facies and the magnetic susceptibility (MS) profile along
the MD01-2425 core allow us to define two major sedimentary units. The contact between
these two units from the point of view of facies and MS appears to be sharp; the contact is
marked by a characteristic bioturbated mud bed (figure III.16).

3.4.3-a Upper unit (marine deposits)


It is mainly composed of homogeneous muds belonging to the Mh facies. Interlayered
within these homogeneous muds, we observed thin silty-sandy laminæ (facies Ls) and slightly
laminæted organic muds (facies Mso). This unit has 17.36 m thick in the MD01-2425 core (0
to 17.36 m core depth) (figure III.16), and according to our age-depth curve (see subchapter
3.4.4) it represent approximately the last 12.8 cal kyr BP of sedimentation. A distinctive
feature of this unit is the presence of calcareous nannoplankton, which is abundant in the top
of the unit, and tends to decrease to the bottom. In addition planktonic foraminifera and
radiolaria were observed, although less abundant than nannoplankton. This unit can be
subdivided into three subunits (U1, U2 and U3) based on their vertical facies evolution and
magnetic content. The upper and inunitediate U1 and U2 subunits (figure III.16) are
characterized by a similar distribution of facies, although the behavior of the MS curve
between these subunits is different, allowing its subdivision into two subunits. The subunit U1
characterized high values of MS, and the subunit U2 by low values (see subchapter 3.5.2-d for
detailed information about MS). The U3 subunit presents MS values relatively low, similar to
those present in the subunit U2. However U3t differs from the U1 and U2 subunits for: 1) the
dominance of slightly laminated organic muds (facies Mso), located between 9.8 and 13.9 m
depth (6.2 to 9.5 cal kyr BP according to our age-depth curve); 2) the presence of very thin
silty-sandy laminææs (facies Ls), which show thicknesses generally less than 1mm thick. Due
to the log scale these laminæ are not represented in the simplified log in the figure III.16.

Base on the lithological and biological characteristics we interpret that this upper unit
was deposited under marine conditions, which suggests a connection between the Marmara
and Aegean Sea at least during the last 12.8 cal kyr BP.

74
Figure III.16: Magnetic susceptibility profile and simplified log of the MD01-2425 core. The
upper marine and the inferior non marine units are separates by a reference bed level. The X-
ray radiogram corresponds to this bed level. U1, U2 and U3 are the subunits defined in the
marine upper unit. See cores location in figure III.3 and core photographs in annex B.
Rectangles in black color indicate the stratigraphic position of six sections studied in detail (A
to F).

75
3.4.3-b Lower unit (non-marine deposits)
As the upper unit, the homogeneous muds are the predominant lithology, although
interlayered within them we observed classical turbidites (facies Tuc), homogenites-turbidites
(HmTu), silty-sandy laminæ (facies Ls), as well as sand, silts and mud exhibiting
deformational structures (facies Sd). This unit has 13.32 m thick in the MD01-2425 core
(17.37 to 30.69 m core depth) (figure III.16), and according to our age-depth curve
(subchapter 3.4.4) it represents approximately the period comprising between 12.8 and 17 cal
kyr BP. The muds are slightly more calcareous with respect to upper unit; reversely, the other
lithological and compositional characteristics are almost identical. The biological content of
this unit is variable in depth. Portions with predominance of centric and pennate diatoms were
identified; some of them are characteristics of freshwater. Others parts with presence of
calcareous nannoplankton were also identified; there, diatoms are or absent. These variations
in the biological content indicate fluctuations in the environmental conditions during this
period, where the Aegean Sea was already connected to the Sea of Marmara (a14.7 kyr BP,
according Vidal et al., 2010). Despite this, the presence of freshwater diatoms also indicates a
possible interruption of the connection between both seas, and/or a high input of fresh water
from the Black Sea, reducing significantly the salinity.

3.4.3-c The contact between the upper and lower units


From the biological and lithological point of view, the contact between both units
appears sharp. In the MD01-2425 core was identified a characteristic layer between both
units, which also was identified in the MD01-2429 and MD01-2431 cores. This layer is 5 cm
thick in the MD01-2429 and MD01-2431 cores, and 2cm thick in the MD01-2425 core. Due
to its characteristics and position between the two units this layer has been considered as a
reference bed level (Rbl) (figure III.16) and has been used to correlate the three cores located
in two separate basins, as discussed in the subchapter 3.8. This Rbl is a mud, similar to facies
Mh, but highly bioturbated. It contains some plant fragments and marks the last appearance of
diatoms in the sedimentary column (17.36 m depth). According to our age-depth curve this
Rbl may have been deposited within a 12.8-to-13.4 cal kyr BP time interval. Vidal et al.
(2010) concluded that the marine conditions were established at 12.8 kyr BP, coinciding with
the beginning of the Younger Dryas. This period was characterized in the Mediterranean by
colder and drier climatic conditions (Valsecchi et al., 2012) and in the Sea of Marmara by an
increase in the contribution of terrigenous (allochtonous) organic material (Vidal et al., 2010).
In the MD01-2425 core, the disappearance of diatoms above this level indicates that the
76
connection between the Black Sea and the Sea of Marmara was negligible or inexistent,
establishing the marine conditions in the basin. Combining the results obtained by Vidal et al.
(2010) and our results we consider that the Rbl was deposited around 12.8 cal kyr BP, and
represent the beginning of the Younger Dryas.

The sharp limit between the two units is also clearly observed in the magnetic
susceptibility MS profiles (figure III.16). They display an abrupt change in the concentration
of magnetic content, which will be discussed in subchapter 3.8. On the other hand, in order to
better know the nature of the contact between these two units, high-resolution geochemical
analyses were performed in the MD01-2425 core, along a 1.5 m section, which includes the
Rbl as well as part of the overlying (marine) and underlying (non-marine) units.

The geochemical elemental analyses were carried out using the X-ray fluorescence
(XRF) core scanner. The evolutions of elementary chemical parameters are shown in the
figure III.17. In this figure the profiles of the terrigenous indicators as Si, Al, K, Ti and Zr
show a similar variable behavior along the studied section, not showing an abrupt change
around the Rbl. A similar behavior can be observed in the Ca, Sr, Fe, Cl, and Rb, which show
a variable behavior along the section, but not an abrupt change, is observed around the Rbl.
Moreover, elements such as S, Mo, Br show an increasing tendency towards more marine
environments, which may indicate an increase in the amount of marine organic matter in the
basin, as well as the dominance of anoxic conditions (Rothwell et al., 2006; Ziegler et al.,
2008; Montero-Serrano et al., 2009). These elements also do not show an abrupt change
around the Rbl. This is contrasting with the observations based on the lithological, biological
and magnetic content, that indicate a sharp contact between the upper (marine) and lower
(non-marine) units. The bulk geochemical evolution may indicate a progressive increase of
the water salinity probably associated with the rising of the influence of marine waters from
the Aegean Sea and a negligible or absent contribution of freshwater from the Black Sea as
has been proposed by Vidal et al. (2010) for the period after the Younger Dryas.

77
Figure III.17: XRF geochemical profiles of a 1.5 m long section of the MD01-2425 core. Reference bed level (Rbl) is marked by dashed red line.
The simplified log and the X-ray image of this section are show on the left side. The legend of this log is the same of the figure III.16. Green color
lines represents classicals turbidites (T), brown color represents millimetric silty-sandy laminæ (Ls).

78
3.4.4. Radiocarbon chronology of the Core MD01-2425

The aim of this subchapter is to know the age of instantaneous events, specifically the
age of those seismo-induced. The study of the time distribution of these earthquakes -
associated features is essential to estimate earthquake recurrence and the probability of
occurrence of future earthquakes (Lettis and Kelson, 2000). Different dating methods are
potentially applicable to solve problems in paleoseismology (e.g. dendrochronology, varve
chronology, uranium series), they have been summarized by Noller (2000) and Walker
(2005). The radiocarbon dating is the most widely used. Few other methods are as accurate as
14
C dating in the time span of most interest to paleoseismologists: the Holocene (McCalpin et
al., 2009). In particular, seasonal sedimentation, specifically the varve counting can be of
interest as a dating method (Mörner et al., 2000). But the main disadvantage of this method is
that it can only be used in geologic setting suitable for the deposition of varves.

14
3.4.4-a C samples
For the MD01-2425 core, eighteen (18) samples taken between 4.5 and 30 m depth
14
were measured using Accelerator Mass Spectrometry C (figure III.18). Twelve (12)
measures were done on plant debris and woods fragments taken in coarse to fine turbidites.
Six (6) others were done in particulate organic matter collected from organic rich muds. Their
14
C ages and respective calibrations are summarized in the table III.2.

Due to the absence of radiocarbon ages for the first 4.5 m, and the presence of
numerous samples with overestimated radiocarbon ages, we will consider others ages
previously published in order to construct a more robust chronological frame. We will take
into account the radiocarbon ages obtained by Beck et al. (2007) in the same core (MD01-
2425) (figure III.18), and the ages obtained by correlation between the MD01-2425 core and
nearby cores from the Çinarcik and Central basins (figures III.18, III.19, III.20), acquired by
Beck et al. (2007) and Drab et al. (2012).

14
All C ages are summarized in the tables III.2, III.3, III.4, with their respective
calibrations. The radiocarbon ages show a marked difference depending on the material used
(plant debris and wood fragments, particulate organic matter, shell fragments, and
foraminifers), (figures III.18, III.19, III.20). Very old values at shallow depths (example
figure III.19, cores Klg01 and Klg02) imply a very low sedimentation rate (SR) of ~0.30

79
mm/yr. This low SR differs from those deunitined in previous works for the Sea of Marmara,
which varies between 1 and 3 mm/yr (e.g. Ça÷atay, 2002; Beck et al., 2007; Ça÷atay, et al.,
2012). The difference in ages depending on the material sampled (at similar depths), as well
as the underestimation of the sedimentation rate, indicates that some ages are overestimated.

Ages calculated from foraminifers (benthic) by Drab (2012) in the Klg04 core appear
to be the most realistic (figures III.19, III.20); they do not seem to overestimate the expected
age. Drab et al. (2012) and Drab (2012) collected numerous radiocarbon dating from seven
piston cores from the same sub-basins. They find that the ages obtained from planktonic and
benthic foraminifers are the most consistent. The ages measured on planktonic and benthic
foraminifers don’t show large differences between them, which may be a criteria for limited
reservoir effect.

Radiocarbon ages obtained from plant debris, woods and shell fragments taken in
turbidites generally overestimate the expected age (figure III.18), although four, among
seventeen, measurements seem to be reliable ages (figure III.19). The overestimated ages can
be explained by the fact that the material dated is a reworked organic matter (OM), which
probably was re-transported from the land by the rivers or from the shelf and upper slope and
deposited in the deep basin by turbiditic currents.

Ages calculated from dispersed organic matter (DOM) also overestimate the expected
age, only one, among eleven, measurement shows a reliable age (figure III.20). Figure III.19
displays the radiocarbon ages obtained in the Klg01 and Klg02 cores (this study and Drab et
al., 2012, respectively). The ages appear very old if compared to the foraminifers-based
radiocarbons ages obtained from the Klg04 core (Drab, 2012). The same situation occurs in
the MD01-2425 core (marine and non-marine units), where five measurements show
overestimated ages (figure III.18). These oldest ages can be explained by the fact that these
samples may contain a significant percentage of reworked organic matter of terrestrial origin
(OMTO). This fact has been verified by Tolun et al. (2002). These authors identified
significant amounts of OMTO in sediments of marine and lacustrines origin during the last 18
kyr. Likewise, Ça÷atay et al. (1999) and Aksu et al. (2002) indicate the presence and
increment of OMTO during the last 12 kyr, especially during the deposits of the sapropelic
layers (10.6-6.4 and 3.8-3.2 kyr).

80
An alternative explanation could be the presence of methanotrophic bacteria and their
associates: the sulphate reducing bacteria, in the DOM. The presence of these bacteria is well
known on the Sea of Marmara’s sediments especially along active faults. Complex
chemosymbiotic communities, bacterial mats and a wide variety of carbonate chimneys,
crusts and concretions have been observed in numerous sites with seafloor fluid seepage,
along the main active North Anatolian Fault (Zitter et al., 2008; Ritt et al., 2010; Chevalier et
al., 2011). Field observations (MARMARASCARPS cruise in 2002; MARMARA-VT cruise
in 2004; MARNAUT cruise in 2007) contributed to identify the extensive areas with active
seafloor fluid seepage (Halbach et al., 2004; Armijo et al., 2005; Geli et al., 2008; Zitter et al.,
2008; Tryon et al., 2010), where the tectonic control appears to be the most important factor
for their occurrence (Zitter et al., 2008). The fluids are enriched in methane and others short-
chain hydrocarbons (ethane, propane and iso-butane) (Geli et al., 2008; Zitter et al., 2008;
Tryon et al., 2010). Halbach et al. (2002) suggest a primarily biogenic origin for the gas with
a minor (1% ethane + propane) thermogenic component.

The presence of these methanotrophic bacteria in the DOM can explain the existence
14
of an important reservoir effect in this type of samples. These bacteria concentrate a C
depleted methane, thus resulting in old 14C ages. Fontugne et al. (2009) report the presence of
methanotrophic bacteria in the Black Sea, with a radiocarbon age of a15.000 yrs (due to a 14C
depleted methane). They propose that the important values of reservoir age in the Black Sea
(> 500 yr) can be related to the presence of a percentage of chemoautotrophic matter in the
DOM. A similar situation could be occurring in the Sea of Marmara. Ménot et al. (2010)
indicate that these methane releases did not only have been occurring during the present, and
show evidences of large methane discharges during the last deglaciation. This fact would
explain some of the radiocarbon ages older than expected along the MD01-2525 core.

3.4.4-b Correlation between cores


Due to large amount of overestimated ages along the MD01-2425 core, we will
consider others ages proposed by Beck et al. (2007) and Drab (2012) in order to construct a
more robust geochronological setting. The estimation of the location in depth of these ages
was performed by core correlation. These correlations were carried out using lithological data,
and magnetic susceptibility (MS) records. We will propose four steps of correlation:
1) correlation between the giant pistons cores MD01-2425 and MD01-2431 (Beck et al.
2007), both taken with the same coring conditions (figure III.18);
81
2) correlation between the giant piston MD01-2431 and two gravity piston cores of 4 to
4.5 m long (Klg01 and Klg02) (figure III.19). The Klg cores were performed during
The MARMARASCARPS cruise in 2002 (Armijo et al., 2005);
3) correlation between the giant piston MD01-2425 and the piston core Klg04 (Drab,
2012).
4) integration of all data obtained in the steps 1 to 3 in the MD01-2425 core (figure
III.20).

For the MD01-2425 core we thus select only nine (9) radiocarbon ages (figure III.20,
table III.4). The depth distribution of these ages is realistic with respect to the general
evolution of the infilling. Furthermore, they are consistent with results published by different
authors for main changes in sedimentation or particular layers as:
1) the non-marine/marine change (a12 kyr BP, according to Ça÷atay et al., 2000) which
seems to be clearly represented by a change in the sedimentary facies and a sharp rise in the
MS values toward the non-marine unit;
2) the presence of a organic rich unit between 10 to 13.9 m depth (6.2 to 9.5 cal kyr BP),
which could be equivalent of the sapropel unit defined by Ça÷atay et al. (2000) and Aksu et
al. (2002) in sediment gravity cores between 10.5 and 6 kyr BP; and by Vidal et al. (2010) in
the MD01-2430 core between 11.5 and 7.5 cal kyr BP;
3) the lack of tephra layers, which have been identified in the MD01-2430 and MD01-2428
cores (Beck et al., 2007; Vidal et al., 2010). They have been associated with the Santorini
eruption dated at 22 kyr cal. BP. This would imply that the age of this core must be younger
than 22 kyr.

82
Figure III.18: Correlation between the MD01-2425 and MD01-2431 cores, based on Magnetic
Susceptibility (MS) and lithological data. In italics and bold letter are represented the uncalibrated
14
C ages mentioned by Beck et al. (2007). In blue color are represented the uncalibrated
radiocarbon dating performed from plant debris and wood fragments extracted from fine and coarse
turbidites. In green color dating from the particulate organic matter extracted from organic rich
muds.

83
Figure III.19: Correlation between the MD01-2431, Klg01, Klg02, and Klg04 cores. The correlation between the cores was based in Magnetic
Susceptibility (MS) and lithological data. In blue color is represented the uncalibrated 14C age obtained from a shell fragment extracted from a
turbidite (Drab 2012); in green color the particulate organic matter extracted from organic rich muds (Drab 2012). Red color uncalibrated 14C
ages obtained from benthic foraminifers.

84
Age BP Age BP Age BP
0 10000 20000 30000 40000 0 10000 20000 30000 40000 0 5000 10000 15000
a) b) c)
0 0 0

1000 1000 1000


Depth (cm)

Depth (cm)
Depth (cm)
2000 2000 2000

3000 3000 3000

Figure III.20: a) Sets of uncalibrated Accelerator Mass Spectrometry 14C measurements. In red color are represented the 14C ages performed on the
MD01-2425 core. In blue color radiocarbon ages obtained by correlation with others cores. b) Sample classification in function of material used: in
blue color plant debris and wood fragments collected in turbidites; in red color the particulate organic matter; in green color the
foraminifers; in purple color a shell fragment. c) Most consistent ages selected to construct the age depth model. Color legend is the same to that
used in b.

85
3.4.4-c Calibration of radiocarbon dating
As the samples of the MD01-2425 core were taken in marine and non-marine
subaquatic environments the Marine09 calibration curve (Reimer et al., 2009) and a local
reservoir correction (ǻR) have been used to calibrate the radiocarbon ages (see annex C for
complementary information about the calibration methods).

For the Sea of Marmara, two ǻR were proposed by Siani et al. (2000): ǻR= 71 r 40
and ǻR= 91 r 40 (summarized on the marine reservoir database www.calib.org). To simplify
the calibration we use a single value of ǻR. It is the average value of both ǻR with their
average errors: ǻR= 81 r 56. A constant reservoir correction R= 405 years (Reimer et al.,
2009) and a local reservoir correction ǻR= 81 r 56 were applied to the whole sedimentary
sequence. This assumption is due to the fact that in the Sea of Marmara there are not enough
studies concerning reservoir effect during the last “lacustrine” period; most probably, a longer
reservoir correction should be applied. Recent studies conducted in the Black Sea have shown
that the reservoir effect has varied between 0 and 1500 years since the last 30 kyr. This
variation is also controlled by the water depth (Bahr et al., 2005; Major et al., 2006; Kwiecien
et al., 2008) and the nature of the organic matter (Fontugne et al., 2009). This shows that our
assumption is probably simplistic, but the lack of data does not allow us to make more precise
corrections.

After having selected the calibration curve and the ǻR correction, we test different
calibration software (Oxcal, Bcal and Clam) in order to deunitine the most realistic age
(tables III.2, III.3, III.4). Each of them shows slight variations in the range of calibrated-ages
(even though in each case the same parameters were used) (see annex C for complementary
information).

3.4.1-d. Age-depth curve of the MD01-2425 core


In the MD01-2425 core, 102 instantaneous events have been identified, seventy five
(75) silty-sandy laminæ and classical turbidites (facies Ls and Tuc) and twenty seven (27)
homogenites-turbidites. They represent “event deposits” and they affect the sedimentation rate
calculation. To achieve a reliable age curve it was necessary to substract these event deposits
from the sedimentary column. The total thickness of these deposits reached 701 cm (from
3102 cm of the total thick).

86
Median of Error
Range of ages Weighted Range of ages
Depth in the ages Error Weighted
Laboratory 13 Age Age calibrated average age calibrated
Core d C calibrated (Oxcal) average age Material
code (yr) error (Oxcal) (Bcal) (Bcal)
(cm) (Oxcal) (cal yr BP) (Bcal)
(cal yr BP) (cal yr BP) (cal yr BP)
(cal yr BP) (cal yr BP)

451 SacA 18150 -26.10 6640 30 7231-6917 7074 157 7071 (+) 155 (-) 147 7226 - 6924 PDOM
600 SacA 18151 -26.90 6170 30 6669-6359 6514 155 6516 (+) 148 (-) 139 6664 - 6377 PDOM
667 SacA 18149 -24.30 5225 35 5625-5321 5473 152 5496 (+) 114 (-) 173 5610 - 5323 PDOM
864 SacA 18152 -25.30 7150 30 7660-7432 7546 114 7549 (+) 104 (-) 105 7653 - 7444 PDOM
1137 SacA 18153 -24.70 7720 50 8278-7948 8113 165 8101 (+) 171 (-) 149 8272 - 7952 POM
1296 SacA 18154 -28.00 9775 40 10729-10390 10559,5 169,5 10557 (+) 155 (-) 156 10712- 10401 POM
1344 SacA 18155 -22.80 10660 50 12105-11397 11751 354 11804 (+) 294 (-) 391 12098 - 11431 POM
1535.5 SacA 18156 -26.70 12780 50 14887-13936 14411,5 475,5 14312 (+) 537 (-) 362 14849 - 13950 POM
1715 SacA 18157 -21.30 21950 150 26140-25100 25620 520 25651 (+) 451 (-) 534 26102 - 25117 PDOM
1718.5 SacA 18158 -26.10 23630 170 28540-27075 27807,5 732,5 no results no results no results PDOM
1734.1 SacA 18159 -23.20 18870 100 22295-21523 21909 386 21916 (+) 344 (-) 368 22260 - 21548 PDOM
1860.5 SacA 18160 -28.90 15240 70 18500-17625 18062,5 437,5 18016 (+) 473 (-) 373 18489 - 17643 POM
1933 SacA 18161 -25.70 18480 100 22005-21151 21578 427 21489 (+) 406 (-) 305 21895 - 21184 PDOM
2000 SacA 18162 -15.60 17850 90 21187-20300 20743,5 443,5 20714 (+) 421 (-) 401 21135-20313 POM
2173 SacA 18163 -24.50 15040 60 17999-17257 17628 371 17699 (+) 283 (-) 259 17982-17440 PDOM
2595 SacA 18164 -28.70 14800 70 17796-17046 17421 375 17405 (+) 325 (-) 319 17730 - 17086 PDOM
2603 SacA 18165 -14.10 39800 1100 45324-42063 43693,5 1630,5 no results no results no results PDOM
2810 SacA 18166 -25.60 21290 130 25174-24394 24784 390 24785 (+) 344 (-) 348 25129 - 24437 PDOM

Table III.2: Eighteen radiocarbon ages carried out in the MD01-2425 core (present work).
All calendar ages were calculated using the Marine09 calibration curve and applying the local
variations of the reservoir effect (ǻR) proposed by Siani et al. (2000) for the Sea of Marmara. The
radiocarbons dating were calibrated using Oxcal (Ramsey, 2001) and Bcal (Buck et al., 1999) software.
A slight difference can be observed in the range of calibrated ages.
Material: POM= particulate organic matter extracted from organic rich muds. PDOM: plant debris
and wood fragments extracted from fine and coarse turbidites; SHF= shell fragments extracted from
fine and coarse turbidites. BF: benthic foraminifera.

Range of Median of Error Range of


Weighted
Depth in ages the ages Error Weighted ages
Core Age Age average age
Core Author calibrated calibrated (Oxcal) average age calibrated Material
(yr) error (Bcal)
(cm) (Oxcal) (Oxcal) (cal yr BP) (Bcal) (Bcal)
(cal yr BP)
(cal yr BP) (cal yr BP) (cal yr BP) (cal yr BP)

Klg 01 103 Present work 5900 50 6398-6030 6214 184 6230 (+) 165 (-) 150 6395 - 6080 POM
Klg 01 157 Present work 6270 50 6816-6439 6627,5 188,5 6626 (+) 174 (-) 175 6800 - 6451 POM
Klg 01 168 Present work 6620 50 7233-6867 7050 183 7048 (+) 181 (-) 169 7229 - 6879 POM
Klg 02 185 Drab 2012 3430 20 3359-3031 3195 164 3205 (+) 152 (-) 158 3357 - 3047 POM
Klg 02 352 Drab 2012 5060 20 5479-5121 5300 179 5331 (+) 142 (-) 194 5473 - 5137 POM
Klg 04 62.5 Drab 2012 1060 30 661-485 573 88 575 (+) 87 (-) 86 662 - 489 BF
Klg 04 182 Drab 2012 1480 35 1103-784 943,5 159,5 951 (+) 143 (-) 158 1094 - 793 BF
Klg 04 302 Drab 2012 1740 35 1334-1059 1196,5 137,5 1209 (+) 122 (-) 140 1331 - 1069 BF
Klg 04 432 Drab 2012 3355 30 3299-2925 3112 187 3108 (+) 179 (-) 170 3287 - 2938 SHF
MD01-24231 1786 Beck et al. 2007 12150 65 13735-13325 13530 205 13525 (+) 196 (-) 186 13721 -13339 PDOM
MD01-2425 2731.5 Beck et al. 2007 12850 55 14945-14034 14489,5 455,5 14443 (+) 455 (-) 395 14898-14048 PDOM
MD01-2425 2901 Beck et al. 2007 13250 55 15650-14655 15152,5 497,5 15192 (+) 425 (-) 483 15617-14709 PDOM

Table III.3: Radiocarbon ages carried out in: present work (core Klg01); Drab, 2012 (cores Klg02
and Klg04) and Beck et al. (2007) (cores MD01-2425 and MD01-2431). See legend in table III.2.

87
To create the age-deph curve, nine (9) radiocarbon ages were used (table III.4). After
subtracting the thickness of each event deposits, different curves were calculated using linear,
spline and polynomial interpolations with the utilization of the software Clam (Blaauw, 2010)
(see annex C for complementary information). The different curves are presented in the
figure III.20. The linear interpolation takes into consideration all radiocarbon ages, but
involves abrupt changes in the sedimentation rate, whereas the spline and polynomial
interpolation curves tend to move away or not pass through some radiocarbon ages. Although
all curves have their weaknesses, for the MD01-2425 core, we privilege the linear
interpolation because it takes into consideration all the 14C ages.

Error
Weighted
Weighted Range of ages
Depth Age Age average age
average age calibrated (Bcal) Type
(cm) (yr) error (Bcal)
(Bcal) (cal yr BP)
(cal yr BP)
(cal yr BP)

62,5 1060 30 575 (+) 87 (-) 86 662 - 489 BF


182 1480 35 951 (+) 143 (-) 158 1094 - 793 BF
302 1740 35 1209 (+) 122 (-) 140 1331 - 1069 BF
432 3355 30 3108 (+) 179 (-) 170 3287 - 2938 SHF
667 5225 35 5496 (+) 114 (-) 173 5610 - 5323 PDOM
1137 7720 50 8101 (+) 171 (-) 149 8272 - 7952 POM
1786 12150 65 13525 (+) 196 (-) 186 13721 -13339 PDOM
2731,5 12850 55 14443 (+) 455 (-) 395 14898-14048 PDOM
2901 13250 55 15192 (+) 425 (-) 483 15617-14709 PDOM

Table III.4: Radiocarbon dating used to construct the depth-age model of the MD01-2425 core.

Marine09 calibration curve and applying the local variations of the reservoir effect (ǻR=81 r
The radiocarbons dating were calibrated using the software Bcal (Buck et al., 1999), the

56) proposed by Siani et al. (2000) for the Sea of Marmara. Material: POM= particulate
organic matter extracted from organic rich muds. PDOM: plant debris and wood fragments
extracted from fine and coarse turbidites; SHF= shell fragments extracted from fine and coarse
turbidites. BF: benthic foraminifera.

88
Lineal interpolation Spline interpolation Polynomial interpolation

Figure III.21: Lineal (left plot), spline (central plot), and polynomial (right plot) interpolations generated using the software Clam (Blaauw, 2010); taking
into account the 9 radiocarbon dating (see table 3). The lineal model involves abrupt changes in the sedimentation rate. The spline and polynomial
interpolation models tend to move away or not pass through some radiocarbon ages (red circles).
89
3.4.4-e. Main results
In the age-depth curve without instantaneous events of the MD01-2425 core, it is
possible to identify several changes in the hemipelagic (“background”) sedimentation rate
(figure III.21). It varies between 0.70 and 6.04 mm/yr. Based on these changes in the
sedimentation rate (SR) we distinguish 6 levels. The thickness of each level and their SR are
summarized in the table III.5.

Figure III.22: Age-depth model for the MD01-2425 core (without instantaneous events). Six levels
characterized by differences in the sedimentation rates (SR) are defined. Non marine to marine
transition is indicated by dashed red line.

In order to know the age of each instant event, specifically the age of those
seismically-induced one, we added them (IE) with their respective thickness to the age-depth
curve represented in the figure III.22. It results in a new age-depth curve with instantaneous
events (figure III.23). The small segments represented by a vertical line (parallel to the depth
axis) represent the numerous IE.

In the figure III.23 it can be noticed that: (a) the new age-depth curve show the same
six levels characterized by differences in the SR. (b)The IE with more significant thicknesses

90
are located in the non-marine section, levels 5 and 6 (13525 to 17590 cal yr BP). (c) These
events represent the 44% to 66% of the total deposits of these levels (table III.5), which is
clearly reflected by the increase of the thickness in almost one hundred percent (level 5) or
more (level 6). Although the IE in the marine section (levels 1 to 4) are presented in
considerable numbers (more of 6 instantaneous events for each level), their low thickness (1
mm to 3 cm on average) do not reflect a major change in the curve. A possible explanation for
the increased sedimentation rate in the non-marine environment is presented in the subchapter
3.7. The ages of the instantaneous events and their recurrence intervals are also discussed in
this subchapter.

Figure III.23: Age-depth model for the MD01-2425 core (with instantaneous events). Six levels
characterized by different sedimentation rates are defined. Radiocarbon ages were calibrated using the
software Bcal (Buck et al., 1999) (see table III.4). Right column shows the relative importance of
instantaneous deposits.

91
Thickness Sedimentation
Thickness with Percentage of
without rate without
Level instantaneous instantaneous
instantaneous instantaneous
events events
events events

Level 1 62,5 62,5 1,08 0


Level 2 239,5 222,3 ~ 3,64 7,18
Level 3 365 354,2 ~ 0,80 2,95
Level 4 1119 1064 ~1,40 4,91
Level 5 945,5 526 6,04 44,36
Level 6 512,5 172 0,78 66,43

Table III.5: Major levels defined according to their sedimentation rates in the MD01-2425
core. The percentage of event deposits in the total sedimentary section of each unit is calculated.

92
3.5. The late Quaternary record of the seismotectonic activity in the sediments of the
Sea of Marmara’s eastern basin (core MD01-2425)

3.5.1. Introduction

In order to extract the seismotectonic signal from the of the Sea of Marmara’s
sedimentary archive, specifically from the Çinarcik Basin, we have conducted textural (grain-
size, shape and fabric), magnetic and compositional analyses (estimation of organic matter
and carbonate content by the loss on ignition method) in the sediments belonging to the
MD01-2425 core (following the whole approaches presented in Chapter 1).

3.5.2. Textural analysis of the MD01-2425 core

3.5.2-a. Grain size


The evolution of the grain size parameters (mean size and distribution) in the MD01-
2425 core allows us to divide the sedimentary record in two principal units with different
textural characteristics. The statistical grain size parameters as median, sorting, skewness and
kurtosis are represented in the figure III.24. The biological content and ages of both units
have been discussed in previous subchapters. Taking into account these data, we interpreted
that the upper unit was deposited under marine conditions, while the lower unit was deposited
in non-marine environment.

The upper unit is dominated by muds with mean grain sizes that can vary between 3.2
and 9 µm. Their sorting is poorly (sorting index So: 1.9 to 2.3), with a general symmetrical
distribution (skewness index Sk: 0.07 to -0.05). Grain size frequency curves are leptokurtic to
mesokurtic (Kurtosis K: 0.90 to 1.12). This unit, as was proposed in the subchapter 3.4.3 can
be subdivided into three sub-units (figure III.24). In the three subunits the muds show similar
characteristics (as described above); variations in subunits concern the percentage of very thin
silty-sandy laminæs (facies Ls). The subunit U1 (0 to 350 cm depth) shows the largest amount
of silty-sandy laminæ, and slightly coarser means grain size (9 to 5 µm); it is very poorly
sorted (So between 2.4 to 5), with variable asymmetry (Sk between -0.23 to 0.17). The U2
subunit (350 to 980 cm depth) shows a slightly lower amount of silty-sandy laminæs
compared to the U1. The laminæ show slightly lower mean grain size (9 and 35 µm), and tend
to be better sorted (So between 1.8 and 4.8). The U3 subunit (980 to 1736 cm depth) is

93
characterized by the smallest amount of silty-sandy laminæ, it tends to be better sorted (So
between 2 and 3.2) and presents a more symmetrical particles grain size distribution (Sk
between 0.15 to -0.13). The presence of authigenic gypsum in this last subunit (see annex B)
could slightly modify their textural characteristics.

Figure III.24: The evolution of the textural parameters in the MD01-2425 core. Two units with
different textural behaviors are observed. The upper unit can be subdivide in three subunits U1, U2
and U3. The sampling interval varied between 25 and 5 mm. On the right is represented a simplified
log of the MD01-2425 core with their respective facies.

94
In the lower lacustrine unit, muds represent the dominant lithology. They are similar to
those from the upper unit, but they show a slightly lower mean grain size varying between 2.8
and 8 µm, with asymmetry or asymmetry towards fine grained particles (Sk between 0.09 to-
0.15).

The main difference between the two units is the large amount of instantaneous coarse
to fine deposits (facies Tuc, HmTu, Sd) interlayered within the muds in the lower one. The
fine grain event deposits (Hm) show compositions similar to the muds’ contents. At the
difference, the coarse event deposits have highly variable mean grain sizes (286 to 14 µm),
generally showing asymmetry towards fine grained particles (Sk between -0.54 to -0.1), with
predominant mesokurtic grain size frequency curves (figure III.24).

In the figures III.25, III.26, III.27, we represented the grain size parameters profile
and the average percentage of main grain size of some selected sections along the MD01-
2425 core. These sections were chosen based on: 1) their stratigraphic position; aiming at
studying sediments deposited in marine and non-marine environments; 2) their content of
different types of event deposits of considerable thickness (facies Ls, HmTu, Tuc). Their
stratigraphic position is showed in the figure III.16. These sections show in detail the vertical
evolution of textural parameters.

The figure III.25 shows the high resolution grain size profiles of two sections
belonging to the upper marine unit, these sections are named A and B. These two sections are
composed by deposits type: homogeneous muds (hemipelagites) within very thin silty-sandy
interlayered laminæ. The muds are characterized by abundant silt-size particles (average:
80.36 r 2.81 %), having lesser amounts of clay size (average: 15.36 r 1.76 %) and sand size
(average: 3.27 r 2.20 %) (figure III.26). In general, they present an average mean grain size
of 6.45 r 0.92 Pm. They are poorly sorted (So: 2.16 r 0.16), with a symmetric grain size
distribution (Sk average: 0.036 r 0.043), and mesokurtic grain size frequency curves
(K average: 1.041 r 0.050). On the other hand, the very thin silty-sandy laminæ are mainly
composed by silt-size particles (average: 67.12 r 6.73 %), with considerable percentages of
medium to very fine sand, varying between 42 to 10%; and lesser amounts of clay size
particles (average: 11.76 r 2.87 %) (figure III.26). These laminæ have an average mean grain
size of 13.86 r 5.05 Pm, they are very poorly to poorly sorted (So between 4.98 and

95
Section A

a)

Section B

b)

Figure III.25: Evolution of the textural parameters in two sections belonging to the upper unit
(marine) of the MD01-2425 core. They are composed by deposits type: homogeneous muds (facies
Mh) within very thin interlayered silty-sandy laminæs (facies Ls). On the right side of the figure are
shown photos or X-ray images of the Ls facies. a) Section A. b) Section B. Black arrow indicates a
coarsening-upward followed by a fining-upward succession. The stratigraphic position of the sections
(A and B) are indicated in the figure III.15. Rectangles in blue color (dashed lines) indicate the
location of some events represented in figure III.27.

96
2.28), with an asymmetric grain size distribution toward fine grained or coarse particles (Sk
between 0.22 to -0.38).

The figure III.27 represents three sections belonging to the lower unit (non-marine).
These sections are named D, E and F (stratigraphic position in figure III.16), and are
composed by deposits type: homogeneous muds (hemipelagic-type deposits) interlayered with
homogenites + turbidites. In the non-marine unit, the muds have characteristics similar to
those described in the marine unit; although they can show highest amounts of sand (between
18 and 2%) and clay (28 to 12%) size particles (figure III.25).

The event deposits type homogenites (Hm) have textural characteristics similar to the
homogeneous muds, although, minor differences can be observed. In contrast to the
hemipelagic type deposits (h.d.), the Hm show a slightly better grain size sorting and slight
asymmetry toward the fine grained particles. On the other hand, the coarse turbidites show
different textural characteristic. They are characterized by a significant and variable presence
of sand and silt size particles, with percentages of sand comprising between 84 and 10% and
silt between 79 and 3%; with lesser proportions of clays (average: 7.20 r 5.18%) (figure
III.27). In general, they have a mean grain size comprised between 179 and 10 Pm, showing a
highly variable sorting between well to very poorly sorted (So between 1.3 and 3.8), and an
asymmetric grain size distribution toward the fine grained particles.

Figure III.26: Average percentage of sand, silt and clay size particles presents in different
lithofacies of the MD01-2425 core.

97
Section D

a)

Section E

b)

Figure III.27: Evolution of the textural parameters of three sections belonging to the lower unit (non-
marine) of the MD01-2425 core. They are composed by deposits type: homogeneous muds
(facies Mh) interlayered with Homogenites+turbidites (facies HmTu) a) Section D. b) Section E.
c) Section E. The stratigraphic of the sections are indicated in the figure III.16. Rectangles in blue
color (dashed lines) indicate the location of some events represented in the figures III.29, III.30.

98
Figure III.27 (continued)
Section F

3.5.2-b. Base-to-top binary diagrams: Skewness vs Sorting and CM Passega’s


diagrams
In the figures III.28 and III.29 are represented the bottom-to-top textural evolution of
some instantaneous events from different units. In the figure III.28 are plotted the skewness
index (Sk) vs. sorting index (So) and the CM Passega’s diagrams of four events belonging to
the silty-sandy laminæ facies (Ls).

- Upper unit (marine)


The events 1 and 2 are characterized by a very thin parallel laminætion, as can be
clearly observed in the X-ray images named “e” in the figure III.25b. Additionally the
laminæ belonging to these types of events can show a tendency to increase its grain size
towards the middle part of the layer which contains them (see figure III.25a section A, event
named “a”, and figure III.25b section B event named “e”). These two events (figures III.28-
1, -2) show a So which evolves from bottom to top into poor sorting, due to a progressive
enrichment in fine particles. Likewise, in the CM diagram, these deposits exhibit a slight
increase in the median grain size from the bottom to the middle part of the sequence, reaching

99
toward the top median grain size values similar to those present in the bottom. Otherwise, the
Q99 display a progressive grain size decreasing.

On the other hand, on the right side of the figure III.28 are represented the events 3
and 4, which are characterized by the presence of flaser laminætion, as can be see in the X-ray
images “f” in the figure III.25b. They show a different So vs. Sk and CM bottom-to-top
textural evolution compared to the events 1 and 2. In the So vs. Sk binary diagram, the events
3 and 4 exhibit a progressive evolution of So, which evolves from bottom to top in extremely
poorly to poorly sorted. Sk shows more complex variations, displaying excess of coarse or
fine grain size particles. The CM diagram also shows a complex behavior from the bottom to
the top of the sequence. The median grain size oscillates between 10 to 50 µm, and the Q99 is
slightly variable (600 to 200 µm).

- Lower unit (non-marine)


In the figure III.29 are plotted So vs Sk and CM bottom-to-top textural evolution of
five instantaneous events type HmTu from the non marine unit. The first event is the thickest;
it is constituted by a basal Tu unit of 16 cm thick and a Hm unit of 53 cm thick. The others
HmTu events plotted have lower thicknesses which vary between 21 and 10 cm. In general
the HmTu events (figure III.29 -1 to 5) are characterized by:
(a) a basal Tu unit with complex variations of their sorting, skewness, Q99 and
median grain size (represented with dark to medium blue triangles in the figure III.29
-1).
(b) an inunitediate unit characterized by the presence of to-and-fro sedimentary
structures, with a less variable So and median grain size. Event 1 (located between
2090 and 2166 cm depth) exhibits a clear separation from the coarse basal unit and the
homogeneous upper unit (light blue triangles in the figure III.29-1).
(c) sharp and clear separations between the Hm and the intermediate and basal
Tu unit. The Hm unit shows a highly constant sorting and median grain size. The
skewness and Q99 are almost invariable, but in the event 1 (fuchsia dots in the figure
III.29-1) they are highly variables, with an oscillating behavior.
(d) Hm and the hemipelagic-type deposits (pink rhombuses) plotted in same
areas of the diagrams; nevertheless, the hemipelagites can show a more variable So,
Sk or Q99, compared to the Hm unit.

100
Figure III.28: Bottom-to-top textural evolution of four events belonging to the silty-sandy laminæs facies (Ls). The events are plotted in the So vs
Sk and Q99 vs medians Passega (1964) diagrams. Events 1 and 2 are characterized by a very thin parallel laminætion, whereas Events 3 and 4
show flaser bedding structures. The locations of some of these events are represented in figure III.25 (rectangles in blue color (dashed lines). The
sampling intervals vary from 0.2 to 0.5 cm.

101
102
Figure III.29: Bottom-to-top textural evolution of fiver events type HmTu belonging to the
non-marine section. The events are plotted in the So vs Sk and Q99 vs median diagrams
(Passega, 1964). The locations of some of these events are represented in figures III.25, III.27
(rectangles in blue color (dashed lines). The sampling intervals vary from 0.2 to 0.5 cm.

103
3.5.2-c. Particles shape of silty-sandy layers
A complete particles shape analysis was conducted on the fine grain sediment fraction
(0.6 to 100 µm) of the MD01-2425 core. In the figure III.30a, b are represented the
circularity vs. diameter (Pm) diagrams and the particles images obtained from the upper
homogenous part of two event deposits and their overlying hemipelagic deposits.

In the figure III.30a is represented an HmTu event belonging to the section D


(stratigraphic position on figure III.16, grain size analyses in the figures III.27, III.29-2) and
their overlying hemipelagite. Two samples were studied, one taken at 2685 cm core depth,
corresponding to the Hm unit, other taken at 2674 cm core depth corresponding to the
hemipelagic deposits. The circularity vs. diameter (Pm) diagrams show that both samples are
made of particles with similar circularity and diameters. The particles are clearly grouped into
two distinct zones of the diagram. A group of particles with diameters between 1.5 and 10 Pm
concentrate in the high circularity (>0.85) zone. The other group of particles with diameter
between 0.6 and 7 Pm, are characterized by a more variable circularity (0.4 to 0.7). The
particles images correspond principally to calcite crystals, fragments of diatoms and
fragments of terrigenous components. In general the grains have very angular to sub-angular
edges.

An HmTu event belonging to the section E is represented in the figure III.30b


(stratigraphic position on figure III.16, grain size analyses in the figure III.27). In this
section we studied two samples taken at 2740 cm core deph (Hm unit), and at 2731 cm core
depth (h.d.). These samples show the same behavior as those studied in the section D, but
with a lower dispersion around the circularity values. The figure III.30b shows a group of
particles (diameter between 1.5 and 7 Pm) generally concentrated around circularity values
>0.9, other group (diameter between 0.6 and 6 Pm) is concentrated around circularity values
comprising between 0.4 and 0.6. Shape analyses results represented the figure III.30 do not
discriminate homogenites from hemipelagic type deposits. 40 additional analyzed samples
confirmed this result.

104
Figure III.30: Shape particles analyses conducted in two homogenites and two hemipelagites
belonging to the MD01-2425 core. In the figure are represented the photo of the studied section,
the circularity vs. diameter (µm) diagram and the particles images obtained from the SYSMEX
FPIA – 2100.

105
3.5.2-d. Magnetic fabric
The sediment grain array or fabric is, beside the grain size distribution, the second
major textural property characterizing the settling mechanism. It is controlled by the shape of
the particles (platy, elongated, isotropic) and the hydrodynamic conditions during settling
(Boggs, 2011). As exposed in subchapter 2.4.2, we developed the use of Anisotropy of
Magnetic Susceptibility (AMS) to detect and characterize the sediment fabric, especially for
the silty-clayey layers. Bulk (or volume) magnetic susceptibility (MS), isothermal remanent
magnetization (IRM, or Hard IRM) and anhysteretic remanent magnetization (ARM) were
performed to deunitine the magnetic carriers and their grain size distribution, on order to
strengthen the ASM’s values interpretation.

The MS profile of the MD01-2425 core is represented in the figure III.16 (see
subchapter 3.4.3). This profile shows the evolution of the concentration of magnetic particles
during the last 17 cal kyr BP. It shows a marked difference in the behavior of the MS between
the marine and non-marine units. It is also possible to subdivide the marine unit into different
subunits with different behavior of the magnetic susceptibility curve. The upper marine
subunit (U1), located between 0 and 3.5 m depth shows values high and variable MS values
(mean MS: 34.81 r 7.3 x10-5 S.I.) compared to the lower marine subunits (U2 and U3). The
latter are located between 3.5 and 16.9 m depth, are relatively constant and show low values
of MS (mean MS: 10.10 r 1.8 x 10-5 S.I.). The non-marine unit has highly variable MS
values. In this unit, MS can vary from 300 to 20 x 10-5 S.I., with mean values of 26.88 r 12.2
x 10-5 S.I. The marked difference in the MS profiles between the lower marine and the non-
marine units allows differentiating these two environments using the MS profiles.

As the magnetic fabric of the MD01-2425 core was already presented in subchapter
2.4.2, we will refer to these results and, here, illustrate three others examples located at
different stratigraphic positions, in order to have a more complete vision of the AMS behavior
along the core.

In the figures III.31a, b, c are represented three section of 30 to 90 cm long, one of


them marine environments, the two other sections from the non-marine unit. Their
stratigraphic positions are showed in the figure III.16.

106
The section A (220 to 260 cm depth) presents two (2) very thin silty-sandy laminæ
exhibiting parallel and flaser laminætion (facies Ls), which are intercalated within continuous
hemipelagic-type sediments (figure III.31a). In this section the average value of the mean
(volume) magnetic susceptibility (Km) is 257 ± 53 x 10-6 S.I.; highest values appear in the
upper silty-sandy laminæ (442.75 x 10-6 S.I.). The section shows a slightly variable magnetic
foliation (F) with a maximum values in the Ls laminæ (F=1.076). The magnetic lineation,
shape parameter and minimum susceptibility axis are highly constant, with a very low
magnetic lineation (L= 1.005±0.003), oblate magnetic ellipsoid shapes (0.63<T<0.96), and a
minimum susceptibility axis very close to the vertical (76< Inc Kmin <90°).

The section C (1995 to 2025 cm depth) comprises deformed sand, silts and muds
(facies Sd). It shows a of 143 ± 97 x 10-6 S.I. Km average; highest values appear in the coarse
grained deformed sand-silty layers (365 x 10-6 S.I.). The magnetic foliation, lineation, shape
parameter, and minimum susceptibility axis are moderately to highly variables. The maximum
F values are observed in the mud located immediately above of the deformed layer of sand-
silt (F= 1.68), the magnetic ellipsoid shape varies between prolate to oblate (-0.74 <T< 0.82),
and the minimum susceptibility axis shows Kmin inclinations highly variable (from 2° to 82°).

The section E (2715 to 2810 cm depth) presents 4 HmTu type event deposits,
intercalated within continuous hemipelagic-type deposits. This section has the same behavior
with the sections studied in the subchapter 2.4.2. The Km shows a tendency to increase with
the increase of the grain size; highest values were measured in the Tu units (mean: 387 r 229
x 10-6 S.I.). The magnetic foliation is highly variable, with highest values in the upper
homogeneous unit Hm (1.10 to 1.14), while the basal Tu unit shows the lowest (1.10 to 1.06).
Likewise, the difference in the F values between homogenites and hemipelagites is high. The
magnetic lineation is very low and almost invariable (between 1.001 and 1.013). All
sedimentary deposits (HmTu and h.d.) have oblate magnetic ellipsoid shapes (0.71 <T <1)
with a minimum susceptibility axis very close to the vertical (75°< Inc Kmin <88).

The systematic vertical orientation of Kmin, perpendicular to the bedding plane (75° <
Kmin) in the sections A and E (figures III.31a, c) suggest a primary sedimentary fabric,
neither affected by any deformation (natural or during coring and subsampling) nor by
bioturbation. To the contrary, in the section C, the highly variable Kmin inclinations, usually
away from the vertical (mean Kmin: 46 r 25) indicate a clear modifications in the primary
107
fabric, which can be visually identified due to the presence of a ball and pillow sedimentary
structure possibly generated by liquefaction processes.

As discussed in the subchapter 2.4.2., the contrast between, on one hand, the high F
values present in section A’s Ls facies and in section E’s Hm units, and in the surrounding
hemipelagic type deposits, on the other hand, cannot be explained by differences of
compaction. In these sections are observed different values of F in different lithofacies (Ls,
HmTu, h.d.) located at the same depth within the sedimentary column. The MD01-2425 core
does not show any progressive increase of the F values with the depth. Therefore, the marked
differences in the F values between the lithofacies are basically due to grain arrangement (or
grain array); therefore, they are directly related to the different transport and deposition
conditions and mechanisms (grains settlement).

108
Section A

Section C

Section E

109
Figure III.31: Photos and AMS profiles of three sections of the MD01-2425 core. a) Section A.
Two very thin silty-sandy laminæ (facies Ls). b) Section C. Deformed sand, silts and muds (facies
Sd). c) Section E. Four HmTu events intercalated within continuous hemipelagic-type deposits.
The stratigraphic position of the three sections is represented in the figure III.15.

3.5.2-e. Magnetic content


The different mineralogical and granulometric fractions within the magnetic
components of the MD01-2425 core were identified using two types of laboratory remanent
magnetisations: isothermal remanent magnetization (IRM) and anhysteretic remanent
magnetization (ARM). Additionally, the S-ratio from Bloemendal et al. (1992) and HIRM
from Robinson (1986) were calculated. The main results and their interpretations are
discussed in the subchapter 2.4.2. In order to reinforce our interpretation, in the present
subchapter are presented the remanent magnetizations profiles of the three sections (figure
III.32) discussed above (magnetic fabric subchapter).

The remanent magnetizations profiles of the three studied sections shown on figure
III.32. The profiles plotted are: (1) S-ratio, which is used to estimate the magnetic mineralogy
(-e.g., magnetite or hematite/goethite- Bloemendal et al., 1992). (2) HIRM, which is applied
as a measure of the concentration of the high coercivity minerals (Robinson, 1986). (3) JARM
(intensity of the ARM), which measure the concentration of ferromagnetic minerals. It is
grain size dependent, and preferentially responds to small magnetite grain size (<10 µm)
(Stoner et al., 1996). (4) Km/J IRM1T ratio. When magnetic mineralogy is dominantly
magnetite, this ratio depends on the magnetic grain size (Stoner and St-Onge, 2007).

In the three studied sections, the S-ratio values >0.9 suggest a predominance of low-
coercivity mineral component (magnetite). This means that the magnetite is the predominant
magnetic mineral along the whole sedimentary record in the MD01-2425 core. Despite this,
the behavior of the HIRM profile in the sections A and E (figure III.32a, c), as well as in the
profiles studied in the subchapter 2.4.2., indicates that in some parts of the sedimentary record
of the MD01-2425 core very small amounts of high coercivity “hematite type” materials can
also be present. Moreover, the occurrence of relatively high values HIRM in the section C
(figure III.32b), jointly with low values of Km/J IRM1T ratio, may indicate that in this
section Single Domain (SD) magnetite is present. Indeed, SD magnetite grains are usually
characterized by low values of MS and high coercivity (Dunlop, 1981). On the other hand, the
the variations of the magnetic particles size (basically magnetite)

110
Section A

Section C

Section E

111
Figure III.32: Magnetic parameters profiles (JARM, Km/J IRM1T, S ratio and HIRM) of a
selected sections from the MD01-2425 core Magnetic foliation is added to check its relationships
with magnetic content. a) Section A. Two very thin silty-sandy laminææs (facies Ls). b) Section C.
Deformed sand, silts and muds (facies Sd). c) Section E. Four HmTu events intercalated within
continuous hemipelagic-type deposits The stratigraphic position of the three sections is
represented in the figure III.15.

versus the intensity of ARM (JARM) and the Km/J IRM1T ratio is not clear. In the section A
the JARM profile is highly regular indicating a constant grain size of the magnetic particles.
In sections C and E, relatively high values of JARM in the coarser units which decrease
towards fine grain size sediments seem to indicate a decrease of the magnetic particles grain
size toward fine grained deposits. The predominance of magnetite as magnetic mineral along
the entire core, and the low values of JARM in the fine grain size sediments, suggest that the
magnetic foliation in the MD01-2425 core is not associated to size or type of magnetic
minerals. This results concern both hemipelagites or homogenites. It confirms the proposed
interpretation, where high values of ASM foliation are due to grain array, thus to specific
depositional process.

3.6. Estimation of organic matter and carbonate content in the MD01-2425 core
(Loss on ignition)

Loss on ignition (LOI) were performed on mud recovered in the marine and non-
marine units (see subchapter 2.4.2). The evolution of these two parameters versus the depth is
presented in figure III.33. The percentages of organic matter are highly variables in both
units. The OM percentage in the marine sediments is moderately higher (10.12 r 1.80 %) than
in the non-marine sediments (7.76 r 1.61 %). Two peaks are present in the marine section.
The first peak is located between 377 and 498 cm depth (2.8 to 3.7 cal kyr BP), with OM%
values between 13.43 and 12.03%. A second peak located between 1004 and 1067 cm depth
(7.3 and 7.6 cal kyr BP), shows OM% values between 13.65 and 12.77 %. These peaks may
correspond to the sapropelic layers previously defined in the Sea of Marmara. The first peak
is correlated with the sapropelic layer described by Ça÷atay et al. (1999; 2000) for the period
between 2.8-4.9 cal kyr BP, while the second is correlated with the sapropelic layer defined
by Ça÷atay et al. (2000), Aksu et al. (2002) and Vidal et al. (2010) and dated between 7.5 and
11.5 cal kyr BP.

112
As the OM%, the carbonate content is highly variables in both units, but in contrast to
the OM% the highest percentages of carbonate are observed in the non-marine unit (average
of 11.26 r 2.09 %), while lower percentages are presents in the marine unit (average of 9.73 r
0.94 %). This phenomenon was also observed through the study of smear slides.

Figure III.33: Weight percentage of organic matter and carbonate content obtained by the Loss
of ignition (LOI) method. The measures were performed over 56 samples taken in the MD01-2425
core.

113
3.7. Data interpretation and conclusion on depositional processes in the MD01-
2425

3.7.1. Depositional processes

The different observation and measurements presented before lead to identify four
predominant types of sedimentary facies (Mh, Ls, Tuc and HmTu). These compositions,
textures, and layering types are the product of different mechanisms and environments of
transport and deposition.
- Facies Mh and Ls
The hemipelagic type deposits (facies Mh) are composed of massive mud deposits.
They correspond to slow and rather continuous settling down of planktonic bio-produced
and bio-induced particles, mixted with fine-grained terrigenous material. On the other
hand, the very thin silty-sandy laminæ (facies Ls) are characterized by the presence of
parallel or flaser laminætion (figure III.25). In both cases the presence of these types of
sedimentary structures are indicative of different process of transport and deposition,
which is clearly evidenced in the Sk vs So and CM Passega’s diagrams plotted for both
cases in the figure III.28.

The silty-sandy laminæ that exhibit parallel lamination, can also show a slight
increase in the grain size in the middle of the layers (coarsening-upward and fining-upward
succession), (figures III.25a, b, III.28-1, -2). We interpret that these units were deposited
by fall-out from graded turbulent flows which gradually evolved into uniform turbulent
flows (Passega, 1964; Mulder and Alexander 2001). The turbulent flow seems to show a
gradual increase and subsequent decrease in its speed. It is reflected in the capabilities to
transport particles of larger grain size, observed in the middle part of the laminated units.
The thickness of the sequences and their grain size indicate that this facies was deposited
in the distal part of a turbiditic system, suffering gradual flow velocity fluctuations. These
fluctuations are typical of a long-lived turbulent flow, and suggest the existence of a
hyperpycnal system product of hyperpycnal river flooding (Mulder and Alexander 2001;
Zavala et al., 2011).

The silty-sandy laminæ which exhibit flaser lamination show an almost progressive
improvement in their grain size sorting (So evolve from 5 to 3), but with a highly variable

114
skewness (Sk between -0.25 to 0.25), median grain size (M between 10 to 70 Pm) and Q99
(C between 200 and 700 Pm). This is not a random variability and shows a tendency to
oscillate between high and low values. These characteristics, as well as the occurrence of
to-and-fro sedimentary structures, allow us to interpret that they were deposit by processes
of traction and fall-out from oscillation-dominated flows.

The depth of water wherein the studied silty-sandy laminæ were deposited (>1000
m.b.s.l.) allows us to consider that the responsible oscillatory flows are neither the product
of wave action nor the result of tidal currents. In closed or semi-closed deep water basin
(marine or lacustrine) the presence of oscillations currents has been associated to reflected
tsunamis or “seiche effect”. In regions with strong seismic activity, several authors (Sturm
et al., 1995; Chapron et .al., 1999; Beck et al., 2007, 2012; Bertrand et al., 2008; Carrillo et
al., 2008; Ça÷atay et al., 2012) relate these feature to seismic shocks and/or subaqueous
landslides. As the Sea of Marmara is a region seismically very active, this could be an
explanation for the presence of oscillatory flows in this depth of the basin.

- Facies Tuc
The Tuc events are massive or display normal gradation, currents ripples and/or
parallel laminætions. We interpret that they were deposited by processes of traction and
fall-out from turbulent flows (Mutti et al., 1999; Mulder and Alexander 2001). The
massive Tuc deposits indicate a fast deposition, avoiding the formation of sedimentary
structures and a good sorting of the grains (Lowe, 1982; Mutti, 1992).

- Facies HmTu
The HmTu events in the MD01-2425 core are constituted by two well differentiated
units (Hm and Tu), showing in some of case in the middle of both an intermediate unit
with very specific characteristics (see figure 1 subchapter 2.4.2). The basal coarse unit Tu
shows a highly variable So and Sk, with a median grain size and Q99 which show a
progressive diminution from the bottom to the top of the unit (figure III.29). We interpret
that this Tu unit was deposit by processes of traction and fall-out (Passega, 1964; Lowe,
1982; Mutti et al., 1999). It shows a progressive transformation in its mechanism of
transport of sediments, clearly observed in CM diagram (figure III.29). The bottom are
characterized by an important bed load transport of coarse particles (coarse to medium
sand), although the transport by suspensions is not negligible (Passega, 1964). In the
115
middle and upper part of this unit, composed mostly by fine sand and silt, the sediment are
mainly transported in suspension by turbulent flows (Mutti et al., 1999).

The intermediate unit can be clearly seen in the thickest HmTu event deposit
studied between 2090 and 2166 cm core depth. The parameters So vs. Sk and CM
diagrams are represented in the figure III.29-1 (light blue color triangles), the others event
deposits illustrated in the figure III.29 can also develop this intermediate unit (light blue
color triangles), but their small thickness does not allow a clear observation of their
textural evolution compared to the first bigger event. In general the intermediate unit
shows an almost invariable sorting and median grain size (M between 9 to 20 Pm), with a
Q99 which can be highly variable, as can be see in the thickest HmTu event deposit (C
between 100 and 700 Pm).

The textural characteristic and the presence of to-and-fro sedimentary structures in


the thickest HmTu event deposit, allow us to consider two different process of transport
and deposition almost simultaneous. First: the coarse sediments (coarse to fine sand size
Q99) particles) were deposited by processes of traction and fall-out from oscillation-
dominated turbulent flows. Second: the predominant fine grain size sediments (median
highly constants of fine silt grain size) were deposited by a rapid decantation from high
density suspension (HDS). These HDSs have been studied in laboratory (Lamb et al.,
2004), and they are the result of high-density turbulent suspension, generated by the
remobilization of fine sediments (mud) from the bottom of a basin product of wave action.
Moreover, the intermediate unit may be deposited under the same processes, with high
density suspension as predominant process.

The upper homogeneous units (Hm) shows a highly constant distribution of the
grain size parameters, compared to the basal Tu unit and the intermediate unit. They show
a highly constant sorting and median grain size (lower than the intermediate unit,
comprised between 7 and 10 µm), whit a more variable Sk and Q99. This later is highly
variable in the thickest HmTu event deposit (figure III.29-1). Compared to the Tu unit, the
textural parameters in the Hm unit indicate a stabilization of the energy of the depositional
environment, dominated by fall-out from a long-lasting fine-grained and homogenous
suspension (HDS and HS). For its part, the highly variable and oscillating Q99 in the

116
thickest HmTu event deposit indicated the presence of oscillating high velocity currents
able to mobilize particles up to 1 mm.
The hemipelagic type deposits (h.d.) and the Hm units are generally located in the
same region of the CM Passega’s diagrams, despite this; in the majority of the cases the
h.d. do not show high values of Q99, and their samples are much more scattered within the
So vs .Sk and CM diagrams. This indicates that they were deposited in calm conditions.

The magnetic fabric and the shape analysis corroborate the interpretations made
with the grain size parameter for the HmTu events and the hemipelagic-type deposits
(h.d.). The AMS foliation parameter (F) confirms the difference between the lower (Tu)
and the upper (Hm) units. It shows a marked difference between the Hm unit and the
overlying hemipelagic deposit (subchapters 2.4.2 and 3.5.2). This difference is very useful,
because both facies show similar grain size and shape parameters. The magnetic fabric,
specifically the F, indicate different grain array, which can be interpreted as two different
depositional conditions. The low magnetic foliation (“normal” with respect to burial depth)
present in the h.d. is characteristic of a primary sedimentary hemipelagic fabric, where
calm conditions and slow continuous settling prevailed. The strong planar horizontal array
in the Hm unit fits with the hypothesis of a specific settling. The re-suspended fine fraction
is deposited by fall-out under oscillatory conditions (associated to “seiche effect”),
favoring a better arrangement of the particles, mostly with platy shape, in planes parallel to
the bedding surface. On the other hand, the slightly high foliation observed in the silty-
sandy laminæ which exhibit flaser bedding may indicate a depositional mechanism similar
to those observed in the Hm units.

3.7.2. Depositional model for the HmTu events


The following singular characteristics of the HmTu events have to be highlighted:
(1) the presence of two visually distinguishable units (Tu and Hm), which present
very different textural characteristics;
(2) the thickness of the Hm unit, generally or much developed than the Tu unit
(up to 90 % of the wheole event);
(3) the limit between the two units, which is generally represented by an
intermediate unit with sedimentary structures indicating to-and-fro grains
displacements such as low angle asymmetric cross bedding, similar to coastal
tidal “flaser” bedding;

117
(4) the evolution of the textural parameters as median grain size and Q99 in the
Hm unit, indicating the presence of oscillating high velocity currents capable
of mobilizing particles up to 1000 µm.
(5) the specific planar array in the homogenites (high magnetic foliation values).
(6) a singular geometry of a large (8 m thick) HmTu events, identified in the
Central Basin by Beck et al. (2007), using high resolution seismic imagery
and core analysis. This event is characterized by: (1) its spatial distribution
restricted to the deepest parts of the basin (below 1190 m depth); (2) sharp
lateral unitinations; (3) increased thickness with depth; (4) the lack of visible
lateral gradient indicating a source area.

These particular sedimentary event deposits represent subaqueous sediment density


ßows, but they do not show the characteristics of a short-duration classical turbidite or so-
called surge-like turbidite (Bouma, 1962; Mulder and Alexander, 2001). The classical
turbidites do not exhibit sedimentary structures indicatives of bidirectional (oscillatory)
currents, show a progressive decrease of its thickness away from the source, as well as,
show a gradual transition between the bed load and suspended load deposits. The
characteristics of the HmTu events, specifically, the evidence of the existence of
oscillatory currents and the volume and homogeneity of the suspended load lead us to
proposed two alternative depositional models (figure III.34):
a) The HmTu events can be the product of high velocity turbidity current which
undergoes several reflections against the steep slopes of the Çinarcik Basin
with a progressive decrease of their sediment load and flow velocity.
Following the deposition of most of the sand load, the residual silt-mud
probably still moving back and forth across the basin, gradually losing their
velocity, finishing with a phase of fall-out of mud from the ponded
suspension to form a thick, homogeneous mudstone cap (Pickering and
Hiscott, 1985). This model was proposed for the unusual turbidites
(“contained turbidites”) generated by a catastrophic slope failures in the
Cloridorme Formation, Quebec (Pickering and Hiscott, 1985), and is
illustrated in figure III.34a.
b) The second hypothesis is based on the models proposed by Beck et al. (2007)
and Beck (2009) (figure III.34b). In this model the earthquakes are
considered as the mechanism for triggering sediment failures, which evolve

118
into mass wasting or slumps, and which progressively are transformed into
density and turbidity currents. The seismic shock is also supposed responsible
in inducing an oscillatory movement of the whole water mass (reflected
tsunami/seiche phenomena), due a co-seismic bottom displacement, or to a
huge sub-aqueous slump. In the Sea of Marmara these oscillatory movement
of the water mass are responsible for: (1) the increase of the suspended load
by the additionally extraction of the clay-silty matrix from the initial flux,
which will be the major component of the high density and homogenous
suspension (HDS and HS). These suspensions can be maintained for several
days to weeks, as has been reported in the Cariaco Trough and north-west
Haiti after recent major earthquakes (Thunell et al., 2009; McHugh et al.,
2011); (2) the generation of a sharp contact with structures of oscillatory
currents between the two HmTu units; and (3) the creation of specific settling
conditions, producing the specific planar array in the homogenites.

a)

b)

119
Figure III.34: Different models that explain the formation of the association Homogenite +
turbidite (HmTu). a) Model proposed by Pickering and Hiscott (1985) for the formation of
“contained turbidites”. b) Model proposed by Beck et al. (2007), Beck (2009) to explain the
mechanism of formation of the association HmTu. Earthquake-induced acceleration triggers slope
failures and mass-wastings which evolve into turbidite currents. Synchronous earthquake-induced
seiche effect increases segregation of the fine-grained fraction and sustains its suspension.

3.7.3. Possible triggering mechanism of sedimentary HmTu events

According to the depositional model proposed by Beck et al. (2007) for large lakes
or closed marine basins, the triggering mechanism of the homogenites+turbidites deposits
is a combination of “reflected tsunami” and huge slope failure evolving into mass-wasting.
Often, the water mass movement is initiated by direct tectonic bottom modifications (faults
offsets, sudden subsidence), but large redepositional processes (fluidized landslides) are
necessary. On the other hand, huge subaqueous landslides alone may also create water
mass displacement and they are not necessarily related to a significant earthquake. This
uncertainty may be solved when several separated synchronous landslides are detected, and
with evidences for ground accelerations through abrupt pore water increase (liquefaction,
microfracturation, water escape, etc.)

After the 1999 Izmit–Kocaeli earthquake, in the Sea of Marmara the distribution,
morphology and trigger mechanisms of the slope instabilities have been the object of
various studies (e.g. Gazio÷lu et al., 2002; Görür and Ça÷atay, 2010; Özeren et al., 2010;
Zitter et al., 2012). As a result of these studies, Zitter et al. (2012) provide a map of the
mass wasting features, active faults and canyons areas in the Sea of Marmara (figure
III.35a). These authors identified that ~30% of all slopes around the Sea of Marmara
(more than 500 km2) show one or several of these elements. They are more important in
the southern slopes, characterized by higher sediment concentration and less steep slopes;
although these features are not negligible in the northern slope, especially in the Çinarcik
Basin (figure III.35a). This basin is surrounded by important landslides and mass wasting
features (features 25, 27, 28, 29, 30, 32 in the figure III.35a) such as: the Tulza Landslide
Complex (32 km2) located on the northern slope and the Yalova Complex (21 km2) located
on the southeast; among others (figure III.35b).

120
The mass wasting features could be the origin of the HmTu deposits in the MD01-
2425 core. They may have originated in any margins of the basin, including the Central
High slope facing of the Çinarcik Basin (features 25, figure III.35a). The presence of blue
and green amphibole and epidote in all lithofacies identified in the MD01-2425 core
(subchapter 3.4.2) allow us to propose a source of sediments predominantly from the
southeast, south or east of the Çinarcik Basin, as proposed by Okay and Ergun (2005)
based on heavy mineral analysis (subchapter 3.3.4). The possible source of these minerals
are the Karakaya HP metamorphic Complex (see subchapter 3.2.3 and figure III.5).

On the other hand, according to historical observations in the Sea of Marmara,


these sediments failure are generally earthquake-induced, which, in many cases, have
given rise to tsunami events (Görür and Ça÷atay, 2010). These historical records reveal
that 16 tsunami events occurred in the past two millennia, which have connection with
major earthquakes (Ambraseys 2002a). These events indicate the possible presence of
oscillatory currents type “seiche” in the Sea of Marmara, which would be responsible for
the homogenites deposits. Sediments failure induced by sediment overloading or gas
escapes in the Sea of Marmara are not discarded, but they would generate classical
turbidites type deposits.

121
a

28°50’E 29°5’E 29°20’E


40°55’N
b
Figure III.35: Map of the Sea of Marmara with
Avalanche debris
Tulza landslide morphological-tectonic features (from Zitter et al., 2010).
a) Map of the mass wasting features (enumerate 1 to 32), active
faults, canyons and fluid seepage areas in the Sea of Marmara.
Slope gradient b) Slope gradient map of Çõnarcõk Basin, derived from EM300
multibeam data.
40°40’N

Mass wasting features

122
3.7.4. Time distribution of HmTu-tpe sedimentary events in the MD01-2425 core

After the identification and posterior characterization of earthquake-induced HmTu


events in the MD01-2425 core, their temporal distribution were estimated. This allowed
calculating their average time span between successive earthquake (recurrence interval RI). In
the figure III.36 the vertical and temporal distribution of all HmTu events recognized in the
MD01-2425 core are reported, so the sedimentation rates for the hemipelagic deposits (blue
color). The methodology used for the construction of this age vs depth curve is explained in
subchapter 3.4.4..

The figure III.36 shows that the HmTu event deposits are concentrated in the non-
marine section, comprised between 13.1 and 17 cal kyr BP. After this period, a drastic
decrease in the frequence and thickness of coarser instantaneous events is observed, as well as
a decrease in hemipelagic deposits’sedimentation rate (6.04 mm/yr and 0.78) both evidencing
a decrease of the (fine to coarse) terrigenous input into the basin during the marine period.

The difference in the terrigenous contributions and sedimentation rates between the
marine and non-marine unit could be strongly influenced by climatic changes. The non-
marine unit was developed in a lowstand phase during the last glacial period, characterized, in
the Northeastern Mediterranean realm, by dry and cold climate conditions; with absence of
significant tree cover and development of steppic plants (Mudie et al., 2007, Valsecchi et al.,
2012). These climatic conditions are inferred to favor higher erosion and subsequent higher
terrigenous material production (e.g. Macklin et al., 2002; Wegmann and Pazzaglia 2009),
which were transported and deposited on the margins of the basin. This increase in the
accumulation of sediment in the basin margins can directly affect the slope stability, make
easier the development of subaqueous slumps by earthquake triggering. Moreover, Ça÷atay et
al. (2000, 2003, 2009) and Polonia et al. (2002) identified in the Izmit Gulf and on the
northern shelf (through multibeam bathymetric maps and high-resolution seismic profiles) a -
85 m paleoshoreline, which probably was exposed during the last glaciation, when the Sea of
Marmara was disconnected from the Aegean Sea (before a14.7 kyr BP, according Vidal et al.,
2010). During this low stand phase, the exposed shelf suffered an active process of erosion.
Its resulted in high amounts of sediment available to be transported to the bottom of the basin,
and therefore an elevated sedimentation rate. This increase in sedimentation rate has also been
reported for the same period by other authors (Beck et al., 2007; McHugh et al., 2008;
123
Ça÷atay et al., 2009). In the non-marine sequence sedimentation rates can be up to three times
higher compared to the marine period (Ça÷atay et al., 2000; Seeber et al., 2006).

14
Figure III.36: Proposed time distribution of HmTu events based on C dating. Deduced
sedimentation rate is applied to “hemipelagic” constant supply, without the inferred
instantaneous deposits.

124
The marine unit was developed during the warm and humid Holocene period, which
was characterized by the development of Mediterranean forest (Valsecchi et al., 2012). This
fact limits the process of erosion and subsequent terrigenous material production, decreasing
the accumulation of sediment in the basin margins. In combination with this reduction of the
terrigenous production a rapid sea-level rise occurred, resulting in a decrease of the erosion
processes on the shelf limiting the amount of sediment available to be transported to the
bottom of the basin.

In the MD01-2425 core, 27 HmTu events were identified in the non-marine section.
These events occurred during a period between 12.8 and 17 cal kyr BP; representing an
average recurrence interval of a155 yr. On the other hand, in the marine unit, no major mass-
flow event has been clearly identified. However, other event deposit of less important
thickness (facies Ls) have been recognized in this unit. Some of them (35 in total) show
sedimentary structures and textural characteristics which can be considered as earthquake
imprints (flaser lamination, ball and pillow structures, boudinage of laminated layers, etc.). If
we consider that these events were induced by earthquakes, a recurrence interval (RI) of a
365 yr can be estimated.

The RI of major earthquakes (Ms • 6.8) along the northern branch of the North
Anatolian Fault are estimated to be between 200 to 300 yr (in Zitter et al 2012); although it
may vary along the different segments of the NAF. For example, in eastern NAF, near the
Erzincan and Yaylabeli cities the earthquake RI varies between 200 and 685 yr (Barka and
Kadinsky, 1988; Kozaci et al., 2011). In the Gazikoy-Saros segment of the NAF, the RI is
comprised between 250 and 300 yr (Rockwell et al., 2001). Whereas, for the fault segments
lying beneath the eastern, central and western parts of the Sea of Marmara the average RI
varies from 255, 258, and 286 yr respectively (Utkucu et al., 2009); although locally, in some
segment of the fault it may vary from a minimum return period of 166 years to a maximum of
354 yr (Utkucu et al., 2009).

If we compare our RI for the marine and non-marine unit, with the historical RI
deunitined by others authors, the following observation can be done:
- in the non-marine unit the earthquake return period is very short (a155 yr) compared
to the average RI deunitined by Utkucu et al. (2009) for the •inarcik Basin (255 yr);

125
although it is close to the minimum RI deunitined by the same author for this eastern
region (166 yr);
- in the marine unit our average RI is very large (a365 yr) compared to the non-marine
RI, although this value is within the maximum ranges observed in the •inarcik basin.

The difference observed in these apparent recurrence intervals between the marine and
non-marine units could be explained by:
 a higher amount of material available to be transported to the bottom of the
basin during the non-marine period. The high sediment feeding of slopes made them
highly sensitive to local and regional seismic activity, including earthquakes that could
present surface wave magnitudes lower than Ms=6.8.
 a reduced amount of material available to be transported during the marine
period. Thus, only the local and major earthquakes were recorded (Ms • 7).
The recurrence interval in the non-marine unit could have been overestimated, considering as
earthquake induced deposits sediments related to other mechanisms, such as subaqueous
landslides or turbidites produced by hyperpycnal flooding. However, the occurrence of
homogeneous units (Hm) with very distinctive characteristics, associated with coarser
turbiditic units, support our RI estimation and the overestimation is probably minimal.

With respect to intensities or magnitude of the paleo-earthquakes, previous works


(Audemard and De Santis, 1991; Rodríguez-Pascua et al., 2000; McCalpin, 2009 and
references therein; Rodríguez-Pascua et al., 2010) indicate that a Mw > 5 is necessary to
produce noticeable geologic effects (liquefaction, failures, injections, etc.). Additionally, the
compilation of historical data in the Sea of Marmara (Ambraseys and Finkel 1991; Ambraseys
and Jackson 2000; Ambraseys 2002b) indicates that during the last 2 kyr an important number
of major earthquakes (Ms• 6.8) have occurred (see subchapter 3.2), some of them being
generators of submarine land slide and tsunamis. Precise estimation of paleo-magnitudes
based on the data acquired in this work remains speculative. However, the presence of
seismo-induced sedimentary structures and the historical data suggest that a Mw > 5 may be
proposed as a conservative estimation.

126
3.8. Correlation between the Çinarcik and Central basins through the MD01-2425,
MD01-2429 and MD01-2431 cores

In the present work we attempt to correlate the MD01-2425, MD01-2429 and MD01-
2431 cores. High resolution seismic reflection data obtained with the 3.5 kHz device is
available for the Çinarcik and Central basins (see subchapter 3.3.1, figures III.8, III.9). In the
Central Basin, a remarkable acoustically transparent layer up to 8 m thick is identified in
different seismic profiles crossing this basin (figure III.9e) (Beck et al., 2007). This
distinctive layer, dated 16kyr PB, could be an element of correlation between the basins, but it
is not clearly identified in the Tekirda÷, Kumburgaz, Çinarcik basins. Then, the correlation
between these basin using the seismic reflection data is doubtful. Therefore, we use the
magnetic susceptibilities profiles and X-ray images, which are the common data available for
the three cores.

Figure III.37 shows the correlation proposed between the MD01-2425, MD01-2429
and MD01-2431 cores, based on the magnetic susceptibility profiles which reflects the
evolution of the concentration of magnetic particles during at least 17 cal kyr BP. The two
profiles display an evolution similar to that observed in the MD01-2425 core (see subchapter
3.5.2 and figure III.16). They exhibit a clearly observed change in the behavior of MS profile
between the marine and non-marine sections. The middle and lower part of the marine
sections (subunit U2 and U3) located between 3.5 and 16.9 m depth in the MD01-2425 core, 5
to 30.5 m depth in the MD01-2429, and 1.4 to 9.5 m depth in the MD01-2431 show low
values of MS (mean MS of the MD01-2425: 10.10 r 1.8 x 10-5 S.I.; MD01-2429: 21.68 r 0.9
x 10-5 S.I.; and MD01-2431: 15.79 r 11.4 x 10-5 S.I.). The non-marine section presents highly
variable MS between 300 to 20 x 10-5 S.I., with mean values of 26.88 r 12.2 x 10-5 S.I. for the
MD01-2425 core; 24.18 r 2.3 x 10-5 S.I. in the MD01-2429; and 66.90 r 50.9 x 10-5 S.I. in
the MD01-2431.

The detailed study of the X-ray images allowed us to identify a distinctive layer of 5
cm thick in the MD01-2429 and MD01-2431 cores, and 2 cm thick in the MD01-2425 core,
marking the boundary between the marine and non marine units. This layer is constituted by
highly bioturbated muds, which exhibit similar features to those described in the facies Mh. It
marks the first appearance of diatoms in the sedimentary column, and coincides with the main

127
Figure III.37: High-resolution magnetic susceptibility (MS) data of the MD01-2425, MD01-
2429 and MD01-2431 cores. The MS of the MD01-2425 core were collected using a
BartingtonTM MS2 contact sensor each 5 mm. The MS profiles of the MD01-2429 and MD01-
2431 cores were taken from Beck et al. (2007). The radiocarbons dating were calibrated

local variations of the reservoir effect (ǻR=81r56) proposed by Siani et al. (2000). The age
using the software Bcal (Buck et al., 1999), the Marine09 calibration curve and applying the

in black color were discussed in the chapter 3.5. The ages in purple color were taken from
Beck et al.(2007). X-ray pictures of the highly bioturbated bed of mud called reference bed
level.
128
change of the MS profile. This layer has been considered as a reference bed level (Rbl)
(figure III.37) and is interpreted as the end of the connection between the Black and the Sea
of Marmaras (see subchapter 3.4.3-c for complementary information). Actually, the Rbl
marks the boundary between marine and non marine units.

These two elements (MS profiles and the Rbl) allows us to make a rough correlation
between the three cores and consequently between the two basins. Unfortunately we do not
find elements that indicate the unequivocally coexistence of the same instantaneous event in
the two basins. Thus a correlation at smaller scale (between instantaneous events like classical
turbidites or homogenites-turbidites) was not possible. Although this does not imply that
synchronous or coseismic events exist in both basins. On the other hand, a correlation event
by event was made between the MD01-2431 and MD01-2429 cores. It is presented in the
subchapter 3.9.

129
3.9. Identification and measurement of the co-seismic fault offset along the North
Anatolian Fault in the Central Basin through the co-seismic sedimentary episodes

The results obtained from Core MD01-2425 brought precisions on major earthquakes’
frequency, but without any attribution to a specific fault or fault section. At the difference,
based on previous results on the Central Basin and different published studies, we attempt, in
this subchapter, to demonstrate that fault offsets observed on the southern part of the Central
Basin are actually co-seismic using adjacent HmTu deposits. This topic is presented in the
following article draft untitled
“Estimation of successive co-seismic vertical offsets using coeval sedimentary events,
in the Sea of Marmara’s Central Basin”.
Accepted (February 2014) for publication in Natural Hazards and Earth Sciences
Systems / NHESS.

130
1 Title:
2 Estimation of successive co-seismic vertical offsets using coeval sedimentary events. Application to
3 the Sea of Marmara’s Central Basin (North Anatolian Fault).
4
5 Authors:
6 Christian Beck1, Corina Campos1,2, K. Kadir Eriş3, Namık Çağatay4, Bernard Mercier de Lepinay5,
7 François Jouanne1.
8
9 Addresses:
10 (1) Laboratoire ISTerre, UMR CNRS 5275, Université de Savoie/Grenoble-Alpes University,
11 73 376 Le Bourget du Lac, France.
12 beck@univ-savoie.fr
13 (2) Departamento de Ciencias de la Tierra, Universidad Simón Bolívar, Sartenejas, Baruta,
14 Venezuela.
15 (3) Firat University Faculty of Engineering Geology Department, 23100, Elazığ, Turkey
16 (4) Istanbul Technical University EMCOL, 34469, Faculty of Mining Ayazağa, İstanbul,
17 Turkey
18 (5) Geoazur, U.M.R. C.N.R.S 6526, Université de Nice-Sophia-Antipolis, 06560 Valbonne,
19 France
20 Abstract
21 In the deep part of the Sea of Marmara (Turkey), the sedimentation developing upon the
22 North Anatolian Fault is strongly influenced by the associated seismic activity. Specific layers
23 (homogenites-turbidites), representing individual sedimentary events, have been characterized along
24 three giant piston cores retrieved from Çinarcik and Central (or Orta) basins. Analyzed sediments
25 represent the last 12 to 17 ky BP. For a 2 kyr-lasting interval, 11 events could be precisely correlated
26 on both sides of the Central Basin’s southern scarp. For each of them, based on the specific
27 depositional process, the thickness difference between the two sites was considered as a direct
28 estimation of the vertical component of a coeval co-seismic offset. The homogenite (upper) term
29 accounts for the major part of the thickness difference. The 6 most significant values range from 36
30 cm to 144 cm and are likely representing dominantly normal throws, with estimated
31 paleomagnitudes (Mw) ranging from 5.9 to 6.6.
32
33 Key-words: homogenite-turbidite, co-seismic scarp, Sea of Marmara, Late Pleistocene.
34
35 Introduction
1
36 Since several decades, sedimentary archives, either marine or lacustrine, have been explored
37 as potential paleoseismic records, beside previously well-developed onland approaches (in
38 McCalpin, 2009). For the subaqueous records, two major groups of effects can be detected and
39 analyzed: i) in situ post-depositional disturbances (e.g.: Marco and Agnon, 1995; Ken-Tor et al.,
40 2001; Rodriguez-Pascua et al., 2002, 2003), ii) gravity-driven reworking and re-settling of large
41 masses of unconsolidated sediments (e.g.: Adams, 1990; Strasser et al., 2006).
42 Two major questions arise for both groups: 1) how to ensure the earthquake-triggering, 2)
43 how to identify the responsible active structure(s). For in situ disturbances, the first problem is
44 generally solved; in particular, it benefits from analogical and/or numerical modelling (e.g. Moretti
45 et al., 1999; Wetzler et al., 2010). For redepositional processes - which are envisaged in the present
46 work - several recent catastrophic events could be suveyed shortly after their occurrence (Thunell et
47 al., 1999; McHugh et al., 2011; Lorenzoni et al., 2012); the results reinforced the earthquake-
48 induced interpretation proposed for some “homogenite-type” layers (Chapron et al., 1999; Beck et
49 al., 2007).
50 For historical and older events, the seismic origin of a specific layer can be established:
51 - directly, using intrinsic characteristics (texture, origin of components, overall geometry, etc.,
52 (see references in Beck (2009) and Beck et al. (2007));
53 - indirectly, i) on the basis of correlations with reported seismic events (for historical
54 seismicity) (e.g. Siegenthaler et al., 1987; Piper et al. 1992; Chapron et al. 1999; Beck et al.,
55 2012); ii) when detecting the same paleo-event in a large area independantly from local
56 setting (e.g. variable slope dip). This second approach is especially used for deep structures,
57 as subduction, (e.g.: Gracia et al., 2010, Moernaut, 2011; Pouderoux et al., 2012), but, also
58 in some cases, for surface-reaching major faults (e.g. Goldfinger et al, 2007);
59 - combining both types of arguments.
60 Direct relationships between an active structure and earthquake-induced sedimentary events
61 are investigated for active faults reaching a sediment/water interface (sea- or lake-bottom), through
62 high resolution imagery or/and coring. This favourable setting recently permitted detailed analyses
63 of fault activity (offsets, slip rates) through adjacent sedimentation (Carrillo et al., 2006, 2008; Bull
64 et al., 2006; Barnes and Pondard, 2010; Beck et al., 2012). The here-presented work was dedicated
65 to one of these cases: the deep part of the Sea of Marmara (northwestern Turkey) developed along
66 the North-Anatolian Fault.
67 As the sedimentological tools and approaches we used have been previously published (Sari
68 and Çağatay, 2006; Beck et al., 2007, 2012; Uçarkuş, 2010; Çağatay et al., 2012; Drab et al., 2012;
69 Eriş et al., 2012; Campos et al., 2013) only results (data and interpretations) implying
70 paleoseismological aspects will be envisaged here. Detailed sedimentological aspects may be
2
71 consulted in the above-mentioned publications.
72

73
74 Figure 1.- The Sea of Marmara and the North Anatolian Fault: simplified bathymetry and active
75 structures. Location of analyzed core. NAF geometry simplified from Armijo et al. (2002, 2005); GPS
76 kinematics from McClusky et al. (2000), Reilinger atal. (2006).
77
78 I-Tectonic context and data acquisition
79 Being the gateway between the Black Sea and the Aegean Sea, with narrow shallow
80 connections (Figure 1), the Sea of Marmara has become the focus of paleoenvironmental
81 investigations. In particular, Late Quaternary climatic cycles, and especially associated sea level
82 changes, let a strong sedimentary imprint, in shallow parts as well as in deep basins (Çağatay et al.,
83 2000; Major et al., 2006; Vidal et al., 2010; etc.). Different hypothesis (including catastrophic
84 flooding) have been proposed for the last “re-connection” of the three realms through Bosphorus
85 and Dardanelles (Çanakkale Straits) sills (Ryan et al., 1997, 1999; Aksu et al., 1999, Eriş et al.,
86 2007; etc.). For our purpose, the age of the last non marine-to-marine shift of the Sea of Marmara is
87 a key point, both for the chronological frame of recent seismo-tectonic activity and for the change
88 of volume, composition, and behaviour of re-mobilized sediments (impact of water density and
89 circulation).
90
91 I-1- Structural setting and recent seismic activity.
92 The whole circum-Mediterranean areas represent complex and active plate boundaries where
93 subduction and faulting are responsible for high seismic hazards (in Ambraseys, 2009). Among
94 microplates located between the two major Eurasian and African plates, the Anatolian plate (Figure
95 1 insert; McClusky et al., 2000; Flerit et al., 2003; Reilinger et al., 2006) is highly investigated as its
96 boundaries have produced catastrophic earthquakes and represent a high permanent seismic risk.

3
97 More precisely, the northern limit of the Anatolian Plate corresponds to the - right lateral strike slip
98 - North Anatolian Fault (N.A.F. in the following), which northern branch follows the Sea of
99 Marmara from the Izmit Gulf (East) to the Aegean Sea (West) (Barka and Kadinsky-Cade, 1988;
100 Armijo et al. 2002; Polonia et al., 2004; McNeill et al., 2004; Gasperini et al., 2011; etc.). Beside
101 the dominant strike slip displacement, the importance of normal faulting and fast subsidence has
102 been underlined, especially for the Central and Çinarcik Basins (Cormier et al., 2006; Carton et al.,
103 2007).
104 The migration of historical catastrophic ruptures along the N.A.F. has been intensively
105 investigated aiming to understand past and present stress distribution, and to improve seismic risk
106 assessment (Toksöz et al., 1979; Ambraseys and Jackson, 2000; Ambraseys, 2002; Hubert-Ferrari et
107 al., 2002; Altunel et al., 2004; Aksoy et al., 2010; Fraser et al., 2011; Uçarkuş et al., 2011;
108 Meghraoui et al. 2012; etc.). In particular, two destructive ruptured sections have been surveyed
109 (offset and length) respectively West and East of the Sea of Marmara: 1) the Mw7.4 1912 Ganos
110 event, 2) the Mw7.4 1999 Izmit event. As the deep basins of the Sea of Marmara are bounded or
111 crosscut by the N.A.F. (Figure 1), several offshore surveys have been dedicated to analyze its
112 submerged section. Morphological and sedimentary impacts of major recent earthquakes have been
113 searched using: seismic reflection with different resolutions and penetrations, multibeam and side
114 scan sonar, different types of coring, and remote operating vehicles (R.O.V.) (Armijo et al., 2005).
115 The different results concern: 1) deep fluids explusion related to seismo-tectonic activity (Géli et
116 al., 2008; Tary et al., 2012; Burnard et al., 2012; etc.), 2) mass wasting and creep (Zitter et al.,
117 2012; Shillington et al., 2012; 3) deep sedimentation specificities (McHugh et al., 2006; Sari and
118 Çağatay, 2006; Beck et al., 2007; Çağatay et al, 2012; Drab et al., 2012); 4) detection and dating of
119 historical co-seismic scarps (Armijo et al., 2005; Uçarkuş, 2010). Historical tsunamis reports and
120 modelling (Altinok et al., 2011, Hébert et al., 2005) complete these different data, taking into
121 account the fact that these phenomena are not systematically associated to earthquakes (Hornbach et
122 al., 2010). Small size lacustrine basins aligned along the N.A.F. East of Izmit have also been studied
123 for paleosesimicity (Avşar, 2013).
124 The Sea of Marmara, and especially its deep basins, represents a favorable setting for the
125 search of past seismic activity, and, by mean, an essential data source for regional sesmic hazards
126 estimation (cf. Armijo et al., 2005). In the following, we will focus on the Çinarcik and Central
127 basins’ recent sedimentary fills, which we studied aiming: 1) to reconstruct a succession of
128 earthquake-induced sedimentery “events”, and, 2) to use part of this succession to analyze the
129 activity of the fault zone corresponding to the southern limit of the “inner” Central Basin (as named
130 by Uçarkuş, 2010).
131
4
132 I-2. Data acquisition and processing.
133 The here-used data were collected during two cruises: 1) the MARMACORE survey (on
134 R/V MARION-DUFRESNE), 2) the MARMARASCARPS survey (on R/V/ ATALANTE). Three
135 types of cores were retrieved: giant gravity piston cores (CALYPSO device), classical Kullenberg-

136
137 Figure 2.- Age/depth curve of Core MD01-2425 (Çinarcik Basin) displaying major instantaneous
138 deposits (homogenite+turbidite). Red dashed line indicates the limit between non marine (below) and marine
139 sequences (reference layer displayed on Figures 3 and 5); pLGH: pre-Late Glacial Homogenite (Beck et al.
140 2007).
5
141
142 type cores, and very short cores (35 cm) visually picked using the VICTOR R.O.V.. In parallel, high
143 resolution (3.5 kHz) seismic profiles were acquired, and a complete survey with the VICTOR
144 R.O.V. was dedicated to a high precision multibeam bathymetric survey of different deep scarps
145 (Armijo et al, 2005, Uçarkuş, 2010).The second and third types of cores have been analyzed and
146 yielded paleoseismic information (Uçarkuş, 2010; Drab et al., 2012). A preliminary analysis of 7
147 Calypso cores (with length ranging from 22 to 37 m) and 3.5 kHz imagery has been achieved (Beck
148 et al. 2007). 3 of them (MD01-2425 in Çinarcik Basin, MD01-2429 and MD01-2431 in Central
149 Basin, location on Figure 1) were chosen for later detailed laboratory sedimentological analyses on
150 split cores (Eriş et al., 2012, and this work):
151 - sediment composition: microscopic observations, Carbon and carbonates contents (LOI)
152 XRF profiles in selected portions (AVAATECH instrument); bulk magnetic content
153 (BARTINGTON contact sensor with 5 mm measurement interval);
154 - layering and texture (grain array):
155 + XRay pictures (SCOPIX device, Migeon et al, 1998)
156 + detailed grain size analysis (MALVERN Mastersizer 2000): base-to-top paths on binary
157 diagrams for turbidites/homogenites layers (in Beck, 2009; Eriş et al., 2012); particle shape
158 analysis for silt-clay fraction (SYSMEX FPIA-2100 device);
159 + Anisotropy of Magnetic Susceptibility profiles (2 cm interval) on selected portions
160 (Campos et al, 2013), completed with Anhysteretic Remanent Magnetization (ARM) and
161 Isothermal Remanent Magnetization (IRM) (AGICO MFK1-FA Kappabridge, SQUID and
162 2G 760R systems).
14
163 The chronology is based on AMS C calibrated ages: previously published measurements
164 performed in Woods Hole Oceanographic Institution (NOSAMS facility) (in Beck et al, 2007), and
165 a set of new mesurements performed at CEA-Saclay (CNRS-INSU ARTEMIS facility).
166
167 II- Recent sedimentation in the Çinarcick and Central basins of the Sea of Marmara
168 Cores MD01-2425, -2429, and -2431 (location on Figure 1) were respectively retrieved at
169 1215 m, 1230 m, and 1170 m depths, with 31.30 m, 37.30 m, and 26.40 m respective lengths. They
170 respectively represent about 17 kyr, 14 kyr, and 18 kyr BP of continuous deposition. The
171 compositions, layering-types, and the general chronostratigraphy, appear very similar between the
172 three cores, thus we will summarize the results obtained for Core MD01-2425 as a reference. They
173 confirm and complete the investigations previously achieved by Eriş et al. (2012).
174
175 II-1- The post-LGM succession in the Çinarcick basin (Core MD01-2425)
6
176 Figure 2 summarizes the succession within which, especially in the lower (non marine) part,
177 numerous turbidites, often associated to an overlying homogenite, are intercalated. For this reason,
178 we will describe separately these instantaneous sedimentary “events” and the continuous (“back
179 ground”) slow sedimentation. A neat subdivision into two successions appears (see also Eriş et al.,
180 2012): 1) a lower part with a mean high sedimentation rate (about 5.4 mm.y-1) due to abundant
181 intercalations of coarser instantaneous terrigenous inputs; 2) an upper part with lower mean
182 sedimentation rate (1.3 mm.y-1) and few coarser intercalations. The limit (discussed hereafter)
183 roughly corresponds to the transition from non marine (only connection with the Black Sea) to
184 marine (connection with Aegean Sea and Black Sea) setting. The whole core corresponds to the
185 Late Glacial-Holocene period.
186
187 II-1-a- “Back ground” sedimentation.
188 It is represented in the whole core by a hemipelagic silty-clayey mud. Although the word
189 “hemipelagite” should be restricted to marine/oceanic deposits, we also use it for the non marine
190 succession as, in both cases, it is a mixture of clayey-silty terrigenous fraction (clay minerals,
191 quartz, plagioclase, amphibole, pyroxene, fresh micas, opaques) and planctonic biogenic or bio-
192 induced particles (carbonate and silica: calcareous nanoplankton, Diatoms). Additionnal authigenic
193 particles are locally abundant (sulphides, calcite, aragonite, Mn oxydes).
194 The bulk carbonate content ranges from 8 to 10 % in the upper marine part; it reaches 16 %
195 at the limit non marine/marine. Organic Matter (weight % dried sediment) ranges from 4 to 6 % in
196 lower part, and 7 to 14 % in the upper part. The highest values characterize the 1380cm-980 cm so-
197 called “sapropelic” interval. This O.M. enrichment has been previously reported in the different
198 basins of the Sea of Marmara, and in the shallower zone between Tekirdağ and Central basins
199 (Çağatay et al., 2000; Reichel and Halbach, 2007; Beck et al., 2007; Vidal et al., 2010). The
14
200 different proposed ages are in agreement and a 11-to-7.5 kyr BP period (cal C without reservoir
201 correction) can be attributed to this particular episode.
202 We include into the “background” sediments numerous silty-sandy laminated intervals
203 present in the upper marine part. They are 1 to 3 cm thick and display millimetric parallel planar
204 bedding, involving subtle changes in grain size (up to very fine sand) and mineralogy of detrital
205 components. They have been observed in the three basins (Tekirdağ, Central, Çinarcik) with same
206 characteristics and occurrence frequency (Beck et al., 2007). We relate these levels to in situ slight
207 reworking by episodic increase of bottom current velocities. A minor part of these intervals (see
208 hereafter) show slow angle micropragradation (flaser bedding type) and are associated to
209 homogenites, thus included into instantaneous gravity reworking events.
210
7
211 II-1-b- Homogenites+turbidites (HmTu) occurrences.
212 Detailed analyses and characterizations of homogenites and their association with turbidites
213 (here labelled HmTu) has been previously published (references in Chapron et al., 1999; Beck et al.,
214 2007; Campos et al., 2013), and their use as paleo-earthquake indicator underlined. They were
215 initially called “seismoturbidites” in Eriş et al. (2012). On split core surface and Xray pictures, a
216 series of such layers have been visually identified (Figure 2) mostly in non marine lower part of the
217 core. With up to 1 m thickness, they consist of: 1) a basal coarse layer with overall normal graded
218 bedding, sometimes subdivided into second order graded episodes (similar to classical turbidite
219 lower term), 2) strongly homogenous fine-grained (2 to 8 μm mean grain size) interval, lacking
220 internal variation, and displaying an anomalously high magnetic foliation. As the content and
221 particle shapes of the homogene upper term are identical to what is observed in the hemipelagic
222 mud, the AMS contrast is attributed to a particular grain array (Campos et al., 2013) and, by mean,
223 to a specific settling process. The sharp break between the two terms is often preceeded by a thin
224 interval with flaser bedding-type layering indicating to-and-fro (oscillatory) current, and/or by a
225 specific grain-size evolution. We interpreted this transitional interval as a consequence of oscillation
226 of the whole water mass (seiche effect, reflected tsunami), thus an effect of earthquake and/or
227 massive subaqueous landslide.
228 In the upper marine part of the succession (Figure 2), these HmTu “events” are scarce and
229 thinner; they only display a discrete coarser layer and homogene mud with same texture as in the
230 non marine HmTu events. Some classical turbidites were also found in Core MD01-2425. We
231 discarded them in the following, as we could not ensure their earthquake triggering using our
232 criteria; this choice probably minimizes the total number of inferred recorded paleo-earthquakes.
233 The contrast between the two parts of the succession (roughly between Late Glacial and Holocene)
234 is a matter of debate (in Beck et al., 2007). To explain the abundance of terrigenous arrivals in the
235 Late Glacial, we may envisage: either higher storage of sediments in subaqueous deltas and
236 subsequent higher potential for gravity reworking (climatic influence), or more frequent and
237 powerfull earthquakes (tectonic influence)?. A change in water density vertical profile and in
238 circulation may also account for the distribution of bedload and suspended load. A similar problem
239 has been underlined for the post-LGM fill of large peri-alpine lakes (Beck et al., 1996).
240
241 II-1-c- The non marine to marine transition. Age and implications.
242 Due to its importance for the study of the last climatic cycle, the hydrologic evolution of the
243 Sea of Marmara has been intensively surveyed through sedimentation. Biological, chemical,
244 mineralogical, and isotopic proxies, have been analyzed to detect the respective influence of the
245 Black Sea and the Aegean Sea since the Last Glacial Maximum (MIS 2) depending on their surface
8
246 level (Çağatay et al., 2000; Major et al., 2006; Reichel and Halbach, 2007; Eriş et al., 2007; Vidal et
247 al., 2010). The impact of these variations has also been investigated in shallow parts (Çağatay et al.,
248 2003).
249 Based on detailed data from cores taken in Central Basin and between Tekirdağ and Central
250 basins, Reichel and Halbach (2007) proposed a modelling of fresh water and marine water mixing.
251 Their results fit with their detection of a calcite peak (30 %) related to authigenic precipitation and
252 interpreted as the result of a first mixing of bottom anoxic fresh water with surface oxic marine
253 water. According to their radiocarbon dating, this changed occurred at 13 cal kyr BP. Vidal et al.
254 (2010) concluded to a slightly different scenario: beginning of Aegean influence at 14.7 kyr cal BP
255 and progressive increase of the mixing, lasting 2 kyr, and followed at 12.8 kyr cal BP by an
256 increased of terrigenous continental organic material. They relate the latter to the beginning of
257 Younger Dryas. The calcite maximum may correspond to the end of mixing process. The end of
258 significant Black Sea input at 11.5 kyr cal BP (Vidal et al. 2010) may correspond to the end of the
259 calcite-rich episode. Differences may be due to data sets respectively coming from a deep basin
260 floor (Vidal et al., 2010) and from a shallower setting between two basins (Reichel and Halbach,
261 2007).

262
263 Figure 3.- Chronostratigraphic correlations between the Çinarcik Basin (Core MD01-2425) and the
264 Central Basin (Cores MD01-2429 an -2431). Blue rectangles correspond to close up and detailed
265 correlation on Figure 5.
9
266
267 In the here-studied cores, Xray pictures (Figure 3 insert) permit to identify a 2 to 5 cm-thick,
268 conspicuous layer of highly bioturbated mud, overlain by about 1 cm of laminated silty mud
269 (parallel planar bedding). It is marked by an increase of fine-grained plant debris, and an abrupt
270 change in magnetic content with respect to hemipelagic intervals. It also corresponds to the last
271 occurrence of Diatoms frustules fragments. A XRF chemical profile (core scanning) performed
272 across this layer - from 40 cm below to 100 cm above - did not display any abrupt change but rather
273 the beginning of a very progressive increase of Br, Mo, and S, considered as diagnostic for more
274 marine environment. According to our measurements, Carbonate content reaches a maximum just
275 below this level (named “reference layer” in the following) and sharply decreases above.
276 Combining all published results (and our 14C results) we consider our “reference layer”as the
277 Y.D. base and we will use, for the paleo-seismic record discussed hereafter, a 12.8 cal kyr BP age.
278
279 II-2- Correlations between Çinarcik and Central Basins and inside Central Basin.
280 In order to extract a paleoseismic record through HmTu events, we checked: 1) a regional
281 correlation between Central and Çinarcik basins (MD01-2425, -2429, and -2431; see also Beck et
282 al., 2007; Eriş et al., 2012); 2) a more localized correlation on both sides of an active scarp in
283 Central Basin (Figure 4). The catastrophic pre-Late Glacial event (pLGH on Figure 4) detected on
284 high resolution profiles and cored at Site MD01-2431 (Beck et al., 2007) was not reached at Site
285 MD01-2425.
286 Figure 3 displays the general correlation on the basis of high resolution bulk Magnetic
287 Susceptibility (MS) profiles. All other measured parameters (not added here: mineralogy and
14
288 chemistry, biogenic and bio-induced markers), C ages, and the non marine-to-marine change
289 horizon, complete the correlation critieria. For the same Late Glacial part of the succession (from
290 about 16 kyr BP to the marine/no marine limit (our “reference layer”, Figure 2 and 3)),
291 homogenites+turbidites (HmTu) appear more frequent in Core MD01-2425 (Çinarcik Basin) than in
292 Core MD01-2431 (Central Basin). This difference is also mentionned by Drab et al. (2012) for the
293 last 2.5 kyr BP (marine section), based on short piston core; these authors also proposed several
294 event-by-event correlations and attributions to historical earthquakes.
295 Despite the Xray pictures resolution and sampling intervals of the different logs, we could
296 not ensure an “event-by-event” correlation between Çinarcik and Central Basins along the whole
297 non marine section. Conversely, an “event-by-event” correlation within the Central Basin could be
298 proposed for the last 2 kyr preceeding the main hydrologic change (blue rectangles on Figure 3).
299 Figure 4 (A and B) shows the location of Cores MD01-2429 and MD01-2431 and the overall
300 correlations previously proposed by Beck et al. (2007) and Eriş et al. (2012) based on 3.5 kHz
10
301 seismic reflection profile.
302
303 III- Inferred co-seismic sedimentary events on both sides of the southern scarp of the Central
304 Basin
305 The two analyzed sites are on both sides of the southern limit of the “inner” Central Basin
306 (Figure 4-A) with a small relative depth difference (50 m at 1200 mbsl). Figure 4-B displays their
307 position with respect to a major active fault scarp. Previous preliminar observations of these cores
308 had demonstrated that the high difference of mean sedimentation rates between the two sites (Figure
309 4-B) was essentially due to the difference of instantaneous sedimentary events thicknesses (Beck et
310 al., 2007).

11
311
312 Figure 4.- Detailed location of Orta/Central Basin’s long cores. A) Bathymetry taken from Rangin et
313 al. (2001); B) very high resolution seismic profile from MARMARASCARPS survey (Armijo et al. 2005;
314 Uçarkuş, G., 2010). Red dashed line indicates the limit between non marine (below) and marine sequences
315 (also underlined on Figures 3, 4 and 5).
316

12
317 Assuming the synchroneity of the “reference layer” in the two cores, a detailed analysis was
318 later performed downcore, starting immediately below. For the fine-grained “back ground”
319 sedimentation, because of its hemipelagic-type deposition, the same sedimentation rate was
320 assumed at the two sites. The proposed layer-by-layer correlation (Figure 5) is based on: i) precise
321 delimit

13
322 14
323 Figure 5.- X-ray close up of two synchronous portions of Cores MD01-2429 and MD01-2431,
324 displaying individually correlated sedimentary events (homogenite+turbidite).
325
326
327 ation of hemipelagic intervals, with same thicknesses, ii) similarities of subdivisions within
328 HmTu composite layers. The corrrelation could be achieved for a 2 m succession in Core
329 MD01-2431 which appears equivalent to a 6.2 m succession in Core MD01-2429, the whole for a 2
330 kyr duration. 11 homogenite+turbidites events (HmTu) account for the difference. For the thickest
331 ones (HmTu A, C, E, H, K) the homogene upper term accounts for about 90 % of the thickness
332 increase in the deeper site. Assuming an earthquake origin for these sedimentary events, and the
333 tendancy of the associated suspension to settle in deepest areas (in Chapron et al., 1999), we
334 consider the increased fills of the downgoing side as successive “seals off” of the created co-seismic
335 scarps (Figure 6 insert, case 2a). In the Lesser Antilles, Beck et al. (2012) described an active
336 normal fault upon which the sea floor is maintained flat and horizontal, being each co-seismic offset
337 quite exactly compensated by a coeval silty-sandy homogenite (Figure 6 insert, case 2b). We
338 tentatively applied their 2b model to the Central Basin events. 10 of the 11 events were plot on an
339 age vs. thickness difference log (Figure 6). 6 of them led to significant values between 40 cm to 160
340 cm (Figure 6).
341 Although the investigated sediments are recent with a reduced depth-in-core, a possible
342 compaction effect has to be discussed as: i) it concerns mainly clayey-silty material, ii) the
343 thickness of the homogenite term is up to ten times higher on the hanging wall with respect to the
344 footwall. Based on this differential compaction, a 10 % maximum estimate is thus proposed for a
345 correction of the thickness difference (leading to about 44 to 178 cm).
346 The inferred offsets were separated by variable time intervals (100 to 550 yr); if taking into
347 account the 11 events, a mean 180 yr interval is deduced. The time distribution is in agreement with
348 previously published paleoseismic results based on sedimentary record in the same area (Beck et al.,
349 2007; Drab et al., 2012). In the present study, a precise rupturing site could be attributed to the
350 sedimentary events.
351
352 IV- Discussion
353 The proposed used of homogenites+turbidites (HmTu) to analyze subaqueous active faulting
354 along the inner Central Basin led to estimate a set of inferred co-seismic offsets, for a 2 kyr-lasting
355 period. As it is a 2D approach, the results only concern a vertical component (cf. Figure 4 and
356 Figure. 6). The latter may represent either the vertical component of an oblique slip displacement or
357 a subvertical (normal here) one. West of the Central Basin, a historical scarp was observed and
15
358 analyzed (Armijo et al., 2005; Uçarcuş, 2010) displaying locally low angle dipping slickensides.
359 This site belongs to a NAF section with dominant strike slip behaviour. Applying such low angles
360 displacements to account for our vertical offset values, especially the highest, would result in
361 anomalously high lateral offsets (e.g. 15° dip and 144 cm vertical component). Otherwise, the here-

362
363 Figure 6.- Successive inferred individual co-seismic offsets deduced from
364 homogenites+turbidites(HmTu) thickness differences. (insert sketch modified from Beck et al.,
365 2012)
366
367
368 analyzed site is considered as a limit of a pull apart basin (Armijo et al., 2005; Uçarcuş, 2010), and
369 different investigations highlighted the importance of vertical component in Izmit Gulf and
370 Tekirdağ Basin (Cormier et al., 2006; Carton et al., 2007). Based on tsunami modelling applied to
371 the Sea of Marmara, Hébert et al. (2005) conclude to the importance of vertical offset related to
372 faulting or to submarine landslides. In the following, we thus assumed that our estimated values
373 represent dominant vertical throws, i.e. normal offsets.
374 With respect to an approach in terms of paleo-magnitude (MW) of the earthquakes associated
375 to estimated offsets, additionnal data are needed to propose an actual, complete, paloseismic
376 approach: horizontal length and lower limit of rupturing. Nevertheless, we propose estimations for
16
377 two inferred offset values (44 cm and 178 cm; Fig. 6; with compaction effect). We consider:
378 - a 70° mean fault dip as displayed by deep seismic reflection data from Laigle et al. (2008)
379 - two possibilities for the sea bottom rupture horizontal length: 8 km if considering the total
380 length of the SW side of the Inner Central Basin “losange”, or 5 km if considering only the
381 eastern continuous scarp (see on Fig. 4 A)
382 - a brittle/ductile transition at two different depths following the distribution proposed by Inan
383 et al. (2007, in Uçarkuş, 2010): 12 km and 20 km. These values respectively correspond to
384 the western termination and the eastern half of the analyzed scarp.
385 We applied the Moment Magitude Calculator software (Jet Propulsion Laboratory and University of
386 Southern Carolina, 2013) with two current shear modulus values. The results (Table I) show MW
387 comprised between 5.9 and 6.6. The corresponding MO (seismic moment) values, the fault length
388 and the fault surface values, were plotted on two diagrams, respectively from Kanamori and
389 Anderson (1975) and Henry and Das (2001). Our results better fit with intraplate earthquakes
390 distribution and using a 8 km rupture length (thus the whole SW limit of the Inner Central Basin).
Horizontal rupture length
brittle/ductile limit 5 km 8 km
3.0 1011 dyne.cm-2 47 cm 12 km 5.9 Mw 6.1 Mw
shear modulus offset 20 km 6.1 Mw 6.2 Mw
2.5 1011 dyne.cm-2 47 cm 12 km 5.9 Mw 6.0 Mw
shear modulus offset 20 km 6.0 Mw 6.1 Mw
3.0 1011 dyne.cm-2 190 cm 12 km 6.3 Mw 6.5 Mw
shear modulus offset 20 km 6.5 Mw 6.6 Mw
2.5 1011 dyne.cm-2 190 cm 12 km 6.3 Mw 6.4 Mw
391 shear modulus offset 20 km 6.4 Mw 6.6 Mw

392 Table I. - Estimation of moment magnitudes MW for the southwestern scarp of the Central Basin.
393
394 In terms of paleoseismicity, our results only concern the pre-Holocene period. Regarding the
395 thickness difference between the two sites (three times higher on the hanging wall), there is no
396 drastic change at the non marine/marine limit. For the marine (upper) part, a specific sedimentary
397 processe is still driving an overthickening on the hanging wall (“inner” Central Basin) but not
398 enough to compensate the scarp as before (insert Figure 6, Holocene situation). The few thin
399 homogenites+turbidites we observed cannot account for the difference. Nevertheless, the top of the
400 marine part (partly disturbed in our giant piston cores) corresponds to a conspicuous increase in
401 M.S. (Figure 3). A 4m long piston core, taken close to Site MD01-2429, shows a set of
402 turbidites+homogenites (Drab et al., 2012); the corresponding interval is approximately comprised
403 between 2 kyr BP and Present.
404 To explain the remaining thickness difference during the 13 kyr BP-to Present period, we

17
405 favour a hypothesis implying two combined mechanisms: 1) the water vertical density profile led to
406 more hyperpycnal distribution of gravity reworked sediments, 2) coarse material strongly decreased
407 due to change in weathering condition. Checking this hypothesis needs further higher resolution
408 analysis of the sediments (especially the laminated episodes). The present day depth difference
409 between Sites MD01-2429 and MD01-2431 (Figure 5 A and B) - about 50 m - corresponds to about
410 12.5 kyr. The post-“reference layer” interval (the marine part) shows a 20 m “additionnal”
411 thickness, leading to about 70 m of total vertical displacement (approx. 75 m slip with a mean 70°
412 fault plane dip) . This could correspond to a 6 mm.y-1 mean normal offset, which distribution into
413 creep vs. co-seismic increments has to be further discussed. Considering a relatively low number of
414 sedimentary events in the Central Basin with respect to Tekirdağ Basin, Drab et al. (2012)
415 underlined different explanation, including partial creeping along the Central segment. For a longer
416 period, we observe a similar difference between Central Basin and Çinarcik Basin, with evidences
417 of a specific behaviour of the southern limit of the former.
418
419 Conclusions
420 The detailed sedimentological analysis of a sedimentary accumulation bounding a
421 subaqueous active fault confirmed the occurrence of co-seismic offsets through coeval specific
422 events. It permitted to estimate their values and also confirms a dominantly co-seismic behaviour
423 (null or negligible creep) at least for a 2 ky time interval. With up to 1.8 m normal slip values,
424 added to local structural and seismological data, this archive led to propose paleo-magnitude values
425 (MW between 5.9 and 6.6) compatible with historical data. These results bring additionnal arguments
426 for seismic hazard assessment along the central portion of the N.A.F. (Sea of Marmara’s Central
427 Basin).
428
429 Acknowledgements:
430 The presented investigations were possible thanks to CNRS-INSU funding through ISTerre
431 Laboratory and the Universe Sciences Observatory of Grenoble. CNRS-INSU is acknowledged for
432 the access to ARTEMIS national AMS radiocarbon measurement facilities. C. Campos’ PhD thesis
433 and stay in ISTerre Laboratory were funded through Venezuela’s FUNDAYACUCHO Grant N°
434 20093262. We thank Anne-Lise Develle (EDYTEM Laboratory) for XRF profiles performing and
435 help for their interpretation.
436
437 References:
438 Adams, J., 1990. Paleoseismicity of the Cascadian subduction zone: evidence from turbidites
439 off the Oregon–Washington margin. Tectonics, vol. 9, 4:569–583.
18
440 Aksoy, M. E., Meghraoui, M., Vallee, M., Cakir, Z. 2010. Rupture characteristics of the A.D.
441 1912 Murefte (Ganos) earthquake segment of the North Anatolian fault (western Turkey).
442 Geology 38:991-994. doi:10.1130/G13447.1
443 Aksu, A. E., Hiscott, R. N. and Yasar, D., 1999. Oscillating Quaternary water levels of the Marmara
444 Sea and vigorous outflow into the Aegean Sea from the Marmara Sea-Black Sea drainage
445 corridor, Marine Geology, 153, 275-302.
446 Altinok, Y., Alpar, B., Ozer, N., Aykurt, H. 2011. Revision of the tsunami catalogue affecting
447 Turkish coasts and surrounding regions. Natural Hazards Earth System Science 11:273-291.
448 Altunel, E., Meghraoui, M., Akyüz, H. S., Dikbas, A., 2004. Characteristics of the 1912 coseismic
449 rupture along the North Anatolian Fault Zone (Turkey): implications for the expected
450 Marmara earthquake, Terra Nova 16:198-204.
451 Ambraseys, N.N., and Jackson, J.A., 2000. Seismicity of the Sea of Marmara (Turkey) since 1500,
452 Geophysical Journal International 141:F1-F6.
453 Ambraseys, N. N., 2002. The seismic activity in the Marmara Sea region over the last 704 years,
454 Bulletin of the Seismological Society of America 92:1-18.
455 Ambraseys, N. N., 2009. Earthquakes in the Mediterranean and Middle East: A Multidisciplinary
456 Study of Seismicity up to 1900, 947 pp. Cambridge University Press, Cambridge, U. K.
457 Armijo R., Meyer B., Navarro S., King G. C. P., Barka A. A., 2002. Asymmetric slip partitioning in
458 the Sea of Marmara pull-apart: a clue to propagation processes of theNorth Anatolian Fault?
459 Terra Nova 14:80-86.
460 Armijo, R., Pondard, N., Meyer, B., Uçarkus, G., Mercier de Lépinay, B., Malavieille, J.,
461 Dominguez, S., Gutscher, M.-A., Schmidt, S., Beck, C., Çagatay, N., Çakir, Z., Imren, C.,
462 Eriş, K., Natalin, B., Özalaybey, S., Tolun, L., Lefèvre, I., Seeber, L., Gasperini, L., Rangin,
463 C., Emre, O., Sarikavak, K., 2005. Submarine fault scarps in the Sea of Marmara pull-apart
464 (North Anatolian Fault): Implications for seismic hazard in Istanbul. Geochemistry,
465 Geophysics, Geosystems, vol. 6, N° 6, Q06009, doi:10.1029/2004GC000896.
466 Avşar, U., 2013. Lacustrine paleoseismic records from the North Anatolian Fault, Turkey. PhD
467 Thesis memoir, University of Ghent, 209 pp..
468 Barka, A.A., and Kadinsky-Cade, K., 1988. Strike-slip fault geometry in Turkey and its influence
469 on earthquake activity. Tectonics 7:663-684.
470 Barnes, P.M., and Pondard, N. (2010). Derivation of direct on-fault submarine paleoearthquake
471 records from high-resolution seismic reflection profiles: Wairau Fault, New Zealand.
472 Geochemistry, Geophysics, Geosystems 11, Q11013, doi:10.1029/2010GC003254.
473 Beck, C., Manalt, F., Chapron, E., Van Rensbergen, P., De Batist, M., 1996. Enhanced seismicity in
474 early post-glacial period: evidences from the posy-Würm sediments of Lake Annecy,
19
475 northwestern Alps. Journal of Geodynamics, 22(1/2):155-171.
476 Beck, C., 2009. Late Quaternary lacustrine paleo-seismic archives in north-western Alps: Examples
477 of earthquake-origin assessment of sedimentary disturbances. Earth-Science Reviews,
478 96:327–344.
479 Beck, C, Mercier de Lépinay, B., Schneider, J.-L., Cremer, M., Çağatay, N., Wendenbaum,
480 E., Boutareaud, S., Ménot, G., Schmidt, S., Weber, O., Eris, K., Armijo, R., Meyer, B.,
481 Pondard, N., Gutscher, M.-A., and the MARMACORE Cruise Party, J.-L. Turon, L.
482 Labeyrie, E. Cortijo, Y. Gallet, H. Bouquerel, N. Gorur, A. Gervais, M.-H. Castera, L.
483 Londeix, A. de Rességuier, A. Jaouen., 2007. Late Quaternary co-seismic sedimentation in
484 the Sea of Marmara's deep basins. in “Sedimentary Records of Catastrophic Events” (F.
485 Bourrouilh-Le Jan, C. Beck, D. Gorsline, Eds.) Spec. Iss Sedimentary Geology, 199:65–89.
486 Beck, C., Reyss, J.L., Leclerc, F., Moreno, E., Feuillet, N., Barrier, L., Beauducel, F.,
487 Boudon, G., Clément, V., Deplus, C., Gallou, N., Lebrun, J.F., Le Friant, A., Nercessian, A.,
488 Paterne, M., Saurel, J.M., Pichot, T., Vidal, C., 2012. Identification of deep subaqueous co-
489 seismic scarps through specific coeval sedimentation in Lesser Antilles: implication for
490 seismic hazard. In “Subaqueous Paleoseismology” (D. Pantosti Ed.), Natural Hazards and
491 Earth System Sciences, doi:10.5194/nhess-12-1-2012.
492 Burnard, P., Bourlange, S., Henry, P., Géli, L., Tryon, M.D., Natalin, B., Sengör, A.M.C., Özeren,
493 M.S., Çağatay, M.N., 2012. Constraints on fluid origins and migratin velocities along the
494 Marmara Main Fault (Sea of Marmara, Turkey) using helium isotopes. Earth and Planetary
495 Sciences Letters, 341/344:68-78..
496 Bull, J.M., Barnes, P.M., Lamarche, G., Sanderson, D.J., Cowie, P., Taylor, S.K., Dix, J.K..,
497 2006. High-resolution record of displacement accumulation on an active normal fault:
498 implications for models of slip accumulation during repeated earthquakes. Journal of
499 Structural Geology, 28:1146-1166.
500 Çağatay, M. N., Görür, N., Algan, O., Eastoe, C., Tchapalyga, A., Ongan, D., Kuhn,T. and Kuşçu,
501 I., 2000. Late Glacial-Holocene palaeoceanography of the Sea of Marmara: timing of
502 connections with the Mediterranean and the Black Seas. Marine Geology, 167:191-206.
503 Çağatay, M. N., Görür, N., Polonia, A., Demirbağ, E., Sakinç, M., Cormier, M.-H., Capotondi, L.,
504 McHugh, C., Emre, Ö., Eriş, K., 2003. Sea level changes and depositional environments in
505 the Ïzmit Gulf, eastern Marmara Sea, during the Late Glacial-Holocene period. Marine
506 Geology, 202:159-173.
507 Çağatay, N., Erel, L., Bellucci, L.G., Polonia, A., Gasperini, L., Eriş, K.K., Sancar, Ü., Biltekin, D.,
508 Uçarkuş, G., Ulgen, Ü.B., Damci, E., 2012. Sedimentary earthquake records in the İzmit
509 Gulf, Sea of Marmara, Turkey. Sedimentary Geology, 282:347-359.
20
510 Campos, C., Beck, C. Crouzet, C., Demory, F., Van Welden, A., and Eris, K., 2013. Deciphering
511 hemipelagites from homogenites through Magnetic Susceptibility Anisotropy. Paleoseismic
512 implications (Sea of Marmara and Gulf of Corinth). Sedimentary Geology, 292:1-14.
513 Carrillo, E., Audemard, F., Beck, C., Cousin, M., Jouanne, F., Cano, V., Castilla, R., Melo,
514 L., and Villemin, T., 2006. A Late Pleistocene natural seismograph along the Boconò Fault
515 (Mérida Andes, Venezuela): the moraine-dammed Los Zerpa palæo-lake. Bulletin of the
516 French Geological Society, 177(1):3-17.
517 Carrillo, E., Beck, C., Audemard, F.A., Moreno, E., Ollarves, R., 2008. Disentangling Late
518 Quaternary climatic and seismo-tectonic controls on Lake Mucubají sedimentation (Mérida
519 Andes, Venezuela) in M De Batist and E. Chapron Eds. “Lake systems: sedimentary
520 archives of climate change and tectonic”, Palaeogeography, Palaeoclimatology,
521 Palaeoecology, 259:284–300.
522 Carton, H., Sing, S.C., Hirn, A., Bazin, S., de Voogd, B., Vigner, A., Ricolleau, A., Cetin, S.,
523 Oçakoğlu, N., Karakoç, F., and Sevilgen, V., 2007. Seismic imaging of the three-
524 dimensional architecture of the Çinarcik Basin along the North Anatolian Fault. Journal of
525 Geophysical Research, vol. 112, B060101, doi:10/1029/2006JB00548, p. 1-17.
526 Chapron, E., Beck, C., Pourchet, M., and Deconinck, J.-F., 1999. 1822 earthquake-triggered
527 homogenite in Lake Le Bourget (NW Alps). Terra Nova, 11:86-92.
528 Cormier, M.-H., Seeber, L., McHugh, C.M.G., Polonia, A., Çağatay, M.N., Emre, O., Gasperini, L.,
529 Görür, N., Bortoluzzi, G., Bonatti, E., Ryan, W.B.F., Newman, K.R., 2006. The North
530 Anatolian fault in the Gulf of Izmit (Turkey): Rapid vertical motion in response to minor
531 bends of a non-vertical continental transform, Journal of Geophysical Research 111:
532 B04102, doi:1029/2005JB003633.
533 Drab, L., Hubert Ferrari, A., Schmidt, S., and Martinez,P., 2012. The earthquake sedimentary
534 record in the western part of the Sea of Marmara, Turkey. Natural Hazards and Earth
535 System Sciences, Special Issue “Subaqueous Paleoseismology” (D. Pantosti Edt.), doi:
536 10.5194/nhess-12-2012, 1235-1254.
537 Eris¸ K.K., Ryan, W.B.F., Çağatay, N., Sancar, U., Lericolais, G., Menot, G., and Bard, E., 2007.
538 The timing and evolution of the post-glacial transgression across the Sea of Marmara shelf
539 south of Istanbul. Marine Geology, 243(1-4):57–76,.
540 Eriş, K., Çağatay, N. Beck, C., Mercier de Lepinay, B., Campos, C., 2012. Late-Pleistocene to
541 Holocene sedimentary fills of the Çınarcık Basin of the Sea of Marmara. Sedimentary
542 Geology, 281:151-165.
543 Flerit, F., Armijo, R., King, G.C.P., Meyer, B., Barka, E., 2003. Slip partitioning in the Sea
544 of Marmara pull-apart determined from GPS velocity vectors. Geophysical Journal
21
545 International, 154:1–7.
546 Fraser, J., Vanneste, K., and Hubert-Ferrari, A., 2010. Recent behaviour of the North Anatolian
547 Fault: insights from an integrated paleoseismological data set. Journal of Geophysical
548 Research, vol. 115, doi:10/1029/JB006982, p. 1-27.
549 Gasperini, L., Polonia, A., Çağatay, M. N., Bortoluzzi, G., Ferrante, V., 2011. Geological slip rates
550 along the North Anatolian Fault in the Marmara Sea. Tectonics 30:TC6001.
551 Géli, L., Henry, P., Andre, C., Zitter, T., Çağatay, N., Mercier de Lepinay, B., LePichon, X.,
552 Sengor, AMD, Gorur, N., Natalin, B., Ucarkus, G., Ozeren, S., Volker, D., Gasperini, L.,
553 Bourlanger S., and the MarNaut Scientific Party 2008. Gas emissions and active tectonics
554 within the submerged section of the North Anatolia Fault zone in the Sea of Marmara. Earth
555 and Planetary Science Letters 274:34-39.
556 Gracia, E., Vizcaino, A., Escutia, C., Asioi, A., Rodes, A., Pallas, R., Garcia-Orellana, J. Lebreiro,
557 S., Goldfinger, C. 2010. Holocene earthquake record offshore Portugal (SW Iberia): testing
558 turbidite paleoseismology in a slow-convergence margin. Quaternary Science Reviews
559 29:1156-1172.
560 Goldfinger, C., Morey, A.E., Nelson, C.H., Gutierez-Pastor, J, Johnson, J.E., Karabanov, E.,
561 Eriksson, A. and shipboard scientific party, 2007. Rupture lengths and temporal history of
562 significant earthquakes on the offshore and north coast segments of the Northern San
563 Andreas Fault based on turbidite stratigraphy, Earth and Planetary Science Letters, 254:9-
564 27.
565 Hébert, H., Schindelé, F., Altinok, Y., Alpar, B., Gaziogluc, C., 2005. Tsunami hazard in the
566 Marmara Sea (Turkey): a numerical approach to discuss active faulting and impact on the
567 Istanbul coastal areas. Marine Geology, 215:23- 43.
568 Henry, C., and Das, S., 2001. Aftershocks of large shallow earthquakes: faults dimensions,
569 aftershock area expansion and scaling relations. Geophysical Journal International,
570 147:272-293.
571 Hornbach, M. J., Braudy, N., Briggs, R.W., Cormier, M.-H., Davis, M.B., Diebold, J.B.,
572 Dieudonne, N., Douilly, R., Frohlich, C., Gulick, S.P.S., Johnson, H.E., Mann, P., McHugh,
573 C.M.G., Ryan-Mishkin, K., Prentice, C.S., Seeber, L., Sorlien, C.C., Steckler, M.S.,
574 Symithe, S.J., Taylor, F.W., and Templeton, J., 2010. High tsunami frequency as a result of
575 combined strike-slip faulting and coastal landslides, Nature Geoscience 3:783-788.
576 Hubert-Ferrari, A. l., Armijo, R., King, G., Meyer, B. and Barka, A., 2002. Morphology,
577 displacement, and slip rates along the North Anatolian Fault, Turkey. Journal of
578 Geophysical Research, Vol. 107, no. B10, pp. 2235.
579 Jet Propulsion Laboratory and University of Southern Carolina, 2013. Moment Magitude
22
580 Calculator. http://quakesim.org/tools/moment-magnitude-calculator
581 Kanamori, H., and Anderson, D.L., 1975. Theoritical basis of some empirical relations in
582 seismology. Bulletin of the Seismological Society of America, 65:1073-1095.
583 Ken-Tor, R., Agnon, A., Enzel, Y., Stein, M., 2001. High-resolution geological record of historic
584 earthquakes in the Dead Sea basin, Journal of Geophysical Research 106:2221-2234.
585 Laigle, M., Becel, A., de Voogd, B. a., Hirn, A., Taymaz, T. and Ozalaybey, S., 2008. A first deep
586 seismic survey in the Sea of Marmara: Deep basins and whole crust architecture and
587 evolution. Earth and Planetary Science Letters, Vol. 270, 3-4:168-179.
588 Lorenzoni, L., Benitez-Nelson, C. R., Thunell, R. C., Hollander, D., Varelan, R., Astor,
589 Y.,Audemard, F. A., Muller-Karger, F. E., 2012. Potential role of event-driven sediment
590 transport on sediment accumulation in the Cariaco Basin, Venezuela. Marine Geology.
591 307/310:105–110, doi:10.1016/j.margeo.2011.12.009
592 Major, C., Goldstein, S.L., Ryan, W.B.F., Lericolais, G., Piotrowski, A.M., and Hajdas, I., 2006.
593 The co-evolution of Black Sea level and composition through the last deglaciation and its
594 paleoclimatic significance. Quaternary Science Reviews, 25(17-18) :2031 – 2047, 2006.
595 Marco, S., and Agnon, A., 1995. Prehistoric earthquake deformations near Masada, Dead Sea
596 graben. Geology, 23:695–698.
597 McCalpin, J.M., (2009) Paleoseismology. International Geophysics Series, Vol. 95, Academic
598 Press, ISBN: 978-0-12-373576-8, 798 pp.
599 McNeill, L. C., Mille, A., Minshull, T.A., Bull, J.M., Kenyon, N.H., Ivanov, M., 2004. Extension of
600 the North Anatolian Fault into the North Aegean Trough: Evidence for transtension, strain
601 partitioning, and analogues for Sea of Marmara basin models, Tectonics, Vol. 3, 2:1-12.
602 doi:10.1029/2002TC001490.
603 Meghraoui, M., Aksoy, M. E., Akyuz, H. S., Ferry, M., Dikbas, A., Altunnel, E., 2012.
604 Paleoseismology of the North Anatolia Fault at Guzelkoy (Ganos segment, Turkey): Size
605 and recurrence time of earthquake ruptures west of the Sea of Marmara. Geochemistry,
606 Geophysics, Geosystems 13:1-26.
607 McClusky, S., Bassalanian, S., Barka, A., Demir, C., Ergintav, S., Georgiev, I., Gurkan, O.,
608 Hamburger, M., Hurst, K., Hans-Gert, H.-G., Karstens, K., Kekelidze, G., King, R., Kotzev,
609 V., Lenk, O., Mahmoud, S., Mishin, A., Nadariya, M., Ouzounis, A., Paradissis, D., Peter,
610 Y., Prilepin, M., Reilinger, R., Sanli, I., Seeger, H., Tealeb, A., Toksöz, M.N., Veis, G.,
611 2000. Global positioning system constraints on plate kinematics and dynamics in the eastern
612 Mediterranean and Caucasus. Journal of Geophysical Research 105, 5695–5719.
613 McHugh, C.M.G., Seeber, L., Cormier, M.-H., Dutton,J., Cagatay, N., Polonia, A., Ryan,
614 W.B.F., and Gorur, N., 2006, Submarine earthquake geology along the North Anatolia fault
23
615 in the Marmara Sea, Turkey: A model for transform basin sedimentation. Earth and
616 Planetary Sciences, v. 248, p. 661–684, doi:10.1016/j.epsl.2006.05.038.
617 McHugh, C., Seeber, L., Braudy, N., Cormier, M.-H., Davis, M.B., Diebold, J.B., Dieudonne,
618 N., Douilly, R., Gulick, S.P.S., Hornbach, M.J., Johnson, H.E. III, Ryan Miskin, K., Sorlien,
619 C., Steckler, M., Symithe, S.J., Templeton, J., 2011. Offshore sedimentary effects of the 12
620 January 2010 Haiti earthquake. Geology, vol. 39, 8:723-726; doi: 10.1130/G31815.1.
621 Migeon, S., Weber, O., Faugères, J.C., Saint Paul, J., 1998. A new X-ray imaging system for core
622 analysis. Geomarine Letters, 18, 251-255.
623 Moernaut, J., 2011. Sublacustrine landslide processes and their paleoseismological significance:
624 revealing the recurrence rate of giant earthquakes in South-Central Chile. PhD Thesis,
625 University of Ghent, 274 pp..
626 Moretti, M., Alfaro, P., Caselles, O., Canas, J.A., 1999. Modelling seismites with a digital
627 shaking table. Tectonophysics, 304:369:383.
628 Piper, D.J.W., Cochonat, P., Ollier, G., Le Drezen, E., Morrison, M, Baltzer, A., 1992.
629 Evolution progressive d’un glissement rotationnel en un courant de turbidité : cas du séisme
630 de 1929 des Grands Bancs (Terre Neuve). Comptes-Rendus de l’Académie des Sciences,
631 Paris, 314:1057-1064.
632 Polonia, A., Gasperini, L., Amorosi, A., Bonatti, E., Bortoluzzi, G., Çağatay, M. N., Capotondi, L.,
633 Cormier, M.-H., Görür, N., McHugh, C. M. G., Seeber, L., 2004. Holocene slip rate of the
634 North Anatolian Fault beneath the Sea of Marmara, Earth and Planetary Science Letters
635 227:411-426.
636 Pouderoux, H., Lamarche, G., Proust, J.-N. 2012. Building a 18,000 Year-Long Paleo-Earthquake
637 Record from detailed Deep-Sea Turbidite Characterization in Poverty Bay, New Zealand.
638 Natural Hazards Earth System Science 12:2077-2101
639 Rangin, C., E. Demirbag, C. Imren, A. Crusson, A. Normand, E. Le Drezen, and A. Le Bot, Marine
640 Atlas of the Sea of Marmara (Turkey), IFREMER/GENAVIR, 2001.
641 Reichel, T., and Halbach, P., 2007. An authigenic calcite layer in the sediments of the Sea of
642 Marmara. A geochemical marker horizon with paleoceanographic significance. Deep-Sea
643 Research II 54 (11/13): 1201–1215.
644 Reilinger, R. E., McClusky, S., Vernant, P., Lawrence, S., Ergintav, S., Cakmak, R., Ozener, H.,
645 Kadirov, F., Guliev, I., Stepanyan, R., Nadariya, M., Hahubia, G., Mahmoud, S., Sakr, K.,
646 ArRajehi, A., Paradissis, D., Al-Aydrus, A., Prilepin, M., Guseva, T., Evren, E., Dmitrotsa,
647 A., Filikov, S. V., Gomez, F., Al-Ghazzi, R. and Karam, G., 2006. GPS constraints on
648 continental deformation in the Africa-Arabia-Eurasia continental collision zone and

24
649 implications for the dynamics of plate interactions, Journal of Geophysical Resarch, Vol.
650 111, no. B05411, p. 1-26, doi:10.1029/2005JB004051.
651 Rodriguez-Pascua, M.A., Calvo, J.P., De Vicente, G., Gòmez-Gras, D., 2002. Soft-sediment
652 deformation structures interpreted as seismites in lacustrine sediments of the Prebetic Zone,
653 SE Spain, and their potential use as indicators of earthquake magnitudes during the Late
654 Miocene. Sedimentary Geology, 135(1/4):117–135.
655 Rodriguez-Pascua, M.A., De Vicente, G., Calvo, J.P. and Perez-Lopez, R. 2003. Similarities
656 between recent seismic activity and paleoseismites during the late Miocene in the external
657 Betic Chain (Spain): relationship by the b value and the fractal dimension. Journal of
658 Structural Geology, 25: 749.
659 Ryan, W.B.F., Pitman III, W.C., Major, C.O., Shimkus, K., Moskalenko, V., Jones, G.A. Dimitrov,
660 P., Gorür, N., Sakinc, M. and Yüce, H., 1997. An abrupt drowning of the Black Sea shelf,
661 Marine Geology, 138:119-126.
662 Ryan, W.B.F., and Pitman III, W.C., 1999. Noah’s Flood. The New Scientific Discoveries About
663 Events That Chenged History. Toutchstone Book, Simon &Schuster Ed., New York.
664 Sari E., and Çağatay, M. N., 2006. Turbidites and their association with past earthquakes in the
665 deep Cinarcik Basin of the Marmara Sea. Geo-Marine Letters 26:69-76.
666 Shillington, D.J., Seeber, L., Sorlien, C.C., Steckler, M.S., Kurt, H., Dondurur, D., Cifici, G., Imren,
667 C., Cormier, M.-H., McHugh, C.M.G., Gurcay, S., Poyraz, D., Okay, S., Atgin, O., Diebold,
668 B., 2012. Evidence for widespread creep on the flanks of the Sea of Marmara transform
669 basin from marine geophysical data. Geology, 40:439-42, doi:10.1130/G32652.1
670 Siegenthaler, C., Finger, W., Kelts, K., Wang, S., 1987. Earthquake and seiche deposits in
671 Lake Lucerne, Switzerland. Eclogae geologicae Helvetiae, 80:241-260.
672 Strasser, M, Anselmetti, F.S., Fäh, D., Giardini, D., Schnellmann, M., 2006. Magnitudes and
673 source areas of large prehistoric northern Alpine earthquakes revealed by slope failures in
674 lakes. Geology, 34 (12) 1005-1008.
675 Tary, J.B., Géli, L., Guennou, C., Henry, P., Sultan, N., Cagatay, N., and V. Vidal, 2012.
676 Microevents produced by gas migration and expulsion at the seabed: a study based on sea
677 bottom recordings from the Sea of Marmara. Geophysical Journal International, 190:993-
678 1007.
679 Thunell, R., Tappa, E., Valera, R., Llano, M., Astor, Y., Muller-Karger, F. and Bohrer, R. 1999.
680 Increased marine sediment suspension and fluxes following an earthquake. Nature, 398:
681 233-236.
682 Toksöz, M.N., Shakal, A.F., Michael, A.J. 1979. Space-time migration of earthquakes along the

25
683 North Anatolian fault zone and seismicity gaps. Pure and Applied Geophysics, 924
684 117:1258-1270.
685 Uçarkuş, G., 2010. Active faulting and earthquake scarps along the North-Anatilian Fault in the Sea
686 of Marmara. PhD Thesis, ITU, Eurasian Institute of Earth Sciences, Istanbul, 173 pp..
687 Uçarkuş, G., Çakır, Z., Armijo, R., 2011. Western Termination of the Mw 7.4, 1999 Izmit
688 Earthquake Rupture: Implications for the Expected Large Earthquake in the Sea of
689 Marmara. Turkish Journal of Earth Sciences. doi:10.3906/yer-0911-72.
690 Vidal, L., Menot, G., Joly, C., Bruneton, H., Rostek, F., Çağatay, N, Major, C., and Bard, E. 2010.
691 Hydrology in the Sea of Marmara during the last 23 ka : Implications for timing of Black
692 Sea connections and sapropel deposition. Paleoceanography, 25. doi
693 :10.1029/2009PA001735.
694 Wetzler, N., Marco, S., and Heifetz, E., 2010. Quantitative analysis of shear-induced turbulence in
695 lake sediments. Geology, vol. 38, 4:303-306.
696 Zitter, T., Grall, C., Henry, P., Özeren, M.S., Çağatay, M.N., Şengor, A.M.C., Gasperini, L., Mercier
697 de Lépinay, Géli, L., 2012. Distribution, morphology and triggers of submarine mass
698 wasting in the Sea of Marmara. Marine Geology, 329/331:58-74.

26
3.10. Conclusions of chapter

The paleoseismological analysis done in the present work allowed identifying the
paleoearthquakes signal and the co-seismic fault offset in the sedimentary record of the
Çinarcik and Central basins of the Sea of Marmara. Additionally, the lithological,
biological, chemical and magnetic analysis give details about the last change from non-
marine to marine environments in the Sea of Marmara. Then, the main conclusions reached
are outlined below.

The age-depth curve built for the MD01-2425 core indicates that the sedimentary
record recovered in this core represents the last 17 cal kyr BP. The vertical distribution of
facies and concentration of magnetic minerals along this core allowed us to define two
major sedimentary units, which coincide with the last major environmental changes in the
Marmara Basin. The upper unit represents approximately the last 12.8 cal kyr BP, and due
to their fossil content (presence of calcareous nannoplankton, planktonic foraminifera and
radiolarian) it is considered as representing full marine conditions, which suggests a
connection between the Sea of Marmara and the Aegean Sea during the last 12.8 cal kyr
BP. The lower unit represent a period comprised between 12.8 and 17 cal kyr BP. The
fossil content of this unit is variable, being identified with the depth areas with
predominance of centric and pennate diatoms, some of them characteristics of freshwater,
and others areas with presence of calcareous nannoplankton; being the diatoms scarce or
absent. These variations in the fossil content indicate fluctuations in the environmental
conditions during this period, where the Aegean Sea was already connected to the Sea of
Marmara (a14.7 kyr BP., according Vidal et al., 2010). However, the presence of
freshwater diatoms also indicates a possible interruption of the connection between both
seas, and/or a high input of fresh water from the Black Sea, reducing significantly the
surface salinity.

The change between these two units from biological and lithological points of view,
seems to be clearcut, although geochemical elemental analyses do not show abrupt
changes. A layer with particular characteristics was identified between these two units. It
was identified in the Central and Çinarcik basins, and was considered as a reference bed
level. It helped us in conjunction with the MS profiles to establish the correlation between
the two basins. According to our age-depth curve this reference bed level was deposited

157
during 13.4-12.8 cal kyr BP. The end of this period coincides with the onset of the
Younger Dryas (colder and drier climatic conditions in the Mediterranean, Valsecchi et al.,
2012). The disappearance of diatoms above this level indicates that the connection between
the Black Sea and the Sea of Marmara was negligible or null, as shown by Vidal et al.
(2010).

The identification of the earthquake-related traces recorded on the MD01-2425 core


was based on visual, compositional, sedimentological, and especially magnetic analysis
(AMS). We applied the Sorting vs. Skewness and Passega (1964) binary diagrams, as well
as, the AMS to differentiate seismo–induced deposits type (Homogenites + Turbidites)
from the “normal” hemipelagic-type sediments. The application of these tools in the
sediments of the MD01-2425 core demonstrates its usefulness to identify this type of
seismo-induced deposits.

The difference in the terrigenous contributions and sedimentation rates between the
marine and non-marine unit could be explained by the influenced of climatic changes. The
non-marine unit was developed in a lowstand phase during the last glacial period,
characterized, in the Northeastern Mediterranean realm, by dry and cold climate
conditions; with absence of significant tree cover and development of steppic plants
(Mudie et al., 2007, Valsecchi et al., 2012). These climatic conditions and lowstands are
inferred to favor higher erosion and subsequent higher terrigenous material production
(e.g. Macklin et al., 2002; Wegmann and Pazzaglia 2009), which were transported and
deposited on the margins of the basin, affecting the slope stability, making easier the
development of subaqueous slumps by earthquake triggering. On the other hand, the
marine unit was developed mainly during a warm and humid Holocene period, which was
characterized by the development of Mediterranean forest (Valsecchi et al., 2012). It
limited the process of erosion and subsequent terrigenous material production, decreasing
the accumulation of sediment in the basin margins. Additionally, the rapid sea-level rise
resulting in a decrease of the erosional and depositional processes on the platform and
basin margins. Both processes result in lower sedimentation rates and limited amounts of
coarser sediment available to be transported to the bottom of the basin.

We interpret the majority of homogenites + turbidites (HmTu) deposits from


MD01-2425 core as earthquake-induced. According to our depositional model in the Sea

158
of Marmara the triggering mechanism of these HmTu deposits are the earthquake-related
slope failures and mass-wasting, which could be originated in margins of any tectonic
basin. The presence of blue and green amphibole and epidote in these deposits suggests a
predominant source of sediments from the southeast, south or east of the Çinarcik Basin,
where the metamorphic Karakaya Complex crops out.

In the MD01-2425 core, 27 HmTu events were identified in the non-marine section,
representing an average earthquake recurrence interval of a155 yr. In the marine unit this
type of event have not been recognized, however, 35 instantaneous event of less important
thickness and with textural characteristics which can be considered as earthquake imprints
(flaser laminætion, ball and pillow structures, boudinage of laminated layers, etc.) were
recognized. These events allowed estimate a recurrence interval of a 365 yr.

In the Central Basin of the Sea of Marmara, the study of correlated HmTu events at
opposite sides of a segment of the NAF allows the estimation of the vertical co-seismic
offset for a continuous period of 2 kyr (13 to 15 cal kyr BP). We interpreted that the
significant values of the vertical offset, up to 144 cm, represent a dominant vertical
(normal) throw. These results only concern the southern boundary fault of the Central
Basin considered as a fast subsiding pull-apart structure. This result is in agreement, at
least for this zone, with the hypothesis of the Sea of Marmara born and evolving as a pull-
apart basin along a segmented NAF (e.g. Barka and Kadinsky-Cade, 1988; Westaway,
1994; Wong et al., 1995; Armijo et al., 1999; 2002). We propose paleo-magnitudes (Mw)
ranging between Mw: 5.9 and 6.6 for the earthquakes associated to estimated offsets, for a
pre-Holocene period. For the last 12 kyr BP, we consider the activity as similar to before,
although the sedimentary recording is poor due to changes in sedimentation regime
induced by climatic conditions. These results may also represent a contribution to the
discussion about the Sea of Marmara’s historical and present seismic gap.

159
CHAPTER 4

The Late Pleistocene/Holocene sedimentary record


of the central Gulf of Corinth

4.1. Introduction

The Gulf of Corinth, in eastern Greece, is one of the most seismic regions in
Europe, and also one of the fastest continental rifting in the world. and provides an ideal
site for investigating in situ faulting and seismicity and for developing efficient seismic
hazard reduction procedures (the Corinth Rift Laboratory-CRL, Cornet et al., 2004). Both
modern and historical seismicity attest to the activity of the Rift (the Gulf and the coastal
areas). Several Mw>6 events caused severe damages and casualties in urban areas
(Rondoyanni and Koukis, 1989; Ambraseys and Jackson, 1990). Different authors
attempted to reconstruct the historical earthquake record of the Gulf for the last 2.5 kyr,
(Galanopoulos, 1953; Papadopoulos, 1998; Ambraseys and Jackson, 1990; Papazachos and
Papazachou, 1997; Papadopoulos, 2003). These works are completed by the
paleoseismological studies conducted in trenches and with offshore surveys (Collier et al.,
1998; Koukouvelas et al., 2001 Pantosti et al., 2004; Moretti et al., 2004; Lykousis et al.,
2007; van Welden, 2007; Campos et al., 2014), providing information about historic and
prehistoric record.

The present work aims to contribute to subaqueous paleoseismological


investigation in the Central part of the Gulf of Corinth, through the sedimentary archive
contained in giant piston cores. These cores were recovered by the R/V MARION-
DUFRESNES cruise as part of the GEOSCIENCES-II program, October 2001, and have
been preliminarily studied by Moretti et al. (2004) and Lykousis et al. (2007) with the
objective of: a) deunitining the subsidence and sedimentation rate in the Gulf of Corinth, b)
quantifying the slip rate on the Antikyria fault (northern marginal fault of the central Gulf
of Corinth), and c) calibrate part of seismostratigraphic units defined form seismic profiles
collected by R/V AEGAEO.

160
A sedimentological study has been conducted on the MD01-2478 core (central part
of the Gulf of Corinth) by van Welden (2007) in his PhD thesis studies. In order to develop
the paleoseismological approach, we will focus on the study of the MD01-2477 and
MD01-2479 cores, both recovered in the deeper part of the basin. Our objective is the
identification of the earthquake-related traces recorded in both cores, applying the
methodology proposed in subchapter 2.4.2. Especially, we applied the Anisotropy of
Magnetic Susceptibility (AMS) and sedimentological analysis in order to identify the
Homogenites + Turbidites (HmTu) deposits, deciphering them from the “normal”
hemipelagic-type sedimentation. As for the Sea of Marmara (chapter 3), we also analyzed
the last change from non-marine to marine environment and its impact on the transport and
deposition dynamics. On the other hand, we also analyze, but with less details, the MD01-
2481 core (collected in the southeast margin of the gulf) in order to study the slope
instabilities expressed as mass wasting deposits and debris flow.

4.2. Geological and tectonic setting

The Aegean region is located in the eastern Mediterranean and forms part of the
interface between the overriding Eurasian plate and the subducting African plate. In the
actual configuration the western and southern boundaries of the Aegean region are formed
by the Hellenic subduction system (figure IV.1) (Gealey, 1988; Stampßi and Borel, 2004,
Doutsos and Kokkalas, 2001). This region is characterized by an Alpine collisional belt
which has suffered a significant extension, associated to the post-orogenic extensional
collapse, at the southernmost margin of the European plate above a clearly identifiable
subducting African plate (McKenzie, 1970; Caputo et al., 1970). Extension above the
Aegean subduction (in back arc position) was clearly identified and modeled by Angelier
(1978), and recently related to slab roll-back. Following Jolivet et al. (2013), the Corinth
Rift begun to develop between 8 and 5 MY above the plunging mantle slab.

The Greek so-called Hellenic Chain is part of the huge and complex Tethyan
collisional belt, which extends from Pyrenees in Spain to the Tibetan plateau and farther
East. During the Mesozoic and Cenozoic times the Greek orogenic system, or the
Hellenids, was part of the Dinaride-Tauride branch of the Alpine system (Cavazza et al.,
2004, Agostini et al., 2009).

161
Figure IV.1: Main structural features of the Hellenic Arc and Trench system (from Doutsos and
Kokkalas, 2001). Gulf of Corinth is located inside the black box. Motion vectors of the African
plate and the Aegean microplate are taken from Kahle et al. (1998). The Aegean microplate is

motion of the African plate is about a10 mm/yr (in the Ionian Sea). a) Schematic map
moving up to 35 mm/yr to the southwest with respect to Eurasia, whereas the relative northward

summarizing the geodynamic framework around the eastern Mediterranean Sea together with the
major plates involved in collision process (after McKenzie, 1972). KF: Kefallonia Fault, NAF:
North Anatolian Fault, EAF: East Anatolian Fault, DSF: Dead Sea Fault, HT: Hellenic Trench,
P: Peloponnesus, C: Crete.

The Hellenids have been subdivided into various geotectonic zones based on their
lithological and structural characteristics (figure IV.2) (Mountrakis, 1984; Robertson and
Dixon, 1984). These zones form three major groups: the Hellenic Hinterland, the Inner
Hellenids, and the Outer Hellenids. The zones that belong to the Hellenic Hinterland are:
the Serbo-Macedonian and the Rhodope Massifs. They represent pre-Mesozoic complex
basement, which has been metamorphosed to eclogite, amphibolite, and/or greenschist
facies, and which is intruded by Tertiary granitoids (summarized by Burg et al., 1996). The
Inner Hellenids encompass: the Circum-Rhodope, the Vardar, the Pelagonian, the Attic-
Cycladic and the sub-Pelagonian zones (figure IV.2). They are constituted of metamorphic
rocks of pre-alpine protoliths, alpine sediments and igneous rocks that have been formed

162
from Mesozoic to Palaeogene magmatic activity (Schermer, 1993; Walcott, 1998). The
Outer Hellenids include from west to east: the Paxon (pre-Apulian), the Ionian, Gavrovo-
Tripolitza, the Olonos-Pindos and the Parnassos-Ghiona zones, respectively (figure IV.2) .
The study area (framed in a black box in the figure IV.2) comprises only sedimentary
rocks belonging to Outer Hellenic Zone. With the exception of the Olonos-Pindos Zone
which is represented by Eocene flysch sequences (Neumann and Zacher, 2004), the
remaining Outer Hellenic units are made of non-metamorphosed Mesozoic to Tertiary
platform carbonates (in Dercourt et al., 1977).

Figure IV.2: Simplified map of the Hellenide, (from Zachariadis, 2007). Gulf of Corinth is
located inside the black box.
The mechanisms that controlled the present-day Aegean Sea extensional regime
The Hellenic Arc subduction is considered to have operated since the Early- to Mid
Miocene (Angelier, 1978; Jolivet et al., 1994, 2013; Le Pichon et al., 1995; Walcott and

163
White, 1998). This extension is supposed to be linked to the subduction of African oceanic
lithosphere along the Hellenic arc and its associated roll back, which also contribute to the
rotation of the Anatolian–Aegean plate (Le Pichon et al., 1995; McClusky et al., 2000,
Sodoudi et al., 2013). The roll-back of the subducting slab is caused by the rapid
subduction of the dense subducted lithosphere relative to the overall convergence rate
(Royden, 1993). It causes a retreat of the subduction zone and an outward migration of the
overriding plate (Royden, 1993). From late Miocene time onward the extensional basins
developed in the entire Aegean is combined with the westward expulsion of Anatolian
plate and the propagation of the North Anatolian Fault into the Aegean region since about
5 Ma (Armijo et al., 1996). This westward expulsion is caused by the final collision of
Arabia and Eurasia, and was accommodated by the development of the North Anatolian
and East Anatolian Fault Zones (McKenzie, 1978).

In the figure IV.3 are represented the main active faults and earthquake fault plane
solutions for the Aegean region (Nyst and Thatcher, 2004). The Normal faulting (red fault
plane solutions and fault traces) is widespread on mainland Greece and in western Turkey
but rare elsewhere. Strike-slip faulting (green solutions and faults) is largely confined to
the northern Aegean Sea and along the North Anatolian Fault. Reverse faulting (blue)
occurs near the Hellenic Trench, in the Ionian Sea (western Greece), and northwest into
Albania.

The Gulf of Corinth – or Corinth Rift – is directly related to Aegean geodynamics,


for which different models have been proposed, as summarized by Nyst and Thatcher
(2004). Le Pichon et al. (1995) proposed a two-blocks model with clockwise rotation of
central Greece and counterclockwise rotation of the South Aegean and Anatolia (figure
IV.4a), resulting in a westward increasing extension across the Gulf of Corinth. Armijo et
al. (1996) suggested that present day extension is localized in the northern Aegean and
central Greece, and that the rifting is associated with the strike slip faulting of the two
branches of the North Anatolian Fault (northern and southern) (figure IV.4b). Finally,
Nyst and Thatcher (2004) proposed, based on to GPS kinematics, that the Aegean region is
made of four rigid microplates (Anatolian, South Aegean, Central Greece, and South
Marmara microplates) (figure IV-4c). The microplate motions, especially the clockwise
rotation of central Greece, induce the SW translation of the Peloponnese and South
Aegean, causing the aN-S extension and opening of the Gulf of Corinth. This extension

164
increases from a5 mm/yr or less beyond its eastern edge to a15 mm/yr in the central gulf
and reaches a20 mm/yr in the northwest.

Figure IV.3: Aegean region, showing active faults and earthquake fault plane solutions (from
Nyst and Thacther, 2004).

165
a b

Figure IV.4: Present Day kinematic models of Aegean region (from Nyst and Thacther,
2004). a) after Le Pichon et al. (1995), in yellow color the Greece plate, in blue color South
Aegean and Anatolia plates, circular arrows represent clockwise and counterclockwise
rotation (from Nyst and Thatcher (2004). b) after Armijo et al. (1996), in light yellow colors
are represented the slow extension zones, in dark yellow the main rift zones, circular
arrows represent clockwise and counterclockwise rotation. c) following Nyst and Thatcher
(2004) with the stable microplates and their approximate boundaries inferred from GPS
results. Predicted relative motions across microplate boundaries are shown by arrows that
indicate the motion of the south bounding block with respect to the north bounding block.

166
4.3. The Gulf of Corinth

The Gulf of Corinth is situated at the western margin of the Aegean extensional
province, in central Greece (figure IV.1), within one of the most actively extending and
highly seismic regions in the world (e.g. Armijo et al., 1996; Briole et al., 2000). It is a
semi-enclosed marine basin, of 115 km long and approximately 30 km wide. This N120°E
elongated basin separates the Peloponnese (south) from continental Greece (north), and is
developed roughly perpendicular to the N160°E trending fold and thrust belt of the Alpine
Hellenids mountain chains (Sorel et al., 2000; Lykousis et al., 2007) (figure IV.2). In the
west the gulf is connected to open marine Ionian Sea through the Rion-Antirion straits.
This straits is 2 km wide and 65 m deep. In the east, the gulf is linked with the Saronikos
Gulf and the western Aegean Sea by an artificial channel (The Corinth Canal) of 6.3 km
long, 21 m wide and 8 m deep.

The Gulf of Corinth has been considered as an active graben/rift bounded by E-W
striking “en échelon” faults, which are located onshore and offshore (figure IV.5) (Brooks
and Ferentinos, 1984; Papatheodorou and Ferentinos, 1993; Armijo et al., 1996;
Sakellariou et al., 2007; Stefatos et al., 2002; McNeill et al., 2005; Lykousis et al., 2007;
Bell et al., 2009). These faults cut the folded and thrusted tectonic units of the Hellenids,
which include mainly Mesozoic and Tertiary phyllites, ophiolites, unmetamorphosed flysh,
and thick platform carbonates. The youngest faulted sediments are of Miocene age,
suggesting that rifting here began post 15 Ma, after the formation of the Hellenids (IGME,
1983; Armijo et al., 1996).

The Gulf of Corinth reaches at its centre a maximum depth of 900 m, and it is
characterized by steep north and south dipping slopes (Stefatos et al., 2002) (figure IV.5).
The southern margin is very steep (slope gradient up to 20°) with a narrow and relatively
steep (3–6°) shelf, while the northern margin displays lower slope gradients (5–10°) and
extended shelves especially offshore Antikyra, Itea and Eratini bays (Lykousis et al.,
2007). The focal mechanism studies indicate a main N-S direction of extension (Bernard et
al., 1997, and references therein) (figure IV.3). Geodetic data show that the active
extension is focused offshore, increasing from 11 mm/yr in the central part of the rift (near
Xylocastro) to 16 mm/yr in its western part (near Aigion) (Avallone et al., 2004). These
data show up to 1.5 mm/yr uplift rate in the southern margin since the late Middle

167
Pleistocene (Keraudren and Sorel, 1987; Armijo et al., 1996; De Martini et al., 2004;
Palyvos et al., 2005; Lykousis et al., 2007; Ford et al., 2012). In the gulf, most of the
seismicity is located between 4 and 13 km depth (Hatzfeld et al., 2000).

The structure, stratigraphy and evolution of the Gulf of Corinth have been
extensively investigated by several authors (e.g., Brooks and Ferentinos 1984; Higgs 1988;
Ori, 1989; Sorel, 2000; Westaway, 2002; Moretti et al., 2003; Lykousis et al., 2007). One
of the last works is published by Taylor et al. (2011). Based on bathymetric and depth
seismic reflection data the authors show the depth structure and stratigraphy of the gulf.
Their main results are summarized in the figure IV.6. This figure reveals that the Gulf of
Corinth is an asymmetric rift, with “u” or “w” shaped basin. The Gulf shows an eastward
deeping floor and crustal thinning, from a40 km thick in the western Gulf to a25 km thick
in the eastern (Tiberi et al., 2001, Sachpazi et al., 2007). It shows also a clearly southwards
inclined basement, with a 3 km-maximum sediment, the thickest sediments accumulations
being located along the southern coastline.

According to Taylor et al. (2011), the gulf is controlled in the south by right-
stepping, “en-echelon” N-dipping faults as: the Hellenic, Ankrata, Derveni, Xylocastro,
and Sithas faults (figure IV.6). These faults systems are biplanar to listric with
surface/seafloor angles of a35° in the central sector and 45° to 48°ƕ in the east, but
decreasing in dip and/or intersecting a low or shallow angle fault of 15° to 20° in the
basement in the central sector and 19° to 30° in the east (Taylor et al. 2011). Whereas, the
major S-dipping border faults were active along the northern margin of the central Gulf
early in the rift history, and remain active in the western Gulf and in the subsidiary Gulf of
Lechaio, but unlike the southern border faults, are without major footwall uplift (Taylor et
al., 2011).

168
Figure IV.5: Simplified geodynamic setting of the Gulf of Corinth. The locations of major onshore - offshore faults and bathymetry were taken
from Moretti et al. (2003) and Stefatos et al. (2006). The relief from NASA SRTM. Motion vectors of the North Anatolian fault after Le Pichon et al.
(1995), and of the Aegean and African plates relative to Eurasia after McClusky et al. (2000). W.ANT.F: west Antikyra fault. Stars in blue and red
color show the location of the MD cores. Lines in red color show the location of two shallow seismic profiles interpreted by Moretti et al., (2004)
and Lykousis et al. (2007) (figure IV.7).

169
170
Figure IV.6: Depth seismic interpreted MCS data (from Taylor et al., 2011). Western (L27),
central (L25, L15, and L48), and eastern (L41, L37 and L06) sectors. H: Heliki fault; WCF:
West Channel fault; S/NEF: S and N Eratini faults; ESB: Eratini subbasin; GAL: Galaxidi fault;
D: Derveni fault; S: Sithas fault; X: Xylocastro fault; DFG: Derveni fault graben; VRA:
Vrachonisida fault; K: Kiato fault; L: Loutraki fault; P: Perahora fault; Per: Perahora
Peninsula; VRA: Vrachonisida fault.

The structure and stratigraphy of the Gulf of Corinth have also been studied at
shallow seismic depths (e.g. Lykousis et al., 1998; Sakellariou et al., 1998; Perissoratis et
al., 2000; Stefatos et al., 2002). In the figures IV.7a, b are represented two shallow-
penetration seismic profiles interpreted by Moretti et al. (2004) and Lykousis et al. (2007).
These profiles have an orientation N-S and pass the location of the MD01-2477, MD01-
2478, MD01-2479 and MD01-2481 cores. In these cores were recovered sediment from the
last 21 kyr BP, covering the latest marine - non-marine transition (see subchapter 4.5).
Some of them was studied in the present work. In these profiles the south N-dipping and
north S-dipping faults identified in the depth seismic profiles can be also observed (figure
IV.6).

Moretti et al. (2004) and Lykousis et al. (2007) concluded from these shallow
seismic profiles and from additional data, that these faults have a steep dip, comprised
between 40° and 60° (e.g. Aigion 60°N, Helike 60°N, Pisia 65°N, Shinos 45°N). These
profiles imaged about 500 m of sediments, which represent the last 250 kyr. The authors

171
indicate that the southern marginal slopes of the Central Gulf display a rather structureless
subbottom configuration with discontinuous and chaotic internal reflectors indicating
intense downslope sediment strata deformation, mass gravity flows (slumps, translational
slides, debris flows, etc.) and Gilbert-type fan delta in the lowermost slope. In contrast, the
seismic profiles from the northern margin show a thicker and relatively undisturbed
sediment column, although slumping is widely distributed along the slope.

Figure IV.7: Shallow seismic profiles with the location of the MD01-2477, MD01-2478,
MD01-2479 and MD01-2481 piston cores. a) F1: west-Antikyra fault. b) Two mega-slump are
present (> 2 km long) (from Moretti et al., 2004). For location of the seismic profiles and
cores see figure IV.5.
172
4.3.1. Seismic activity in the Gulf of Corinth

The Gulf of Corinth is tectonically one of the most active areas of Europe. A great
number of historical earthquakes occurred in this region, some of them reaching a
magnitude greater than Ms > 6 (figure IV.8). Severe damages in urban areas (Rondoyanni
and Koukis, 1989; Ambraseys and Jackson, 1990) and casualties have been caused. The
period from 1810 onwards is relatively well documented. However, further back in time it
becomes increasingly difficult to find data (Ambraseys and Jackson, 1990). Despite this,
different authors have tried to reconstruct the earthquake historical record of the Gulf of
Corinth for the last 2.5 kyr, (Galanopoulos., 1953; Papadopoulos, 1998; Ambraseys and
Jackson, 1990; Papazachos and Papazachou, 1997; Papadopoulos, 2003) (figure IV.8).

Figure IV.8: Earthquake epicenters in Greece during 550 BC to 1899 AD with shallow
focal depth h ” 60 km (from Burton et al, 2004).

The historical observations in the Gulf of Corinth give evidence of earthquake-


induced nearshore sediments failure, such as: the 373 BC, 1817 and 1861 events
(Papadopoulos, 2003 and references therein), the 1965 event (Ambraseys, 1967), the 1981,
1989 and 1995 events (Perissoratis et al., 1984; Papadopoulos, 2003; Ambraseys and

173
Synolakis, 2010). These events can generate high amplitude tsunami waves (Dominey-
Howes, 2002; Papadopoulos, 2003; Ambraseys and Synolakis, 2010). Some authors have
attempted to complete this historical record whit paleoseismological studies conducted in
trenches (Collier et al., 1998; Pantosti et al 2004, Koukouvelas et al., 2001) and
submarines researches (van Welden, 2007; Campos et al., 2013), which provide
information about historic and prehistoric record.

4.3.2. Oceanography and Quaternary sedimentation in the Gulf of Corinth

The modern sedimentological and oceanographic characteristics in the Gulf of


Corinth have been studied by different authors (Anderson and Carmack, 1973; Brooks and
Ferentinos, 1984; Piper et al., 1990; Papatheodorou and Ferentinos, 1993; Poulos et al.,
1996). Poulos et al. (1996) summarizes these sedimentological and oceanographic
characteristics, which are represented in the figure IV.9. These authors indicate the
presence of a thermocline layer located between 0 an 60 m depth, associated with a rapid
temperature decrease. This layer is separated from an almost uniform bottom water mass
(> 100 m depth) by a transitional zone placed between 60 to 100 m depth, in which the
temperature decreases gradually with depth. They signaled that the salinity remains almost
constant (38.5 r 0.1 psu) throughout the whole water column. At water depths higher than
600 m, the dissolved oxygen varies from 1.9 to 2.6 ml/l, underlining the absence of anoxic
conditions. Near-bed water circulation is associated with slow-moving currents, with 8
cm/s observed maximum speed. These currents are not sufficient to resuspend bottom fine-
grained sediments, but may displace previously-suspended clay and silt. These currents are
influenced by climatic conditions (atmospheric pressure changes). Recently acquired high-
resolution seismic data (Beckers et al, 2013) depicted strong erosion/lack of sedimentation
in the western termination of the Gulf of Corinth.

Poulos et al. (1996) and Perissorastis et al. (2000) indicate that the modern
sedimentation within the gulf is controlled by the input of terrigenous material discharged
by the rivers along its coastline, and hemipelagic processes (biogenic and /or bio-induced
material originating from the water column). The largest river feeding the gulf is the
Mornos, in the northwest, while, a series of short steep rivers enter in the southern gulf
draining the mountains of the northern Peloponesos region. Perissorastis et al. (2000)
proposed an average sedimentation rate for suspended sediment of about 0.25 mm/yr.

174
Poulos et al. (1996) indicate that the sediments presents in the central abyssal plain show
textural and compositional similarities to those present in the southern slope and rise. They
interpret that most of the sedimentary material derives from the high relief of the
Peloponesos and is discharged by the numerous ephemeral rivers and streams in the south
shelves. This material is transported then seawards over the narrow shelf and the steep
slope by gravity driven mass flows.

Figure IV.9: Schematic representation of the oceanographical and sedimentological


characteristics and processes acting within the Gulf of Corinth (from Poulos et al., 1996).

During late Quaternary the sedimentation of the Gulf of Corinth has been
controlled by the input of terrigenous material supplied by the rivers, as well as by the sea-
level changes. The gulf has been subjected to both marine and non-marine conditions;
isolation has been governed by sea-level stands, in relation to the bathymetric sill across
the Rion Straits (a -60 m, at present) (Piper et al., 1990), which was relatively stable during
the late Pleistocene (Perissorastis et al., 2000). When the sea level fell below the level of
the Rion Strait sill, the Gulf was transformed into a “lake” or, if the sea level was a few
meters above the Rion Strait sill, into a marine area with high fresh water content. When
sea level was higher, normal open marine conditions were restored (Perissorastis et al.,
2000). The last marine - non-marine transition is considered to have occurred between 12

175
kyr BP (Perissoratis et al., 2000; Collier et al., 2000) and 13.2 kyr BP (Moretti et al., 2004;
Lykousis et al., 2007).

176
4.4.The Late Quaternary sedimentary infill of the Gulf of Corinth

The study of the Late Quaternary sedimentary infill of the Gulf of Corinth was
performed through the sediment core analysis of three long piston cores recovered during
the GEOSCIENCES-II program, October 2001. In this campaign 5 long Calypso piston
cores were recovered (Dannielou, 2002) (figure IV.5 and table IV.1), and have been
preliminary studied by Moretti et al. (2004), Lykousis et al. (2007) and van Welden (2007).
The later author studied in detail the MD01-2478 core with paleoseismological objectives.
We dedicated our detailed investigations to three of theses cores: MD01-2477, MD01-2478
and MD01-2481, and will summarize their main lithological characteristic in this
subchapter.

Water Length
Core Latitude Longitude
depth (m) (m)
MD01-2477 38°13.28' 22°33.53' 867 20.08
MD01-2478 38°11.68' 22°33.51' 861 2.97
MD01-2479 38°13.83' 22°33.63' 837 19.6
MD01-2480 38°09.02' 22°47.57' 852 29.73
MD01-2481 38°05.49' 22°51.40' 847 15

Table IV.1: Location of 5 long Calypso piston cores were recovered by the R/V MARION-
DUFRESNES cruise as part of the GEOSCIENCES-II program, october 2001 In. italic letter the
cores studied in the present work.

4.4.1. Lithologic description of the MD01-2477, MD01-2478 and MD01-2481


cores.

Three cores are studied in the present work. The MD01-2477 core is the longest
one (figure IV.10, table IV.1, and annex D), it presents the most complete sedimentary
record compared to the other cores. The MD01-2481 core is the medium length and
contains chaotic deposits (figure IV.10, table IV.1, and annex F). The MD01-2478 core is
the shorter; it is 2.97 m long (figure IV.10, table IV.1, and annex E). Its short length is
due to a lack of penetration during the recovery (Moretti et al., 2004).

177
Through visual descriptions (color, grain-size, structure and contacts) and
microscopic observation of coarse sediments (binocular) and fine-grained part (80 smear
slides), we identified six (6) different sedimentary facies. The results and observations
obtained by van Welden (2007) and Moretti et al (2004) were also taken in consideration
for the facies definition. Three (3) muddy facies have been identified: homogeneous mud
(no, or very weak, lamination), laminated mud, and deformed mud (mega slump) facies.
Also, three coarse grain facies were defined: classical turbidite, homogenite-turbidite and
mud-clast breccia facies.

4.4.1-a- Muddy facies


Facies Mh (Homogeneous mud): Homogeneous or poorly laminated calcareous
muds, light olive grey to dark yellowish brown. They are can form few centimetres to one
meter thick layers. The grains are angular to sub-rounded with variable sphericity (high
and low). They comprise terrigenous components as carbonates and siliciclastics (detrital
carbonates being the most abundant) (figure IV.11); as well as by biogenic and bio-
induced particles such (calcareous nannoplankton, few planktonic foraminifers, sponges
spicules, algues, diatoms frustules and bio-induced micrite). Traces of aragonite crystals
(abundant in the top of the non-marine section), traces of micas, glauconite (marine
section), pyrite (predominant in the non-marine section) and flakes of organic matter are
present. In this facies the terrigenous and bio-induced precipited carbonates components
are the more abundant (80 to 98%). Sediments belonging to this facies are interpreted as
hemipelagic deposits (h.d.) or hemipelagites. Although, this unit is only applied to deep
marine muds, in order to simplify, the non-marine deep muds will also be called
hemipelagic-type deposits. The Mh facies dominates in the three cores (figure IV.10,
annex D, E and F). It represents the 60% to 66% of the total sediments in each core.

Facies Ml (Laminated mud): Thinly laminated calcareous mud. They display sharp
vertical changes of color from white, olive grey, brownish olive grey, greyish yellow to
brownish black. The laminæ are 1 to 2 mm thick; they form thin layers of few centimetres
(1.5 to 3 cm). In contrast with the facies Mh, this facies present a higher percentage of
aragonite acicular crystals (< 4 %); being observed a bed of 2 cm thick at 13.55 m depth
with 70% aragonite (figure IV.11). These laminated muds have also been studied by
Moretti et al. (2004) and Lykousis et al. (2007); they underlined the aragonite precipitation.
As the sediments containing aragonite also contain a terrigenous fraction, these authors

178
proposed that the aragonite was precipitated in shallow oversaturated water by rapid
evaporation and subsequently transported to deeper environments under the action of
turbidity flows. The Ml facies shows a low amount of diatoms, although, the 2 cm thick
bed of aragonite doe not contains biologenic elements. This facies is only present in the
MD01-2477 core, specifically, in the top of the non-marine section (between 13.55 to
16.61 m depth) (figure IV.10, annex D). In this core it represents 3% of the whole section.

Figure IV.10: Facies, magnetic susceptibility profile, radiocarbon ages and simplified log of the
MD01-2477, MD01-2478 and MD01-2481. The radiocarbon ages were calibrated using the Bcal
(Buck et al., 1999) software. In yellow color is represented the possible correlation between the
MD01-2477 and MD01-2478 cores. See cores location in figure IV-5 and cores photographs in
annex B, C and D. Rectangles in blue color indicate the stratigraphic position of four sections
studied in detail.

179
Facies Md (Deformed mud): Muddy units in oblique position or folded. This unit
is texturally and compositionally similar to the facies Mh. They are predominantly light
olive grey and olive grey color. Due to their high level of deformation this facies can be
interpreted as a slump (figure IV.11). This facies is only present in the MD01-2481 core
(figure IV.10, annex F), It has 4.76 m thick and represents 31% of the sediments of the
core.

4.4.1-b- Coarse facies


Facies Tuc (Classical turbidite): medium sands to coarse silts, which grade upward
to very fine silts or muds. They are predominantly dark yellowish brown or olive gray
color, show sharp or erosive bases, are massive or present rare current ripples and slight
normal gradation. Rare inverse gradation were observed. This facies forms very thin layer
from 0.5 to 2 cm thick (figure IV.11). The grains are angular to sub-rounded with variable
sphericity (high and low). Their components (terrigenous and biogenic) are similar to those
described in the facies Mh, but the percentage of siliciclastic can be slightly higher
compared to the homogeneous mud facies (30 to 15%). This facies can be interpreted as a
classical turbidites as defined by Bouma (1962), and Mutti and Ricci Lucci (1978).
Whereas, the exceptional layers that show inverse gradations probably are the product of
hyperpycnal river flooding, as the hyperpycnal turbidites defined by Mulder and
Alexander, (2001). This Tuc facies is present in the three cores. It represents the 20%
(MD01-2478) to 0.1% (MD01-2481) of the whole sedimentary sequence (figure IV.10).

Facies HmTu (Homogenite-Turbidites): this facies is composed by two units,


generally with a sharp contact and oscillation sedimentary structures in between (see
example in figure 1 subchapter 2.4.2). The basal unit (Tu) is coarse grained. It has similar
characteristics to the facies Tuc, but in general, the top of this unit show an intermediate
unit with sedimentary structures indicative of to-and-fro particles displacement. The upper
unit (Hm) is a homogeneous calcareous mud of gray olive color (figure IV.11). The Hm
unit is textural and compositionally similar to the Mh facies, being not easy differentiate
between them. Nevertheless, the following characteristic are useful in the identification of
the Hm unit: 1) its association with the basal Tu unit; 2) the presence of oscillation
sedimentary structures in between; and in some case 3) its characteristic color. However,
the best tool to differentiate the Hm unit from the facies Mh is the AMS, specifically the
magnetic foliation (see subchapter 2.4.2 and 4.7.2). This facies was recognized in the

180
MD01-2477 and MD01-2478 cores. It represents 30% of the whole sedimentary sequence
in the MD01-2477 core, and the 20% in the MD01-2478 core (figure IV.10).

Facies Mcb (Mud–clast breccia): The clasts are constituted of calcareous muds.
They have variable grain size from 10 cm to 1 cm of diameter, and usually are angular
(figure IV.11). The clasts are composed of sediments similar to those described in the
facies Mh and Ml. They are embedded within a mud matrix similar to the facies Mh. The
Mcb facies is only present in the MD01-2481 core. It has 0.54 m thick and represents 3.4
% of the all sedimentary sequence (figure IV.10).

181
a b

c d

Figure IV.11: Illustration of different lithofacies defined in the MD01-2477, MD01-2478 and
MD01-2481 cores. 1). Macroscopic photographs of fresh split cores (scale in cm). 2) Optical
microscope observation of smear slides of different facies evidencing the presence of:
calcareous nanoplankton, abundant calcite crystal and some quartz (microphoto a facies Mh
and microphoto b facies HmTu). Aragonite needles and crystal of calcite and quartz,
(microphoto c, facies Ml). Crystals of calcite and quartz (microphoto d, facies Tuc).

182
4.5. Stratigraphy of the MD01-2477 and MD01-2481 cores

The vertical distribution of facies and the Magnetic Susceptibility (MS) profile
along the MD01-2477 and MD01-2481 cores allow us to define different sedimentary units
along these cores.

- The MD01-2477
This core can be divided into two major units: an upper and a lower unit. The upper
unit contains classical turbidites (Tuc facies) and homogenites-turbidites (HmTu)
interlayered within homogeneous mud (Mh facies). This unit is 13.5 m thick (0 to 13.5 m
core depth) (figure IV.10). According to our age-depth curve (see subchapter 4.6.2), it
represents approximately the last 11.7 cal kyr BP of sedimentation. This unit is
characterized by the presence of calcareous nannoplankton, with planktonic foraminifera in
less proportion. It can be subdivided into two subunits: U1 and U2 (figure IV.10). The
upper subunit U1 (0 to 11.7 m depth) is characterized by a greater amount of event
deposits (Tuc and HmTu facies), compared to the lower one (U2 between 11.7-13.5 m
depth). The values of Magnetic Susceptibility (MS) are highly constants and higher in U1
compared to the U2 subunit (see subchapter 4.7.1 for detailed information about MS).
Based on the biological characteristics we consider this upper unit as deposited under
marine conditions, which probably were well established starting from 11.7 cal kyr BP.

The lower unit contains classical turbidites (facies Tuc) and homogenites-turbidites
(HmTu) interlayered within homogeneous mud (Mh facies) and thinly laminated
calcareous mud (facies Ml). This unit is 6.58 m thick (13.5 to 20.08 m core depth) (figure
IV.10), and according to our age-depth curve (subchapter 4.6.2) it represents
approximately the 11.7-17 cal kyr BP period. The most distinctive elements of this unit
are: 1) the occurrence of diatoms; and 2) the presence in its upper part (between 13.55 to
16.61 m depth) of thinly laminated calcareous mud, which contain acicular aragonite
crystals. These aragonite crystals probably precipitated in shallow water, during the Late
Pleistocene (~ 13.2 to 14.53 cal kyr BP), as was documented for the Gulf Corinth by
Lykousis et al. (2007), or for the Gulf of Pagassitikos by Karageorgis et al. (2013). They
were subsequent transported to the bottom of the basin by turbidity currents. Two probable
mechanisms could produce the precipitation of aragonite in the Gulf of Corinth:

183
(1) Karageorgis et al. (2013) propose that for this period of time in Greece
prevailed cold and dry climatic conditions, favoring the elevated rates of
evaporation and consequently the aragonite precipitation.
(2) A physicochemical precipitation by mixing of fresh and sea waters. In the Gulf
of Corinth the last marine - non-marine transition is considered to have
occurred between 12 kyr BP (Perissoratis et al., 2000; Collier et al., 2000) and
13.2 kyr BP (Moretti et al., 2004; Lykousis et al., 2007). This period also
coincides with the occurrence of aragonite in the MD01-2477 core. The
aragonite may have precipitated in this period of transition, characterized by the
entry of sea water from the Ionian Sea, through the the Rion-Antirion strait,
within the isolated Gulf of Corinth (brackish water), favoring the mixing of the
brackish and sea waters and precipiting aragonite.
Base on the biological (presence of diatoms) and mineralogical (aragonite crystals)
evidences we consider this lower carbonate-rich unit to have been deposited in non-marine
and transitional (marine to non-marine) environments.

- The MD01-2481
This core can be divided into three major units (figure IV.10), all of them deposited
under marine conditions during the last 7.5 cal kyr BP. The upper (0 to 2.3 m depth) and
lower (6.5 to 15 m depth) units are composed by a small number of classical turbidites (Tuc
facies) interlayered within homogeneous mud (Mh facies). These layers show similar
lithological characteristics, and exhibit relatively constant values of MS, compared to the
middle unit. This last unit is characterized by the presence of highly deformed deposits
belonging to the mud-clast breccia and mud deformed facies (Mcb and Md facies), and
exhibit values of MS highly variables (see subchapter 4.7.1 for detailed information about
MS).

184
4.6. Chronostratigraphic framework

A chronostratigraphic framework for the studied cores has been developed by


Moretti et al. (2004), Lykousis et al. (2007) and van Welden (2007). In the present work
we will use the data proposed by these authors to define our age-depth curve and to
calculate the sedimentations rates. The dating were done in seven samples with plant debris
and woods fragments, taken in coarse to fine turbidites, as well as, one sample with
particulate organic matter collected from organic rich muds

4.6.1. Analytical results

14
The Accelerator Mass Spectroscopy C results have been calibrated in the same
way than those from the Marmara Sea. In the subchapter 3.4.4-c these processes were
explain in detail. To construct the age-depth curve six (6) radiocarbon ages were used
(table IV.2). The ages taken at 6.5 and 20.07 m depth were discarded because they
overestimate the expected age. The 6.5 m depth sample was collected in a turbidite
belonging to HmTu facies. Therefore, the organic matter from this sample is reworked.
The sample collected at 20.07 m depth belongs to a 2.32 m thick layer (17.75 to 20.07 m
depth) (see figure IV.10 and annex D). This layer is made of homogeneous mud, with
highly constant values of Magnetic Susceptibility (MS), anisotropy of magnetic
susceptibility (AMS) and grain size distribution. This layer does not show deformation or
sedimentary structures, and can represent an event deposit, probably the upper unit (Hm)
of a big HmTu association. In the Gulf of Corinth, when the sediments have low values of
MS (< 50 x 10-5 S.I.) and very low values of magnetic foliation (figure IV.12), the AMS
can not be used as a tool for interpretation if this deposit is a homogenite (Hm).
1.1
Figure IV.12: Magnetic foliation vs
1.08 magnetic susceptibility plot of all very
fine grain samples studied in the
Foliation

1.06 MD01-2477 core. In blue color are


represented the Homogenites samples,
1.04 in pink color the “normal”
hemipelagic type deposits. In general,
1.02 samples with MS> 50 x 10-5 S.I show
higher values of foliation compared to
1 the hemipelagic deposits.

0 50 100 150 200 250 300 350


Magnetic susceptibility
10 -5 S.I.

185
Thus, as the age of the 20.07 m sample is overestimated, and is unclear if the 2.32 m thick
deposit represents a normal continuous or event deposit, we will not take into account this
age and this unit for the construction of our age-depth curve.

All radiocarbon ages were calibrated by the Marine09 calibration curve (Reimer et
al., 2009). To calibrate this data it is also necessary to know the value of the local
correction of the marine reservoir age (ǻR). In the Gulf of Corinth the ǻR value and the
reservoir correction (R) are not well-constrained. Severals authors (e.g. Heezen et al.,
1966; Stiros et al., 1992; Soter, 1998; Pirazzoli et al., 2004; Palyvos et al., 2005) proposed
and/or use different values of R and ǻR for the present day marine conditions. These
parameters been calculated for the Aegean or the Mediterranean seas. The ǻR values used
varies between ǻR = -80 yr (Stiros et al., 1992) and ǻR = 380 yr (Soter, 1998). The range
between these values includes the ǻR deunitined by Reimer and McCormac (2002) for
Aegean Sea. They proposed a ǻR = 35 r 50 yr for Nauplia, Greece (at 70 km to the Gulf of
Corinth) and ǻR = 143 r 41 yr for Piraeus, Greece (at 30 km to the Gulf of Corinth).

In order to test how the different values of ǻR can modified the range of calibrate
ages, we calibrated the radiocarbon age belonging to the MD01-2477 core with the
program OxCal v4.1 (Ramsey, 2001) using different ǻR (see table IV.2). The range of
calibrated ages obtained for each ǻR are very different, especially if we compare the
opposed values (ǻR = -80 yr and ǻR = 380 yr). To calibrate the radiocarbon ages of the
Gulf of Corinth we have chosen the ǻR values proposed by Reimer and McCormac (2002),
because these values are more consistent with the values given for the Mediterranean
region (ǻR= 35 ± 70 yr according to Siani et al. (2000) and ǻR=58 ± 85 yr by Reimer and
McCormac (2002)), for the Marmara Sea (ǻR= 71 r 40 and ǻR= 91 r 40 yr by Siani et al.
(2000)) and for the Black Sea (ǻR=75 ± 65 yr according to Siani et al. (2000)). To use a
single value of ǻR we calculated the average value of both ǻR with their average errors,
then we used a ǻR= 89 r 58.

186
Error
Range of Range of Range of Weighted Range of ages
Weighted
Depth ages calibrated ages calibrated ages calibrated average age calibrated
Age Age average age
in core (Oxcal) (Oxcal) (Oxcal) (Bcal) (Bcal) Type
(yr) error (Bcal)
(m) (cal yr BP) (cal yr BP) (cal yr BP) (cal yr BP) (cal yr BP)
(cal yr BP)
(Delta R=-80) (Delta R=380) (Delta R=89) (Delta R=89) (Delta R=89)
(Delta R=89)

a
2.63 3190 40 3236 -2966 2698 - 2408 3071 - 2730 2892 (+) 172 (-)157 3064 to 2734 PDOM
a
3.9 3355 34 3401 -3210 2835 - 2690 3301 - 2907 3097 (+)183 (-)184 3280 to 2913 PDOM
a
5.16 5590 40 6185 - 5950 5655 - 5465 6082 - 5701 5877 (+)185 (-)171 6052 to 5706 PDOM
a
6.5 12420 60 14120 - 13786 13673 - 13319 13990 - 13493 -- -- -- PDOM
a
11.07 8500 50 9392 - 9077 8760 - 8440 9251 - 8763 9014 (+)225 (-)235 9239 to 8779 PDOM
a
16.33 12470 60 14171 - 13816 13726 - 13375 14040 - 13614 13831 (+)207 (-)213 14038 to 13618 PDOM
a
17.66 14000 100 17049 - 16550 16688 - 15291 16946 - 15987 16593 (+)335 (-)375 16928 to 16218 PDOM
b
20.08 21133 110 25108 - 24448 24692 - 23871 25005 - 24300 -- -- -- POM

Table IV.2: Results of radiocarbon dating carried out by van Welden (2007) (denoted by the
letter a) and Lykousis et al. (2007) (denoted by the letter b). The calendar ages were calculated
using the Marine09 calibration curve and applying different values of the local variations of the
reservoir effect (ǻR) proposed for the Gulf of Corinth and surrounding areas. The radiocarbons
dating were calibrated using Oxcal (Ramsey, 2001) and Bcal (Buck et al., 1999) software.
Samples in italic overestimate the expected age. Type of organic matter: PDOM= plant
debris and wood fragments extracted from fine and coarse turbidites; POM= particulate
organic matter extracted from organic rich muds.

4.6.2. Age-depth curve

After selection of the calibration curve and the ǻR correction we proceed to


construct the age-depth curve. Due to the abundance of reworked deposits in the longer
MD01-2477 core, and the limited or inexistent amount of radiocarbon ages for the MD01-
2478 and MD0-2481 cores (table IV.3), we will focus on the construction of the age-depth
curve for the MD01-2477 core. In this core sixty-seven (67) event deposits (32 turbidites
and 35 homogenites-turbidites) have been identified and removed from the sedimentary
column in order to obtain the normal sedimentation. The cumulated thickness of these
event deposits is 608 cm, approximately the 34.4% of the total thickness of the
sedimentary column.

Range of ages
Depth Age Age calibrated
Core
(cm) (yr) error (cal yr BP)
(Delta R=89)

MD01-2481 7.65 28540 230 32985-31550


MD01-2481 15.57 7320 40 7846-7564

Table IV.3: Results of radiocarbon dating carried out by Lykousis et al. (2007) in the
MD01-2481 core. The calendar ages were calculated using the software Oxcal (Ramsey,
2001) with the Marine09 calibration curve and applying the local variations of the reservoir
effect (ǻR) proposed by Reimer and McCormac (2002).

187
After the elimination of the instantaneous deposits we tested different age-depth
curves (linear, spline and polynomial) with the utilization of the software Clam (Blaauw,
2010). Figure IV.13 displays the different curves. The lineal interpolation takes into
consideration all radiocarbon ages, but involves abrupt changes in the sedimentation rate
(SR), whereas the spline and polynomial interpolation curves tend to move away or not
pass through some radiocarbon ages, specifically where a sudden change in the
sedimentation rate is observed (between 2 and 4 m depth). Although all curves have their
weaknesses, for the MD01-2477 core we privilege the linear interpolation because it takes
into consideration all the 14C ages.

4.6.3. Mains results

Figure IV.14 represents the age-depth curve for the MD01-2477 core without
instantaneous events. In this curve it is possible to identified five (5) levels characterized
by changes in the SR which varies between 0.32 to 3.68 mm/yr. The thickness of each
level and their SR are summarized in the table IV.4. After calculating the SR were added
the instantaneous event (IE) to the age-depth curve represented in the figure IV.14, as was
explained in the subchapter 3.4.4-e. It results in a new age-depth curve with instantaneous
events (figure IV.15). In the figure IV.15 and table IV.4, it can be observed that the event
deposits are distributed along the whole section, although the higher concentration are
observed in the levels 2 (marine section), 4 and 5 (marine and non-marine sections).
According to this age-depth curve the M01-2477 core represents the last 17 cal kyr BP of
sedimentation. A slightly higher SR is observed in the hemipelagic deposits of marine
environment (0.32 to 3.68 mm/yr) compared to those deposited in the non-marine one
(0.31 to 0.82 mm/yr). These differences (SR and recurrence intervals of instantaneous
events) are discussed in the subchapter 4.11.

188
Lineal interpolation Spline interpolation Polynomial interpolation

Figure IV.13:: Lineal (left plot), spline (central plot), and polynomial (right plot) interpolations generated using the software Clam
(Blaauw, 2010), and taking into account 6 radiocarbon dating (see table IV.2). The lineal curve involves abrupt changes in the
sedimentation rate. The spline and polynomial interpolation curves tend to move away or not pass through some radiocarbon ages
(red circles)
189
Figure IV.14: Age-depth curve for the MD01-2477 core (without instantaneous events. Five levels
characterized by differences in the sedimentation rates (SR) are defined.

Figure IV.15: Age-depth curve for the MD01-2425 core (with instantaneous events). Five levels
characterized by different sedimentation rates are defined. The radiocarbons dating were calibrated
using the software Bcal (Buck et al., 1999) (see table IV.2).

190
Thickness Sedimentation
Thickness with Percentage of
without rate without
Level instantaneous instantaneous
instantaneous instantaneous
events events
events events

Level 1 263 212 0.73 19.4


Level 2 127 73 3.68 42.5
Level 3 126 91 0.32 27.8
Level 4 1117 694 ~0.75 37.9
Level 5 133 88 0.31 33.8

Table IV.4: Major levels defined according to their sedimentation rates in the MD01-2477
core. Thickness in cm; instantaneous events percentage calculated for each level.

191
4.7.Magnetic properties of sediments of the MD01-2477, MD01-2478 and MD01-
2481 cores

In order to study the evolution of the concentration of magnetic particles


throughout the MD01-2477, MD01-2478 and MD01-2481 cores, high resolution Magnetic
Susceptibility (MS) measurements were performed. In addition, in the MD01-2477 and
MD01-2481 cores the magnetic fabric (orientation and direction of alignment of the
magnetic minerals that constitute the sediment) was studied through Anisotropy of
Magnetic Susceptibility (AMS) analysis. Finally, in order to distinguish different
mineralogical and granulometric fractions within the magnetic components in the MD01-
2477 core, two types of laboratory remanent magnetizations were artificially imparted:
isothermal remanent magnetization (IRM) and anhysteretic remanent magnetization
(ARM).

4.7.1. Magnetic Susceptibility (MS)

The high resolution Magnetic Susceptibility (MS) profiles of the studied cores
(MD01-2477 and MD01-2478) are represented in the figure IV.10. The longer core
(MD01-2477) presents the more complete profile; it covers at least the last 17 kyr
sedimentation (see subchapter 4.6 for detailed information) and shows the variations of the
concentration of magnetic particles for the marine and non-marine sections. In this core the
marine section can be divided into two subunits with different behavior of the MS curve.
The upper subunit (0 to 11.7 m depth) show highly constant MS values, with a mean value
of 23.92 r 6.27 x 10-5 S.I. The lower subunit (11.7 to 13.5 m depth), as the upper unit,
show highly constant MS values, but the values are lower (mean: 11.20 r 6.67 x 10-5 S.I.).

The marine to non-marine transition, and the non-marine section presents highly
variable MS values, varying from 1000 to 1 x 10-5 S.I., with a mean value of 34.03 r
101.68 x 10-5 S.I. Therefore, the MS curve behavior allows the differentiation between
marine and non-marine environments. Additionally, this MS curve permits to correlate this
core with other long piston cores taken in the Gulf of Corinth Basin, such as the MD01-
2479 and MD01-2480, studied by Moretti et al. (2004), and van Welden (2007).

192
The MD01-2478 core is the shorter one, it covers only the upper part of the marine
section. As for MD01-2477 core, the values of the MS are highly constants, with a 26.99 r
12.21 x 10-5 S.I. mean. The coincidences in the behavior between these two curves permit
their correlation (figure IV.10).

The MD01-2481 core represents only the marine section, and penetrated chaotic
deposits (figures IV.10, IV.16). They have a particular signature in the MS curve: high
variability if compared with lower and upper normal deposits. MS values vary from 46.63
to 0.66 x 10-5 S.I., (mean=16.39r 9.51 x 10-5 S.I) whereas, the upper (mean=20.20 r 3.38 x
10-5 S.I) and lower (mean=28.54 r 6.85 x 10-5 S.I) deposits are more constant, and presents
values in the same order than for the marine section of the MD01-2477 and MD01-2478
cores.

4.7.2. Anisotropy of Magnetic Susceptibility (AMS)

In the Gulf of Corinth the AMS analysis were conducted with two objectives:
1) to confirm the potential for paleoseismic recording of its sedimentary infill as
previously underlined (Moretti et al., 2004; Lykousis et al., 2007, Campos et
al., 2011, Campos et al., 2013). Additionally, the MD01-2477 core shows the
presence of numerous HmTu events with similar characteristics to those that
have been interpreted as earthquake-triggered (e.g. Sturm et al., 1995; Chapron
et al., 1999; Beck et al., 2007). The AMS in this core has been applied in order
to identify and characterized the probable earthquake trigger deposits.
Specifically, we want to identify and characterize the instantaneous
homogeneous deposits (Hm), and attempt to distinguish them from normal
hemipelagic - type deposits (h.d.), using the methodology presented in the
subchapter 2.4.2. This will allow us to measure the thickness of the hemipelagic
intervals, which will be used to estimate the earthquake recurrence time interval
(see subchapter 4.11).
2) 2) The AMS was also applied with the aim of evaluating its degree of
sensitivity to detect deformation in soft sediments as submarines slumps. The
AMS has been previously used in the study of deformation of sediment, where
several authors (e.g. Housen and Kanamatsu, 2003; Schwehr and Tauxe, 2003;

193
Meissl et al., 2011) have shown that the magnetic fabric is extremely sensitive
to the strain (Parés, 2004). A mud clast breccia and a mud deformed units
(mega slump) have been identified in the MD01-2481 core between 230 and
750 cm depth (figure IV.10, annex F), Thus, the AMS has been tested in these
layers to identify sediment deformation.

4.7.2-a Anisotropy of Magnetic Susceptibility of the MD01-2477 core


Methodology, main results and interpretations of the AMS measures in the MD01-
2477 core are discussed in the subchapter 2.4.2 (article published in Sedimentary
Geology). Additionally, in the present subchapter we will illustrate two others examples in
order to complete the data.

- Results
In the figure IV.16 are represented two sections of 70 and 50 cm long, where
continuous AMS analysis were performed. The location of the sections is indicated in the
figure IV.10. The section B (960 to 1030 cm depth) presents five (5) event deposits (two
Tuc and three HmTu) intercalated within continuous hemipelagic-type sediments. The
section D (1390 to 1450 cm depth) presents only two (2) events deposits (type HmTu)
intercalated within hemipelagic deposits.

The section B (figure IV.16a) has the same behavior as the sections studied in the
subchapter 2.4.2. The MS shows a tendency to increase with the decrease of the grain size.
The highest values of MS are present in the hemipelagites, and upper Hm units (mean: 215
r 20 x 10-6 S.I.). The lowest values are observed in the Tuc and Tu units (mean: 152 r 47 x
10-6 S.I.). In section B (figure IV.16b), at the opposite, MS is higher in the coarse interval.
The magnetic foliation (F) is variable in each unit. The upper Hm units present the highest
values (mean: 1.066 r 0.006), while the basal coarser units shows the lowest (mean: 1.031
r 0.01). Moreover, the magnetic lineation throughout the section remains very low and
relatively constant (mean: 1.004 r 0.002). Likewise, all sedimentary deposits (Tuc, HmTu
and h.d.) have oblate magnetic ellipsoid shapes (0.54<T<1) with a minimum susceptibility
axis very close to the vertical (80°< Inc kmin <90°).

194
Section B

Section D

Figure IV.16: Photos and AMS profiles of two sections of the MD01-2477 core. a) Section B. b)
Section D. Stratigraphic position of these sections are indicated in figure IV.10. Sampling
intervals each 2 cm.

195
The section D (figure IV.16b), shows the highest values of F in the upper Hm unit,
presenting the lowest values in the coarse units (Tu). Additionally, it shows oblate
magnetic ellipsoid shapes (0.06<T<0.98) with a minimum susceptibility axis very close to
the vertical (60°< Inc kmin <88°). It differs from the previous studied sections by their very
low MS (mean: 76 r 30 x 10-6 S.I.). Likewise, this low concentration of magnetic particles
affects directly the AMS. However, the F in the Hm units shows the highest values, but
they not exceed 1.05 (mean=1.030 r 0.007).

- Interpretations
The main result presented in this subchapter (section B and D of the core); as well
as those shown in the subchapter 2.4.2, demonstrate that highest values of F are always
found in the Hm unit. We interpret that these variations of F between different units can
not be explained by differences of compaction, because all depositional events (Tuc, HmTu
and h.d.) in each section are located at the same depth within the sediment column. The
regular oblate deposit fabrics and the systematic vertical orientation of kmin (perpendicular
to the bedding plane) suggest that these magnetic parameters are diagnostic of primary
sedimentary fabric (or grain array), not being affected by any deformation (natural or
during coring and subsampling) or strong bioturbation. The marked difference between
high F values in the Hm units and low values in the hemipelagic-type deposits (both with
the same grain size) reflect different settlement process that originated these deposits.
Otherwise, the results obtained in the section D (figure IV.16b) indicate that the sediments
with low Magnetic Susceptibility and low F values are not ideal for the study of the
magnetic fabric.

4.7.2-b Anisotropy of Magnetic Susceptibility of the MD01-2481 core


- Results
In the MD01-2481 core a total of 22 AMS punctual samples were measured. In the
figure IV.17 are represented the MD01-2481 simplified log, a general grain size profile, as
well as, the MS, F, L and kmin inclination profiles. As was discussed above (subchapter
4.7.1) the chaotic deposits have a particular signature in the MS curve. Moreover, the F is
moderately variable along the mud deformed unit (mean: 1.027 r 0.01). The highest values
of F (mean= 1.064 r 0.04) are observed below this unit (800 to 900 cm depth). All samples
present relatively low and constant L values (mean: 1.01 r 0.01), whereas the inclination

196
angle of the kmin axis and T have variable values along the whole sequence. The
variability of the different magnetic parameters allows us to defined four units.

The first unit represented only by two samples (between 0 and 120 cm depth)
shows kmin inclinations values away from the vertical (63° and 50°). This unit
corresponds to the first section, which is slightly disturbed. They have a relatively low bulk
density and high water content that increase their tendancy to suffer minor to moderate
disturbance during the coring processes (Glew et al., 2001). The second unit (between 150
and 230 cm depth) shows two samples with kmin inclinations very close to the vertical
direction (78° and 85°). The third unit (between 230 and 750 cm depth) corresponds to a
mud deformed section (mega slump). It shows kmin inclinations highly variable (from 24°
to 81°), this variability is dependent of the position of the sampled layer with respect to the
horizontal. The figure IV.18 shows a section (40 cm thick) of this mega slump. In this
section different layers present different angles with respect to the horizontal. Two AMS
samples were taken in different beds with different dip. The beds with dips near to the
horizontal show kmin inclination closer to the vertical (73° to 81°), whereas, the samples
taken in beds with high dips show kmin inclinations away from the vertical (24° to 53°).
Finally, the fourth unit (between 750 and 1150 cm depth) exhibits kmin inclinations values
clustered around the vertical direction (mean: 79 r 4).

The units 1 and 3 show modifications in the primary fabric, in both cases these
modifications are related to different deformational processes. The disturbances in the first
unit most probably result from the low consolidation of sediments and subsequent
coring/sampling influence. Otherwise, the presence of a mega slump in the third unit
indicates that the variations in the kmin inclinations are associated to displacement,
movement and folding of cohesive muds. On the other hand, the kmin inclinations very
close to the vertical direction (78° and 85°) in the units 2 and 4, suggest the presence of a
primary planar deposition fabric, which is expected in undeformed sediments.

197
Figure IV.17: Simplified log of the MD01-2481 core, with a median, magnetic susceptibility,
magnetic lineation, foliation and kmin inclination profiles. The deformed mud (mega slump) and
the clast breccia facies are delimited by blue and yellow rectangles, respectively.

-Interpretation of results

The high values of F present below the mega slump unit (800 to 900 cm depth); can
be explained by an anomalous process of compaction associated with an increase of
lithostatic load (increase of vertical stress), probably induced by the deposits of the mega-
slump sequence.

198
Figure IV.18: 40 cm thick section belonging to the deformed mud facies (mega slump). Two AMS
samples were taken (black squares). A bed with dip near to the horizontal (upper section) shows a
kmin inclination close to the vertical (80°), whereas the sample taken in a bed with high dip show
a kmin inclination away to the vertical (46°).

199
4.7.3. Magnetic contents and their possible influence on magnetic fabric in the
MD01-2477 core

The different mineralogical and granulometric fractions within the magnetic


components were studied through two types of laboratory remanent magnetisations:
isothermal remanent magnetization (IRM) and anhysteretic remanent magnetization
(ARM). The main results and their interpretations are discussed in the subchapter 2.4.2. In
order to complete these data, in the present subchapter are presented the remanent
magnetizations profiles of the section B (960 to 1030 cm depth), see figures IV.10,
IV.16a, IV.19.

The rock magnetic parameters of the section B (figure IV.19a) show similar
behavior that the sections studied in the subchapter 2.4.2. However, the reduced thickness
of some instantaneous events (1 mm to 3 cm thick), implies a smoothing in the parameters
measurements. This differs from the examples presented in the subchapter 2.4.2, where the
high thickness of sedimentary events (Tu, HmTu) permits a clear characterization of the
mineralogical and granulometric fractions within the magnetic components.

The section B (figure IV.19a) presents five (5) instantaneous events: two turbidites
(named TucA and TucB), and three HmTu events (named HmTu 1, HmTu 2, and HmTu 3)
intercalated within continuous hemipelagic-type sediments. As for sections described in
the subchapter 2.4.2 the high S-ratio values (> 0.9) suggest a predominance of low-
coercivity mineral component (magnetite). Although, the high values of HIRM in the
upper part of the turbidites and the the bottom of the Hm units, and the slight increase of
the Km/J IRM1T ratio with the decrease of the grain size, indicate a minor proportion of
high-coercivity minerals (hematite, goethite), probably more abundant in the coarse
fraction.

The intensity values of ARM are generally lower in coarse grain size (Tu and Tuc)
and higher in smaller grain size units (Hm and hemipelagites), showing a tendency to
increase with the decrease of the grain size. The figures IV.19b, c, d show the relation
between the JARM and F of three HmTu events. The figure IV.19b represents the HmTu
1 event and its overlying h.d. deposit (major event of the section B). In the figures IV.19c,
d are illustrated two examples of two others instantaneous events (HmTu + hemipelagic-

200
type deposits) studied in the subchapter 2.4.2. The HmTu event represented in the figure
IV.19c is the largest event identified in the MD01-2477 core (70 cm, located in the section
C, between 1040 and 1110 cm core depth, see location in figure IV.10), while the event
represented in the figure IV.19d is the smaller one (26 cm, located in the section A,
between 662 and 688 cm core depth, see location in figure IV.10).

In the figures IV.19b, c, d the ARM intensity and the F parameter appear to
increase progressively with the decrease of the grain size. Nevertheless, the Hm units and
the hemipelagic-type deposits with relatively similar grain size and ARM intensity have
clearly different values of F (in figure IV.19b F = 1.061 r 0.001 in Hm and F = 1.055 r
0.02 in hemipelagites. in figure IV.19 c: F = 1.068 r 0.005 in Hm and F = 1.043 r 0.006 in
hemipelagites). This clear separation between high F values in homogenites and lower
values in hemipelagites and turbidites suggest that the F is not associated to volume MS or
to type of magnetic minerals. Therefore, this result confirms the interpretation proposed,
where high values of ASM foliation are due to specific depositional process.

Section B

201
Figure IV.19: a) Magnetic parameters profiles (JARM, Km/J IRM1T, S ratio and HIRM) of a selected portion from the MD01-2477 core (named section D
in the figure IV.10). The sampling interval is 2 cm. F is added to check its relationships with magnetic content. b). F versus JARM for the HmTu 1 event
and its overlying hemipelagic-type deposit (see location in a). c, d) Foliation versus JARM for two HmTu events, taken from subchapter 2.4.2.

202
4.8. High-resolution geochemistry of the MD01-2477core

The chemical composition analyses of sediments in the MD01-2477 core were


made with the objective of:
1) to deunitine a possible geochemical imprint of Late Quaternary sea level
changes. Specifically, during the transition between the highstand sea level phases
(associated to the Holocene glacio-eustasy between 12 kyr BP to Present), and the
lowstand phase (during the last glacial period between 70 to 12 kyr BP) (Collier et
al., 2000; Lykousis et al., 2007; Bell et al., 2008). The Gulf of Corinth is connected
with the Ionian Sea via the Rion-Antirion sill (a65 m depth) and during glacial
periods, sea level drops below this level resulted in a disconnected basin (“lake”);
during high sea-level (interglacial periods) open marine conditions are established
(Perissoratis et al., 2000);
2) to characterize the chemical composition of the event deposits, in particular the
HmTu events, and check differences vs. similarities between HmTu and the normal
hemipelagic-type deposits.

Measurements were performed on a 3 m long section (between 1200 and 1500 core
depth), where presumably the marine to non-marine transition is placed (determination
14
based on C data and lithological characteristics). Beside, several other sections - with
characterized HmTu and Tuc units - were processed. (figure IV.10).

4.8.1. High-resolution geochemical analyses of the marine non-marine


transition in the MD01-2477core

The evolution of the chemical parameters and different significant elements ratios
in the marine to non-marine transition of the MD01-2477 core are represented in the
figures IV.20 and IV.21. The figure IV.20 shows chemical elements profiles with general
regular behavior over the 3 m section (marine environments toward the top and non-marine
toward the bottom). The Si, Al and K elements are highly variable along the entire section,
whereas, the Ni, Cu, Zn and Zr are only slightly variable and do not show important
changes between both environments. The Ca and Sr show a more variable behavior and
relatively higher values tower the non-marine section. The Pb, S, Cl and Mn are highly
constant, with some peaks both in the marine (Pb, S, Cl) and non-marine sections (S, Mn).

203
The Fe and Br tend slightly to increase toward the top of the section, which may indicate
increasing marine conditions toward the top with a predominance of oxidizing
environment (Rothwell et al., 2006; Ziegler et al., 2008). Otherwise, the higher
concentration of peaks with S toward non-marine environments can be interpreted as a
predominance of anoxic conditions, which was confirmed through the visual observation
and smear slide analysis, where microcrystalline pyrite was identified. Similar observations
have been made by Moretti et al. (2004) and van Welden (2007) based on visual and
microscopic observations (smear slide and Scanning Electron Microscope).

We used the elemental ratios, as the chemical elements do not show any abrupt and
net shift between the marine and non-marine environments. The Si/Al and Si/K ratios show
a strong correlation with the grain size profiles. The Ca/Si ratio profile is highly variable,
showing a predominance of carbonate fraction toward the non-marine environments. In
general, the hemipelagic deposits exhibit an increased proportion of Ca. The Ca/Fe and
Ca/Zr ratio profiles are less variable in the marine section, with a predominance of
carbonate particles toward the non-marine environments. The relatively higher proportions
of Ca in the non-marine environments (particularly in the hemipelagic deposits), compared
to Zr indicative of terrigenous source (Cuven et al., 2010), or compared to Fe indicative of
reworked material from shallow water source (Rothwell et al., 2006), is interpreted as an
increase of biogenic and bio-induced carbonate minerals in this unit. Moreover, the little
variable relation in the Ca/Fe and Ca/Zr ratios toward the marine environments indicate
that this carbonate has mainly a terrigenous source. In addition, the Sr/Ca ratio profile
show high values of Sr compared to Ca in the laminated mud units. Sr is probably
associated with the presence of aragonite described in the subchapter 4.4.1-a.

In summary, the geochemical elemental analyses do not show an abrupt shift


between the marine and non-marine environments, to the contrary, the change seems to be
gradual. Although there is no clear limit between both environments, the predominance of
anoxic conditions, the increase of bio-induced and terrigenous carbonates toward the non-
marine section, and the occurrence (at 1355 m depth) of laminated mud with aragonite-rich
levels, suggest that below 1355 m depth the studied section was deposited under
transitional to non-marine conditions.

204
Figure IV.20: XRF geochemical profiles of a 3 m long section of the MD01-2477 core, where is placed the marine non-marine transition. A
simplified log of this section is show in the left.

205
Figure IV.21: XRF elements ratios of a 3 m long section of the MD01-2477 core. A simplified log
of this section is show in the left.

4.8.2. High-resolution geochemistry of event deposits of the MD01-2477 core

In the figures IV.22a, b, c, d are represented the pictures, median grain size, XRF
profiles of Rb and K, and different elements ratios (Si/K, Si/Al, Si/Ti, Ca/Si, Ca/Ti and
Ca/Fe) for the sections A, B and C. The sections A and B include several instantaneous
events (type HmTu or Tuc) intercalated within continuous hemipelagic-type sediments,
while. The entire section C is a single HmTu event.

The instantaneous events (Tuc and HmTu) and the hemipelagic-type deposits show
a similar behavior in each of the three studied sections. In all sections the Si/Al and Si/K
ratios show a strong correlation with the grain size profiles. This means that these ratios
can be used as a proxy of grain size evolution in the Gulf of Corinth. In the studied
sections, only the hemipelagic deposits show that this relation tends to be slightly variable

206
and different from the granulometric curve. This relation between Si, Al and K indicate an
abundance of detrital rock-forming minerals in the coarse units. The Si has two possible
sources: terrigenous silicates or primary biogenic productivity (Rothwell et al., 2006).
Nevertheless, the similarities in the profiles of Si/Al, Si/K and Si/Ti ratios suggest that all
these elements are mainly derived from terrigenous aluminosilicates (phyllites).

Section A

Figure IV.22: Picture, median grain size, XRF profiles of Rb and K, and different elements
ratios (Si/K, Si/Al, Si/Ti, Ca/Si, Ca/Ti and Ca/Fe) for the sections A, B and C of the MD01-2477
core. Stratigraphic position is indicated in the figure IV.10. a) Chemical analysis made in the
section A. b) Very high resolution measures (each 1mm), made in the 660-690 m depth HmTu
event present in the section A. c) Chemical analysis made in the section B. d) Chemical analysis
made in the section C.

207
Figure IV.22 (Continued)
Section B

Section D

The Ca/Si ratio profiles show an abundance of Ca, in particular in the coarse units
and hemipelagic deposits. Microscopic observation (binocular loupe and smear slides)
made in the sediments of the MD01-2477 core (all granulometric fractions) show that they
are mainly carbonates (98 to 70%). Terrigenous calcitic grains and micrite (more abundant

208
in fine grained deposits) are dominant. Biogenic calcareous debris and microfossils
represent a minor fraction. These observations are confirmed with the XRF results, were
the high values of the Ca/Fe and Ca/Ti relations in the basal coarser turbiditic units are
related with the high concentration of coarse carbonate particles in these units. The slightly
variable Ca/Fe and Ca/Ti ratios along the Hm units indicate a more uniform distribution of
regular grain size carbonate particles of terrigenous source, although carbonate bio-induced
precipitated crystals are also present (smear slides observations). High Ca values
characterize the hemipelagic type deposits, if compared to Ti, indicative of terrigenous
source (Cuven et al., 2010). High values of Fe in turbidites indicate shallow water source
(Rothwell et al., 2006).

The Rb and K profiles, both indicatives of detrital clays (Rothwell et al., 2006;
Cuven et al., 2010) are plotted for all sections. In the figure IV.22 the Rb and K
progressively increases with decreasing grain size, in all sections. The maximum values are
observed in the Hm units. The hemipelagic type deposits show values of Rb and K slightly
lower than in the Hm unit. This indicates a minor proportion of detrital clays in the
hemipelagic deposits.

The hemipelagic-type deposits show a slightly irregular behavior in most of the


plotted curves (figure IV.22), indicating a more heterogeneous composition than the
HmTu and Tuc deposits. Hemipelgic-type intervals show higher contribution of bio-
induced precipited carbonates compared to the instantaneous deposits.

The XRF analyses show that this method applied in conjunction with other
methods, such as sedimentological and magnetic analysis can help to differentiate
hemipelagic type deposits from instantaneous Hm deposits. It can not be achieved using
only granulometric analysis (see grain size profiles (median) in the studied sections, figure
IV.22). Both units (Hm and hemipelagites) with similar grain size slightly differ in their
content of detrital clays (slightly more abundant in the Hm units), and Ca bio-induced
(slightly more abundant in the hemipelagites). This difference illustrates that Hm deposits
can be the result of reworking processes: re-suspension and re-concentration of fine
terrigenous material transported by turbidity current and in situ deep sediment (by
earthquake-induced tsunami-seiche effect). On the other hand, the hemipelagic sediments
were deposited in calm conditions, with high biological productivity. This setting favors

209
the precipitation of carbonate through the removal of carbon dioxide from water by
photosynthesizing phytoplankton, such as diatoms, dinoflagellates and coccoliths (Boggs,
2011).

4.9. Estimation of organic matter and carbonate content in the MD01-2477 core
(Loss on ignition)

The deunitination of weight percentage of organic matter content in sediments of


the MD01-2477 core by Loss of ignition (LOI) method was made by van Welden (2007)
during the development of his PhD thesis (unpublished data). He carried out forty (40) LOI
measures in marine and non-marine sediments, both coarse (turbidites) and muds
(homogenites and hemipelagic deposits) sediments. The results are presented in figure
IV.23. They reveals a content of organic matter (OM) that vary between 2% and 12%,
presenting a higher percentage of OM in fine grain size sediments, compared to coarse
sediments. However, high OM values can be observed in some turbidites. In general, the
percentage of OM in the marine sediments tend to be slightly greater (average rate of 8%)
than non-marine (average rate of 6.4%) sediments. The OM weight percentage values
obtained by van Welden (2007) are close to those obtained by Poulos et al (1996), which
indicates values of OM between 5 to 7% for modern marine sediments of the Gulf of
Corinth.

Van Welden (2007) deunitined the carbonate percentage applying the Trentesaux et
al (2001) method. This method deduced the percentage of carbonates in sediments through
microgranulometric analysis (it is based on the study of the evolution of the obscuration
during the addition of HCl into the tank of a micro particle size analyzer MALVERNTM
Mastersizer S). Van Welden (2007) obtained about 30% values for coarse sediments
(turbidites) in both marine and non-marine environments. Higher values up to 50 to 70%
were also obtained for the fine grained muds (hemipelagites and homogenites). In general
the average percentage of carbonate in the non-marine muds was higher than 65%, while
for the marine muds was around 45%. These values are close to those reported by Lykousis
et al. (2007) for non-marine (mean about 55%), and marine (30 to 45%) sediments, as well
as, for those obtained by Poulos et al. (1996) for modern marine deposits (30 to 60%),
using the LOI method. Likewise, these values are consistent whit the high-resolution

210
geochemical analyses and visual carbonate percentage estimations (smears slides) made in
the present work.

We focused our study in the deunitination of percentage of OM content in the event


deposits, in particular the HmTu events, to test a possible additional characteristic.

The table IV.5 summarizes the percentage of OM obtained for ten (10) samples
studied in the MD01-2477 core. The samples were performed in four (4) Hm or HmTu
events; and their overlying hemipelagic-type deposit (these events were defined by their
visual, textural and magnetic features). Three of these examples (1, 2 and 3) are located in
the section A (see stratigraphic position in the figure IV.10), for which we have magnetic
(subchapters 2.4.2, 4.7.2), geochemical (subchapter 4.8.2), and textural (subchapter 4.10)
parameters.

Figure IV.23: Weight percentage of


organic matter content obtained by the
Loss of ignition (LOI) method. The
measures were performed over 40
samples taken in the MD01-2477 core.
Modified from van Welden (2007).
Unpublished data.

211
In the coarse (Tu) sediments, OM values are low (mean: 5.38 r 0.28), whereas the
higher values are present in the fine grained sediments (Hm or h.d.) (mean: 9.04 r 0.50).
These percentages coincide with these reported by van Welden (2007) and Poulos et al.
(1996). These percentages allow differentiating between coarse and fine grained sediments,
not being observed a marked difference in the percentage of OM between hemipelagic-
type deposit and homogenites.

% Organic
Type of Core depth
Example matter
Sedimentary deposit (cm)
(LOI 550°)
Hemipelagite 643 9,75
1
Homogenite 647 9,11
Hemipelagite 656 8,58
2 Homogenite 670 9,39
Turbidite 687 5,18
Hemipelagite 692 8,96
3
Homogenite 698 8,13
Hemipelagite 365 9,13
4 Homogenite 380 9,30
Turbidite 394 5,58

Table IV.5: Weight percentage of organic matter content obtained by the Loss of
ignition (LOI) method (this work). The measures were performed over 10 samples taken
in the MD01-2477 core.

212
4.10. Interpretation of depositional processes in the MD01-2477 core through
sedimentary texture analysis

In the present subchapter we examine the textural characteristics, such as: the grain
size parameters, shape and fabric of instantaneous (Tuc and HmTu) and normal
(hemipelagites) deposits, with the aim of understanding the processes of transport and
deposition that originated them and the environmental conditions that prevailed during its
deposition. The sedimentary fabric is controlled by the shape of the particles (platy or
elongated) and principally by the grains orientation, which indicate flow velocities and
hydraulic conditions at the deposition site (Boggs, 2011). In the MD01-2477 core the
sedimentary fabric was studied through the magnetic fabric. The main results obtained for
the MD01-2477 core are presented in subchapters 2.4.2 and 4.7.2. In this subchapter we
integrate all the textural data in order to perform a more complete interpretation of the
sedimentary processes. Moreover, complementary textural data and their interpretation will
be presented in the subchapter 4.11 (article in press in the Annals of Geophysics journal).

In the MD01-2477 core, we studied in detail three sections (figure IV.24), the large
amount of grain size data is presented in statistical form. In order to know the bottom-to-
top textural evolution of some instantaneous events, the skewness vs sorting index binary
diagrams and the Q99 (C) vs medians (M) Passega (1964) diagrams were plotted.

4.10.1. Results

4.10.1-a Grain size


The sedimentary assemblage of the MD01-2477 core both in the marine and non-
marine section show mostly two types of sediments, based on their granulometry:
(1) fine sediments (hemipelagic-type deposits or homogenites) characterized by a
dominance of silt-sized particles (83 – 75%), with lesser amounts of clay size (23 –
17% ) and sand (8 – 0%) particles. These fine sediments (mean: 4 to 7.5 µm) are
poorly sorted (sorting index (So) between 1.95 and 2.6), showing in some case a
symmetry (Sk index between 0.035 and -0.089) or an asymmetry toward fine
grained particles grain size distribution (Sk between -0.10 and -0.24);
(2) coarse sediments (turbidites) characterized by a dominance of sand (95 - 1%) or
silt size (88 - 5%) grains, with lesser amounts of clay size (17 - 0%) particles. The

213
sediments are moderately to very poorly sorted (So index between 1.4 and 5.5),
presenting a strong asymmetry toward fine grained particles (Sk: -0.11 to -0.57).

The figure IV.24a, b, c show the high resolution grain size profiles for three
sections of the MD01-2477 core (see stratigraphic position on figure IV.10). They were
chosen for their content of large number of well defined instantaneous events (most of
them HmTu events) with significant thicknesses (> 5cm). To achieve a complete
characterization of these events and their surrounding sediments, their chemical and
magnetic parameters were also measured with high-resolution spacing. (subchapters 2.4.2,
4.7.2, 4.7.3 and 4.8.2).

- Homogenites
In the selected sections (figure IV.24) the homogenites (Hm) present an average
mean grain size of 5.59 r 0.50 µm. They are characterized by an abundance of silt-sized
particles (average: 80.21 r 3.99 %), with lesser amounts of clay size (average: 19.03 r 1.60
%) and negligible sand (average: 0.76 r 0.46 %) particles. They are poorly sorted (So
average: 2.13 r 0.04) and the grain size distribution reveal an asymmetry toward fine
grained particles (Sk average: -0.12 r 0.03). In general their grain size frequency curves
are mesokurtic (Kurtosis (K) average: 0.94 r 0.02).

- Hemipelagites and turbidites


The hemipelagic deposits (h.d.) exhibit textural features similar to those described
in the homogenites, but generally have a higher standard deviation. Their average mean
grain size is slightly greater than Hm deposits (5.68 r 0.80 µm). The percentages of silt
and clay grain size are comparable to those of the homogenites (average of silt: 78.24 r
4.65 %, and clay: 19.64 r 2.17 %), although the sand size is slightly higher (average: 2.11
r 1.46 %). The sediments are poorly sorted (So average: 2.30 r 0.04), and generally do not
have excess of fine or coarse particles, exhibiting symmetric grain size distribution (Sk
average:- 0.04 r 0.05). The frequency curves usually are mesokurtic (K average: 0.97 r
0.03).

214
Figure IV.24: Textural characterization of three sections of the MD01-2477 core. a) Section A.
b) Section B. c) Section C. The stratigraphic position is located in the figure IV.10. Rectangles
in blue color (dashed lines) indicate the location of the events represented in Figure IV.25.

215
Figure IV.24 (continued)

4.10.1-b Skewness vs sorting and CM Passega’s diagrams


In the figure IV.25a, b c are represented the bottom to top textural evolution of
four (4) instantaneous events; three (3) HmTu events and one (1) turbidite (Tuc). Due to
the thin thickness (1 mm to 2 cm thick) that characterizes the Tuc deposits, it was not
possible to construct numerous skewness vs sorting and CM Passega’s diagrams. The
events illustrated in the figure IV.25 belong to the sections A, B and C were chosen by
their thickness (> 4 cm) (see location in the figure IV.24).

The event deposits types HmTu and Tuc show different characteristics in their
bottom to top grain size distribution. The HmTu events are characterized for: (1) a basal Tu
unit (dark to medium blue triangles in the figure) with complex variations of their sorting,
skewness, Q99 and median grain size; (2) an intermediate unit (light blue triangles in the
figure), characterized by the presence of to-and-fro sedimentary structures, generally well
differentiated from the coarse basal unit and the homogeneous upper unit. In contrast to the
basal unit, the intermediate unit is characterized by a slightly variable sorting and median
grain size, with a more variable skewness and Q99; (3) a sharp and clear separation
between the Hm unit and the intermediate and basal Tu unit. The Hm unit (pink dots in the
figure) show a highly constant sorting, skewness, and median grain size, but a slightly
variable Q99; and (4) the hemipelagic deposits and the Hm unit are presents in the same

216
regions of the diagram. Nevertheless, the hemipelagic deposits show more variable
skewness and sorting compared to the Hm unit.

Figure IV.25: Bottom to top textural evolution of four instantaneous events represented in the
skewness vs sorting and Q99 vs medians Passega (1964) diagrams. a) (1, 2) Two HmTu events
belonging to section A. b) Classical turbidite from section B. c) HmTu event from section C. The
locations of the events are represented in figure IV.24 (rectangles in blue color (dashed lines)).
The sampling intervals vary from 0.2 to 1 cm.

217
Figure IV.25 (continued)

The Tuc event represented in the figure IV.25b shows a highly variable sorting, but
with a less variable skewness, Q99 and median grain size. These characteristics allow
differentiate the Tuc event from the basal Tu unit associated to the HmTu events. Due to
the thickness of the Tuc events in the MD01-2477 core, more representations of the bottom
to top textural evolution were not possible. However, in the figure IV.26 are presented the
CM Passega’s diagrams of two classical turbidites belonging to the MD01-2479 core made
by Van Welden (2007) (unpublished data). These layers (yellow triangles and red squares)
show a bottom to top progressive decrease of the Q99 and of the median grain size. In
general the samples are located parallel to the C=M line. This last characteristic has also
been noted by Passega (1964) in the classical turbidites of the Marnoso-arenacea

218
Formation, Italian Apennines. The HmTu events of the MD01-2477 core do not show the
same characteristics that the classical turbidites identified in the MD01-2479 core,
evidencing different transports and depositions.

Figure IV.26: Bottom-to-top textural evolution of two classical turbidites of the MD01-2479 core
(from van Welden, 2007 -unpublished data-). The samples are plotted in the Passega (1964)
diagram.

On the other hand, Lignier (2001, in Beck (2009)) and Arnaud et al. (2002), in a
study conducted on a small intramontane lake NW Alps, identified two different paths of
grain size distribution in the skewness vs sorting and CM Passega (1964) diagrams. They
distinguished what they named “flood turbidite” (bypassing of tributary discharged) and
“slump turbidite” (foreset collapse evolving into turbidity current). The binary diagrams
showed in the figure IV.27a, b.

The bottom to top grain size distribution obtained in the HmTu events of the
MD01-2477 core (figure IV.25) is very similar to that obtained in the “slump turbidite”
defined by Beck (2009), which show a bottom characterized by high variation of sorting,

219
skewness, Q99 and median grain size. While, in the top of the turbidite prevails a slightly
variable sorting and median grain size, with a more variable skewness and Q99. On the
other hand, the HmTu event differs from the “flood turbidites”. The last show a
progressive decrease of the Q99 and the median grain size, with samples located sub-
parallels to the C=M line. Based on this comparison, we can interpret that the HmTu
events of the MD01-2477 core probably are product of the deposition of slump turbidites.

Figure IV.27: Bottom-to-top textural evolution of two types of turbidites detected in Lake
Anterne (from Beck, 2009). The samples are plotted in the Skewness vs. Sorting and Passega
(1964) diagram, with a sampling interval of 2mm. The two turbidites (flood turbidite and slump
turbidite) show distinct base-to-top paths.

220
4.10.1-c Particles shape
The particles shapes analyses conducted in the fine grain sediments of the MD01-
2477 show very similar results for all studied samples (either homogenites as
hemipelagites). In the figure IV.28a, b are represented the circularity vs diameter (µm)
diagrams and the particles images obtained from the SYSMEXTM FPIA – 2100 of two
instantaneous events and their overlaying hemipelagic deposits. In the figure IV.28a we
represented an HmTu event from the section A (see stratigraphic position on figure IV.10,
textural analyses in the figures IV.24a, IV.25a2) and the overlying hemipelagite. In this
section were studied two samples, one taken at 675 cm core depth, corresponding to the
Hm unit, other taken at 662 cm core depth corresponding to the hemipelagic deposits. As
indicated by circularity vs diameter (µm) diagrams, the samples are similar. In both
diagrams can be clearly identified that the particles are grouped into two distinct zones of
the diagram. A group of particles (diameters between 1.5 and 10 µm) is concentrated in the
high circularity (>0.8) zone. The other group of particles (diameter between 0.6 and 7 µm)
has a higher scattering and are characterized by more variable circularity (0.4 to 0.8). The
particles images correspond principally to fragments of terrigenous components,
predominantly carbonates and bio-induced precipitated crystals of calcite. In general they
have very angular to sub-angular edges.

In the figure IV.28b is represented an HmTu event belonging to the section B (see
stratigraphic position on figure IV.10, textural analyses in the figure IV.24b). In this
section were also studied two samples taken at 1020 cm core deph (Hm unit), and at 1006
cm core depth (h.d.). These samples show the same characteristics that those described in
the the section A. Additionally in this section some ~15 µm diameter diatoms frustules
were identidied. All samples studied along the MD01-2477 core exhibit the same
characteristics; in particular, homogenites and hemipelagic-type deposits are not
ditinguishable.

221
Figure IV.28: The shape particles analyses conducted in two homogenites and two

section, the circularity vs diameter (Pm) diagrams and the particles images obtained from the
hemipelagites of the MD01-2477 core. In the figure are represented the pictures of the studied

Sysmex FPIA – 2100.

222
4.10.2. Interpretation and conclusion of depositional processes in the MD01-
2477 core

Based on the textural parameters studied in the sediments of MD01-2477 core


(grain size, shape and fabric), three sedimentary facies and corresponding layering were
defined (Tuc, HmTu, and Mh), which we inferred as produced by different depositional
processes.

- Hemipelagites
The hemipelagic type deposits (facies Mh) are massive (non stratified) mud
deposits; we interpret them as slow almost contiuous settling down from the upper part of
the water column.

-Facies Tuc
The only Tuc event of the MD01-2477 core represented in the figure IV.24 (b) was
deposited by processes of traction and fall-out from turbulent flows (Mutti et al., 1999;
Mulder and Alexander 2001). The depositional process was fast, avoiding the formation of
sedimentary structures and a good selection of the grains (Lowe, 1982; Mutti, 1992).
Moreover, the two classical turbidites studied by van Welden (2007) in the MD01-2479
core (figure IV.26), show a more complex and continuous evolution of transport and
depositional processes.

- Facies HmTu
In the MD01-2477 core, the HmTu encompasses two well separated units (Hm and
Tu), showing in between an intermediate unit with very specific characteristics. We
interpret that the basal coarse unit (Tu) was deposited by processes of traction and fall-out
(Passega, 1964; Lowe, 1982; Mutti et al., 1999). It shows a progressive transformation of
the transport mechanism of the sediments, clearly observed in CM diagram (figure IV.26).
The base is characterized by an intense near-bed transport of coarse particles (coarse and
medium sand) forming a traction load in a density flows (Passega, 1964, Mulder and
Alexander, 2001). In the middle and upper part of this unit, composed mostly by fine sand
and silt, the sediment are mainly transported in suspension by turbulent flows (Mutti et al.,
1999). The intermediate unit displays an almost invariable sorting, a moderately changing
median grain size (M between 10 to 20 µm), and a highly variable Q99 (C between 50 and

223
500 µm). Adding the presence of to-and-fro sedimentary structures, these data point out to
two different process of transport and deposition, which took place simultaneously, after
the rapid immobilization of the coarse fraction. First, the clayey silt (with minor fine to
medium sand) involved in the to-and-fro sedimentary structures were deposited by a
combination of bottom traction and fall-out from the base of a turbulent flow, the whole
submitted to periodic reversal. Second, the fine grain sediments (fine silt) were deposited
by a rapid decantation from high density suspension (HDS), as is proposed by Lamb et al.
(2004). These authors indicate that the HDS are the result from the remobilization of fine
sediments (mud) from the bottom of the basins, product of wave action; which generates a
high-density turbulent suspension.

The upper homogeneous unit (Hm) shows a highly constant distribution of the grain
size parameters. It shows a lower median grain size (M average 6.08 r 0.54) compared to
the intermediate unit described above. The textural parameters indicate a stabilization of
the energy of the depositional environment, dominated by fall-out from a long-lasting fine-
grained homogenous suspension (HS).

As was discussed in the subchapter 3.7.1 for the Marmara Sea, the depth of water
wherein were deposited the sediments of the MD01-2477 core (>800 m depth below the
sea level) doesnot allow the development of wave action. Therefore, the oscillating
component and the high segregation of fine-graind fraction result from a kind of “reflected
tsunami” (which may evolve into a seiche). The latter is produced by one (or several)
subaqueous lansdslides (i.e. mass wasting) evolving into density currents, the initial
landslides being generally earthquake-induced. The mobilization of the whole water mass
may be directly due to a tectonic seafloor modification (e.g. fault offset) and/or to a
landslide triggered by the same seismic event.

Both model are extensively explained and illustrated in the subchapter 3.7.2 and
figure III.34.

The magnetic fabric and the grain shape analyses corroborate the interpretations
made with the grain size parameter. The AMS foliation parameter (F) confirms the
difference between the lower (Tu) and the upper (Hm) units (see subchapter 2.4.2 and
4.7.2). This diference is of our interest, because although both show similar grain size and

224
shape parameters, the magnetic fabric (specifically the F parameter) indicate different grain
array, which can be interpreted as two different depositional conditions. We consider the
hemipelagic-type deposits’ low magnetic foliation (“normal” with respect to burying
depth) as a primary sedimentary fabric (initial settling without later removing). In contrast,
the strong planar horizontal array in the Hm unit fits with the hypothesis of a specific
settling (Chapron et al, 1999; Beck, 2009 and references therein, Campos et al., 2013),
where the re-suspended fine fraction is deposited by fall-out under oscillatory conditions
(associated to “seiche effect”), favoring a better arrangement of the particles, mostly with
platy shape, in planes parallel to the bedding surface.

225
4.11. Late Quaternary paleoseismic sedimentary archive from deep central
Gulf of Corinth: time distribution of inferred earthquake-induced layers

After visual identification and characterization of the HmTu events along the MD01-
2477 core, we attempted to deunitine the temporal distribution of such events, in order to
propose an average time span between these inferred earthquake-induced HmTu events
(recurrence interval). This subchapter corresponds to the following article:
“Late Quaternary paleoseismic sedimentary archive from deep central Gulf of Corinth:
time distribution of inferred earthquake-induced layers”, by C. Campos, et al..
Special Issue “Earthquake Geology: Science, Society and Critical Facilities”(S. Barba
Edr.). ANNALS OF GEOPHYSICS, 56, 6, 2013, S0670; doi: 10.4401 / ag-6626.

226
ANNALS OF GEOPHYSICS, 56, 6, 2013, S0670; doi:10.4401/ag-6226

Special Issue: Earthquake geology

Late Quaternary paleoseismic sedimentary archive from deep


central Gulf of Corinth: time distribution of inferred
earthquake-induced layers
Corina Campos1,2,*, Christian Beck2, Christian Crouzet2, Eduardo Carrillo3,
Aurélien Van Welden4,2, Efthymios Tripsanas5

1 Universidad Simón Bolívar, Departamento de Ciencias de la Tierra, Sartenejas, Baruta, Venezuela


2 Grenoble-Alpes University, Laboratoire ISTerre, UMR CNRS 5275, CISM, Le Bourget du Lac, France
3 Universidad Central de Venezuela, Instituto de Ciencias de la Tierra, Caracas, Venezuela
4 Fugro Survey AS, Skøyen, Oslo, Norway
5 Institute of Oceanography, Hellenic Centre for Marine Research (HCMR), Anavyssos, Greece

Article history
Received November 16, 2012; accepted December 20, 2013.
Subject classification:
Homogenite, Paleoseismicity, Anisotropy of magnetic susceptibility, Corinth.

ABSTRACT
Marco and Agnon 1995, Levi et al. 2006], ii) slope fail-
A sedimentary archive corresponding to the last 17 cal kyr BP has been ures and gravity redeposition, associated, or not, to
studied by means of a giant piston core retrieved on board R/V MAR- tsunamis or seiche effects. Our investigations concern
ION-DUFRESNE in the North Central Gulf of Corinth. Based on previ- the second case.
ous methodological improvements, grain-size distribution and Magnetic Mass wastings (triggered, or not, by seismic
Susceptibility Anisotropy (MSA) have been analysed in order to detect shocks) evolving into gravity flowing (density/turbid-
earthquake-induced deposits. We indentified 36 specific layers -Homogen- ity currents) may provide large areas for spreading and
ites+Turbidites (HmTu) - intercalated within continuous hemipelagic- produce “classical” turbidites (with the differents terms
type sediments (biogenic or bio-induced fraction and fine-grained defined by Bouma [1962] or Mutti and Ricci Lucchi
siliciclastic fraction). The whole succession is divided into a non-marine [1978]; Piper and Normark [2009]). Conversely, when
lower half and a marine upper half. The “events” are distributed through occurring in more confined (or restricted) basins (lakes
the entire core and they are composed of two terms: a coarse-grained or isolated marine basins) re-depositional processes
lower term and an upper homogeneous fine-grained term, sharply sep- may result into specific complex layers, due to reflec-
arated. Their average time recurrence interval could be estimated for tions on steep slopes and/or to oscillations of the
the entire MD01-2477 core. The non-marine and the marine sections whole water mass (reflected tsunami, seiche effect).
yielded close estimated values for event recurrence times of around 400 More often, a significant seismo-tectonic activity is re-
yrs to 500 yrs. sponsible for the second type of sedimentary “event”,
the reason for which it is particularly searched for pa-
1. Introduction leoseismic purpose and contribution to seismic hazard
Following different studies achieved in tectonically estimation.
active areas, lake and restricted marine basins have The main characteristics of these sedimentary
demonstrated their potential for the sedimentary events are: i) a sharp limit at the top of a coarse graded
recording of seismic shocks [e.g. Hempton and Dewey lower part, ii) a structureless highly homogenous fine-
1983, Siegenthaler et al. 1987, Van Loon et al. 1995, grained upper part [Sturm et al. 1995, Siegenthaler et
Syvitski and Schafer 1996, Mörner 1996, Chapron et al. al 1987, Chapron et al. 1999, Beck et al 2007]. Addi-
1999, Shiki et al. 2000, De Batist et al. 2002]. Two major tionally, a mixed term-indicating to-and-fro particle dis-
mechanisms may account for this recording: i) in situ placements-is often observed in between; it may be
disturbances (micro-fracturing, micro-folding; lique- directly visible trough X-ray imaging [Beck et al. 2007]
faction, injection; [e.g. Rodriguez-Pascua et al. 2000, or be detected with high-resolution grain-size evolu-

S0670
CAMPOS ET AL.

tion (which is proposed for the here-presented exam- tribution along a long core from the Gulf of Corinth
ples). The basal layer is coarser-grained and may vary [Moretti et al. 2004, Lykousis et al. 2007].
from very thin silty laminae to thick sandy/gravelly
normal graded layer as the lower term of a turbidite 2. Geological setting
sensu Bouma [1962] and Mutti & Ricci Lucchi [1978]; it The Aegean region is one of the most active exten-
may consist of several pulses with an overall grain-size sional continental regions around the world [McKen-
decrease. The upper unit is a fine-grained homoge- zie1972, 1978, Le Pichon et al. 1981]. It is moving up to 30
neous term, named “homogenite” (Hm) in reference mm/yr to the southwest with respect to Eurasia and pro-
to the concept of Kastens and Cita [1981]. After the dis- gressively migrates to the south [McClusky et al. 2000,
covery of the Mediterranean tsunami-induced fine- Jolivet 2001]. Within this framework, the Gulf of Corinth,
grained “homogenite” [Kastens and Cita 1981, Cita et a semi-enclosed marine basin located in Central Greece
al. 1996], similar homogeneous layers have been re- (Figure 1), represents one of the most recent extensional
ported in lakes and closed marines basins and have features in the area. It is an active half-graben and is
been associated to subaqueous earthquake-effects bounded by E-W striking, en echelon faults located on-
[Sturm et al. 1995, Chapron et al. 1999, Beck et al. shore and offshore [Brooks and Ferentinos 1984, Armijo
2007, Bertrand et al. 2008, Carrillo et al. 2008, Çağatay et al. 1996, Sakellariou et al. 2001]. The focal mechanism
et al. 2012]. This interpretation (especially the impor- solutions indicate a main direction of extension N-S
tance of a long-lasting suspension settling under oscil- [McKenzie 1978, Jackson et al. 1982, Bernard et al. 2006].
latory conditions at the beginning) has been confirmed Geodetic data show that the active extension is focused
by recent observations made immediately after major offshore, increasing from 11 mm/yr in the central part of
earthquakes in the Cariaco Trough [Thunell et al. the rift (near Xylocastro) to 16 mm/yr in its western part
1999] and north-west Haiti [McHugh et al. 2011]. They (near Aigion) [Avallone et al. 2004]. The gulf has a maxi-
indicate the presence of a long-lasting “cloud” of fine mum uplift rate of 1.3 mm/yr in the southemargin
particles that remained in suspension for several [Armijo et al. 1996]. It is 115 km long, ∼30 km wide, with
weeks, and which is responsible for the deposition of a 900 m maximum depth (Figure 1). It is characterized by
homogeneous layers. Thus, in this work we will focus steep north and south dipping slopes (8°-25°) [Stefatos et
on the occurrences of such associations homogen- al. 2002]. To the west, the gulf is separated from the open
ite+turbidite (named here HmTu) and their time-dis- marine Ionian Sea by the Rion-Antirion strait (2 km wide

22 00 22 30 23 00

Itea 20 Km
Rion
Antirion

200 400

Kar-
600
W-ANT F.
parelli
Aigion

MD01-2477
GULF O
F CORIN
TH
Akrata
Co
20 30
rin th F
Xylocastro F. . 800
Apu

Eurasian Plate
North Anatolian Fault
Xylocastro
lian
Pla

40 25 mm/yr
te

Anatolian Plate Megara


Aegean
Sea
ULF
M

Corinth OS G
ed

ONIC
ite

He SAR
rra

30 mm/yr
Arabian Plate

lle
a Fault
nea

n ic 2500 m
n

35 Faults
R

Trench
id
ge

Dead Se

Major towns
10 mm/yr
200
Water Depth (m)
0m
African Plate Core location

22 00 22 30 23 00

Figure 1. Simplified geodynamic setting of the Gulf of Corinth with the location of the MD01-2477 core. The locations of major faults and the
bathymetry are from Moretti et al. [2003] (200 m isobaths intervals); onshore relief is an extract from NASA/SRTM. Motion vectors of the North
Anatolian fault after Le Pichon et al. [1995], and of the Aegean and African plates relative to Eurasia after McClusky et al. [2000].

2
PALEOSEISMIC RECORD OF GULF OF CORINTH

and 65 m deep). To the east, it is linked with the Saronikos was applied to selected ones. Two of them (blue rec-
Gulf and the western Aegean Sea through an artificially- tangle on Figure 2) are presented and discussed here: i)
dredged channel (the Corinth Canal), which is 6.3 km the thickest one, which permitted a high resolution
long, 21 m wide and 8 m deep. analysis with respect to the thickness, ii) a thinner one
more representative of the whole “events”.
3. Methodology
The paleoseismic record in the Gulf of Corinth 3.1 Grain size and shape analysis
was studied using the MD01-2477 long piston core, re- The grain size-analyses were conducted using a
covered during a R/V MARION-DUFRESNES cruise laser diffraction microgranulometer MALVERNTM
as part of GEOSCIENCES-II program, carried out in Mastersizer 2000, which has a 0.02 to 2000 m range.
October 2001. This Calypso piston core was taken in Measurements were performed in the ISTerre labora-
the central part of the Gulf of Corinth at 38°13.28’N, tory, Savoie University, France, with sampling intervals
22°33.53’E and 867 m water depth. The core is 20.08 m varying from 0.5 to 2 cm. The distribution parameters
long (Figure 2), and it is supposed to represent the last have been calculated following Folk and Ward [1957]:
25.3 cal. kyr BP of sedimentation [Moretti et al. 2004, skewness index (Sk), sorting index (So), kurtosis. Sk
Lykousis et al. 2007]. Shipboard processing included and So were plotted on simple binary diagrams. Per-
GEOTEK core logger profiles and splitting; further lab- centiles 99 (Q99=C) and medians (Q50=M) were plot-
oratory samplings and measurements were done on ted on the CM Passega [1964] diagram. Both diagrams
half cores horizontally stored at 4°C. were used, as proposed in Beck [2009], to display and
To characterize and discriminate homogenites analyse bottom-to-top paths. Shape analyses were per-
from other fine-grained deposits, analysis of the tex- formed on silt-clay fraction using a Sysmex FPIA-2100
tural parameters is the most common tool [Chapron et size and shape particle image analyzer, with a sampling
al. 1999, Beck et al. 2007, Carrillo et al. 2008, Bertrand interval of ∼ 50 cm.
et al. 2008], especially those depending on particles ar-
rays. The combination of grain size, undrained shear 3.2 Magnetic properties
stress, and magnetic fabric has been used to get a well- High-resolution bulk magnetic susceptibility data
constrained characterization of the homogenites and (Figure 2) was collected using a BartingtonTM MS2
their temporal distribution [Chapron et al. 1999, Car- contact sensor every 5 mm. The magnetic susceptibil-
rillo et al. 2008, Bertrand et al. 2008, Beck 2009]. Recent ity anisotropy (MSA) samples were collected into 8
works show the importance of the application of the cm3 non-magnetic plastic boxes and a total of 263
Magnetic Susceptibility Anisotropy (MSA) in the study samples were analysed. The measurements were car-
and identification of post-depositional disturbances re- ried out in the CEREGE Magnetism Laboratory (Aix
lated to earthquake triggering [Levi et al. 2006, Mörner - Marseille University, France) using the MFK1-FA
and Sun 2008]. When previously settled deposits in- Kappabridge of AGICO (spinning specimen method).
clude hemipelagites, the fine-grained resuspended frac- For each sample, the system reconstructs the AMS
tion (incipient homogenite) may be very similar tensor ellipsoid, defined by three eigenvectors (kmax,
(grain-size and composition) to these hemipelagites. kint and kmin) [Hrouda 1982, Tarling and Hrouda
MSA appears to be a useful tool to establish this dis- 1993]. This susceptibility ellipsoid represents the com-
tinction [Campos et al. 2011, Campos et al. 2013], and, bined result of the susceptibility anisotropy produced
thus, precisely measure the thickness of a hemipelagic by individual grain shape and/or crystallography in a
interval. The latter represents the time elapsed between sample [Joseph et al. 1998]. The shape of the MSA el-
two successive reworking events; its duration may thus lipsoid is illustrated by several parameters defined by
be considered as an earthquake recurrence time inter- Jelinek [1981], such as L = kmax/kint and F = kint/kmin.
val [Adams 1990, Gorsline et al. 2000, Goldfinger et al. Only the major events (6 to 74 cm thick) were
2003, Huh 2004]. sampled for textural (every 5 mm) and MSA (every 20
The here-used age/depth curve is based on this as- mm) studies; two of them were selected (A and B in-
sumption: sedimentation rates estimated from tervals, position on Figure 2) as representative of the
hemipelagic thicknesses, zero-time applied to ho- whole set. Figure 3 and 4, respectively, present the re-
mogenites+turbidites (HmTu) “events” [e.g. Goldfin- sults for a 74 cm thick event (30 cm thick basal Tu
ger 2009, McCalpin 2009 and references therein, Beck et term and 44 cm thick upper Hm term) and a 17 cm-
al. 2012] thick event (8 cm thick basal Tu term and 9 cm thick
After visual identification of specific sedimentary upper Hm term). For the first event, a total of 185
“events” (HmTu), a detailed textural characterization measurements were performed (148 for grain-size dis-

3
CAMPOS ET AL.

tribution and 37 for MSA parameters). diocarbon Laboratory (Poland). The measures were
performed on plant debris and wood fragments from
3.3 Chronology coarse to fine turbidites (7 measures, Table 1) and par-
Two sets of Accelerator Mass Spectroscopy 14C ticulate organic matter collected from organic rich
measurements are used: 1) published preliminary dat- muds (1 measure). Radiocarbon ages were calibrated
ing performed by Moretti et al. [2004] and Lykousis et using OxCal-v4 software [Bronk Ramsey 2001]. As the
al. [2007] in BETA ANALYTIC Laboratory (USA). 2) samples were taken in subaquatic environments
A new set of radiocarbon ages obtained at Poznan Ra- (marines and non-marine), we use the Marine09 cali-

4
PALEOSEISMIC RECORD OF GULF OF CORINTH

1030
1.2 1.6 2 2.4 2.8 3.2 3.6 us to make0.8more precise
1.2 1.6 corrections.
2 2.4 Results are sum-
1 1.04 1.08 1.12
Sorting Kurtosis
marized in table in figure 2. They
1 and represented Foliation
1040
h.d
correspond to 95% of confidence (2σ).

1050
4. Results
1060
4.1 Sedimentological observations and sedimentary
Hm
1070 events recognized in the MD01-2477 core
Depth (cm)

The MD01-2477 sediment infill is Holocene-Late


1080
HmTu
Pleistocene and it represents both marine and non-ma- HmTu

rine environments. The marine - non-marine transition


1090
is considered to have occurred between 12 kyr BP
1100 [Perissoratis et al. 2000, Collier et al. 2000] and 13.2 kyr
BP [Moretti et al. 2004, Lykousis et al. 2007]. The Tu
stud-
1110
ied core is composed of olive grey (5Y5/2) to dark yel-
lowish brown (10YR4/2) calcareous muds interlayering
1120
with millimetric to decimetric layers of yellowish
0 100 200 300 400 500 -0.6 -0.4 -0.2 0 0.2 1 1.04 1.08 1.12
Mean µm brown (10YR4/2) sand and silt.Lineation
Skewness In the non-marine sec-
HmTu.: re-depositional event
tion, few centimetres of laminated mud (olive grey,
Homogenite (Hm)
a h.d.: Hemipelagic deposits (homogenite + turbidite)
dusky yellow, greyish Turbidite (Tu)
yellow and brownish black (5Y4/1,
5Y6/4, 5Y8/4, 5YR2/1)) are present. These layers have
0.2 Hemipelagic deposits
Homogenite
been interpreted as varves by Moretti et al. [2004] and
10000
Intermediate term Lykousis et al. [2007].
Turbidite (1092 - 1097)
Turbidite (1097 - 1102) Microscopic observations of coarser sediments
Rolling
0 Turbidite (1102 - 1107)
Turbidite (1107 - 1112)
(binocular)
2000 and fine-grained smear slides show sedi-
Bottom suspension
Rolling
B
T
Turbidite (1112 - 1116,5) ments principally composed of terrigenous compo-
Bottom (B) to top (T) 1000
HmTu nents (carbonates and siliciclastics), being the carbonates
Skewness

C=Q99 (µm)

Graded suspension
500
-0.2 3 the most abundant. Aragonite crystal (abundant in the
top of the non-marine
Pelagic suspension
section) and traces of micas,
2
glauconite (marine section), pyrite (predominant in the
100 6
non-marine section) and flakes of organic matter are
=M

-0.4 70
C

50
present. Biogenic and bio-induced particles mainly
1 compose the silty/clayey fraction (nannofossil,
-0.6 foraminifers,
10
diatoms and bio-induced micrite). In this
core all granulometric fractions have angular to
1.2 1.6 2 2.4 2.8 3.2 3.6 1 10 100 1000
b Sorting
rounded
c grains with medium M=Median sphericity.
(µm)

Figure 3. Textural characterization of Homogenite+Turbidite (HmTu) B (1035-1115 cm; location on Figure 2). a) main granulometric pa-
rameters and MSA values; b) Sk index vs. So index bottom-to-top path; c) bottom-to-top path on Passega [1964]’s diagram. All textural pa-
rameters show a sharp break between turbiditic lower part and overlying homogenite.

bration curve [Reimer et al. 2009], and the local ma- In the MD01-2477 core, two types of deposits
rine reservoir correction (∆R) proposed by Reimer linked to density currents and reworking processes have
and McCormac [2002] for the Aegean Sea. They pro- been identified, intercalated within the predominant
posed a ∆R =35±50 yr for Nafplio, Greece (at 70 km mud. These depsosits are:
from the Gulf of Corinth) and ∆R =143±41 yr for Pi- 1) Turbidites (Tu); they are present as 0.5 to 2 cm
raeus, Greece (at 30 km from the Gulf of Corinth). To thick layers, made of silt to very fine sand. They can
use a single value of ∆R we calculate the average show sharp (sometimes clearly erosive) bases, slight
value of both ∆R with their average errors: ∆R= normal grading typical of classical turbidites defined by
89±58. We applied this ∆R to the whole sedimentary Bouma [1962], and Mutti & Ricci Lucchi [1978]. Oth-
sequence (marine and non-marine conditions; erwise, they locally present inverse gradations;
Holocene and Late Pleistocene). This assumption is 2) Homogenite+Turbidite associations (HmTu):
probably simplistic, but the lack of data does not allow They are composed by two main terms (see introduc-

5
CAMPOS ET AL.

0.8 1 1.2 1.4 1.02 1.04 1.06


2 3 4
Kurtosis Foliation
Sorting
1005

h.d.
h.d
1010

1015
Depth (cm)

Hm
1020

HmTu
HmTu

1025

Tu
1030

0 50 100 150 200 250 -0.6 -0.4 -0.2 0 1 1.02 1.04 1.06
Mean µm Skewness Lineation

HmTu.: re-depositional event


a h.d.: Hemipelagic deposits
(homogenite + turbidite)
Turbidite (Tu) Homogenite (Hm)

0
Hemipelagic deposits 10000
Homogenite
Turbidite (top)
Turbidite (medium)
Bottom suspension
Turbidite (medium) Rolling
Turbidite (bottom) Graded
1000
C=Q99 (µm)

Turbidite (bottom) suspension


-0.2 Bottom to top HmTu Uniform
Skewness

suspension

Pelagic
suspension

-0.4 100
=M
C

-0.6 10

1 2 3 4 1 10 100 1000
Sorting M=Median (µm)
b c
Figure 4. Textural characterization of Homogenite+Turbidite (HmTu) A (1005-1030 cm; location on Figure 2). a) main granulometric pa-
rameters and MSA values; b) Sk index vs. So index bottom-to-top path; c) bottom-to-top path on Passega [1964]’s diagram.

tion) with a sharp contact. At this level, oscillation sed- cm thick) were identified through sedimentological ob-
imentary structures, that may be direcly eye-visible in servations. We will focus on the two selected events
between [Beck et al. 2007], were detected here through (see 3.2), detailing first the larger one (Event B, Figure
high resolution microgranulometric profiles and MSA 3) and then comparing the results with the smaller
following the methodology proposed by Campos et al. one’s (Event A, Figure 4). Other events, with interme-
2013. In the following paragraphs, these textural prop- diate thicknessses, are similar to the here-presented
erties - used as criteria to address relationships between ones, and are not detailed hereafter.
sedimentation and seismicity - are detailed. For Event B, The basal term (Tu) has a main grain
size which progressively decreases from 500 to 9 µm.
4.2 Textural properties of the HmTu events The parameters as sorting (2.364± 0.418), skewness (-
In the MD01-2477 core, 36 HmTu events (6 to 74 0.369 ± 0.120) and kurtosis (1.321± 0.340) are highly

6
PALEOSEISMIC RECORD OF GULF OF CORINTH

Depth in core (m) Age Age Calibrated Material


in core (m) (yr) error 14C age (2 σ)

2.63 3190 40 3071-2730 PDOM

3.9 3355 34 3301-2907 PDOM

5.16 5590 40 6082-5701 PDOM

6.5 12420 60 13990-13493 PDOM

11.07 8500 50 9251-8763 PDOM

16.33 12470 60 14040-13614 PDOM

17.66 14000 100 16946-15987 PDOM

20.08a 21133 110 25005-24300 POM

Table 1. Radiocarbon dating from the MD01-2477 core. The sample denoted by the letter “a” was carried out by Lykousis et al. [2007]. The
14C ages were calibrated using OxCal software [Bronk Ramsey 2001], the Marine09 calibration curve [Reimer et al. 2009]; and applying the

local marine reservoir correction (∆R) proposed by Reimer and McCormac [2002] for Aegean Sea. Samples in italic overestimate the ex-
pected age. PDOM= plant debris and wood fragments extracted from fine and coarse turbidites; POM= particulate organic matter extracted
from organic rich muds.

variable. Their variability gradually decreases to the top we propose to call it “to-and-fro” interval.
of this unit (Figure 3a). Contrary to Tu, the upper Hm An additional remark concerns the Hm and the
term (44 cm thick) have parameters highly constant hemipelagites, which are not separated, especially on
(mean = 5.486± 0.470 µm, sorting = 2.758 ± 0.987, the CM diagram, and which may indicate similar com-
skewness = -0.150 ± 0.034 and kurtosis = 0.945 ± 0.035). position and slow settling from suspension.
As the Hm term, the above hemipelagic deposits show In the studied section (Figure 3a), the MSA meas-
almost constant parameters (mean = 5.860 ± 1.167 µm, urements show a low anisotropy (mostly oblate type el-
sorting = 2.346 ± 0.431, skewness = -0.066 ± 0.043 and lipsoid), excepted in the homogenite. The magnetic
kurtosis = 0.964 ± 0.027). The grain-size parameters do lineation is very low and relatively constant (L= 1.005 ±
not clearly separate the hemipelagic deposits from the 0.002). At the difference, magnetic foliation displays sig-
upper homogeneous unit; at the opposite, the limit is nificant variations and anomalously high values with
neat on magnetic foliation profile (see after). The bot- respect to expected compaction in the Hm term (F=
tom-to-top grain-size distribution displays a three-step 1.069 ± 0.005) compared to the basal Tu term (F= 1.032
evolution both on skewness/sorting and CM (Passega’s) ± 0.011). The hemipelagic deposits (F= 1.047 ± 0.009)
diagrams (Figure 3-b and 3-c): (1) Basal turbidite (dark appear significantly lower than for the Hm term, with
blue triangles). (2) Intermediate term (light blue trian- a neat limit. This difference has been previously under-
gles). (3) Homogenite (pink dots). Red arrows underline lined by different authors [Beck 2009, Campos et al.
detail evolutions within the three terms. The following 2013] in marine and lacustrine settings. These authors
has to be underlined: have shown magnetic foliation in Hm up to 1.090, in
- a neat separation between three steps: especially similar sediments with same age and depth in cores
between Hm and underlying terms well evidence on with the same coring device.
both diagrams; The second - thin-HmTu (Event A, Figure 4), has
- a neat separation between Tu and intermediate been analyzed following the same sampling intervals
term; and measurements as the first - thick - one. The general
- complex variations within the Tu term bottom-to-top evolution appears similar, although not
- the individualisation of the intermediate terms so complex within the lower term. The intermediate
with internal variations corresponding to oscillations; term is present; the fact of not displaying the “to-and-

7
CAMPOS ET AL.

fro” signature may be due either to its actual lack, or to sets. The here-obtained paths clearly resemble their
a too low sampling frequency unable to evidence it. “slump turbidite” path, but with a significant differ-
The characteristics displayed by the two here-pre- ence, the intercalation of the intermediate “to-and-
sented HmTu sedimentary “events” may be extended fro” term.
to the 36 HmTu occurring along the MD01-2477 core. The MSA foliation parameter (F) also confirms
the difference between the lower (Tu) and the upper
5. Discussion (Hm) terms; a major difference also appears between
the Hm term and the overlying hemipelagic deposit.
5.1 Interpretation of the sedimentary HmTu event In summary, the here-used textural characteriza-
The analyzed textural parameters show a clear tion points out a particular and complex depositional
difference between the two main terms which con- mechanism involving, for each event, density/turbid-
stitute the HmTu deposit. The basal coarse term ity current (mass wasting and fluidisation), to-and-fro
shows that this unit was deposit by processes of trac- bottom current (related to water mass oscillation),
tion and fall-out [Passega 1964, Lowe 1982, Mutti et and quiet long-lasting settling of fine-grained ho-
al. 1999]. This term in the CM diagram (Figure 3c, 4c) mogenous suspension. Seismic shocks and/or sub-
shows a progressive transformation of the transport aqueous landslides in closed basins may account for
mechanism of the sediments. The base is character- the whole process, especially the whole water column
ized by an intense near-bed transport of coarse parti- movement (constrained tsunami or seiche effect).
cles (coarse sand) forming a traction load, in a density This interpretation has been proposed for different
flow [Passega 1964, Mulder and Alexander 2001]. In marine or lacustrine basins developed in areas with
the middle and upper part of this term, composed frequent and strong seismic activity [Sturm et al.
mostly by fine sand and silt, the sediments are mainly 1995, Chapron et .al. 1999, Beck et al. 2007, 2012,
transported in suspension by turbulent flows [Mutti Bertrand et al. 2008, Carrillo et al. 2008, Çağatay et
et al. 1999]. Furthermore, the high variability in the al. 2012].
skewness and sorting (Figure 3b) indicate fluctuations
in the depositional dynamic. Additionally, the pres- 5.2 Triggering mechanism of sedimentary HmTu
ence of two fining-upward sequences within the events
same basal unit could reflect different pulses of the HmTu associations were found throughout the
gravity current, as described by Shiki et al. [2000] and whole MD01-2477 (Figure 2). These units, resulting
Nakajima and Kanai [2000] in seismo-turbidites from from gravity re-depositional processes, represent a
Japan. specific evolution of slope failures and mass-wasting.
The upper homogeneous term shows a highly The Gulf of Corinth is well known for the numerous
constant distribution of the textural parameters (Fig- submarines gravitational mass movements as slides,
ure 3b, c, Figure 4b, c), indicating a stabilization of slumps and debris/mud flows [Brooks and Ferentinos
the energy of the depositional environment, and fall- 1984, Papatheodorou and Ferentinos 1997, Hasiotis
out from a suspended load. Three interesting points et al. 2002, Lykousis et al. 2007]. In the Gulf, the main
have to be explained: the sharp separation of the Hm mechanisms responsible for these observed mass
term, its thickness and extreme homogeneity, and the movements are: (1) the frequent seismic activity in
intermediate “to-and-fro” term. As previously pro- the region, (2) the development of rapid prograding
posed [Chapron et al 1999, Beck et al 2007], we ex- prodeltas associated with high sedimentation rates of
plain the latter as fluctuations within a high-density the numerous river mouths, (3) the steep slopes and
suspension whose relative stability is due to lateral os- (4) the presence of gas-charged sediments [Hasiotis
cillatory displacements of the whole water column; et al. 2002, Lykousis et al. 2009].
coeval partial settling is inferred. This intermediate Additionally, historical observations in the Gulf
episode is considered responsible for an increased “ex- of Corinth give evidence of earthquake induced
traction” of the finer-grained fraction, which will be nearshore sediment failure. Several may be cited,
the major component of a long-lasting (almost sta- among others: the 373 BC, 1817 and 1861 events [Pa-
ble) suspension. Based on base-to-top evolutions on padopoulos 2003 and references therein], the 1965
the two types of binary diagrams, Arnaud et al. [2002] event [Ambraseys 1967], the 1981, 1989 and 1995
and Lignier [2001], in: Beck [2009] identified, in a lake events [Perissoratis et al. 1984, Papadopoulos 2003,
infill, two different paths deciphering “flood tur- Ambraseys and Synolakis 2010]. Beside, few aseismic
bidites” from “slump turbidites”, respectively related submarines gravitational slide have been reported as
to tributary flooding and to collapse of deltaic fore- the 1963 and 1996 events [Galanopoulos et al. 1964,

8
PALEOSEISMIC RECORD OF GULF OF CORINTH

Papadopoulos 2003]. Despite the presence of sedi- MD01-2477 core. The last 2.32m (17.75 to 20.07 m
ment failure of aseismic origin, earthquakes remain depth) are made of a homogeneous mud displaying
the main cause of submarine failure. In the Gulf of highly constant MS, MSA, and grain size distribution.
Corinth the earthquakes and the sediment failures of This basal section does not show deformation or sed-
seismic or aseismic origen can generate high ampli- imentary structures. Due to its characteristics (high
tude tsunami waves [Dominey-Howes 2002, Pa- homogeneity and absence of a coarser sediments
padopoulos 2003, Ambraseys and Synolakis 2010]. layer), this section could represent the upper term
These authors have reported a high frequency oscil- (Hm) of a major HmTu association whose lower part
lation (amplitude of 20-30 cm) immediately after the was not retrieved. The measured low values of MSA
1981 earthquake, with a gradual attenuation during may be related either to the low content of magnetic
four days after the earthquake [Papadopoulos 2003]. minerals in this interval (low MS values) or to an ac-
In this study, the gravity-reworked deposits tual weak magnetic foliation. For these reasons, this
HmTu present in the MD01-2477 core occured far- 2.32 m thick interval is not taken in account for the
ther back in time, when no historical data are avail- statistical distribution.
able. Therefore, it is not possible to correlate them In the MD01-2477 core, the magnetic susceptibil-
with the documented historical earthquakes. Despite ity curve and the occurrence of the reworking events
this, several arguments led us to consider HmTu de- (Tu, HmTu) are different between the marine and
sposits as earthquake-induced: (1) the drilling site lo- non-marine succession (Figure 2). In the non-marine
cation in the basin plain, avoiding coarser fluxes at section the values of MS are highly variable compared
tributary river mouths, (2) the high seismic activity in to the marine section. Higher values of MS are pres-
the region, (3) the well known earthquake-induced ent in the non-marine section (>1000 S.I.x10-5). The
sediment failures during the last 2000 yrs (4) and the reworking events are thinner in the non-marine sec-
similarities of the HmTu deposits with previous doc- tion, compared to the marine one. But the HmTu de-
umented earthquake-triggered homogenites+tur- posits in both environments represent more or les the
bidites deposits in marine and lacustrine same fraction of the total sedimentation (∼32.5%).
environments. In the non-marine section the total mean sedi-
The specific planar array in the homogenites mentation rate (hemipelagic + instantaneous de-
(high magnetic foliation values), the sharp contacts posits) is about 0.68 mm/yr, whereas the hemipelagic
within HmTu, and the sedimentary structures typi- deposits have a sedimentation rate between 0.31 and
cal of to-and-fro currents fit with the hypothesis of 0.82 mm/yr. In the marine section, the total main rate
major earthquakes inducing coeval gravity reworking of sedimentation is higher: about 1.1 mm/yr, show-
and seiche effects [Chapron et al. 1999, Beck 2009 and ing rates of sedimentation in the hemipelagic deposits
references therein, Campos et al. 2013]. In these cases, between 0.32 and 3.68 mm/yr.
the seiche effect is responsible for: (1) the increase of In the non-marine section the total mean sedi-
suspended load by the additional extraction of the mentation rate (hemipelagic + instantaneous de-
clayey-silt matrix from the initial flux, (2) the genera- posits) is about 0.68 mm/yr, varying from 0.31 and
tion of a sharp contact between the two HmTu main 0.82 mm/yr (using hemipelagic deposits only). In the
terms by oscillatory bottom currents, (3) and the cre- marine section, the mean sedimentation rate is about
ation of specific settling conditions, producing the 1.1 mm/yr, varying from 0.32 and 3.68 mm/yr.
specific planar array (phyllosilicates homogeneous This difference in the rates of sedimentation be-
orientation) in the homogenites. tween the marine and non-marine section can be ex-
plained by a climatic influence. The non-marine
6. Time distribution of the sedimentary “events” section was developed in a lowstand phase during the
On Figure 5, we represent the vertical and tem- last glacial period (70 to 12 kyr BP) [Collier et al. 2000,
poral distribution of all HmTu events recognized in Lykousis et al. 2007, Bell et al. 2008] characterized, in
the MD01-2477 core, and the sedimentation rates for the Mediterranean realm, by a dry and cold climate,
the hemipelagic deposits (blue colour). The age vs. with absence of significant tree cover (presence of
depth graph was constructed using six radiocarbon steppe vegetation) [Leeder et al. 1998, Collier et al.
ages, assuming a constant sedimentation rate in be- 2000]. These climatic conditions are inferred to favour
tween, which is most probably oversimplified. The higher mechanical erosion and subsequent higher ter-
mean rates of sedimentation - and the estimation of rigenous material production. However, due to lower
the main HmTu recurrence time interval - were esti- transport capacity of the rivers (associated to the dry
mated for the first 17.75m (0 to 17 kyr BP) of the conditions), these sediments were probably accumu-

9
CAMPOS ET AL.

Age cal yr BP
0 2000 4000 6000 8000 10000 12000 14000 16000 18000

0
Legend
S.R: 0.73 mm/yr Cal. AMS 14C age
S.R: Sedimentation rate
HmTu: Homogenite (Hm)+(Tu) Turbidite
263 cm = 3064 - 2734 cal yr BP
300 S.R: 3.68 mm/yr
390 cm = 3280 - 2913 cal yr BP

S.R: 0.32 mm/yr


516 cm = 6052 - 5706 cal yr BP

600
Depth cm

S.R: 0.93 mm/yr


900

1107 cm = 9239 - 8779 cal yr BP

1200

S.R: 0.82 mm/yr

1500

1633 cm = 14038 - 13618 cal yr BP

S.R: 0.31 mm/yr


1766 cm = 16928 - 16218
1800 cal yr BP

Figure 5. Proposed time distribution of HmTu events based on 14C dating. Deduced sedimentation rate is applied to “hemipelagic” constant
supply, without the supposed instantaneous deposits.

lated in the fluvial system or in the margin of the 36 HmTu events are distributed through the
basin as deltas. Otherwise, during the transgressive to MD01-2477 core, corresponding to the last 17 cal kyr
highstand phases associated to the Holocene glacio- BP. At the top of the core (0 to 2 m depth) and at the
eustasy (12 to 0 kyr BP) [Collier et al. 2000, Lykousis base of the marine section (11.2 to 12.8 m depth) no
et al. 2007, Bell et al. 2008], these sediments were re- HmTu were detected, although abundant thin tur-
mobilized and transported from the fluvial systems to bidites (0.5 to 1.5 cm thick) are present. The later ones
the margin of the basin. It provides a greater amount are not thick enough to apply the here-used textural
of sediments available to be transported to the deep analyses, and no clear evidences for seismic origin can
basin by seismic destabilization, compared to the last be assumed.
glacial period. Similar conditions were reported for To define the recurrence time interval between
post-glacial lakes in the French Alps and the Andes two HmTu events, an age/depth curve (Figure 5) was
[Beck et al. 1996, Carrillo et al. 2008]. Despite this, constructed. In the 17 to 11.7 cal kyr BP period (non-
rapid tectonic uplift, especially the southern side of marine section and in the marine and non-marine
the Gulf of Corinth (see Paragraph 2), has also con- transition), 13 events were identified, representing a
trolled the sediment supply during the whole Quater- ∼400 yr average recurrence interval. In the marine
nary. This basin combined uniform regional uplift of section (11 cal kyr BP to Present), 22 events were rec-
0.3 mm yr-1 [Collier et al. 1992] with 1.3 mm yr-1 south ognized, representing a ∼500 yr average recurrence in-
footwall uplift [Armijo et al. 1996]. terval. According to chronological precision, the

10
PALEOSEISMIC RECORD OF GULF OF CORINTH

difference between the sections is not significant. part of the Gulf of Corinth. Among used criteria to dis-
These recurrence intervals are compatible (same cuss depositional processes (instantaneous and contin-
order) with, or slightly larger than those proposed for uous) MSA appeared to be a useful tool to identify the
the last 2 kyr period concerning the Aeghion, Skinos instantaneous homogeneous deposits (Hm), and to dis-
and Eliki Faults using palaeoseismological trenches. tinguish them from normal hemipelagic deposits. This
The later ones displayed respective recurrence inter- textural parameter also permits to precisely measure
vals of 360 yr [Pantosti et al. 2004], 330 yr [Collier et the time-equivalent of the hemipelagic interval which
al. 1998], and 200 to 600 yr [McNeill et al. 2005]. separates two events inferred as co-seismic (HmTu).
Using the HmTu events may both represent: In this core, 36 earthquake-induced HmTu (ho-
- an over-estimation of the number of recorded mogenite associated to turbidite) were detected, inter-
earthquakes, as part of them may be only due to sub- calated within a continuous sedimentary recording.
aqueous lanslides. Some historical examples occurred They can be explained by the combination of: 1) a con-
on delta fronts on the southern coast with tsunami im- stant seismo-tectonic activity in the Gulf of Corinth
pact on the northern coast. In recently retrieved short during the Late Pleistocene-Holocene. 2) terrigenous
cores [Mortier 2012, Beckers et al. 2013] one of these sedimentary feeding of shelves and upper slopes dur-
events [1963] displays a clear signature; ing the same time.
- an under-estimation; first, we did not include clas- These HmTu events have a minimum average re-
sical turbidites in our list; second, some homogenites currence interval between 17 to 11.7 cal kyr BP of ∼400
may be not detected when a basal coarser layer is lack- yr; whereas, the marine sequence, show a minimum av-
ing. As described in the Sea of Marmara [Beck et al. erage recurrence interval between of ∼500 yr. A 6.0 to
2007, Eriş et al. 2012] or in Lake Le Bourget [Chapron 7.0 Mw approximate magnitude is considered as re-
et al. 1999] a homogenite may just have a very thin sponsible for these sedimentary events.
laminae of silt/very fine sand at its base, only visible In order to improve these paleoseismological stud-
on a high resolution X-ray picture; the homogenite it- ies, future investigations should include in situ geot-
self would possibly be characterized with detailed pro- echnical measurements to assess slope failure potential
file of MSA. These types of events have not yet been [Strasser et al. 2006, Stegmann et al. 2007]. A search of
investigated here. a method to discriminate earthquake-related HmTu
With respect to intensities or magnitude of in- and HmTu only due to landslides is also needed.
volved paleo-earthquake, previous works [e.g. Aude-
mard and De Santis 1991, Rodrı́guez-Pascua et al.2000, Acknowledgements. The here-presented investigations were
McCalpin 2009 and references therein, Rodríguez-Pas- funded through different grants and facilities: i) University of
Savoie’s AAP-2012-16 AGRASM Grant, dedicated to ASM sedimen-
cua et al., 2010] indicate that a Mw > 5 or 5.5 earth- tological application, ii) CNRS-INSU funding through ISTerre Lab-
quake is necessary to produce noticeable geologic oratory. iii) CNRS-INSU ARTEMIS national radiocarbon AMS
effects, especially pore water pressure sudden increase measurement program. These investigations are also associated to
(driving force for liquefaction, failures, injections, French Agence Nationale pour la Recherche (ANR)’s SISCOR
Project dedicated to seismic hazards related in the Gulf of Corinth
etc.). Additionally, the compilation of historical data rifting.C. Campos’s PhD thesis and stay in ISTerre Laboratory is
in the Gulf of Corinth [e.g. Ambraseys 1967, Pa- funded through Venezuela’s FUNDAYACUCHO Grant N°
padopoulos 2003, Ambraseys and Synolakis 2010 and 20093262. We also acknowledge INQUA for the grant which per-
references therein] indicates, for the last 2.3 kyr, sur- mits her attend the 2nd INQUA-ICGP 567 International Workshop
on Active Tectonics, Earthquake Geology, Archaeology and Engi-
face wave magnitudes comprised between Ms= 7.0 neering. Authors are grateful to these different institutes and gov-
and Ms=6.0. Most of them have generated submarine ernments, and to two anonymous reviewers who greatly helped to
landslides and tsunamis. Based on all data, a precise improve the initial manuscript.
estimate of paleo-magnitudes remains speculative;
Mw> 6 and reaching 7 may be proposed as a realistic References
estimate. Adams, J. (1990). Paleoseismicity of the Cascadia Sub-
duction Zone: Evidence from turbidites off the Ore-
7. Conclusion gon-Washington Margin, Tectonics, 9(4), 569-583.
Based on high resolution analyses of layering, Ambraseys, N. (1967). The earthquakes of 1965–66 in
composition, and, overall, textures (magnetic suscepti- the Peloponnesus, Greece; a field report, Bulletin of
bility, and Magnetic Susceptibility Anisotropy, grain-size the Seismological Society of America, 57, 1025-
distribution), the sedimentary succession sampled 1046.
through the MD01-2477 core is considered as a paleo- Ambraseys, N. and C. Synolakis (2010). Tsunami Cata-
seismic archive for the last 17 cal kyr BP, for the central logs for the Eastern Mediterranean, Revisited. Jour-

11
CAMPOS ET AL.

nal of Earthquake Engineering, 14(3), 309-330. K. De Rycker, P. Bascou, and D. Strivay (2013). High
Armijo, R., B. Meyer, G. C. P. King, A. Rigo and D. Pa- energy environment offshore deposits in the west-
panastassiou (1996). Quaternary evolution of the ern Gulf of Corinth, Greece, Geophysical Research
Corinth Rift and its implications for the Late Ceno- Abstracts, Vol. 15, EGU2013-4877, EGU General As-
zoic evolution of the Aegean, Geophysical Journal sembly 2013.
International, 126(1), 11-53. Bell, R. E., L. C. McNeill, M. J. Bull and T. J. Henstock
Arnaud, F., V. Lignier, M. Revel, M. Desmet, C. Beck, (2008). Evolution of the offshore western Gulf of
M. Pourchet and F. Charlet (2002). Flood and earth- Corinth, Geological Society of America Bulletin,
quake disturbance of 210Pb geochronology (Lake 120(1-2), 156-178.
Anterne, NW Alps), Terra Nova, 14(4), 225-232. Bernard, P., H. Lyon-Caen, P. Briole, A. Deschamps, F.
Audemard, F. A. and F. De Santis (1991). Survey of liq- Boudin, K. Makropoulos, P. Papadimitriou, F.
uefaction structures induced by recent moderate Lemeille, G. Patau, H. Billiris, D. Paradissis, H. Cas-
earthquakes, Bulletin of the International Associa- tarède, O. Charade, A. Nercessian, D. Avallone, F.
tion of Engineering Geologists, 44, 5-16. Pachiani, J. Zahradnik, S. Sacks and A. Linde (2006).
Avallone, A., Briole, P., Agatza-Balodimou, A. M., Bil- Seismicity, deformation and hazard in the western
liris, H., Charade, O., Mitsakaki, C., Nercessian, A., rift of Corinth: New insights from the Corinth Rift
Papazissi, K., D. Paradissis and G. Veis (2004). Analy- Laboratory (CRL), Tectonophysics, 426, 7-30.
sis of eleven years of deformation measured by GPS Bertrand, S., F. Charlet, E. Chapron, N. Fageland and
in the Corinth Rift Laboratory area. Comptes Ren- M. De Batist (2008). Reconstruction of the
dus Geosci. 336, 301-311. Holocene seismotectonic activity of the Southern
Beck, C. (2009). Late Quaternary lacustrine paleo-seis- Andes from seismites recorded in Lago Icalma,
mic archives in north-western Alps: Examples of Chile, 39°S, Palaeogeography, Palaeoclimatology,
earthquake-origin assessment of sedimentary dis- Palaeoecology, 259(2-3), 301-322.
turbances, Earth-Science Reviews, 96(4), 327-344. Bouma, A.H. (1962). Sedimentology of Some Flysch
Beck, C., F. D. Manalt, E. Chapron, P. V. A. N. Rensber- Deposits. Elsevier Publishing, 168 p.
gen and M. D. E. Batist (1996). Enhanced Seismicity Bronk Ramsey, C (2001). Development of the Radio-
in the Early Post-Glacial Period: Evidence from the carbon calibration program OxCal, Radiocarbon,
Post-Würm Sediments of lake Annecy, Northwest- Proceedings of 17th International 14C Conference,
ern Alps. Journal of Geodynamics, 22(112), 155-171. 43, 355-363.
Beck, C., B. Mercier de Lépinay, J.-L. Schneider, M. Cre- Brooks, M. and G. Ferentinos (1984). Tectonics and sed-
mer, N. Ça atay, E. Wendenbaum, S. Boutareaud, G. imentation in the Gulf of Corinth and the Zakyn-
Ménot, S. Schmidt, O. Weber, K. Eris, R. Armijo, B. thos and Kefallinia channels, Western Greece,
Meyer, N. Pondard, M.-A. Gutscher, and the MAR- Tectonophysics, 101(1-2), 25-54.
MACORE Cruise Party J.-L. Turon, L. Labeyrie, E. Çağatay, N., L. Erel, L. G. Bellucci, A. Polonia, L.
Cortijo, Y. Gallet, H. Bouquerel, N. Gorur, A. Ger- Gasperini, K. K. Eri , Ü. Sancar, D. Biltekin, G.
vais, M.-H. Castera, L. Londeix, A. de Rességuier, Uçarku , U. B. Ülgen and E. Damcı (2012). Sedi-
A. Jaouen (2007). Late Quaternary co-seismic sedi- mentary earthquake records in the zmit Gulf, Sea
mentation in the Sea of Marmara’s deep basins, in of Marmara, Turkey, Sedimentary Geology, 282,
Sedimentary Records of Catastrophic Events F. 347–359.
Bourrouilh-Le Jan, C. Beck, D. Gorsline (Editors) Campos, C., C. Beck, C. Crouzet and E. Carrillo (2011).
Spec. Iss., Sedimentary Geology, 199, 65–89. Characterization of Late Pleistocene-Holocene earth-
Beck, C., J.-L. Reyss, F. Leclerc, E. Moreno, N Feuillet quake-induced “homogenites” in the Sea of Marmara
and GWADASEIS Cruise Scientific Party: L. Barrier, through magnetic fabric. Implications for co-seismic
F. Beauducel, G. Boudon, V. Clément, C. Deplus, N. offsets detection and measurements, Proceedings 2nd
Gallou, J.-F. Lebrun, A. Le Friant, A. Nercessian, M. INQUA-IGCP-567 International Workshop on Active
Paterne, J.-M. Saurel, T. Pichot, C. Vidal (2012). Tectonics, Earthquake Geology, Archaeology and En-
Identification of deep subaqueous co-seismic scarps gineering, Corinth, Greece, 19-21.
through specific coeval sedimentation, in Lesser An- Campos, C., C. Beck, C. Crouzet, D. François, A. van
tilles: implication for seismic hazard D. Pantosti (Ed- Welden, and K. Eris. (2013) Deciphering
itor), Natural Hazards and Earth System Sc., Special hemipelagites from homogenites through Magnetic
Issue: Subaqueous Paleoseismology, 12, 1755-1767. Susceptibility Anisotropy. Paleoseismic implications
Beckers, A., C. Mortier, C. Beck, A. Hubert-Ferrari, J.- (Sea of Marmara and Gulf of Corinth), Sedimen-
L. Reyss, E. Tripsanas, D. Sakellariou, M. De Batist, tary Geology, 292, 1-14.

12
PALEOSEISMIC RECORD OF GULF OF CORINTH

Carrillo, E., C. Beck, F. A. Audemard, E. Moreno and in motion without shock, Annales Géologiques des
R. Ollarves (2008). Disentangling Late Quaternary Pays. Helléniques, 16, 93-110.
climatic and seismo-tectonic controls on Lake Mu- Goldfinger, C., C. H. Nelson and J. E. Johnson (2003).
cubají sedimentation (Mérida Andes, Venezuela), Holocene Earthquake records from the Cascadia
Palaeogeography, Palaeoclimatology, Palaeoecol- subduction zone and northern San Andreas Fault
ogy, 259(2-3), 284-300. based on precise dating of offshore turbidites, An-
Chapron, E., C. Beck, M. Pourchet and J.-F. Deconinck nual Review of Earth and Planetary Sciences, 31(1),
(1999). 1822 earthquake-triggered homogenite in 555-577.
Lake Le Bourget (NW Alps), Terra Nova, 11(2-3), Goldfinger, C. (2009). Subaqueous Paleoseismology, in
86-92. Paleoseismology J. McCalpin (Editor), 2nd edition,
Cita, M. B., A. Camerlenghi and B. Rimoldi (1996). Elsevier, 119-169.
Deep-sea tsunami deposits in the eastern Mediter- Gorsline, D. S., T. D. Diego and E. H. Nava-Sanchez
ranean: New evidence and depositional models, (2000). Seismically triggered turbidites in small mar-
Sedimentary Geology, 104(1-4), 155-173. gin basins: Alfonso Basin, Western Gulf of Califor-
Collier, R., M. R. Leeder, P. J. Rowe and T. C. Atkinson nia and Santa Monica Basin, California Borderland,
(1992). Rates of tectonic uplift in the Corinth and Science, 135, 21-35.
Megara Basins, central Greece, Tectonics, 11(6), Hasiotis, T., G. Papatheodorou, G. Bouckovalas, C.
1159-1167. Corbau and G. Ferentinos (2002). Earthquake-in-
Collier, R., D. Pantosti, G. D’Addezio, P. M. De Mar- duced coastal sediment instabilities in the western
tini, E. Masana and D. Sakellariou (1998). Paleoseis- Gulf of Corinth, Greece, Marine Geology, 186(3-4),
micity of the 1981 Corinth earthquake fault: Seismic 319-335.
contribution to extensional strain in central Greece Hempton, M. R. and J. F. Dewey (1983). Earthquake-in-
and implications for seismic hazard, Journal of Geo- duced deformational structures in young lacustrine
physical Research, 103(B12), 30001-30019. sediments, East Anatolian Fault, southeast Turkey,
Collier, R., M. R. Leeder, M. Trout, G. Ferentinos, E. Tectonophysics, 98(3-4), 7-14.
Lyberis and G. Papatheodorou (2000). High sedi- Hrouda, F. E. K. (1982). Magnetic anisotropy of rocks
ment yields and cool, wet winters: Test of last gla- and its application, Surveys in Geophysics, 5, 37-82.
cial paleoclimates in the northern Mediterranean, Huh, C.-A. (2004). Linkages between turbidites in the
Geology, 28(11), 999-1002. southern Okinawa Trough and submarine earth-
De Batist, M., J. Klerkx, Y. Imbo, A. Giralt, V. Lignier, quakes, Geophysical Research Letters, 31(12), 2-5.
C. Beck and D. Delvaux. (2002). Bathymetry and Jackson, J. A., J. Gagnepain, G. Houseman, G. C. P.
sedimentary environments of Lake Issyk-Kul, Kyr- King, P. Papadimitriou, C. Soufleris and J. Virieux
gyz Republic (Central Asia): a large, high-altitude, (1982). Seismicity, normal faulting and the geomor-
tectonic lake, in Lake Issyk-Kul: its natural environ- phological development of the Gulf of Corinth
ment J. Klerkx and B. Imanackunov (Editors), (Greece): the Corinth earthquakes of February and
NATO Science Series IV, Kluwer Academic Publ, 13, March 1981, Earth and Planetary Science Letters,
101-124. 57(2), 377-397.
Dominey-Howes, D. (2002). Documentary and Geo- Jelinek, V. (1981). Characterization of the magnetic fab-
logical Records of Tsunamis in the Aegean Sea Re- ric of rocks, Tectonophysics, 79(3-4), 63-67.
gion of Greece and their Potential Value to Risk Jolivet, L. (2001). A comparison of geodetic and finite
Assessment and Disaster Management, Natural strain pattern in the Aegean, geodynamic implica-
Hazards, 25(3), 195-224. tions, Earth and Planetary Science Letters, Volume
187(1-2), 95-104.
Eri , K., N. Çağatay by K. Eriş, N. Çağatay, B. Mercier Joseph, L. H., D. K. Rea and B. A.Van der Pluijm (1998).
de Lepinay and C. Campos (2012). Late-Pleistocene Use of Grain Size and Magnetic Fabric Analyses to
to Holocene sedimentary fills of the Çınarcık Basin Distinguish Among Depositional Environments, Pa-
of the Sea of Marmara, Sedimentary Geology, 281, leoceanography, 13(5), 491-501.
151-165. Kastens, K. and M. B. Cita (1981). Tsunami-induced
Folk R. L. and W. C. Ward (1957). Brazos River bar: a sediment transport in the abyssal Mediterranean
study in the significance of grain size parameters, Sea, Geological Society of America Bulletin, 92(11),
Journal of Sedimentary Petrology, 27, 3-26. 845-857.
Galanopoulos, A. G., P. Comninakis and N. Delibasis Le Pichon, X., J. Angelier, M. F. Osmaston and L. Stegena
(1964). A tsunami generated by an earth slump set (1981). The Aegean Sea [and Discussion], Philosoph-

13
CAMPOS ET AL.

ical Transactions of the Royal Society of London, Se- McKenzie, D. (1978). Active tectonics of the Alpine-Hi-
ries A, Math. Phys. Sc., 300(1454), 357-372. malayan belt: the Aegean Sea and surrounding re-
Le Pichon, X., N. Chamot-Rooke, S. Lallemant, R. gions, Geophys. J. Royal Astr. Soc., 55(1), 217-254.
Noomen and G. Vets (1995). Geodetic determination McNeill, L. C., R. Collier, P. M. De Martini, D. Pantosti
of the kinematics of central Greece with respect to and G. D’Addezio (2005). Recent history of the
Europe: Implications for eastern Mediterranean tec- Eastern Eliki Fault, Gulf of Corinth: geomorphol-
tonics, Journal of Geophysical Research, 100(B7), ogy, palaeoseismology and impact on palaeoenvi-
12675-12690. ronments, Geophysical Journal International,
Leeder, M. R., T. Harris and M. J. Kirkby (1998). Sedi- 161(1), 154-166.
ment supply and climate change: implications for Moretti, I., D. Sakellariou, V. Lykousis and L. Micarelli
basin stratigraphy, Basin Research, 10(1), 7-18. (2003). The Gulf of Corinth: an active half graben?,
Levi, T., R. Weinberger, T. Aïfa, Y. Eyal and S. Marco Journal of Geodynamics, 36(1-2), 323-340.
(2006). Earthquake-induced clastic dikes detected by Moretti, I., V. Lykousis, D. Sakellariou, J.-Y. Reynaud, B.
anisotropy of magnetic susceptibility, Geology, Benziane and A. Prinzhoffer (2004). Sedimentation
34(2), 69-72. and subsidence rate in the Gulf of Corinth: what we
Lignier, V. (2001). Les sédiments lacustres et l'enreg- learn from the Marion Dufresne’s long-piston coring,
istrement de la palæoséismicité. Etude comparative Comptes-Rendus Geosciences, 336(4-5), 291-299.
de différents cas dans le Quaternaire des Alpes Mörner, N-A. (1996). Liquefaction and varve deforma-
Nord-Occidentales et du Tien-Shan Kyrghize, Un- tion as evidence of paleoseismic events and
published Doct, Thèse, Université de Savoie, 2 vol., tsunamis. The autumn 10,430 BP case in Sweden,
330 pp. Quaternary Science Review, 15(8), 939-948.
Lowe, D. R. (1982). Sediment gravity flows; II, Deposi- Mörner, N-A. and G. Sun (2008). Paleoearthquake de-
tional models with special reference to the deposits formations recorded by magnetic variables, Earth
of high-density turbidity currents, Journal of Sedi- and Planetary Science Letters, 267(3-4), 495-502.
mentary Research, 52(1), 279-297. Mortier, C. (2012). Etude sédimentologique dans la par-
Lykousis, V., D. Sakellariou, I. Moretti and H. Kaberi tie occidentale du Golfe de Corinthe (Grèce). MSc.
(2007). Late Quaternary basin evolution of the Gulf Memoir, University of Liège, 129 p.
of Corinth: Sequence stratigraphy, sedimentation, Mulder, T. and J. Alexander (2001). The physical char-
fault–slip and subsidence rates, Tectonophysics, acter of subaqueous sedimentary density flows and
440(1-4), 29-51. their deposits, Sedimentology, 48(2), 269-299.
Lykousis, V., G. Roussakis and D. Sakellariou (2009). Mutti E. and F. Ricci Lucchi (1978). Turbidites of the
Slope failures and stability analysis of shallow water northern Apennines: introduction to facies analy-
prodeltas in the active margins of Western Greece, sis, American Geological Institute, Reprint Series
northeastern Mediterranean Sea, International Jour- 3, 127-166.
nal of Earth Sciences, 98(4), 807-822. Mutti, E., R. Tinterri, E. Remacha, N. Mavilla, S. An-
Marco, S. and A. Agnon (1995). Prehistoric earthquake gella and L. Fava (1999). An introduction to the
deformations near Masada, Dead-Sea graben, Ge- analysis of ancient turbidite basins from an outcrop
ology, 23(8), 695-698. perspective: American Association of Petroleum
McCalpin, J. (2009). Paleoseismology, 2nd edition, El- Geology, Continuing Education Course Note Series,
sevier, 119-169. 39, 95.
McClusky, S., S. Balassanian, A. Barka, C. Demir, S. Nakajima, T. and Y. Kanai (2000). Sedimentary features
Ergintav, I. Georgiev and O. Gurkan (2000). Global of seismoturbidites triggered by the 1983 and older
Positioning System constraints on plate kinematics historical earthquakes in the eastern margin of the
and dynamics in the eastern Mediterranean and Japan Sea, Sedimentary Geology, 135(1-4), 1-19.
Caucasus, Journal of Geophysical Research, Pantosti, D., P. M. De Martini, I. Koukouvelas, L. Stam-
105(B3), 5695-5719. atopoulos, N. Palyvos, S. Pucci and F. Lemeille
McHugh, C. M., L. Seeber, N. Braudy, M.-H. Cormier, (2004). Palaeoseismological investigations of the Ai-
M. B. Davis, J. B. Diebold and N. Dieudonne (2011). gion Fault (Gulf of Corinth, Greece), Comptes-Ren-
Offshore sedimentary effects of the 12 January 2010 dus Geosciences, 336(4-5), 335-342.
Haiti earthquake, Geology, 39(8), 723-726. Papadopoulos, G. A. (2003). Tsunami Hazard in the
McKenzie, D. (1972). Active Tectonics of the Mediter- Eastern Mediterranean: Strong Earthquakes and
ranean Region, Geophysical Journal of the Royal Tsunamis in the Corinth Gulf, Central Greece, Nat-
Astronomical Society, 30(2), 109-185. ural Hazards, 29(3), 437-464.

14
PALEOSEISMIC RECORD OF GULF OF CORINTH

Papatheodorou, G. and G. Ferentinos (1997). Subma- ogy, 135, 7-9.


rine and coastal sediment failure triggered by the Siegenthaler, C., W. Finguer, K. Kelts and W. Suming
1995, M, = 6.1 R Aegion earthquake, Gulf of (1987). Earthquake and seiche deposits in Lake
Corinth, Greece, Holocene, Marine Geology, Lucerne, Switzerland, Ecloae Geologiae Helvetae,
137(3), 287-304. 80(1), 241-260.
Passega, R. (1964). Grain Size Representation by CM Stefatos, A., G. Papatheodorou, G. Ferentinos, M.
Patterns as a Geological Tool, SEPM, Journal of Leeder, and R. Collier (2002). Seismic reflection im-
Sedimentary Research, Vol. 34(4), 830-847. aging of active offshore faults in the Gulf of
Perissoratis, C., D. Mitropoulos and I. Angelopoulos Corinth: their seismotectonic significance, Basin Re-
(1984). The role of earthquakes in inducing sedi- search, 14(4), 487-502.
ment mass movements in the eastern Corinthiakos Stegmann, S., M. Strasser, F.S. Anselmetti and A. Kopf
Gulf: an example from the February 24–March 4 ac- (2007). Geotechnical in situ
tivity, Marine Geology, 55, 35-45. characterisation of subaquatic slopes: The role of pore
Perissoratis, C., D. J. Piper and V. Lykousis (2000). Al- pressure transients versus frictional strength in land-
ternating marine and lacustrine sedimentation dur- slide initiation, Geophysical Research Letters, 34,
ing late Quaternary in the Gulf of Corinth rift basin, L07607.
central Greece, Marine Geology, 167(3-4), 391-411. Strasser, M., F.S Anselmetti, D. Fäh, D. Giardini and M.
Piper, D. J. W. and W. R. Normark (2009). Processes Schnellmann (2006). Magnitudes and source areas
That Initiate Turbidity Currents and Their Influence of large prehistoric northern Alpine earthquakes re-
on Turbidites: A Marine Geology Perspective, Jour- vealed by slope failures in lakes, Geology, 34 (12),
nal of Sedimentary Research, 79(6), 347-362. 1005-1008.
Reimer, P. J and F. G. McCormac (2002). Marine radio- Sturm, M., C. Siegenthaler and R.A. Pickrill (1995). Tur-
carbon reservoir corrections for the Mediterranean bidites and “homogenites”: a conceptual model of
and Aegean Seas, Radiocarbon, 44, 159-166. flood and slide deposits, Publication of the Interna-
Reimer, P. J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. tional Association of Sedimentologists, 16th Re-
Beck, P. G. Blackwell, C. Bronk Ramsey, C. E. Buck, gional Meeting, Paris, 22:40.
G. S. Burr, R. L. Edwards, M. Friedrich, P. M. Grootes, Syvitski, J. P. M and C. T. Schafer (1996). Evidence for
T. P. Guilderson, I. Hajdas, T. J. Heaton, A. G. Hogg, an earthquake-triggered basin collapse in Saguenay
K. A. Hughen, K. F. Kaiser, B. Kromer, F. G. McCor- Fjord, Canada, Sedimentary Geology, 104(1-4),
mac, S. W. Manning, R. W. Reimer, D. A. Richards, J. 127-153.
R. S. Southon Talamo, C. S. M. Turney, J. van der Tarling, D. H and F. E. K. Hrouda (1993). The Magnetic
Plicht and C. E. Weyhenmeyer (2009). IntCal09 and Anisotropy of Rocks, Chapman & Hall, London,
Marine09 radiocarbon age calibration curves, 0- 218 p.
50,000 years cal BP, Radiocarbon, 51(4), 1111-1150. Thunell, R., E. Tappa, R. Varela, M. Llano, Y. Astor, F.
Rodríguez-Pascua, M. A., V. H. Garduño-Monroy, I. Is- Muller-Karger and R. Bohrer (1999). Increased ma-
rade-Alcántara and R. Pérez-López (2010). Estima- rine sediment suspension and fluxes following an
tion of the paleoepicentral area from the spatial earthquake, Nature, 398, 233-236.
gradient of deformation in lacustrine seismites Van Loon, A. J., K. Brodzikowski and T. Zielinski
(Tierras Blancas Basin, Mexico), Quaternary Inter- (1995). Shock-induced resuspension deposits from a
national, 219(1-2), 66-78. Pleistocene proglacial lake (Kleszczow Graben, cen-
Rodríguez-Pascua, M. A., J. P. Calvo, G. De Vicente and tral Poland), Journal of Sedimentary Research,
D. Gómez-Gras (2000). Soft-sediment deformation 65(2), 417-422.
structures interpreted as seismites in lacustrine sed-
iments of the Prebetic Zone, SE Spain and their po-
tential use as indicators of earthquake magnitudes
during the Late Miocene, Sedimentary Geology,
135(1-4), 117-135.
Sakellariou, D., V. Lykousis and D. Papanikolaou (2001).
Active faulting in the Gulf of Corinth, Greece, F. *Corresponding author: Corina Campos,
Universidad Simón Bolívar, Departamento de Ciencias de la
Briand (Editor), 36th CIESM Congress, 43.
Tierra, Sartenejas, Baruta, Venezuela; email: campossc@usb.ve
Shiki, T., M. B. Cita and D. S. Gorsline (2000). Sedi-
mentary features of seismites, seismo-turbidites and © 2013 by the Istituto Nazionale di Geofisica e Vulcanologia. All
tsunamiites - an introduction, Sedimentary Geol- rights reserved.

15
4.12. Conclusions of chapter

In the present subchapter, through the study of three long cores, and the integration of
data from previous works we attempted to answer two major questions.
-First, we tried to precise the transition between the non-marine and marine environment
during the Late Quaternary sea level change, by meam of its impact on sedimentation
(dynamics of transport and deposition, sediment composition);
-Second, we attempted to detect the impact of the seismo-tectonic activity (related to the
Corinth rifting) on the sedimentary fill in the deep central Gulf.
Both questions converge to a main target: to develop paleosismic archives, adding
informations to the previous contributions done by deifferent teams.

The age-depth curve constructed for the MD01-2477 core indicates that the
sedimentary record recovered in this core represents the last 17 cal kyr BP. In this core the
Magnetic Susceptibility curve show changes in the concentration of magnetic minerals in both
environments, the marine section shows highly constants MS values, whereas the non-marine
section shows higher variability (including the highest values among the entire core).
Aragonite crystals were observed in two kinds of accurrences: 1) several layers of thinly
laminated calcareous mud (between 13.55 to 16.61 m depth), 2) a unique 2 cm thick layer
containing almost 70% of aragonite crystals (at 13.55 m depth). We thus infer the
predominance of non-marine environment below 13.55 m depth-in-core. On the other hand,
the evolution of OM concentration and several elements underline a gradual transition rather
than an abrupt change. Anoxic conditions, with a slightly higher presence of carbonates
prevailed during the non-marine episode, while, oxic conditions with a slight increase of
terrigenous contribution dominated the marine environments. According to the age-depth
curve proposed in the present work, the change between marine and non-marine conditions
may have occurred between 11.2 and 12.2 cal kyr BP.

The marine environment shows variable sedimentations rates and slightly higher (0.32
to 3.68 mm/yr) than for the non-marine environments (0.31 and 0.82 mm/yr). This difference
may be explained by climatic influence. The non-marine section was developed during the
last glacial period (70 to 12 kyr BP) characterized, in the Mediterranean realm, by a dry and
cold climate, with absence of significant tree cover, which favour the mechanical erosion and
subsequent a higher production of terrigenous material (Leeder et al., 1998; Collier et al.,

242
2000). However, the lower transport capacity of the rivers associated to the dry conditions,
favors the accumulation of sediments in the fluvial system or deltas. During the Holocene
glacio-eustasy (12 to 0 kyr BP) sediments were remobilized and transported from the fluvial
systems and incorporated to the margin of the basin (mainly the coarse sediments) or in the
water column. It provides a greater amount of sediments available to be transported to deep
basin by seismic destabilization.

Despite the slight increase of available terrigenous materials to be transported during


the marine conditions, MD01-2477 doesnot indicate a marked difference in the conditions of
transport and deposition in both environments. A continuous record of instantaneous events,
(earthquake-induced or not) is observed along the entire core, resulting of: 1) constant seismo-
tectonic activity in, and around, the Gulf of during the Late Pleistocene-Holocene; 2) an
important amount of sediments easily transportable from the margins to the deep basin trough
gravity reworking.

Along the entire core, 36 HmTu events were characterized. In units of paloseismic
archive, this total has to be considered as:
- either a minimum, considering that some earthquakes may be not recorded, due to low
landsliding potential (see Strasser et al., 2006; Stegmann et al., 2007);
- or a maximum taking in account the occurrence of large landslides inducing tsunamis, and
for which no seismic shock has been reported.
These HmTu events show a minimum average recurrence interval of a400 yr in the non-
marine sequence, and a minimum average recurrence interval of a500 yr in the marine one.

Based on visual description and high resolution textural analyses (grain size, shape
and fabric) were identified different transport and depositional mechanisms contributing to the
sedimentary infilling of the deep central Gulf of Corinth.

In the MD01-2481 core, mud-clast breccias have been identified, corresponding to the
mega slump identified on seismic profile (Moretti et al., 2004; Fig. IV.7). Our textural
analyses (AMS) precise the reworking features. These types of deposits are well know in the
Gulf of Corinth (Brooks and Ferentinos, 1984; Papatheodorou and Ferentinos, 1997; Hasiotis
et al., 2002; Lykousis et al., 2007). Four main mechanisms may be responsible - separated or
combined - for these observed mass movements: (1) the frequent significant seismic activity

243
in the region, (2) the development of rapidly prograding prodeltas associated with high
sedimentation rates of the numerous river mouths, (3) the steep slopes; and (4) the presence of
gas-charged sediments.

On the other hand, in the MD01-2477 core were identified instantaneous (or “event”)
deposits such as: classical turbidites and the asociation HmTu, not detectable at the seismic
imegnig scale. They represent turbiditic gravity flows with different driving processes and
different final settling. The visual observation, as well as, the skewness vs sorting and the CM
Passega (1964) diagrams of the HmTu events demonstrate the presence of two well
differentiated units (Hm and Tu), generally showing in the middle of both an intermediate
unit with to-and-fro sedimentary structures. We consider these sedimentary structures as
associated to oscillating currents driven by reflected tsunamis with possible “seiche effect”.
As previously proposed for the Sea of Marmara (Chapter III) seismic shocks and/or
subaqueous landslides are inferred triggering events.

Despite different depositional processes, Hm unit and hemipelagic-type sediments


deposits show similar grain size distribution and shape parameters. Only the magnetic fabric
reflects a neat difference between them, indicating different grain arrays, and, by mean,
different settling conditions. The low magnetic foliation present in the hemipelagites is
characteristic of a primary sedimentary fabric. On the other hand, the strong planar horizontal
array in the Hm unit confirm the hypothesis of a specific settling, where the re-suspended fine
fraction is deposited by fall-out under oscillatory conditions (“seiche effect”).

From the compositional point of view a slight difference can be seen between the Hm
unit and the hemipelagic deposits. From a qualitative point of view, all components are
identical; on the other hand, terrigenous clays are slightly more abundant in the Hm units,
whereas bio-induced precipited carbonates are slightly more abundant in the hemipelagites.
This difference can be explained, for the Hm units:
- either by a shallower provenance of fine terrigenous material transported by turbidity
currents,
- or by a deep in situ origin (same initial composition as hemipelagites with possible addition
of turbidites’content during the fluidization of a previous accumulation) and segregation
processes during re-suspension and re-concentration.

244
CHAPTER 5

Conclusions
Comparison of the two investigated sites.

The Late Quaternary sedimentary archives of the Sea of Marmara and of the central
Gulf of Corinth were investigated through high resolution sedimentological core analyses,
aiming to extract paleoseismic informations. We focused our work on the characterization of
specific deposits considered as related to (moderate to strong) local/regional seismicity. Our
interpretations are proposed taking into account the evolution of depositional environment
(climate, salinity, etc.), and we attempted to provide a reliable chronology.

The age-depth curve constructed for the Sea of Marmara (MD01-2425 core) and the
Gulf of Corinth (MD01-2477 core) indicates that the sedimentary record recovered by these
cores represents the last 17 cal kyr BP. In both regions the vertical distribution of facies, the
concentration of magnetic minerals and the biogenic/bio-induced content led to define a major
subdivision which coincides with the last major environmental changes: end of the LGM and
transition to the Holocene, and change from non-marine to marine environment (figure V.1).
In the Sea of Marmara, this limit is marked by an abrupt change in the biological, lithological
and magnetic (MS) contents, as well as by a progressive change in the chemical elements
ditribution. We propose that the last change in the environmental conditions of this basin
occurred around 12.8 cal kyr BP. On the other hand, in the Gulf of Corinth the studied
elements around the marine non-marine passage vary more gradually and the change seems to
have occurred between 11.2 and 12.2 cal kyr BP (table V.1).

In the Sea of Marmara and in Gulf of Corinth, the tentative identification of


earthquake-related traces recorded on the studied cores was based on visual, compositional,
sedimentological and magnetic analysis (MS and AMS). Among all used tools, two of them
demonstrated their usefulness to identify the shock-induced “Homogenite+Turbidite” (HmTu)
deposits: 1) the textural analysis, specifically, the bottom-to-top textural evolution figured out
by the path on Sorting vs. Skewness diagram and on Passega (1964)’s binary diagram; and 2)
the Anisotropy of Magnetic Susceptibility. Both tools demonstrated their efficiency to

245
differentiate: 1) fine-grained deposits resulting from re-suspension and re-settling under
oscillatory conditions, and 2) non reworked (primary) hemipelagic-type deposits (marine or
non marine) resulting from a unique settling from planktonic production and variable amount
of terrigenous suspended load.

The sedimentation rate and the distribution of the HmTu-type composite layers along
the sedimentary record in both basins are different. In the Sea of Marmara, the HmTu events
are concentrated in the non-marine section (Late Pleistocene). During this period, coarse
event deposits are more frequent and thicker, inducing a high mean sedimentation rate. In the
marine section (Holocene) the sedimentation rates can be three times lower, and the HmTu
units have a negligible thickness and are scarce. In contrast, in the Gulf of Corinth the HmTu
units are distributed along the entire sedimentary record, although the sedimentation rate and
thickness of the event deposits are slightly higher in the marine section (figure V.1).

Both basins underwent similar climatic conditions during the Late Quaternary, thus we
propose that the difference in sedimentation rates and the record of the inferred earthquake-
induced events in both basins during the Late Quaternary was mainly controlled by:
- the tectonic context; in the Gulf of Corinth the continuous amount of sediments available
to be transported to the deep basin is controlled by the rapid tectonic uplift of the Gulf’s
bounding margins, especially on the south-eastern side (Armijo et al., 1996); climatic
fluctuations are slightly modulating this influence through erosional processes and
sediment transport. In contrast, the predominant strike-slip regime along the Sea of
Marmara results in slight uplift (Yaltirak et al., 2002; Paluska et al 1998) and low creation
of relief; therefore, climatic fluctuations have a predominant influence on sediment
supply;
- the water vertical density variations; in the Sea of Marmara, a mostly homogeneous water
density profile was established during the non-marine stage, favoring the development of
underflows (hyperpycnal flow) able to transport large amount of sediments to deep water,
through density and turbiditic flows. During the marine stage, the changes in the
circulation and vertical density stratification of the basin (highly stratified), promoted by
the connection with the Aegean and Black seas, result in a change in the pattern of
distribution of sediments in the basin. This stage favored the development of interflows
(mesopycnal flows), decreasing the volume of sediment transported directly to the deep

246
Figure V.1: Synthetics logs and MS profiles of Cores MD01-2477 (Gulf of Corinth) and
MD01-2425 (Sea of Marmara). The sedimentary record recovered by these cores represents the
last 17 cal kyr BP. In the Sea of Marmara, the HmTu events are concentrated in the non-marine
section (Late Pleistocene). While, in the Gulf of Corinth the HmTu events are distributed along
the entire sedimentary record.

247
basin floor. In the Gulf of Corinth, a less pronounced water density stratification resulted
in a more homogeneous distribution of sediments during the Late Quaternary, with a
higher probability of develop underflows currents.

We considered the record of tsunami-induced instantaneous events (HmTu) as


a reliable paleoseismic archive (see discussion about non earthquake-induced landslides) in
both basins. We thus used these results for an estimation of the average earthquake recurrence
interval; this estimation only concerns events able to trigger massive subaqueous landslides
and/or other seafloor abrupt modification. In the Çinarcik Basin of the Sea of Marmara, the
minimum average recurrence interval estimated for the non-marine section was a155 yr,
whereas, for the marine section was a 365 yr. While for the eastern part of the Gulf of Corinth
a minimum average recurrence interval of a400 yr for the non-marine section and a500 yr in
the marine one was deunitined (table V.1).

Likewise, in the Central Basin of the Sea of Marmara, the study of correlated HmTu
events at opposite sides of a segment of a pull-apart basin along the NAF allows the
estimation of the vertical co-seismic offset for a continuous period of 2 kyr (13 to 15 cal kyr
BP). Our study underlines the usefulness of this type of earthquake induced deposits to
estimate time distribution of major co-seismic offsets along active faults simultaneously with
the amounts of their vertical component. Specifically in the Central Basin, we interpreted that
the significant values of the vertical offset, up to 144 cm, imply a dominant vertical (normal)
throw, implying a dominant transtensional tectonic regime. The hypothesis that this fault
segment is a dominantly strike slip fault was discarded because our important values of
vertical offsets would result in anomalously high lateral offset. This makes us favor the
hypothesis of a dominant normal behavior for this local branch of the NAF, favoring the
models that proposed that the Sea of Marmara evolve as a pull-apart basin along a segmented
NAF (e.g. Barka and Kadinsky-Cade, 1988; Westaway, 1994; Wong et al., 1995; Armijo et
al., 1999; 2002). The pull-apart complex geometry and the seismogenic behavior of the
Central Basin thus imply specific tsunami triggering, which have an implication regarding
hazards assessment in the Istanbul city area.

248
Sea of Marmara
Gulf of Corinth
(Çinarcik Basin)
Change from non-marine Between 11.2 and
Around 12.8 cal kyr BP
to marine enviroment 12.2 cal kyr BP
Earthquake recurrence interval
~ 400 yr ~155 yr
(non-marine enviroment)
Earthquake recurrence interval
~ 500 yr ~ 365 yr
(marine enviroment)

Table V.1: Earthquake recurrence intervals and estimate date for the last non-marine to marine
environmental change in the Gulf of Corinth and Sea of Marmara

To improve the present contribution to the paleoseismic record of the Sea of Marmara
and the Gulf of Corinth, we propose two groups of investigations:
- concerning sedimentary processes; it appears necessary to perform high resolution
Anisotropy of Magnetic Susceptibility profiles on the very thin silty-sandy laminas
(facies LS) and overlying and underlying fine-grained deposits; this could precisely
define the oscillatory mechanism and, may be, its time elapsing. On the other hand, a
high-resolution study of the evolution of the chemical composition along the entire
cores (by using X-ray fluorescence core scanner) will allow better understaning of the
major chemical changes occurred in the basin during the last non-marine to marine
passage;
- concerning the attribution of each reworking event to a specific fault segment activity,
this appears difficult using few localized long archives (e.g. MD cores in Çinarcik
Basin and Gulf of Corinth); nevertheless, they gave a broad appraisal of regional
instability. Conversely, the method developed for the Central Basin of the Sea of
Marmara should be applied to a maximum of similar sites. Each of them should be
dedicated to a selected fault segment and comprise a set of selected long coring, in a
way similar to the surveys performed, for the last century, with very short cores
(Armijo et al, 2005; Uçarcuú, 2010).

249
References

A
Adams, J., 1990. Paleoseismicity of the Cascadia Subduction Zone: Evidence from turbidites
off the Oregon-Washington Margin, Tectonics, 9 (4), 569-583.

Agostini, S., Doglioni, C., Fabrizio, I., Manetti, P., Tonarini, S., 2010. On the geodynamics of
the Aegean rift. Tectonophysics 488, Issues 1-4, 7-21.

Aksoy, M. E., Meghraoui, M., Vallee, M., Cakir, Z., 2010. Rupture characteristics of the A.D.
1912 Murefte (Ganos) earthquake segment of the North Anatolian fault (western
Turkey). Geology 38:991-994. doi:10.1130/G13447.1

Aksu, A. E., Hiscott, R. N., Yaúar, D., 1999. Oscillating Quaternary water levels of the
Marmara Sea and vigorous outflow into the Aegean Sea from the Marmara Sea -Black
Sea drainage corridor. Mar. Geol., 153, 275-302.

Aksu, A. E., Calon, T. J., Hiscott, R. N., 2000. Anatomy of the North Anatolian fault zone in
the Marmara Sea, western Turkey; extensional basins above a continental transform,
Geophysical Society of America Today, 10 (6), 3-7.

Aksu, A. E., Hiscott, R. N., Kaminski, M. A., Mudie, P. J., Gillespie, H., Abrajano, T., Yaúar,
D., 2002a. Last glacial-Holocene paleoceanography of the Black Sea and Marmara
Sea: stable isotopic, foraminiferal and coccolith evidence, Marine Geology 190, Issues
1-2, 119-149.

Aksu, A. E., Hiscott, R. N., Mudie, P. J., Rochon, A., Kaminski, M. A., Abrajano, T., Yaúar,
D., 2002b. Persistent Holocene outflow from the Black Sea to the eastern
Mediterranean contradicts Noah’s flood hypothesis. GSA Today, 12 (5), 4-10.

Aksu, A. E., Hall, J., Yaltirak, C., 2005. Miocene to Recent tectonic evolution of the Eastern
Mediterranean: new pieces of the old Mediterranean puzzle. Marine Geology 221, 1-
440.

Altinok, Y., Alpar, B., Ozer, N., Aykurt, H. 2011. Revision of the tsunami catalogue affecting
Turkish coasts and surrounding regions. Natural Hazards Earth System Science
11:273-291.

Altunel, E., Meghraoui, M., Akyüz, H. S., Dikbas, A., 2004. Characteristics of the 1912
coseismic rupture along the North Anatolian Fault Zone (Turkey): implications for the
expected Marmara earthquake, Terra Nova 16:198-204.

Ambraseys, N., 1967. The earthquakes of 1965-66 in the Peloponnesus, Greece; a field report,
Bulletin of the Seismologistsl Society of America, 57, 1025-1046.

250
Ambraseys, N. N., 1970, Some characteristic features of the North Anatolian fault:
Tectonophysics, v. 9, p. 143–165, doi: 10.1016/0040-1951(70)90014-4.

Ambraseys, N. N., 1988. Engineering seismology: Earthquake engineering and structural


dynamics. J. Int. Assoc. Earthquake Eng. 17, 1-105.

Ambraseys, N. N., 2002a. Seismic sea-waves in the Marmara Sea region during the last 20
centuries. J Seismol 6, 571-578.

Ambraseys, N. N., 2002b. The seismic activity of the Marmara Sea Region over the last 2000
years. Bull Seismol Soc Am 92, 1-18.

Ambraseys, N. N., 2009. Earthquakes in the Mediterranean and Middle East: A


Multidisciplinary Study of Seismicity up to 1900, 947 pp. Cambridge University
Press, Cambridge, U. K.

Ambraseys, N. N., Jackson, J. A., 1990. Seismicity and associated strain of central Greece
between 1890 and 1988. Geophys. J. Int 101, 663-708.

Ambraseys, N. N., Finkel, C. F., 1991. Long-unit seismicity of Istanbul and of the Marmara
Sea region. Terra Nova 3,527-539.

Ambraseys, N. N., Finkel, C. F., 1995. The Seismicity of Turkey and Adjacent Areas: A
Historical Review, 1500–1800: Istanbul. Muhittin Salih Eren, 240 pp.

Ambraseys, N. N., Jackson, J. A., 1997. Seismicity and strain in the Gulf of Corinth (Greece)
since 1694. J. Earthq. Engineer. 1, 433-474.

Ambraseys, N. N., Jackson, J. A., 2000. Seismicity of Sea of Marmara (Turkey) since 1500.
Geophys J Int 141, F1-F6.

Ambraseys, N., Synolakis, C., 2010. Tsunami Catalogs for the Eastern Mediterranean,
Revisited. Journal of Earthquake Engineering, 14(3), 309-330.

Anderson, J. J., Carmack, E. C., 1973. Some physical and chemical properties of the Gulf of
Corinth. Estuar. Coastal Shelf Sci. 1, 195-202.

Angelier, J., 1978. Tectonic evolution of the Hellenic Arc since the Late Miocene.
Tectonohysics, 49:23-36.

Armijo, R., Meyer, King, G. C. P., Rigo, A., Papanastassiou, D., 1996. Quaternary evolution
of the Corinth Rift and its implications for the Late Cenozoic evolution of the Aegean.
Geophysical Journal International, 126 (1), 11-53.

Armijo, R., Meyer, B., Hubert, A., Barka, A., 1999. Westward propagation of the north
Anatolian fault into the northern Aegean: timing and kinematics. Geology 27, 267-
270.

251
Armijo,R., Meyer, B.,Navarro, S., King, G., Barka, A., 2002. Asymmetric slip partitioning in
the Sea of Marmara pull-apart: a clue to propagation processes of the Anatolian Fault
?.Terra Nova 14, 80-86.

Armijo, R., Pondard, N., Meyer, B., Uçarkuú, G., de Lepinay, B.M.,Malavieille, J.,
Dominguez, S., Gustcher, M.A., Schmidt, S., Beck, C., Ça÷atay, N., Cakir, Z., Imren,
C., Eris, K., Natalin, B., Ozalaybey, S., Tolun, L., Lefevre, I., Seeber, L., Gasperini,
L., Rangin, C., Emre, O., Sarikavak, K., 2005. Submarine fault scarps in the Sea of
Marmara pull-apart (North Anatolian Fault): implications for seismic hazard in
Istanbul. Geochemistry Geophysics Geosystems 6, Q06009, 29,
doi:10.1029/2004GC000896

Arnaud, F., Lignier, V., Revel, M., Desmet, M., Beck, C., Pourchet, M., Charlet, F., 2002.
Flood and earthquake disturbance of 210Pb geochronology (Lake Anterne, NW Alps).
Terra Nova, 14(4), 225-232.

Audemard, F. A., De Santis, F., 1991. Survey of liquefaction structures induced by recent
moderate earthquakes. Bulletin, International Association of Engineering Geologists
44, 5-16.

Avallone, A., Briole, P., Agatza-Balodimou, A. M., Billiris, H., Charade, O., Mitsakaki, C.,
Nercessian, A., Papazissi, K., Paradissis, D., Veis, G., 2004. Analysis of eleven years
of deformation measured by GPS in the Corinth Rift Laboratory area. Comptes
Rendus Geosci. 336, 301-311.

Avúar, U., 2013. Lacustrine Paleoseismic Records from the North Anatolian Fault, Turkey.
PhD Thesis, Ghent University. Faculty of Sciences, 209 pp.

B
Bahr A, Lamy F, Arz H, Kuhlmann H, Wefer G. 2005. Late glacial to Holocene climate and
sedimentation history in the NW Black Sea. Marine Geology 214 (4), 309-22.

Barka, A. A., 1992. The North Anatolian fault zone (supplement VI). Ann. Tectonicae 1, 164-
195.

Barka, A., 1999. The 17 August 1999 Içzmit earthquake. Science 285, 1858-1859.

Barka, A., Kadinsky-Cade, K., 1988. Strike-slip fault geometry in Turkey and its influence on
earthquake activity. Tectonics 7, 663-684.

Barnes, P.M., Pondard, N., 2010. Derivation of direct on-fault submarine paleoearthquake
records from high-resolution seismic reflection profiles: Wairau Fault, New Zealand.
Geochemistry, Geophysics, Geosystems 11, doi:10.1029/2010GC003254.

Barrier, E., Chamot-Rooke, N., Giordano, G., 2004. Geodynamic Map of the Mediterranean.
Sheet 1, Tectonics and Kinematics, CGMW-UNESCO.

Bécel, A., Laigle, M., de Voogd, B., Hirn, A., Taymaz, T., Yolsal-Cevikbilen, S., Shimamura,
H., 2010. North Marmara Trough architecture of basin infill, basement and faults,
from PSDM reflection and OBS refraction seismics. Tectonophysics 490, 1-14.

252
Beck, C., 2009. Late Quaternary lacustrine paleo-seismic archives in north-western Alps:
Examples of earthquake-origin assessment of sedimentary disturbances. Earth-Science
Reviews, 96(4), 327-344.

Beck, C, Campos, C., Eriú, K., Ça÷atay, N., Mercier de Lepinay, B., Jouanne, F., 2014.
Estimation of successive co-seismic vertical offsets using coeval sedimentary events.
Application to the Sea of Marmara’s Central Basin (North Anatolian Fault). Natural
Hazards and Earth Science Systems, discussion forum, nhess 2014-39.

Beck, C., Manalt, F. D., Chapron, E., Rensbergen, P. V. A. N., Batist, M. D. E., 1996.
Enhanced Seismicity in the Early Post-Glacial Period: Evidence from the Post-Würm
Sediments of lake Annecy, Northwestern Alps. Journal of Geodynamics, 22(112),
155-171.

Beck, C., Mercier de Lépinay, B., Schneider, J.-luc., Cremer, M., Ça÷atay, N., Wendenbaum,
E., Boutareaud, S., Ménot, G., Schmidt, S., Weber, O., Eris, K., Armijo, R., Meyer,
B., Pondard, N., Gutscher, M.-A., MARMACORE Cruise Party: Turon, J.-L.,
Labeyrie, L., Cortijo, E., Gallet, Y., Bouquerel, H., Gorur, N., Gervais, A., Castera,
M.-H., Londeix, L., de Rességuier, A., Jaouen, A., 2007. Late Quaternary co-seismic
sedimentation in the Sea of Marmara’s deep basins. In: Bourrouilh-Le Jan, F., Beck,
C., Gorsline, D., (Eds.), Sedimentary Records of Catastrophic Events Spec. Issue,
Sedimentary Geology 199, 65-89.

Beck, C., Reyss, J.-L., Leclerc, F., Moreno, E., Feuillet, N., GWADASEIS Cruise Scientific
Party: Barrier, L., Beauducel, F., Boudon, G., Clément, V., Deplus, C., Gallou, N.,
Lebrun, J.-F., Le Friant, A., Nercessian, A., Paterne, M., Saurel, J.-M., Pichot, T.,
Vidal, C., 2012. Identification of deep subaqueous co-seismic scarps through specific
coeval sedimentation. In: Pantosti, D., (Ed.), Lesser Antilles: implication for seismic
hazard, Natural Hazards and Earth System Sc., Special Issue: Subaqueous
Paleoseismology 12, 1755-1767.

Beckers, A., Bodeux, S., Beck, C., Hubert-Ferrari, A., Tripsanas, E., Sakellariou, D., De
Batist, M., De Rycker, K., Bascou, P., and Versteeg, W., 2013. Late Pleistocene to
Present - normal and strike slip - faulting in the western Gulf of Corinth; data from
high resolution seismic reflection SISCOR surveys. Geophysical Research Abstracts,
Vol. 15, EGU2013-4877, 2013, EGU General Assembly 2013.

Beckers, A., Mortier, C., Beck, C., Hubert-Ferrari, A., Reyss, J.-L., Tripsanas, E., Sakellariou,
D., De Batist, M., De Rycker, K., Bascou, P., Strivay., D., 2013. High energy
environment offshore deposits in the western Gulf of Corinth, Greece. Geophysical
Research Abstracts, Vol. 15, EGU2013-4877, EGU General Assembly 2013.

Bell, R. E., McNeill, L. C., Bull, M. J., Henstock, T. J., 2008. Evolution of the offshore
western Gulf of Corinth. Geological Societyof America Bulletin, 120 (1-2), 156-178.

Bell, R., McNeill, L., Bull, J. M., Henstock, T. J., Collier, R. E. L., Leeder, M. R., 2009. Fault
architecture, basin structure and evolution of the Gulf of Corinth Rift, central Greece.
Basin Research 21, 824-835.

253
Bernard, P., Briole, P., Meyer, B., Lyon-Caen, H., Gomez, J.-M., Tiberi, C., Berge, C., Catin,
R., Hatzfeld, D., Lachet, C., Lebrun, B., Deschamps, A., Courboulex, F., Laroque, C.,
Rigo, A., Massonet, D., Papadimitriou, P., Kassaras, J., Diagourtas, D., Makropoulos,
K., Veis, G., Papazisi, E., Mitsakaki, C., Karakostas, V., Papadimitriou, P.,
Papanastassiou, D., Chouliaras, G., Stavrakakis, G., 1997. The Ms = 6.2, June 15,
1995 Aigion Earthquake (Greece): evidence for low angle normal faulting in the
Corinth Rift. Journal of Seismology 1, 131-150.

Bernard, P., Lyon-Caen, H., Briole, P., Deschamps, A., Boudin, F., Makropoulos, K.,
Papadimitriou, P., Lemeille, F., Patau, G., Billiris, H., Paradissis, D., Castarède, H.,
Charade, O., Nercessian, A., Avallone, D., Pachiani, F., Zahradnik, J., Sacks, S.,
Linde, A., 2006. Seismicity, deformation and hazard in the western rift of Corinth:
New insights from the Corinth Rift Laboratory (CRL). Tectonophysics 426:7-30.

Bertrand, S., Charlet, F., Chapron, E., Fageland, N., De Batist, M., 2008. Reconstruction of
the Holocene seismotectonic activity of the Southern Andes from seismites recorded
in Lago Icalma, Chile, 39°S. Palaeogeography, Palaeoclimatology, Palaeoecology,
259 (2-3), 301-322.

Bes¸ S. T., Sur, I. H., Ozsoy, E., Latif, M A., Oguz, T., Unluata. U., 1994. The circulation and
hydrography of the Marmara Sea. Progress in Oceanography, 34 (4), 285-334.

Beúiktepe, S. T., Sur, H., Ozsoy, E., Latif, M. A., Oguz, T., Unluata, U., 1994. The circulation
and hydrography of the Marmara Sea. Progress in Oceanography 34 (4), 285-334.

Blaauw, M., 2010. Methods and code for ‘classical’ age-modelling of radiocarbon sequences.
Quaternary Geochronology, doi:10.1016/j.quageo.2010.01.002

Bloemendal, J., King, J. W., Hall, F. R., Doh, S.-J., 1992. Rock magnetism of Late Neogene
and Pleistocene deep-sea sediments relationship to sediment source, diagenetic processes,
and sediment lithology. Journal of Geophysical Research 97, 4361-4375.

Blott, S. J., 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.

Boggs, S. 2010. Principles of Sedimentology and Stratigraphy. Pearson Prentice Hall, Upper
Saddle River, N.J., 585 pp.

Borradaile, G., 1988. Magnetic susceptibility, petrofabrics and strain. Tectonophysics, 156:1-
20.

Bouma, A. H., 1962. Sedimentology of Some Flysch Deposits. Elsevier Publishing, 168 pp.

Bowman, D., Korjenkov, A., Porat, N., 2004. Late-Pleistocene seismites from Lake Issyk-
Kul, the Tien Shan range, Kyrghyzstan. Sedimentary Geology 163, 211-228.

Briole, P., Rigo, A., Lyon-Caen, H., Ruegg, J. C., Papazissi, K., Mitsakaki, C., Balodimou,
A., Veis, G., Hatzfeld, D., Deschamps, A., 2000. Active deformation of the Corinth

254
rift, Greece: Results from repeated Global Positioning System surveys between 1990
and 1995. Journal of Geophysical Research 105 (B11), 25605-25625.

Brooks, M., Ferentinos, G., 1984. Tectonics and sedimentation in the Gulf of Corinth and the
Zakynthos and Kefallinia channels, Western Greece. Tectonophysics 101 (1-2), 25-54.

Buck, C. E., Cavanagh, W. G., Litton, C. D. 1996. Bayesian approach to interpreting


archaeological data. Chichester, New York, Brisbane, Toronto, Tokyo, Singapore:
John Wiley & Sons, 382 pp.

Buck C. E., Christen J. A., James G. N., 1999. BCal: an online Bayesian radiocarbon
calibration tool. Internet Archaeology 7. (http://intarch.ac.uk/journal/issue7/buck/).

Bull, J. M., Barnes, P. M., Lamarche, G., Sanderson, D. J., Cowie, P., Taylor, S. K., Dix, J.
K.., 2006. High-resolution record of displacement accumulation on an active normal
fault: implications for models of slip accumulation during repeated earthquakes.
Journal of Structural Geology 28, 1146-1166.

Burg, J-P., Ricou. L-E., GodfIriaux, I., Dimov, D., Khin, L., 1996. Syn-metamorphic nappe
complex in the Rhodopc Massif. Struccure and kinematics. Terra. Nova 8, 6-15.

Burnard, P., Bourlange, S., Henry, P., Géli, L., Tryon, M. D., Natalin, B., Sengör, A. M. C.,
Özeren, M.S., Ça÷atay, M. N., 2012. Constraints on fluid origins and migratin
velocities along the Marmara Main Fault (Sea of Marmara, Turkey) using helium
isotopes. Earth and Planetary Sciences Letters 341/344, 68-78.

Burton, P. W., Xu, Y., Qin, C., Tselentis, G-A., Sokos, E. 2004. A catalogue of seismicity in
Greece and the adjacent areas for the twentieth century. Tectonophysics, Volume 390,
Issue 1-4, 117-127.

C
Ça÷atay, M. N., Algan, O., Sakinc¸ M., Eastoe, M. J., Egesel, L., Balkis, N., Ongan, D.,
Caner. H., 1999. A mid-late holocene sapropelic sediment unit from the southern
marmara sea shelf and its palaeoceanographic significance. Quaternary Science
Reviews, 18 (4-5), 531-540.

Ça÷atay, M. N., Görür, N., Algan, O., Eastoe, C., Tchapalyga, A., Ongan, D., Kuhn, T.,
Kuúcu. I., 2000. Late Glacial-Holocene palaeoceanography of the Sea of Marmara :
timing of connections with the Mediterranean and the Black Seas. Marine Geology,
167 (3-4), 191-206.

Ça÷atay, M. N., Görür, N., Polonia, A., Demirba÷, E., Sakinç, M., Cormier, M.-H.,
Capotondi, L., McHugh, C., Emre, Ö., Eriú, K., 2003. Sea level changes and
depositional environments in the Ïzmit Gulf, eastern Marmara Sea, during the Late
Glacial-Holocene period. Marine Geology 202, 159-173.

Ça÷atay, M.N., Eris, K., Ryan, W. B. F., Sancar, U., Polonia, A., Akcer, S., Biltekin, D.,
Gasperini, L., Görür, N., Lericolais, G., Bard, E., 2009. Late Pleistocene-Holocene
evolution of the northern shelf of the Sea of Marmara. Marine Geology 265 (3-4), 87-
100.

255
Ça÷atay, M. N., Erel, L., Bellucci, L. G., Polonia, A., Gasperini, L., Eriú, K. K., Sancar, Ü.,
Biltekin, D., Uçarkuú, G., Ülgen, U. B., Damcõ, E., 2012. Sedimentary earthquake
records in the øzmit Gulf, Sea of Marmara, Turkey. Sedimentary Geology 282, 347-
359.

Campos, C., Beck, C., Crouzet, C., Carrillo, E., 2011. Characterization of Late Pleistocene-
Holocene earthquake-induced “homogenites” in the Sea of Marmara through magnetic
fabric. Implications for co-seismic offsets detection and measurements. Proceedings
2nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake
Geology, Archaeology and Engineering, Corinth, Greece, 19-21.

Campos, C., Beck, C., Crouzet, C., François, D., van Welden, A., Eris, K., 2013. Deciphering
hemipelagites from homogenites through Magnetic Susceptibility Anisotropy.
Paleoseismic implications (Sea of Marmara and Gulf of Corinth), Sedimentary
Geology 292, 1-14.

Campos, C., Beck, C., Crouzet, C., Carrillo, E., van Welden, A., Tripsanas, E., 2013. Late
Quaternary paleoseismic sedimentary archive from deep central Gulf of Corinth: time
distribution of inferred earthquake-induced layers. Annals of Geophysics, 56/6:1-15,
doi:10.4401 / ag-6226.

Caputo, M ., Panza, G. F., Postpischl, D., 1970. Deep structure of the Mediterranean basin. J.
Geoph. Res. 1, 4919-4923.

Carrillo, E., Audemard, F., Beck, C., Cousin, M., Jouanne, F., Cano, V., Castilla, R., Melo,
L., and Villemin, T., 2006. A Late Pleistocene natural seismograph along the Boconò
Fault (Mérida Andes, Venezuela): the moraine-dammed Los Zerpa palæo-lake.
Bulletin of the French Geological Society 177(1), 3-17.

Carrillo, E., Beck, C., Audemard, F.A., Moreno, E., Ollarves, R., 2008. Disentangling Late
Quaternary climatic and seismo-tectonic controls on Lake Mucubají sedimentation
(Mérida Andes, Venezuela). In: De Batist, M., Chapron, E., (Eds.), Lake systems:
sedimentary archives of climate change and tectonic. Palaeogeography,
Palaeoclimatology, Palaeoecology 259, 284-300.

Carton, H., Singh, S. C., Hirn, A., Bazin, S., de Voogd, B., Vigner, A., Richolleau, A., Cetin,
S., Oçakoglu, N., Karakoç, F., Sevilgen, V., 2007. Seismic imaging of the three
dimensional architecture of the Çinarcik Basin along the North Anatolian Fault.
Journal of Geophysical Research 112, B06101, 1-17

Cavazza, W., Roure, F., Ziegler, P., 2004. The Mediterranean Area and the Surrounding
Regions: Active Processes, Remnants of Former Tethyan Oceans and Related Thrust
Belts. In: Cavazza, W., Roure, F., Spakman, W., Stampfli, G. M., Ziegler, P., (Eds.).
The TRANSMED Atlas, Springer Verlag, Berlin, 1-29.

Chapron, E., Beck, C., Pourchet, M., Deconinck, J. -F., 1999. 1822 earthquake-triggered
homogenite in Lake Le Bourget (NW Alps). Terra Nova 11 (2-3), 86-92.

256
Chevalier, N., Bouloubassi, I., Birgel, D., Crémière, A., Taphanel, M-H., Pierre, C., 2011.
Authigenic carbonates at cold seeps in the Marmara Sea (Turkey): A lipid biomarker
and stable carbon and oxygen isotope investigation. Marine Geology 288, Issues 1-4,
112-121.

Chen, F., Siebel, W., Satir, M., Terzio÷lu, N., Saka, K., 2002. Geochronology of the Karadere
basement (NW Turkey) and implications for the geological evolution of the østanbul
Zone. International Journal Earth Sciences 91, 469-481.

Cita, M. B., Camerlenghi, A., Rimoldi, B., 1996. Deep-sea tsunami deposits in the eastern
Mediterranean: New evidence and depositional models. Sedimentary Geology 104 (1-
4), 155-173.

Cita, M. B., Rimoldi, B., 1997. Geological and geophysical evidence for a Holocene tsunami
deposit in the eastern Mediterranean deep-sea record. In: Hancock, Michetti , A. M.,
(Eds.), Paleoseismology; understanding past earthquakes using Quaternary geology.
Journal of Geodynamics 24, 293-304.

Collier, R., Leeder, M. R., Rowe, P. J., Atkinson, T. C., 1992. Rates of tectonic uplift in the
Corinth and Megara Basins, central Greece. Tectonics 11 (6), 1159-1167.

Collier, R., Pantosti, D., D’addezio, G., De Martini, P. M., Masana, E., Sakellariou, D., 1998.
Paleoseismicity of the 1981 Corinth earthquake fault: Seismic contribution to
extensional strain in central Greece and implications for seismic hazard. Journal of
Geophysical Research 103 (B12), 30001-30019.

Collier, R., Leeder, M. R., Trout, M., Ferentinos, G., Lyberis, E., Papatheodorou, G., 2000.
High sediment yields and cool, wet winters: Test of last glacial paleoclimates in the
northern Mediterranean. Geology 28 (11), 999-1002.

Collombat, H., 1993. Etude des propriétés magnétiques des sédiments non consolidés:
anisotropie et erreurs d’inclinaison paléomagnétique. PhD Thesis, J. Fourier
University Grenoble, France, 214 pp.

Cormier, M.-H., Seeber, L., McHugh, C. M. G., Polonia, A., Ça÷atay, M. N., Emre, O.,
Gasperini, L., Görür, N., Bortoluzzi, G., Bonatti, E., Ryan, W. B. F., Newman, K. R.,
2006. The North Anatolian fault in the Gulf of Izmit (Turkey): Rapid vertical motion
in response to minor bends of a non-vertical continental transform. Journal of
Geophysical Research 111, B04102, doi:1029/2005JB003633.

Cornet, F. H., Bernard, P., Moretti, L., 2004. The Corinth Rift Laboratory. Compte Rendus
Geosciences 336 (4-5), 235-241.

Croudace, I. W., Rindby, A., Rothwell, R. G., 2006. ITRAX: description and evaluation of a
new X-ray core scanner. In: Rothwell, R.G., (Ed.), New ways of looking at sediment
cores and core data. Geological Society of London 267, 51-63.

Cuven, S., Francus, P., Lamoureux, S. F., 2010. Estimation of grain size variability with
micro X-ray fluorescence in laminated lacustrine sediments, Cape Bounty. Canadian
High Arctic 44, 803-817.

257
D
Dannielou, B., 2002. Les rapports de campagnes à la mer MD124/ geosciences2, à bord du
Marion Dufresne. Institut Polaire Français, Paul-Emile Victor, ECO/2002/02, 75 pp.

De Batist, M., Klerkx, J., Imbo, Y., Giralt, A., Lignier, V., Beck, C., Delvaux, D., 2002.
Bathymetry and sedimentary environments of Lake Issyk-Kul, Kyrgyz Republic
(Central Asia): a large, high-altitude, tectonic lake. In : Klerkx, J., Imanackunov, B.,
(Eds.), Lake Issyk-Kul: its natural environment, NATO Science Series IV, Kluwer
Academic Publication 13, 101-124.

De Martini, P. M., Pantosti, D., Palyvos, N., Lemeille, F., McNeill, L., Collier, R. E. L., 2004.
Slip rates of the Aigion and Eliki Faults from uplifted marine terraces, Corinth Gulf,
Greece. Compte-Rendus. Geosciences 336, 325-334.

Dercourt, J., Aubouin, J., Savoyat, E., Desprairies, A., Terry, J., Vergely, P., Mercier, J.L.,
Godfriaux, I.., Ferrière, J., Fleury, J.J., Celet, P., et Clément, B., 1977. Compte-Rendu
de la réunion Extraordinaire de la Société Géologique de France en Grèce. Bulletin de
la Société Géologique de France, (7), t. XIX, 1 :5-70.

Dercourt, J., Zonenshain, L. P., Ricou, L. E., Kazmin, V. G., Le Pichon, X., Knipper, A. L.,
Grandjacquet, C., Sbortshikov, I. M., Geyssant, J., Lepvrier, C., Pechersky, D. H.,
Boulin, J., Sibuet, J. C., Savostin, L. A., Sorokhtin, O., Westphal, M., Bazhenov, M.
L., Lauer, J. P., Biju-Duval., B., 1986. Geological evolution of the Thetys belt from
the Atlantic to the Pamirs since 20 the Lias. Tectonophysics 123, 241-315.

Dewey, J. F., Pitman, W. C., Ryan, W. B. F., Bonnin, J., 1973. Plate tectonics and the
evolution of the alpine system. Geological Society of America Bulletin 84 (10), 3137-
3180.

Dewey, J. F., ùengör, A. M. C., 1979. Aegean and surrounding region: complex multiplate
and continuum tectonics in a convergent zone. Geological Society of America Bulletin
90, 84-92.

Dewey, J. F., Hempton, M. R., Kidd, W. S. F., Xaro÷lu, F., Xengfr, A. M. C., 1986.
Shortening of continental lithosphere: the neotectonics of eastern Anatolia—a young
collision zone. In: Coward, M. P., Ries, A. C. (Eds.). Collision Tectonics, Geological
Society Special Publication, vol. 19, 3-36.

Dixon, J. E., Robertson, A. H. F., 1984. The geological evolution of the Eastern
Mediterranean. Spec. Publ. Geol. Soc. 17, 824 pp.

Demory, F., Oberhänsli, H., Nowaczyk, N.R., Gottschalk, M., Wirth, R., Naumann, R., 2005.
Detrital input and early diagenesis in sediments from Lake Baikal revealed by rock
magnetism. Global and Planetary Change 46 (1-4), 145-166.

Dominey-Howes, D., 2002. Documentary and Geological Records of Tsunamis in the Aegean
Sea Region of Greece and their Potential Value to Risk Assessment and Disaster
Management. Natural Hazards 25 (3), 195-224.

258
Doig, R. 1990. 2300 yr history of seismicity from silting events in Lake Tadousac,
Charlevois, Quebec. Geology 18, 820-823.

Doutsos, T., Kokkalas, S., 2001. Stress and deformation patterns in the Aegean region.
Journal of Structural Geology 23, Issues 2-3, 455-472.

Drab, L., 2012. Etude multidisciplinaire le long de la Faille Nord Anatolienne, Turquie :
Paléosismologie marine et paléomagnétisme en Mer de Marmara. Etude
géomorphologique du décalage de la rivière Kõzõlõrmak par utilisation des isotopes
cosmogéniques. PhD Thesis, University of Paris-Sud and Ecole Normale Supérieure,
250 pp..

Drab, L., Hubert-Ferrari, A., Schmidt, S., Martinez, P., 2012. The earthquake sedimentary
record in the western part sea of marmara, turkey. In: Pantosti, D., (Ed.), Natural
Hazards and Earth System Science, Special Issue “Subaqueous Paleoseismology”12
(4), 1235-1254.

Dunlop, D. J., 1981. The rock magnetism of fine particles. Physics of the Earth and Planetary
Interiors 26, Issues 1-2, 1-26.

E
Ergin, M., Bodur, M. N., Ediger, V., 1991. Distribution of surficial shelf sediments in the
northeastern and southwestern parts of the Sea of Marmara: strait and canyon regimes
of the Dardanelles and Bosporus. Mar. Geol. 96, 313-340.

Ergün, M., Özel, E., 1995. Structural relationship between the Sea of Marmara Basin and the
North Anatolian fault zone. Terra Nova 7, 278-288.

Eriú, K. K., Ryan, W. B. F., Ça÷atay, M. N., Sancar, U., Lericolais, G., Menot, G., Bard, E.,
2007. The timing and evolution of the post-glacial transgression across the Sea of
Marmara shelf south of Istanbul. Marine Geology, 243 (1-4), 57-76, 2007.

Eriú, K. K., Ça÷atay, N., Beck, C., de Lépinay, M. B., Campos, C., 2012. Late-Pleistocene to
Holocene sedimentary fills of the Çinarcik Basin of the Sea of Marmara. Sedimentary
Geology, 281, 151-165.

F
Flerit, F., Armijo, R., King, G. C. P., Meyer, B., Barka, E., 2003. Slip partitioning in the Sea
of Marmara pull-apart deunitined from GPS velocity vectors. Geophysical Journal
International 154, 1-7.

Fraser, J., Vanneste, K., Hubert-Ferrari, A., 2010. Recent behavior of the North Anatolian
Fault : Insights from an integrated paleoseismological data set. Journal of Geophysical
Research 115 (B9), B09316.

Folk R. L., Ward, W. C., 1957. Brazos River bar: a study in the significance of grain size
parameters. Journal of Sedimentary Petrology 27, 3-26.

259
Ford, M., Rohais, S., Williams,E.A., Bourlange, S., Jousselin, D., Backert, N., Malartre, F.,
(in press). Tectono-sedimentary evolution of the western Corinth rift (Central Greece).
Basin Research.

G
Galanopoulos, A. G., 1953. Katalog der Erdbeben in Griechland für die zeit von 1879 bis
1892. Ann. Geol. Pays Hellen. 5, 144-229.

Galanopoulos, A. G., Comninakis, P., Delibasis N., 1964. A tsunami generated by an earth
slump set in motion without shock. Annales Géologiques des Pays. Helléniques 16,
93-110.

Gasperini, L., Polonia, A., Ça÷atay, M. N., Bortoluzzi, G., Ferrante, V., 2011. Geological slip
rates along the North Anatolian Fault in the Marmara Sea. Tectonics 30, TC6001.

Gazio÷lu, C., Gokasan, E., Algan, O., Yucel, Z., Tok, B., Dogan, E., 2002. Morphologic
features of the Marmara Sea from multi-beam data. Marine Geology 190 (1-2), 397-
420.

Gazio÷lu, C., Yucel, Z.Y., Dogan, E., 2005. Morphological features of major submarine
landslides of Marmara Sea using multibeam data. Journal of Coastal Research 21 (4),
664-673.

Gealey, W. K., 1988. Plate tectonic evolution of the Mediterranean-Middle East region,
Tectonophysics, 155, 285-306.

Geli, L., Henry, P., Zitter, T., Dupré, S., Tryon, M., Ça÷atay, M. N., Mercier de Lépinay, B.,
Le Pichon, X., Sengor, A. M. C., Görur, N., Natalin, B., Uçarkus, G., Volker, D.,
Gasperini, L., Burnard, P., Bourlange, S., The MARNAUT Scientific Party, 2008. Gas
emissions and active tectonics within the submerged section of the North Anatolian
Fault zone in the Sea of Marmara. Earth and Planetary Science Letters 274, 34-39.

Glew, J. R., Smol, J. P., Last, W. M., 2001. Sediment core collection and extrusion. In: Last,
W. M., Smol, J. P., (Eds.), Tracking environmental change using lake sediments.
Volume 1: basin analysis, coring, and chronological techniques. Kluwer Academic
Publishers, Dordrecht, 73-105.

Gracia, E., Vizcaino, A., Escutia, C., Asioi, A., Rodes, A., Pallas, R., Garcia-Orellana, J.,
Lebreiro, S., Goldfinger, C., 2010. Holocene earthquake record offshore Portugal (SW
Iberia): testing turbidite paleoseismology in a slow-convergence margin. Quaternary
Science Reviews 29, 1156-1172.

Goldfinger, C., Nelson, C. H., Johnson, J. E. 2003. Holocene Earthquake records from the
cascadia subduction zone and northern San Andreas Fault based on precise dating of
offshore turbidites. Annual Review of Earth and Planetary Sciences 31 (1), 555-577.

Goldfinger, C., Morey, A. E., Nelson, C. H., Gutierrez-Pastor, J., Johnson, J. E., Karabanov,
E., Chaytor, J., Eriksson, A., Shipboard Scientific Party., 2007. Rupture lengths and
temporal history of significant earthquakes on the offshore and north coast segments

260
of the Northern San Andreas Fault based on turbidite stratigraphy. Earth and Planetary
Science Letters 254, 9-27.

Goldfinger, C., 2009. Subaqueous Paleoseismology. In: McCalpin, J., (Ed.), Paleoseismology,
2nd edition, Elsevier, 119-169.

Gorsline, D. S., Diego, T. D., Nava-sanchez, E. H., 2000. Seismically triggered turbidites in
small margin basins: Alfonso Basin, Western Gulf of California and Santa Monica
Basin, California Borderland. Science 135, 21-35.

Görür, N., Okay, A. I., 1996. Fore-arc origin of the Thrace Basin, northwest Turkey.
Geologische Rundschau 85, 662-668.

Görür, N., Ça÷atay, M.N., 2010. Geohazards rooted from the northern margin of the Sea of
Marmara since the late Pleistocene: a review of recent results. Natural Hazards 54 (2),
583-603.

Grácia, E., Vizcaino, A., Escutia, C., Asioli, A., Rodes, A., Pallas, R., Garcia-Orellana, J.,
Lebreiro, S., Goldfinger, C., 2010. Holocene earhquake record offshore Portugal (SW
Iberia): testing turbidite paleoseismology in a slow-convergence margin. Quaternary
Science Reviews 29, 1156-1172.

Halbach, P., Kusçu, I., Inthorn, M., Kuhn, T., Pekdeger, A., Seifert, R., 2002. Methane in
sediments of the deep Marmara Sea and its relation to local tectonic structures. In: Gö
rür, N., Papadopulos, G.A., Okay, N. (Eds.), Integration of Science Research on the
1999 Turkish and Greek Earthquakes, NATO Science Series IV Earth and
Environmental Sciences 9, 71-85.

Halbach, P., Holzbecher, E., Reichel, T., Moche, R., 2004. Migration of the sulphatemethane
reaction zone in marine sediments of the Sea of Marmara — can this mechanism be
tectonically induced? Chemical Geology 205, 73-82.

Hasiotis, T., Papatheodorou, G., Bouckovalas, G., Corbau, C., Ferentinos, G., 2002.
Earthquake-induced coastal sediment instabilities in the western Gulf of Corinth,
Greece. Marine Geology, 186 (3-4), 319-335.

Hamilton, N., Rees, A. I., 1970. The use of magnetic fabric in paleocurrent estimation. In:
Runcorn, S. K., (Ed.), Paleogeophysics, Academic Press, New York, 445-463.

Hancock, P. L., Barka, A. A., 1981. Opposed shear senses inferred from neotectonic
mesofractures systems in the North Anatolian fault zone. Journal of Structural
Geology 3, 383-392.

Hatzfeld, D., Karakostas, V., Ziazia, M., Kassaras, I., Papadimitriou, E., Makropoulos, K.,
Voulgaris, N., Papaioannou, C., 2000. Microseismicity and faulting geometry in the
Gulf of Corinth (Greece). Geophys. J. Int. 141, 438-456.

261
Hébert, H., Schindelé, F., Altinok, Y., Alpar, B., Gaziogluc, C., 2005. Tsunami hazard in the
Marmara Sea (Turkey): a numerical approach to discuss active faulting and impact on
the Istanbul coastal areas. Marine Geology, 215, 23- 43.

Heegaard, E. 2003. Mixed effect age-depth routine for R. Retrieved 26 January 2005 from
http://www.bio.uu.nl/ palaeo/ Congressen/Holivar/Literature_Holivar2003.htm or
http://www.

Heezen, B. C., Ewing, M., 1952. Turbidity currents and submarine slumps, and the 1929
Grand Banks earthquake. Am. J. Sci. 250, 849-873.

Heezen, B. C., Ewing, M., Johnson, G.L., 1966. The Gulf of Corinth floor. Deep-Sea Res. 13,
381- 411.

Hempton, M. R., Dewey, J. F., 1983. Earthquake-induced deformational structures in young


lacustrine sediments, East Anatolian Fault, southeast Turkey. Tectonophysics 98,
Issues 3-4, T7-T14.
Hempton, M.R., 1985. Structure and deformation history of the bitlis suture near lake Hazar,
southern Turkey. Geol. Soc. Amer. Bull. 96, 233-243.

Hesse, R., 1977. Turbiditic and non-turbiditic mudstone of Cretaceous flysch sections of the
East Alps and other basins. Sedimentology 22, 387-416.

Higgs, B., 1988. Syn-sedimentary structural controls on basin deformation in the Gulf of
Corinth, Greece. Basin Res. 1, 155-165.

Hiscott. R. N., Aksu, A. E., 2002. Late Quaternary history of the Marmara Sea and Black Sea
from high-resolution seismic and gravity core studies. Marine Geology 190 (1-2), 261-
282.

Hiscott, R. N., Aksu, A. E., Mudie, P. J., Marret, F., Abrajano, T., Kaminski, M. A., Evans, J.,
Çakõro÷lu, A. I., Yaúar, D., 2007. A gradual drowning of the southwestern Black Sea
shelf: Evidence for a progressive rather than abrupt Holocene reconnection with the
eastern Mediterranean Sea through the Marmara Sea Gateway. Quat. Int., 167-168,
19-34.

Hornbach, M. J., Braudy, N., Briggs, R.W., Cormier, M.-H., Davis, M.B., Diebold, J.B.,
Dieudonne, N., Douilly, R., Frohlich, C., Gulick, S.P.S., Johnson, H.E., Mann, P.,
McHugh, C.M.G., Ryan-Mishkin, K., Prentice, C.S., Seeber, L., Sorlien, C.C.,
Steckler, M.S., Symithe, S.J., Taylor, F.W., and Templeton, J., 2010. High tsunami
frequency as a result of combined strike-slip faulting and coastal landslides, Nature
Geoscience 3, 783-788.

Hounslow, M., Bootes, P. A., Whyman, G., 1990. Remanent magnetization of sediments
undergoing deformation in the Barbados accretionary prism: ODP Leg 110. In: Moore,
J.C., Mascles, A., et al., (Eds.), Proceedings of the Ocean Drilling Program, Scientific
Results 110, 379-391.

262
Housen, B. A., Kanamatsu, T., 2003. Magnetic fabrics from the Costa Rica margin: sediment
deformation during the initial dewatering and underplating process. Earth and
Planetary Science Letters 206, Issues 1-2, 215-228.

Hrouda, F. E. K., 1982. Magnetic anisotropy of rocks and its application. Geophysical
Surveys 5, 37-82.

Huh, C. A., Su, C. C., Liang, W. T., Ling, C. Y., 2004. Linkage between turbidites in the
southern Okinawa Trough and submarine earthquakes. Geophysical Research Letters
31, L12304. 1-4.

Hubert, A., Armijo, R., King, G., Gasse, F., Barka, A. A., 1997. Slip rate of the North
Anatolian fault, Turkey. Eos (Transactions, American Geophysical Union) 78, F715.

Hubert-Ferrari, A., Barka, A., Jacques, E., Nalbant, S. S., Meyer, B., Armijo, R., Tapponnier,
P., King, G. C. P., 2000. Seismic hazard in the Marmara Sea region following the 17
August 1999 Izmit earthquake. Nature 404, 6775, 269-273.

Hubert-Ferrari, A., Armijo, R., King, G.C.P., Meyer, B., Barka, A., 2002. Morphology,
displacement and slip rates along the North Anatolian Fault (Turkey). J. geophys. Res.
107, B10, ETG9–1 - ETG9–33.

I
IGME, 1983. Geological Map of Greece, scale 1:500 000, Institute of Geology and Mineral
Exploration, Athens.

Imren, C., Le Pichon, X., Rangin, C., Demirbag, E., Ecevitoglu, B., Görür, N., 2001. The
North Anatolian Fault within the Sea of Marmara: A new interpretation based on
multi-channel seismic and multi-beam bathymetry data. Earth Planet. Sci. Lett. 186,
143-158.

J
Jackson, J. A., Gagnepain, J., Houseman, G., King, G. C. P., Papadimitriou, P., Soufleris, C.,
Virieux, J., 1982. Seismicity, normal faulting and the geomorphological development
of the Gulf of Corinth (Greece): the Corinth earthquakes of February and March 1981.
Earth and Planetary Science Letters 57 (2), 377-397.

Jacobshagen V., Wallbrecher, E., I984. Pre-Neogene nappe structure and metamorphism of
the North Sporades and the southern Pelion peninsula. In: Robertson, A. H. F., Emeis,
K. C., Richter, C., Camerlenghi, A. (Eds.). Proceeding of the Ocean Drilling Program,
Scientific Results, vol. 160, 591-602.

Jelinek, V., 1981. Characterization of the magnetic fabric of rocks. Tectonophysics, 79 (3-4),
63-67.

Jolivet, L., 2001. A comparison of geodetic and finite strain pattern in the Aegean,
geodynamic implications. Earth and Planetary Science Letters, Volume 187 (1-2), 95-
104.

263
Jolivet L., Brun, J-P., Gautier, P., Lallemant, S., Patriat, M., 1994. 3D-kinematics of extension
in the Aegean region from the early miocene to the Present, insights from the ductile
crust. Bu11. Soc. geol. France 165, 195-209.

Jolivet, L., Faccenna, C., Huet, B., Labrousse, L., Le Pourhiet, L., Lacombe, O., Lecomte, E.,
Burov, E., Denèle, Y., Brun, J.-P., Philippon, M., Paul, A., Salaün, G., Karabulut, H.,
Piromallo, C., Monié, P., Gueydan, F., Okay, A., Oberhänsli, R., Pourteau, A., Augier,
R., Gadenne, L., Driussi, O., 2013. Aegean tectonics: Strain localisation, slab tearing
and trench retreat. Tectonophysics 597–598 (2013) 1–33

Joseph, L. H., Rea, D. K., Van der Pluijm, B. A., 1998. Use of Grain Size and Magnetic
Fabric Analyses to Distinguish Among Depositional Environments. Paleoceanography
13(5), 491-501.

K
Kahle, H.-G., Straub, C., Reilinger, R., McClusky, S., King, R., Hurst, K., Kastens, K., Cross,
P., 1998. The strain rate field in the eastern Mediterranean region, estimated by
repeated GPS measurements. Tectonophysics 294, 237-252.

Kaminski, M. A., Aksu, A., Box, M., Hiscott, R. N., Filipescu, S., Al-Salameen, M., 2002.
Late Glacial to Holocene benthic foraminifera in the Marmara Sea: implications for
Black Sea-Mediterranean Sea connections following the last deglaciation. Marine
Geology, 190 (1-2), 165-202.

Kastens, K., Cita, M. B., 1981. Tsunami-induced sediment transport in the abyssal
Mediterranean Sea. Geological Society of America Bulletin 92 (11), 845-857.

Karageorgis, A.P., Kanellopoulos, T. D., Mavromatis, V., Anagnostous, C. Koutsopoulou, E.,


Schmidt, M., Pavlopoulos, K., Tripsanas, E. K., Hallberg, R. O., 2013 Authigenic
carbonate mineral formation in the Pagassitikos palaeolake during the latest
Pleistocene, central Greece. Geo-Marine Letters 33, Issue 1, 13-29

Keraudren, B., Sorel, D., 1987. The terraces of Corinth (Greece) - A detailed record of
eustatic sea-level variations during the last 500000 years. Marine Geology 77, 99-107.

Kent, D.V., Lowrie, W., 1975. On the magnetic susceptibility anisotropy of deep-sea
sediments. Earth and Planetary Sciences Letters 28, 1, 1-12.

Ken-Tor, R., Agnon, A., Enzel, Y., Stein, M., 2001. High-resolution geological record of
historic earthquakes in the Dead Sea basin, Journal of Geophysical Research
106:2221-2234.

Klinger, Y., Sieh, K., Altunel, E., Akoglu, A., Barka, A., Dawson, T., 2003. Paleoseismic
evidence of characteristic slip on the western segment of the North Anatolian Fault,
Turkey. Bull Seismol Soc Am 93, 2317-2332.

Koukis, G., Rondoyanni, T., 1989. Neotectonics and recent seismic activity in the Eastern
Corinthian Gulf, Greece. Geotechnical implications. Bulletin de l'Association
française pour l'étude du quaternaire 26-3, 171-180.

264
Koukouvelas, I. K., Stamatopoulos, L., Katsonopoulou, D., Pavlides, S., 2001. A
paleoseismological and geoarchaeological investigation of the Eliki Fault, Gulf of
Corinth, Greece. J. Struct. Geol. 23, 531-543.

Kozacõ, Ö., Dolan, J F., Robert, C., Finkel, A., 2009. A late Holocene slip rate for the central
North Anatolian fault, at Tahtaköprü, Turkey, from cosmogenic 10Be geochronology:
Implications for fault loading and strain release rates. Journal of Gephysical Research:
Solid Earth 114 (B1), B01405.

Kozacõ, Ö., Dolan, J F., Yönlü, Ö., Ross, D., Hartleb, R. D., 2010. Paleoseismologic evidence
for the relatively regular recurrence of infrequent, large-magnitude earthquakes on the
eastern North Anatolian fault at Yaylabeli. Turkey Lithosphere 3, 37-54.

Kwiecien, O., Arz, W. H., Lamy, F., Wulf, S., Bahr, A., Röhl, U., Haug, G. H., 2008.
Estimated reservoir ages of the Black Sea since the last glacial. Radiocarbon 50, 99-
118.

L
Laigle, M., Bécel, A., de Voogd, B., Hirn, A., Taymaz, T., Ozalaybey, S., Cetin, S., Galvé,
A., Karabulut, H., Lépine, J.C., Saatçilar, R., Sapin, M., Shimamura, H., Murai, Y.,
Singh, S., Tan, O., Vigner, A., Yolsal, S., 2008. A first deep seismic survey of the Sea
of Marmara: Deep basins and whole crust architecture and evolution. Earth and
Planetary Science Letters 270, 168-179.

Lamb, M. P., D’Asaro, E., Parsons, J., D., (2004). Turbulent structure of high-density
suspensions formed under waves. J. Geophys. Res. 109, C12026.

Last, W. 2001. Textural analysis of lake sediments. In: Last, W. W., Smol, J. P. (Eds.)
Tracking environmental change using lake sediments. Kluwer academic publishers,
Dordrecht, 41-81.

Leeder, M. R., Harris, T., Kirkby, M., J., 1998. Sediment supply and climate change:
implications for basin stratigraphy. Basin Research 10 (1), 7-18.

Le Pichon, X., Angelier, J., Osmaston, M. F., Stegena, L., 1981. The Aegean Sea.
Philosophical Transactions of the Royal Society of London, Series A, Math. Phys.
Sciences 300 (1454), 357-372.

Le Pichon, X., Chamot-Rooke, N., Lallemant, S., Noomen, R., Vets, G., 1995. Geodetic
deunitination of the kinematics of central Greece with respect to Europe: Implications
for eastern Mediterranean tectonics. Journal of Geophysical Research 100 (B7),
12675-12690.

Le Pichon, X., Sengör, A. M. C., Demirbag, E., Rangin, C., Imren, C., Armijo, R., Görür, N.,
Ça÷atay, N., de Lépinay, B. M., Meyer, B., Saatçllar, R., Tok, B., 2001 The active
main marmara fault. Earth and Planetary Science Letters, 192 (4), 595 - 616.

Lettis, W. R., Kelson, K. I., 2000. Applying geochronology in paleoseismology. In: Noller,
J.S., Sowers, J.M., Lettis, W.R., (Eds.), Quaternary geochronolgy: methods and

265
applications, AGU reference shelf 4: American Geophysical Union, Washington D.C,
479-496.

Levi, T., Weinberger, R., Aïfa, T., Eyal, Y., Marco, S., 2006. Earthquake-induced clastic
dikes detected by anisotropy of magnetic susceptibility. Geology, 34 (2), 69-72.

Lignier, V., 2001. Les sédiments lacustres et l'enregistrement de la palæoséismicité. Etude


comparative de différents cas dans le Quaternaire des Alpes Nord-Occidentales et du
Tien-Shan Kyrghize. PhD Thesis, Université de Savoie, 2 vol., 330 pp.

Londeix, L., Herreyre, Y., Turon, J. L., Fletcher, W., 2009. Last glacial to holocene hydrology
of the marmara sea inferred from a dinoflagellate cyst record. Review of Palaeobotany
and Palynology, 158 (1-2), 52-71.

Lorenzoni, L., Benitez-Nelson, C. R., Thunell, R. C., Hollander, D., Varelan, R., Astor,
Y.,Audemard, F. A., Muller-Karger, F. E., 2012. Potential role of event-driven
sediment transport on sediment accumulation in the Cariaco Basin, Venezuela. Marine
Geology 307/310, 105–110, doi:10.1016/j.margeo.2011.12.009

LØvlie, R., Lowrie, W., Jacobs, M., 1971. Magnetic properties and mineralogy of four deep-
sea cores. Earth and Planetary Sciences Letters 15, 157-168.

Lowe, D. R., 1982. Sediment gravity flows; II, Depositional models with special reference to
the deposits of high-density turbidity currents. Journal of Sedimentary Research 52
(1), 279-297.

Lykousis, V., Anagnostou, Ch., 1994. Sedimentological and paleogeographic evolution of the
Saronic Gulf during the Late Quaternary. Bull. Geol. Soc. Greece XXVIII/1, 501-510
(in Greek).

Lykousis, V., Sakellariou, D., Papanikolaou, D., 1998. Sequence stratigraphy in the N. margin
of the Gulf of Corinth: implications to Upper Quaternary basin evolution. Bull. Geol.
Soc. Greece XXXII/2, 157–164.

Lykousis, V., Sakellariou, D., Moretti, I., Kaberi, H., 2007. Late Quaternary basin evolution
of the Gulf of Corinth: Sequence stratigraphy, sedimentation, fault-slip and subsidence
rates. Tectonophysics 440 (1-4), 29-51.

Lykousis, V., Roussakis, G., Sakellariou, D., 2009. Slope failures and stability analysis of
shallow water prodeltas in the active margins of Western Greece, northeastern
Mediterranean Sea. International Journal of Earth Sciences 98 (4), 807-822.

M
Macklin, M.G., Fuller, I.C., Lewin, J., Maas, G.S., Passmore, D.G., Rose, J., Woodward, J.C.,
Black, S., Hamlin, R.H.B., Rowa, J.S., 2002. Correlation of fluvial sequences in the
Mediterranean basin over last 200 ka and their relationship to climate change. Quat.
Sci. Rev. 21, 1633-1641.

266
Major C, Ryan W, Lericolais G, Hajdas I. 2002. Constraints on Black Sea outflow to the Sea
of Marmara during the last glacial-interglacial transition. Marine Geology 190 (1-
2):19-34.

Major, C., Goldstein, S. L., Ryan, W. B. F., Lericolais, G., Piotrowski, A. M., Hajdas, I.,
2006. The co-evolution of black sea level and composition through the last
deglaciation and its paleoclimatic significance. Quaternary Science Reviews, 25(17-
18), 2031 - 2047.

Malvern instruments: www.malvern.co.uk

Marco, S., Agnon, A., 1995. Prehistoric earthquake deformations near Masada, Dead-Sea
graben. Geology 23 (8), 695-698.

McCalpin, J., 2009. Paleoseismology, International Geophysics Series 95, Elsevier, 2nd
edition, ISBN: 978-0-12-373576-8, 798 pp.

McCave, I. N., Bryant, R. J., Cook, H. F., Coughanowr, C. A. 1986. Evaluation of laser-
diffraction-size analyzer for use with natural sediments. Journal of Sedimentary
Petrology 56 (4), 561-564.

McCormac, F.G., Hogg, A.G., Blackwell, P.G., Buck, C.E., Higham, T.F.G., Reimer, P.J.,
2004. SHCal04 southern hemisphere calibration, 0 -11.0 cal kyr BP. Radiocarbon 46,
1087-1092.

McClusky, S., Balassanian, S., Barka, A., Demir, C., Ergintav, S., Georgiev, I., Gurkan, O.,
Hamburger, M., Hurst, K., Kahle, H., Kastens, K., Kekelidze, G., King, R., Kotzev,
V., Lenk, O., Mahmoud, S., Mishin, A., Nadariya, M., Ouzounis, A., Paradissis, D.,
Peter, Y., Prilepin, M., Reilinger, R., Sanli, I., Seeger, H., Tealeb, A., Toksöz, M.N.,
and Veis, G., 2000. Global positioning system constraints on plate kinematics and
dynamics in the eastern Mediterranean and Caucasus. Journal of Geophysical
Research 105 (B3), 5695-5719.

McClusky, C. M. G., Reilinger, R., Mahmoud, S., Ben Sari, D., Tealeb, A., 2003. GPS
constraints on Africa (Nubia) and Arabia plate motions. Geophysical Journal
International 155, 126-138.

McHugh, C. M. G., Seeber, L., Cormier, M-H., Dutton, J., Ça÷atay, N., Polonia, A., Ryan, W.
B. F., Görür, N., 2006. Submarine earthquake geology along the North Anatolia Fault
in the Marmara Sea, Turkey: A model for transform basin sedimentation, Earth Planet.
Sci. Lett., 248, 661-684.

McHugh, C. M. G., Gurung, D., Giosan, L., Ryan, W. B. F., Mart, Y., Sancar, U., Burckle, L.,
Ça÷atay, M. N., 2008. The last reconnection of the Marmara Sea (Turkey) to the
World Ocean: a paleoceanographic and paleoclimatic perspective. Marine Geology
255 (1-2), 64-82.

McHugh, C. M., Seeber, L., Braudy, N., Cormier, M.-H., Davis, M. B., Diebold, J. B.,
Dieudonne, N., Douilly, R., Gulick, S.P.S., Hornbach, M.J., Johnson, H.E. III, Ryan
Miskin, K., Sorlien, C., Steckler, M., Symithe, S.J., Templeton, J., 2011. Offshore

267
sedimentary effects of the 12 January 2010 Haiti earthquake. Geology, 39 (8), 723-
726.

McKenzie D., 1970. The plate tectonics of the Mediterranean region. Nature 226, 239-243.

McKenzie, D., 1972. Active Tectonics of the Mediterranean Region. Geophysical Journal of
the Royal Astronomical Society 30 (2), 109-185.

McKenzie, D., 1978. Active tectonics of the Alpine—Himalayan belt: the Aegean Sea and
surrounding regions. Geophys. J. Royal Astr. Soc. 55 (1), 217-254.

McNeill, L. C., Mille, A., Minshull, T.A., Bull, J.M., Kenyon, N.H., Ivanov, M., 2004.
Extension of the North Anatolian Fault into the North Aegean Trough: Evidence for
transtension, strain partitioning, and analogues for Sea of Marmara basin models.
Tectonics 3, 2:1-12, doi:10.1029/2002TC001490.

McNeill, L. C., Collier, R., De Martini, P. M., Pantosti, D., D’Addezio, G., 2005. Recent
history of the Eastern Eliki Fault, Gulf of Corinth: geomorphology, palaeoseismology
and impact on palaeoenvironments. Geophysical Journal International 161 (1), 154-
166.

McNeill, L. C., Cotterill, C. J., Henstock, T. J., Bull, J .M., Stefatos, A., Collier, R. E. L.,
Papatheodorou, G., Ferentinos, G., Hicks, S. E., 2005. Active faulting within the
offshore western Gulf of Corinth, Greece: implications for model of continental rift
deformation. Geology 33, 241-244.

Meghraoui, M., Aksoy, M. E., Akyuz, H. S., Ferry, M., Dikbas, A., Altunnel, E., 2012.
Paleoseismology of the North Anatolia Fault at Guzelkoy (Ganos segment, Turkey):
Size and recurrence time of earthquake ruptures west of the Sea of Marmara.
Geochemistry, Geophysics, Geosystems 13, 1-26.

Meissl, S., Behrmann, J. H., Franke, C., 2011. Magnetic fabrics in Quaternary sediments,
Ursa Basin, northern Gulf of Mexico record transport processes, compaction and
submarine slumping. Marine Geology 286, Issues 1-4, 51-64.

Ménot, G., Bard, E., 2010. Geochemical evidence for a large methane release during the last
deglaciation from Marmara Sea sediments. Geochimica et Cosmochimica Acta 74,
Issue 5, 1537-1550.

Mestel, R., 1997. Noah's Flood. New Scientist 156 (2102), 24-27.

Michetti, A. M., Hancock, P. L., 1997. Paleoseismology: understanding past earthquakes


using Quaternary geology. Journal of Geodynamics 24 (1), 3-10.

Migeon, S., Weber, O., Fauge`res, J.C., Saint Paul, J., 1998. A new X-ray imaging system for
core analysis. Geomarine Letters 18, 251-255.

Moernaut, J., 2011. Sublacustrine landslide processes and their paleoseismological


significance: revealing the recurrence rate of giant earthquakes in SouthͲCentral Chile.
PhD Thesis, University of Ghent, 274 pp.

268
Monecke, K., Anselmetti, F.S., Becker, A., Sturm, M., Giardini, D., 2004. The record of
historic earthquakes in lake sediments of central Switzerland. Tectonophysics 394:
21–40

Montenat, C., Barrier, P., d’Estevou, P. O., Hibsch, C., 2007. Seismites: An attempt at critical
analysis and classification. Sed. Geol. 196, 5-30.

Moretti, M., Alfaro, P., Caselles, O., Canas, J.A., 1999. Modelling seismites with a digital
shaking table. Tectonophysics 304, 369-383.

Moretti, I., Sakellariou, D., Lykousis, V., Micarelli, L., 2003. The Gulf of Corinth: an active
half graben?. Journal of Geodynamics 36 (1-2), 323-340.

Moretti, I., Lykousis, V., Sakellariou, D., Reynaud, J.-Y., Benziane, B., Prinzhoffer, A., 2004.
Sedimentation and subsidence rate in the Gulf of Corinth: what we learn from the
Marion Dufresne’s long-piston coring. Comptes-Rendus. Geosciences 336 (4-5), 291-
299.

Mörner, N-A., 1996. Liquefaction and varve deformation as evidence of paleoseismic events
and tsunamis. The autumn 10,430 BP case in Sweden. Quaternary Science Review 15
(8), 939-948.

Mörner, N-A., Tröften, P. E., Sjöberg, R., Grant, D., Dawson, S., Bronge, C., Kvamsdal, O.,
Sidén, A., 2000. Deglacial paleoseismicity in Sweden: the 9663 BP Iggesund event.
Quaternary Science Reviews 19, Issues 14-15.

Mörner, N-A. Sun, G., 2008. Paleoearthquake deformations recorded by magnetic variables.
Earth and Planetary Science Letters 267 (3-4), 495-502.

Mortier, C., 2012. Etude sédimentologique dans la partie occidentale du golfe de Corinthe
(Grèce), Ms Sc, Memoir, University of Liège, 129 pp.

Mountrakis, D., 1984. Structural evolution of the Pelagonian Zone in Northwestern


Macedonia, Greece. In: Dixon, J. E., Robertson, A. H. F., (Eds.). The geological
evolution of the Eastern Mediterranean. Geol. Soc. Lond. Spec. Publ. 17, 581-590.

Mountrakis, D., 1986. The Pelagonian Zone in Greece: a polyphase-deformed fragment of the
Cimmeran Continent and its role in the geotectonic evolution of the eastern
Mediterranean. Journal of Geology 94, 335-347.

Mudie, P. J., Marret, F., Aksu, A. E., Hiscott, R. N., Gillespie, H., 2007. Palynological
evidence for climatic change, anthropogenic activity and outflow of Black Sea water
during the late Pleistocene and Holocene: Centennial- to decadal-scale records from
the Black and Marmara Seas. Quaternary International 167-168, 73-90.

Mulder, T., Alexander, J., 2001. The physical character of subaqueous sedimentary density
flows and their deposits. Sedimentology 48 (2), 269-299.

Mulder, T., Zaragosi, S., Razin , P., Grelaud, C., Lanfumey, V., Bavoil, F., 2009. A new

269
conceptual model for the deposition process of homogenite: Application to a
Cretaceous megaturbidite of the western Pyrenees (Basque region, SW France).
Sedimentary Geology 222 (2009) 263–273.

Muller J. R., Aydin, A., 2005. Using mechanical modeling to constrain fault geometries
proposed for the northern Marmara Sea. Journal of Geophysical Research: Solid
Earth 110 (B3), B03407.

Mutti E., Ricci Lucchi., 1978. Turbidites of the northern Apennines: introduction to facies
analysis. American Geological Institute, Reprint Series 3, 127-166.

Mutti, E., 1992. Turbidite sandstones. Milan, Italy, Agip, 588 Instituto di Geologia,
Universita` di Parma, San 589 Donato Milanese, 275 pp.

Mutti, E., Tinterri, R., Remacha, E., Mavilla, N., Angella, S., Fava, L., 1999. An introduction
to the analysis of ancient turbidite basins from an outcrop perspective. American
Association of Petroleum Geology, Continuing Education Course Note Series 39, 95
pp.

N
Nakajima, T., Kanai, Y., 2000. Sedimentary features of seismoturbidites triggered by the 1983
and older historical earthquakes in the eastern margin of the Japan Sea. Sedimentary
Geology 135 (1-4), 1-19.

Neumann, P., Zacher, W., 2004. The Cretaceous sedimentary history of the Pindos Basin
(Greece). Int. J. Earth Sci. 93, 119-131.

Neuser, R. D., Simon, M., Richter, D. K., 1982. Die “Neogen”und Quartaren Grobzyklen im
Bereich des Kanals von Korinth (Griechenland). Bochumer Geol. U. Geotechn. Arb.
8, 53-145.

Noda, A., Tuzino, T., Kanai, Y., Furukawa, R., Uchida, J., 2008. Paleoseismicity along the
southern Kuril Trench deduced from submarine-fan turbidites. Marine Geology 254,
73-90.

Noller, J. S., Sowers, J. M., Lettis, W. R., 2000. Quaternary geochronology: Methods and
applications. Am. Geophys. Union, Reference Shelf 4, 582 pp.

Noquet, J.-M., 2012. Present-day kinematics of the Mediterranean: a comprehensive overview


of GPS results. Tectonophysics, 579:220-242.

Nyst, M., Thatcher, W., 2004. New constraints on the active tectonic deformation of the
Aegean, J. Geophys. Res.109, B11406.

O
Obermeier, S. F., Martin, J. R., Frankel, A. D., Youd, T. L., Munson, P. J., Munson, C. A.,
Pond, E. C., 1993. Liquefaction evidence for one or more strong Holocene
earthquakes in the Wabash Valley of southern Indiana and Illinois, with a preliminary
estimate of magnitude. U.S. Geol. Surv. Prof. Paper 1536, 27 pp.

270
Obermeier, S. F., 1996. Use of liquefaction - induced features for paleoseismic analysis. An
overview of how seismic liquefaction features can be distinguished from other features
and how their regional distribution and properties of source sediment can be used to
infer the location and strength of Holocene paleo-earthquakes. Engineering Geology
44, 1-76.
Obermeier, S. F., Pond, E. C., Olson, S. M., Green, R. A., Stark, T. D., Mitchell, J. K. 2001.
Paleoliquefaction studies in continental setting; geologic and geotechnical factors in
interpretations and back-analysis. Open-File Report - U. S. Geological Survey, Reston,
VA, United States, 72 pp.
Obermeier, S. F., 2009, Using liquefaction-induced and other soft-sediment features for
paleoseismic analysis. In: McCalpin, J. M., (Ed.) Paleoseismology. International
Geophysics Series 95, 497-564.

Oeschger, H., Siegenthaler, U., Schotterer, U., Guglemann, A., 1975. A box-diffusion model
to study the carbon dioxide exchange in nature, Tellus 27, 168-192.

Okay, A. I., 1989. Tectonic units and sutures in the Pontides, northern Turkey, NATO ASI
series. Series C, Mathematical and physical sciences 259, 109-116.

Okay, A. I., 2008. Geology of Turkey: A synopsis. Anschnitt 21, 19-42.

Okay, A. I., Monié., P., 1997. Early Mesozoic subduction in the Eastern Mediterranean:
evidence from Triassic eclogite in northwest Turkey. Geology 25, 595- 598.

Okay, A. I., Tüysüz, O., 1999. Tethyan sutures of northern turkey. Geological Society,
London, Special Publications 156 (1), 475 pp.

Okay, A. I., Kaslilar-Ozcan, A., Imren, C., Boztepe-Guney, A., Demirbag, E., Kuscu, I., 2000.
Active faults and evolving strike-slip basins in the Marmara Sea, northwest Turkey: A
multichannel seismic reflection study. Tectonophysics 321, 189– 218.

Okay, A. I., Göncüoglu, M. C., 2004. Karakaya Complex: a review of data and concepts.
Turkish Journal of Earth Sciences 13, 77-95.

Okay, N., Ergun, B., 2005. Source of basinal sediments in the Marmara Sea investigated using
heavy minerals in the modern beach sands. Mar. Geol. 216, 1-15.

Ori, G., 1989. Geological history of the extensional basin of the Gulf of Corinth (Miocene-
Pleistocene), Greece. Geology 17, 918-921.

Özeren, M. S., Ça÷atay, M. N., Postacioglu, N., ùengör, A. M. C., Görür, N., Eris, K., 2010.
Mathematical modelling of a potential tsunami associated with a late glacial
submarine landslide in the Sea of Marmara. Geo-Marine Letters 30 (5), 523-539.

P
Palyvos, N., Pantosti, D., De Martini, P.M., Lemeille, F., Sorel, D., Pavlopoulos, K., 2005.
The Aigion-Neos Erineos coastal normal fault system (west Corinth Gulf Rift,
Greece): geomorphological signature, recent earthquake history and evolution. Journal
of Geophysical Research 110, B09302, 1-15.

271
Pantosti, D., De Martini, P. M., Koukouvelas, I., Stamatopoulos, L., Palyvos, N., Pucci, S.,
Lemeille, F., 2004. Palaeoseismological investigations of the Aigion Fault (Gulf of
Corinth, Greece), Comptes-Rendus. Geosciences 336 (4-5), 335-342.

Passega, R., 1964. Grain Size Representation by CM Patterns as a Geological Tool. SEPM.
Journal of Sedimentary Research 34 (4), 830-847.

Papadopoulos, G. A., 1998. A reconstruction of the great earthquake of 373 B.C. in the
western gulf of Corinth. In: Katsonopoulou, D., Soter, S., Schilardi, D., (Eds.). Helike
II-Ancient Helike and Aigialeia, Proc. 2nd Intern. Conf., Aighion, 479-494.

Papadopoulos, G. A., Plessa, A., 2000. Magnitude-distance relations for earthquake-induced


landslides in Greece. Engineering Geology 58, 377-38.

Papadopoulos, G. A., 2003. Tsunami Hazard in the Eastern Mediterranean: Strong


Earthquakes and Tsunamis in the Corinth Gulf, Central Greece. Natural Hazards 29
(3), 437-464.

Papatheodorou, G., Ferentinos, G., 1993. Sedimentation processes and basin-filling


depositional architecture in an active asymmetric graben: Strava graben, Gulf of
Corinth, Greece. Basin Research 5, 235-253.

Papatheodorou, G., Ferentinos, G., 1997. Submarine and coastal sediment failure triggered by
the 1995, M , = 6.1 R Aegion earthquake, Gulf of Corinth, Greece, Holocene. Marine
Geology 137 (3), 287-304.

Paluska, A., Poetsch, T. & Bargu, S. 1989. In: Zschau, J. & Ergu¨nay, O. (eds) Tectonics,
paleoseismic activity and recent deformation mechanism in the Sapanca–Abant region
(NW Turkey, North Anatolian Fault Zone) Proceedings of the Turkish–German
Earthquake Research Project. Istanbul University, 18–32.

Papazachos, B., Papazachou, K., 1997. The Earthquakes of Greece. Editions Ziti,
Thessaloniki, 71–76.

Parés, J. M., 2004 How deformed are weakly deformed mudrocks? Insights from magnetic
anisotropy. Geological Society of London, Special Publications 238, 191-203.

Parke, J. R., Minshull, T. A., Anderson, G., White, R. S., McKenzie, D., Kuscu, I., Bull, J.
M., Görür, N., Sengör, C., 1999. Active faults in the Sea of Marmara, western
Turkey, imaged by seismic reflection profiles. Terra Nova 11, 223-227.

Parsons, T., 2004. Recalculated probability of M • 7 earthquakes beneath the Sea of


Marmara, Turkey, J. Geophys. Res. 109, B05304.

Parsons, T., Stein, R. S., Dieterich, J. H., Barka, A., Toda, S., 2000. Influence of the 17
August 1999 Izmit earthquake on seismic hazards in Istanbul. In The 1999 Izmit and
Düzce Earthquakes: Preliminary Results, 295-310 pp..

272
Perissoratis, C., Mitropoulos, D., Angelopoulos, I., 1984. The role of earthquakes in inducing
sediment mass movements in the eastern Corinthiakos Gulf: an example from the
February 24-March 4 activity. Marine Geology 55, 35-45.

Perissoratis, C., Piper, D.J.W., Lykousis, V., 1993. Late Quaternary sedimentation in the Gulf
of Corinth: the effects of marine-lake fluctuations driven by eustatic sea level changes.
Special Publication Dedicated to Prof. A. Panagos. Nat. Tech. Univ. of Athens, 693-
744.

Perissoratis, C., Piper, D. J., Lykousis, V., 2000. Alternating marine and lacustrine
sedimentation during the late Quaternary in the Gulf of Corinth rift basin, central
Greece: Marine Geology 167, 391-411.

Pettijohn, F. J., Potter, P. E., Siever, E., 1987. Sand and Sandstone. Springer-Verlag, London,
553 pp.

Pickering, K. T., Hiscott, R. N., 1985. Contained (reflected) turbidity currents from the
Middle Ordovician Cloridorme Formation, Quebec, Canada: An alternative to the
antidune hypothesis. Sedimentology 32, 373-394.

Piper, D. J. W., Stamatopoulos, L., Poulimenos, G., Doutsos, T., Kontopoulos, N., 1990.
Quaternary history of the Gulfs of Patras and Corinth, Greece. Z. Geomorphol., 34(4),
451-458.

Piper, D .J. W., Cochonat, P., Ollier, G., Le Drezen, E., Morrison, M, Baltzer, A., 1992.
Evolution progressive d’un glissement rotationnel en un courant de turbidité: cas du
séisme de 1929 des Grands Bancs (Terre Neuve). C.R. Acad. Sc. Paris, 314:1057-
1064.

Piper, D. J. W., Cochonat, P., Morrison, M., 1999. The sequence of events around the
epicentre of the 1929 Grand Banks earthquake: initiation of debris flows and turbidity
current inferred from sidescan sonar. Sedimentology 46 (1), 79-97.

Piper, D. J. W., Normark, W. R., 2009. Processes That Initiate Turbidity Currents and Their
Influence on Turbidites: A Marine Geology Perspective. Journal of Sedimentary
Research 79 (6), 347-362.

Pirazzoli, P. A., Stiros, S. C., Fontunge, M., Arnold, M., 2004. Holocene and Quaternary
uplift in the central part of the southern coast of the Corinth Gulf (Greece). Mar. Geol.
212, 35-44.

Polonia, A., Gasperini, L., Amorosi, A., Bonatti, E., Bortoluzzi, G., Ça÷atay, M. N.,
Capotondi, L., Cormier, M.-H., Görür, N., McHugh, C. M. G., Seeber, L., 2004.
Holocene slip rate of the North Anatolian Fault beneath the Sea of Marmara, Earth and
Planetary Science Letters 227:411-426.

Pondard, N., Armijo, R., King, G. C. P., Meyer, B., Flerit, F., 2007. Fault interactions in the
Sea of Marmara pull-apart (North Anatolian Fault): earthquake clustering and
propagating earthquake sequences. Geophysical J. International 171, 1185-1197.

273
Pouderoux, H., Lamarche, G., Proust, J.-N., 2012. Building a 18,000 Year-Long Paleo-
Earthquake Record from detailed Deep-Sea Turbidite Characterization in Poverty Bay,
New Zealand. Natural Hazards Earth System Science 12, 2077-2101.

Poulos, S. E., Collins, M. B., Pattiaratchi, C., Cramp, A., Gull, W., Tsimplis, M.,
Papatheodorou, G., 1996. Oceanography and sedimentation in the semi-enclosed,
deep-water Gulf of Corinth (Greece). Marine Geology 134, Issues 3-4, 213-235.

R
Ramsey, B. C., 2001. Development of the Radiocarbon calibration program OxCal,
Radiocarbon, 43, 355-363, Proceedings of 17th International 14C Conference.

Rangin, C., Demirba÷, E., ømren, C., Crusson, A., Normand, A., Le Drezen, E., Le Bot, A.,
2001. Marine Atlas of the Sea of Marmara (Turkey). Ifremer Brest Technology
Center, 11 Plates and 1 Booklet.

Reichel, T., and Halbach, P., 2007. An authigenic calcite layer in the sediments of the Sea of
Marmara. A geochemical marker horizon with paleoceanographic significance. Deep-
Sea Research II 54 (11/13): 1201–1215.

Reilinger, R. E., McClusky, S. C., Oral, M. B., King, R. W., Toksfz, M. N., 1997. Global
positioning system measurements in the Arabia–Africa–Eurasia plate collision zone. J.
Geophys. Res. 102 (B5), 9983-9999.

Reilinger, R., McClusky, S., Vernant, P., Lawrence, S., Ergintav, S., Cakmak, R., Ozener, H.,
Kadirov, F., Guliev, I., Stepanyan, R., Nadariya, M., Hahubia, G., Mahmoud, S., Sakr,
K., ArRajehi, A., Paradissis, D., Al-Aydrus, A., Prilepin, M., Guseva, T., Evren, E.,
Dmitrotsa, A., Filikov, S.V., Gomez, F., Al-Ghazzi, R., and Karam, G., 2006. GPS
constraints on continental deformation in the Africa-Arabia-Eurasia continental
collision zone and implications for the dynamics of plate interactions. Journal of
Geophysical Research: Solid Earth 111 (B5), B05411.

Reimer, P. J., McCormac, F. G., 2002. Marine radiocarbon reservoir corrections for the
Mediterranean and Aegean Seas. Radiocarbon 44, 159-166.

Reimer, P. J., Baillie, M. G. L., Bard, E., Bayliss, A., Beck, J. W., Blackwell, P. G., Ramsey,
B. C., Buck, C. E., Burr, G. S., Edwards, R. L., Friedrich, M., Grootes, P. M.,
Guilderson, T. P., Hajdas, I., Heaton, T. J., Hogg, A. G., Hughen, K. A., Kaiser, K. F.,
Kromer, B., McCormac, F. G., Manning, S. W., Reimer, R. W., Richards, D. A.,
Southon Talamo, J. R. S., Turney, C. S. M., van der Plicht, J., Weyhenmeyer, C. E.,
2009. IntCal09 and Marine09 radiocarbon age calibration curves, 0-50,000 years cal
BP. Radiocarbon 51 (4), 1111-1150.

Renfrew, C., 1973, Before Civilisation: The Radiocarbon Revolution and Prehistoric Europe.
London: Jonathan Cape.

Richter, D.K., Neuser, R.D., 1987. Marine Aragonit-Ooide UND BRACKISCHE Mg-Calcit-
Ooide in “Neogen”/Pleistozan- Megazyklen des Kanaleinschnitts von Korinth
(Griechenland). Heidelb. Geowiss. Abh. 8, 184-186.

274
Richter, D. K., Anagnostou, Ch., Lykousis, V., 1993. Aragonitishe Whiting-Ablagerungen in
plio-/pleistozänen Mergelsequenzen bei Korinth (Griechenland). Zbl. Geol. Paläont.
Teil I, Heft 6, 675-688.

Richter, T. O., van der Gaast, S., Koster, B., Vaars, A., Gieles, R., de Stigter, H.C., De Haas,
H., Van Weering, T. C. E., 2006. The Avaatech XRF core scanner: technical
description and applications to NE Atlantic sediments. In: Rothwell, R.G., (Ed.), New
techniques in sediment core analysis. Geological Society of London 267, 39-50.

Ritt, B., Sarrazin, J., Caprais, J-C., Noël, P., Gauthier, O., Pierre, C., Henry, P., Desbruyères,
D., 2010. First insights into the structure and environmental setting of cold-seep
communities in the Marmara Sea, Deep Sea Research Part I. Oceanographic Research
Papers 57, Issue 9, 1120-1136.

Robertson, A. H. F., Dixon, J. E., 1984. Introduction: aspects of the geological evolution of
the Eastern Mediterranean. In: Robertson, A. H. F., Dixon, J. E., (Eds.). The
geological evolution of the eastern Mediterranean. Geo. Soc. Of London, Speccial
Publication 17, 1-74.

Robertson, A. H. F., Clift, P. D., Degnan, P. J., Jones, G., 1991. Palaeogeographic and
palaeotectonic evolution of the eastern Mediterranean Neotethys. In: Channell, J. E.
T., Winterer, E. L., Jansa, L. F., (Eds.). Palaeogeography and Paleoceanography of
Tethys: Palaeogeog., Palaeoclim., Palaeoecol. 87, 289-343.

Robertson, A. H. F., 1998. Mesozoic–Tertiary tectonic evolution of the easternmost


Mediterranean area: integration of marine and land evidence. In: Robertson, A. H. F.,
Emeis, K. C., Richter, C., Camerlenghi, A. (Eds.). Proceeding of the Ocean Drilling
Program, Scientific Results, vol. 160, 723– 782.

Robinson, S.G., 1986. The late Pleistocene palaeoclimatic record of North Atlantic deepsea
sediments revealed by mineral magnetic measurements. Physics of the Earth and
Planetary Interiors 42, 22-47.

Rochette, P., Jackson, M., Aubourg, C., 1992. Rock magnetism and the interpretation of
anisotropy of magnetic susceptibility. Reviews of Geophysics 30, 3, 209-226.

Rochette, P., Vadeboin, F., Clochard, L., 2001. Rock magnetic applications of Halbach
cylinders. Physics of the Earth and Planetary interiors, 126, 109-117.

Rockwell T., Barka, A., Dawson, T., Akyüz ,S., Thorup, K., 2001. Paleoseismology of the
Gaziköy-Saros segment of the North Anatolian Fault Northwestern Turkey:
comparison of the historical and paleoseismic records, implications of regional seismic
hazard, and models of earthquake recurrence. J. Seismol. 5 (3), 433-448.

Rodríguez-Pascua, M.A., Calvo, J.P., De Vicente, G., Gómez Gras, D., 2000. Soft-sediment
deformation structures interpreted as seismites in lacustrine sediments of the Prebetic
Zone, SE Spain, and their potential use as indicators of earthquake magnitudes during
the Late Miocene. Sedimentary Geology 135, 117-135.

275
Rodríguez-Pascua, M.A., De Vicente, G., Calvo, J.P., 2001. Paleoseismological analysis of
late Miocene lacustrine successions in the Prebetic Zone, SE Spain. Acta Geologica
Hispanica 36 (3-4), 213-232.

Rodríguez-Pascua, M. A., De Vicente, G., Calvo, J. P., Perez-Lopez, R. 2003. Similarities


between recent seismic activity and paleoseismites during the late miocene in the
external Betic Chain (Spain): relationship by the b value and the fractal dimension.
Journal of Structural Geology 25, 749-763.

Rodríguez-Pascua, M. A., Garduño-Monroy, V. H., Israde-Alcántara, I., Pérez-López, R.,


2010. Estimation of the paleoepicentral area from the spatial gradient of deformation
in lacustrine seismites (Tierras Blancas Basin, Mexico). Quaternary International 219
(1-2), 66-78.

Rodríguez-Lopez, J., Meléndez, N., Soria, A. R., Liesa, C., Van Loon, A. 2007. Lateral
variability of ancient seismites related to differences in sedimentary facies (the synrift
Escucha Formation, mid-Cretaceous, eastern Spain). Sedimentary Geology 201, 461-
484.

Royden, L. H., 1993. The tectonic expression slab pull at continental convergent boundaries
Tectonics 2 (2), 303-325

Ryan, W. B. F., Pitman III, W. C., Major, C. O., Shimkus, K., Moskalenko, V., Jones, G. A.,
Dimitrov, P., Gorür, N., Sakinç, M., Yüce, H., 1997. An abrupt drowning of the Black
Sea shelf. Marine Geology 138, 119-126.

Ryan, W. B. F., Pitman III, W. C., 1999. Noah’s Flood: The New Scientific Discoveries
About the Event That Changed History. Simon and Schuster, New York. 319 pp.

Ryan, W. B. F., Major, C. O., Lericolais, G., Goldstein, S. L., 2003. Catastrophic flooding of
the Black Sea, Annu. Rev. Earth Planet. Sci. 31, 525-554.

S
Sachpazi, M., Galvé, A., Laigle, M., Him, A., Sokos, E., Serpestsidaki, A., Marthelot, J.-M.,
Pi Alperin, J. M., Zelt, B., Taylor, B., 2007. Moho topography under central Greece
and its compensation by Pn time-units for accurate location of hypocenters: the
example of the Gulf of Corinth 1995 Aigion earthquake, Tectonophyiscs 440, 53-65.

Sakellariou, D., Lykousis, V., Papanikolaou, D., 1998. Neotectonic structure and evolution of
the Gulf of Alkyonides, central Greece. Bull. Geol. Soc. Greece 32, 241-250.

Sakellariou, D., Lykousis, V., Papanikolaou, D., 2001. Active faulting in the Gulf of Corinth,
Greece. In: Briand, F., (Ed.), 36th CIESM Congress Proceeding, 36 pp.

Sakellariou, D., Lykousis, V., Alexandri, S., Kaberi, H., Rousakis, G., Nomikou, P.,
Georgiou, P., Ballas, D., 2007. Faulting, seismic-stratigraphic architecture and Late
Quaternary evolution of the Gulf of Alkyonides Basin, East Gulf of Corinth, Central
Greece. Basin Research 19, 273-295.

276
Sarõ, E., Ça÷atay, M. N., 2006. Turbidites and their association with past earthquakes in the
deep Cõnarcõk Basin of the Marmara Sea. Geo-Marine Letters, 26 (2), 69-76.

Schermer E. R., 1993. Geometry and kinematics of continental basement deformation during
the Alpine orogeny, Mr. Olympos region, Greece. J. Struc. Geol. 15, Issues 3-5, 571-
591.

Schwehr, K., Driscoll, N., Tauxe, L., 2007. Origin of continental margin morphology:
Submarine-slide or downslope current-controlled bedforms, a rock magnetic approach.
Marine Geology 240, 19-41.

Shillington, D.J., Seeber, L., Sorlien, C.C., Steckler, M.S., Kurt, H., Dondurur, D., Cifici, G.,
Imren, C., Cormier, M.-H., McHugh, C.M.G., Gurcay, S., Poyraz, D., Okay, S., Atgin,
O., Diebold, B., 2012. Evidence for widespread creep on the flanks of the Sea of
Marmara transform basin from marine geophysical data. Geology, 40:439-42,
doi:10.1130/G32652.1

Schindler, C., 1998. Geology of NW Turkey: results of the Marmara polyproject. In:
Schindler, C., Pfster, M., (Eds.). Active Tectonics of Northwestern Anatolia—The
Marmara Poly-Project, a multidisciplinary approach by space-geodesy, geology,
hydrology, geothermics and seismology. Verlag der Fachvereine, Zurich.

Schmalz, R. F., 1971. Formation of beach rock at Eniwetok Atoll. In: Bricker, O. P., (Ed.)
Carbonate Cements. Maryland: Johns Hopkins University Press, 17-24 pp..

Schneider, D. A., Kent, D. V., 1988. Inclination anomalies from Indian Ocean sediments and
the possibility of a standing nondipole field. Journal of Geophysical Research 93, B10,
11, 621-630.

Seeber, L., Cormier, M. H., McHugh, C., Emre, O., Polonia, A., Sorlien, C., 2006. Rapid
subsidence and sedimentation from oblique slip near a bend on the North Anatolian
transform fault in the Marmara Sea, Turkey. Geology 34 (11).

Seilacher, A., 1969. Fault-graded beds interpreted as seismites. Sedimentology 13, 155-159.

Selley, R. C., 2000. Applied Sedimentology. Royal School of Mines Imperial College of
Science, Technology, and Medicine London, United Kingdom, 523 pp.

Sengör, A. M. C., 1979. The North Anatolian transform fault: its age, offset and tectonic
significance. Geol. Soc. London 136, 269-282.

ùengör, A. M. C., Kidd, W. S. F., 1979. Post-collisional tectonics of the Turkish- Iranian
plateau and a comparison with Tibet. Tectonophysics 55 (3-4), 361-376.

Sengör, A. M. C., Yõlmaz, Y., 1981. Tethyan evolution of turkey : a plate tectonic approach.
Tectonophysics 75 (3-4), 181-190.

Sengor, A. M. C., Gorur, N., Saroglu, F., 1985. Strike-slip faulting and related basin
formation in zones of tectonic escape: Turkey as a case study. In: Biddle, K. T.,

277
Christie-Blick, N. (Eds.) Strike-slip deformation, Basin formation and sedimentation.
Soc. Econ. Paleont. Min., Sp. Publ. 37, 227–264.

Sengör, A. M. C., Tuysuz, O., Imren, C., Sakinc, M., Eyidogan, H., Gorur, N., Le Pichon, X.,
Rangin, C., 2004. The North Anatolian Fault. A new look. Annual Review of Earth
and Planetary Sciences 33, 37-112.

Shiki, T., Cita, M. B., Gorsline, D. S., 2000. Sedimentary features of seismites, seismo-
turbidites and tsunamiites - an introduction. Sedimentary Geology 135, 7-9.

Siani, G., Paterne, M., Arnold, M., Bard, E., Metivier, B., Tisnerat, N., Bassinot, F., 2000.
Radiocarbon reservoir ages in the Mediterranean Sea and Black Sea. Radiocarbon 42,
271-280.

Siegenthaler, C., Finguer, W., Kelts, K., Suming, W., 1987. Earthquake and seiche deposits in
Lake Lucerne, Switzerland, Eclogae Geologicae Helvetiae 80 (1), 241-260.

Sieh, K. E., (1981). A review of geological evidence for recurrence times for large
earthquakes. In: Simpson, D. W., Richards, P. G., (Eds.). Earthquake Prediction: An
International Review, Maurice Ewing Ser., American Geophysical Union, vol. 4, 181-
207.

Sims, J., 1973. Earthquake-induced structures in sediments of Van Norman Lake, San
Fernando, California. Science 182, 161-163.

Siyako M., Bürkan K. A., Okay, A. I., 1989. Tertiary geology and hydrocarbon potential of
the Biga and Gelibolu peninsulas. Turkish Assoc. Petrol. Geol. Bull. 1, 183-200.

Sodoudi, F., Bruestle, A., Meier, T., Kind, R., Friederich, W., EGELADOS working group.
2013. New constraints on the geometry of the subducting African plate and the
overriding Aegean plate obtained from P receiver functions and seismicity. Solid
Earth Discuss. 5, 427-461.

Sorel, D., 2000. A Pleistocene and still-active detachment fault and the origin of the Corinth-
Patras rift, Greece. Geology 28, 83-86.

Soter, S., 1999. Holocene uplift and subsidence of the Helike Delta, Gulf of Corinth, Greece.
In: Stewart, I., Vita-Finzi, C., (Eds.), Coastal Tectonics. Geol. Soc., London, Spec.
Publ. 146, 41-56.

Stampfli, G. M., 2000. Tethyan oceans. In: Bozkurt, E., Winchester, J.A., Piper, J. D. A.,
(Eds.). Tectonics and Magmatism in Turkey and the Surrounding Area. Geological
Society Special Publications, vol. 173, 1-23.

Stampßi, G. M., Borel, G. D., 2004. The transmed transect in space and time: constraints on
the paleotectonic evolution of the mediterranean domain. In: Cavazza, W., Roure, F.,
Spakmen, W., Stampfli, G. M., Ziegler, P., (Eds.). The TRANSMED Atlas, Springer,
Verlag, Berlin, 53-80.

278
Stefatos, A., Papatheodorou, G., Ferentinos, G., Leeder, M., Collier, R., 2002. Seismic
reflection imaging of active offshore faults in the Gulf of Corinth: their seismotectonic
significance. Basin Research 14 (4), 487-502.

Stefatos, A., Charalambakis, M., Papatheodorou, G., Ferentinos, G., 2006. Tsunamigenic
sources in an active European half-graben (Gulf of Corinth, Central Greece). Marine
Geology 232, 35-47.

Stegmann, S., Strasser, M., Anselmetti, F. S., Kopf, A., 2007. Geotechnical in situ
characterisation of subaquatic slopes: The role of pore pressure transients versus
frictional strength in landslide initiation. Geophysical Research Letters 34, L07607, 1-
5.

Stegmann, S., Strasser, M., Anselmetti, F. S., and Kopf, A., 2007. Geotechnical in situ
characterisation of subaquatic slopes: the role of pore pressure transients versus
frictional strength in landslide initiation, Geophysical Research Letter, 34, L07607,
doi:10.1029/2007GLO29881.

Stein, R. S., Barka, A. A., Dieterich, J. H., 1997. Progressive failure on the North Anatolian
fault since 1939 by earthquake stress triggering. Geophysical Journal International
128, 594-604.

Stiros, S. C., Arnold, M., Pirazzoli, P. A., Laborel, J., Laborel, F., Papageorgiou, S., 1992.
Historical coseismic uplift on Euboea Island, Greece. Earth Planet. Sci. Lett. 108, 109-
117.

Stoner, J. S., Channell, J. E. T., Hillaire-Marcel, C., 1996. The magnetic signature of rapidly
deposited detrital layers from the deep Labrador sea: Relationship to North Atlantic
Heinrich layers. Paleoceanography 11, 309-325.

Stoner, J. S., St-Onge, G., 2007. Chapter Three: Magnetic Stratigraphy in Paleoceanography:
Reversals, Excursions, Paleointensity, and Secular Variation. In: Hillaire-Marcel, C.,
De Vernal, A., (Eds.), Developments in Marine Geology, Elsevier, Volume 1, 99-138.

Strasser, M., Anselmetti, F. S., Fäh, D., Giardini, D., Schnellmann, M., 2006. Magnitudes and
source areas of large prehistoric northern Alpine earthquakes revealed by slope
failures in lakes. Geology 34 (12), 1005-1008.

Straub, C., Kahle, H. G., Schindler, C., 1997. GPS and geologic estimates of the tectonic
activity in the Marmara Sea region, NW Anatolia. J. Geophys. Res. Solid Earth B12,
27587-27601.

Stuiver, M, Polach H. A., 1977. Discussion: reporting of 14C data. Radiocarbon 19 (3), 355-
63.

Stuiver, M., Pearson, G. W., Braziunas, T. F., 1986. Radiocarbon age calibration of marine
samples back to 9000 cal yr BP. In: Stuiver, M., Kra, R. S., (Eds.), Radiocarbon. 28,
980-1021.

279
Stuiver, M., Brazunias, T. F., Becker, B., Kromer, B., 1991. Climatic solar, oceanic and
geomagnetic influences on Late-Glacial and Holocene atmospheric 14C/12C change,
Quaternary Research, 35, 1-24.

Stuiver, M., Reimer, P. J., 1993. Extended C-14 data-base and revised CALIB 3.0 C-14 age
calibration program. Radiocarbon 35, 215-30.

Stuiver, M., Reimer, P. J., Reimer, R., 2005. CALIB Radiocarbon Calibration 5.0.2,
http://radiocarbon.pa.qub.ac.uk.

Sturm, M., Siegenthaler, C., Pickrill, R. A., 1995. Turbidites and “homogenites”: a conceptual
model of flood and slide deposits. Publication of the International Association of
Sedimentologists, 16th Regional Meeting, Paris, 22-40.

Suzanne, P., Lyberis, N., Chorowicz, J., Nurlu, M., Yurur, T., Kasapoglu, E., 1990. La
geometrie de la faille nord anatolienne a partir d'images Land SAT-MSS. Bulletin de
la Societe Geologique de France VI, 589-599.

Syvitski, J. P. M, Schafer, C. T., 1996. Evidence for an earthquake-triggered basin collapse in


Saguenay Fjord, Canada. Sedimentary Geology 104 (1-4), 127-153.

T
Tary, J.B., Géli, L., Guennou, C., Henry, P., Sultan, N., Cagatay, N., and V. Vidal, 2012.
Microevents produced by gas migration and expulsion at the seabed: a study based on
sea bottom recordings from the Sea of Marmara. Geophysical Journal International,
190:993-1007.

Tarling, D. H., Hrouda, F., 1993. The Magnetic Anisotropy of Rocks. Chapman & Hall,
London, 218 pp.

Taylor, B., Weiss, J. R., Goodliffe, A. M., Sachpazi, M., Laigle, M., HIRN, A., 2011. The
structures, stratigraphy and evolution of the Gulf of Corinth rift, Greece. Geophys. J.
Int. 185, 1189-1219.

Taymaz, T., Yilmaz, Y., Dilek, Y., 2007. The Geodynamics of the Aegean and Anatolia:
Introduction. In: Taymaz, T., Yilmaz, Y., Dilek, Y., (Eds.). The Geodynamics of the
Aegean and Anatolia . Geological Society of London Special Publications 291, 1-16.

Thunell, R., Tappa, E., Varela, R., Llano, M., Astor, Y., Muller-Krarger, F., Bohrer, F., 1999.
Increased marine sediment suspension and fluxes following an earthquake. Nature
398, 233-236.

Tiberi, C., Diament, M., Lyon-Caen, H., King, T., 2001. Moho topography beneath the
Corinth Rift area (Greece) from inversion of gravity data. Geophysical Journal
International 145, 797-808.

Tjallingii, R., Röhl, U., Kölling, M., Bickert, T., 2007. Influence of the water content on X-
ray fluorescence core-scanning measurements in soft marine sediments.
Geochemistry, Geophysics, Geosystems, 8 (2), Q02004.

280
Toksöz, M.N., Shakal, A.F., Michael, A.J. 1979. Space-time migration of earthquakes along
the North Anatolian fault zone and seismicity gaps. Pure and Applied Geophysics, 924
117:1258-1270.

Trentesaux, A., Recourt, P., Bout-Roumazeilles, V., Tribovillard, N., 2001. Carbonate Grain-
Size Distribution in Hemipelagic Sediments from a Laser Particle Sizer. Journal of
Sedimentary Research 71 (5), 858-862.

Tryon, M.D., Henry, P., Ça÷atay, M.N., Zitter, T.A.C., Géli, M.L., Gasperini, L., Burnard, P.,
Bourlange, S., Grall, C., 2010. Pore fluid chemistry of the North Anatolian Fault Zone
in the Sea of Marmara: a diversity of sources and processes. Geochemistry,
Geophysics, Geosystems 11.

Turgut S., Türkaslan M., Perinçek, D., 1991. Evolution of the Thrace sedimentary basin and
its hydrocarbon prospectivity. In: Spencer, A. M., (Ed.), Generation, accumulation,
and production of Europe’s hydrocarbons. Spec. Publ. Eur. Assoc. Petrol. Geosci. 1,
415-437.

U
Uçarkuú, G., Çakõr, Z., Armijo, R., 2011. Western Unitination of the Mw 7.4, 1999 Izmit
Earthquake Rupture: Implications for the Expected Large Earthquake in the Sea of
Marmara. Turkish Journal of Earth Sciences. doi:10.3906/yer-0911-72.

Uçarkuú, G., 2010. Active faulting and earthquake scarps along the North Anatolian Fault in
the Sea of Marmara. PhD Thesis, Istanbul Technical University. Turkey, 173 pp.

U.S. Geological Survey, 2000., Implications for Earthquake Risk Reduction in the United
States from the Kocaeli, Turkey Earthquake of August 17, 1999. U. S. Geol. Surv.
Circ., 1193, 64 pp.

Utkucu, M., Kanbur, Z., Alptekin, Ö., Sünbül, F., 2009. Seismic behavior of the North
Anatolian Fault beneath the Sea of Marmara (NW Turkey): implications for
earthquake recurrence times and future seismic hazard. Natural Hazards 50, Issue 1,
45-71.

V
Valsecchi, V., Sanchez Goñi, M. F., Londeix, L., 2012. Climate of the Past Vegetation
dynamics in the Northeastern Mediterranean region during the past 23 000 yr: insights
from a new pollen record from the Sea of Marmara. Climate of the Past 8(6), 1941-
1956.

Van Loon, A. J., Brodzikowski, K., Zielinski, T., 1995. Shock-induced resuspension deposits
from a Pleistocene proglacial lake (Kleszczow Graben, central Poland). Journal of
Sedimentary Research 65 (2), 417-422.

van Welden, A., 2007. Enregistrements sédimentaires imbriqués d’une activité sismique et de
changements paléoenvironnementaux. Etude comparative de différents sites: Golfe de
Corinthe (Grèce), Lac de Shkodra (Albanie/Monténégro), Golfe de Cariaco
(Vénézuela). PhD Thesis, Université de Savoie, France, 287 pp.

281
Vidal, L., Ménot, G., Joly, C., Bruneton, H., Rostek, F., Ça÷atay, M. N., Major, C., Bard, E.,
2010. Hydrology in the Sea of Marmara during the last 23000 years: Implications for
timing of Black Sea connections and sapropel deposition. Paleoceanography 25 (1), 1-
16.

W
Walcott, C. R.,White, S. H., 1998. Constraints on the kinematics of post-orogenic extension
imposed by stretching lineations in the Aegean region. Tectonophysics 298, Issue (1-
3), 155-175.

Walker, M. (2005). Quaternary Dating Methods, 304 p. John Wiley & Sons, New York.

Wegmann, K. W., Pazzaglia, F. J., 2009. Late Quaternary fluvial terraces of the Romagna and
Marche Apennines, Italy. Quat. Sci. Rev. 28, 137-165.

Westaway, R., 2002. The Quaternary evolution of the Gulf of Corinth, central Greece:
coupling between surface processes and flow in the lower continental crust.
Tectonophysics 348 (4), 269-318.

Wetzler, N., Marco, S., Heifetz, E., 2010. Quantitative analysis of shear-induced turbulence in
lake sediments. Geology 38, 4:303-306.

Wilhelm, B., Arnaud, F., Enters, D., Allignol, F., Legaz, A., Magand, O., Revillon, S.,
Giguet-Covex, C., Malet, E., 2012. Does global warming favour the occurrence of
extreme floods in European Alps? First evidences from a NW Alps proglacial lake
sediment record. Climatic Change 113, Issue 3-4, 563-581.

William M. Last 2002, Textural Analysis of Lake Sediments. Developments in


Paleoenvironmental Research Volume 2, 41-81 pp.

Wong, H. K., Luedmann, T., Ulug, A., Görür, N., 1995. The Sea of Marmara: A plate
boundary sea in an escape tectonic regime. Tectonophysics 244, 231-250.

Y
Yaltõrak, C., 2002. Tectonic evolution of the Marmara Sea and its surroundings, Marine
Geology 190 (1-2), 493-529.

Yaltõrak, C., Alpar, B., 2002. Evolution of the middle strand of North Anatolian Fault and
shallow seismic investigation of the southeastern Marmara Sea (Gemlik Bay), Marine
Geology 190 (1-2), 307-327.

Yaltirak, C., Sakinc, M., Aksu, A. E., Hiscott, R. N., Galleb, B., Ulgen, U. B., 2002. Late
Pleistocene uplift history along the southwestern Marmara Sea deunitined from raised
coastal deposits and global sea-level variations. Marine Geology 190.

Yaltirak, C., Sakinç, M., Aksu, A. E., Hiscott, R. N., Galleb, B., Ulgen, U. B., 2002. Late
Pleistocene uplift history along the Southwestern Marmara Sea deunitine from raised
coastal deposits and global sea-level variations. Marine Geology 190 (1-2), 283-305.

282
Yil÷itbaú, E., Kerrich, R., Yilmaz, Y., Elmas, A., Xie, Q. L., 2004. Characteristics and
geochemistry of Precambrian ophiolites and related volcanics from the østanbul-
Zonguldak Unit, Northwestern Anatolia, Turkey: following the missing chain of the
Precambrian South European suture zone to the east. Precambrian Research 132, 179-
206.

Youd, T. L., 1973. Liquefaction, flow, and associated ground failure. U. S. Geol. Surv. 688,
1-12.

Z
Zachariadis, P., 2007. Ophiolites of the eastern Vardar Zone, N. Greece. PhD Thesis,
University of Mainz, Germany, 131 pp
Zavala, C., Arcuri, M., Di Meglio, M., Gamero Diaz, H., Contreras, C., 2011. A genetic facies
tract for the analysis of sustained hyperpycnal flow deposits. In: Slatt, R. M., Zavala,
C., (Eds.), Sediment transfer from shelf to deep water—Revisiting the delivery
system: AAPG Studies in Geology 61, 31-51.

Ziegler, M., Jilbert, T., de Lange, G. J., Lourens, L. J., Reichart, G.-J., 2008. Bromine counts
from XRF scanning as an estimate of the marine organic carbon content of sediment
cores. Geochemistry, Geophysics, Geosystem 9 (5), Q05009.

Zitter, T. A. C., Henry, P., Aloisi, G., Delaygue, G., Ça÷atay, M. N., de Lépinay, B. M., Al-
Samir, M., Fornacciari, F., Tesmer, M., Pekdeger,A., Wallmann,K., Lericolais,G.,
2008. Cold seeps along the main Marmara Fault in the Sea of Marmara (Turkey).
Deep-Sea Research PartI—Oceanographic Research Papers 55 (4), 552-570.

Zitter, T. A. C., Grall, C., Henry, P., Özeren, M. S., Ça÷atay, M. N., ùengör, A. M. C.,
Gasperini, L., de Lépinay, B. M., Géli, L., 2012. Distribution, morphology and
triggers of submarine mass wasting in the Sea of Marmara. Marine Geology 329-33,
58-74.

283
Annex A
Table summarizing the principal earthquakes with magnitude Ms•6.8 occurred along the
North Anatolian Fault Zone in the Marmara region, NW Turkey, after 400 AD. Compiled by
Utkucu et al. (2009) from: Ambraseys and Finkel (1987, 1991, 1995); Papazachos and
Papazachou (1997); Ambraseys (2002).

No. Date Latitude Longitude Ms Region


1 01/04/407 40.9 28.7 6.8 Istanbul
2 25/09/437 40.8 28.5 6.8 Istanbul
3 06/11/447 40.7 30.3 7.2 Izmit
4 460 40.1 27.6 6.9 Erdek
5 25/09/478 40.7 29.8 7.3 Karamürsel
6 484 40.5 26.6 7.2 Gelibolu
7 16/08/554 40.7 29.8 6.9 Izmit
8 14/12/557 40.9 28.3 6.9 Silivri
9 26/10/740 40.7 28.7 7.1 Marmara
10 05/05/824 40.6 26.8 7 Barbaros
11 23/05/860 40.8 28.5 6.8 Marmara
12 09/01/869 40.8 29 7 Marmara
13 25/10/989 40.8 28.7 7.2 Marmara
14 08/01/1010 40.6 27 7.4 Gelibolu
15 23/09/1063 40.8 27.4 7.4 Barbaros
16 09/01/1900 40.4 30 6.8 Izmit
17 1231 41 28.6 6.9 Istanbul
18 01/06/1296 40.5 30.5 7 Geyve
19 18/10/1343 40.7 27.1 6.9 Ganos
20 18/10/1343 40.9 28 7 Ere÷li
21 01/03/1354 40.7 27 7.4 Gelibolu
22 15/03/1419 40.4 29.3 7.2 Bursa
23 10/09/1509 40.9 28.7 7.2 Istanbul
24 10/05/1556 40.6 28 7.1 Erdek
25 18/05/1625 40.3 26 7.1 Saros
26 17/02/1659 40.5 26.4 7.2 Saros
27 25/05/1719 40.7 29.8 7.4 Izmit
28 06/02/1737 40 27 7 Biga
29 29/07/1752 41.5 26.7 6.8 Edirne
30 02/09/1754 40.8 29.2 6.8 Izmit
31 22/05/1766 40.8 29 7.1 Marmara
32 05/08/1766 40.6 27 7.4 Ganos
33 28/02/1855 40.1 28.6 7.1 Bursa
34 21/08/1859 40.3 26.1 6.8 Saros
35 09/02/1893 40.5 26.2 6.9 Saros
36 10/07/1894 40.7 29.6 7.3 Izmit
37 09/08/1912 40.7 27.2 7.3–7.4 Ganos
38 09/08/1912 40.7 27 6.8 Ganos
39 18/03/1953 40.1 27.4 7.1 Gönen
40 26/05/1957 40.7 31 7.1 Abant
41 06/10/1964 40.1 28.2 6.8 Manyas
42 22/07/1967 40.7 30.7 7.2 Mudurnu
43 17/08/1999 29.9 40.7 7.4 Izmit

284
Annex B

Pictures of the MD01-2425 core

285
Section Section Section Section Section Section
I II III IV V VI
0

50
Length cm

100

150
Depth cm 150 300 450 500 650 800
Section Section Section Section Section Section
VII VIII IX X XI XII
0

50
Length cm

100

150
Depth cm 1050 1200 1350 1500 1650 1800
Section Sections Section Section Section Section
XIII XIV-XVI XVII XVIII XIX XX
0

50
Length cm

100

150
Depth cm 1950 2085 2235 2385 2535 2685
Section Section Section Section
XXI XXII XXIII XXV
0

150
Depth cm 2835 2985 3135 3244
Annex C

Calibration of radiocarbon dating and age-depth curve

1. Calibration of radiocarbon dating


The estimation of radiocarbon ages is based on the assumption that the concentration
of 14C in the atmosphere remained constant over time (Stuiver and Polach, 1977). However,
scientists working on tree-ring chronologies found significant discrepancies between the age
of wood based on dendrochronological dating and the age based on radiocarbon (Walker,
2005; Reimer et al., 2009), being younger the radiocarbon ages than the dendrochronological
14
dates (Renfrew, 1973; Walker, 2005). These results evidence that atmospheric C activity
fluctuated over time. The causes of these variations remain to be established, but a major
influence seem to be associated to the changes of the cosmic ray flux produced by changes in
the earth geomagnetic field or by variations in the intensity of solar activity, as well as,
changes in the carbon cycle (Stuiver et al., 1991, Nollerr et al., 2000).

To calibrate the radiocarbon age several currently internationally accepted calibration


curves have been created as: IntCal09 (northern hemisphere atmospheric; Reimer et al.,
2009), Marine09 (for marine dates; Reimer et al., 2009), and SHCal04 (southern hemisphere
atmospheric; McCormac et al., 2004). They provide a comprehensive summary of the current
state of knowledge of past variation in 14C (Reimer et al., 2009). The curves are based mainly
on the dendrochronological records. Away from the limit of continuous dendrochronological
series the calibration is based on matched uranium-series and radiocarbon dates on fossil and
corals, coupled with radiocarbon-dated organic material from laminated (annually
accumulating) marine sediments in the Cariaco Basin, Venezuela (Walker 2005). To calibrate
the radiocarbon age different software have been developed (e.g. Calib by Stuiver and
Reimer, (1993); Oxcal by Ramsey, (2001); Bcal by Buck et al. (1999)). These software are
using the internationally calibration curves mentioned above.

If organic matter was formed in marine or lacustrine environments, it is necessary to


perform a special correction, called: the reservoir correction. The 14C ages of marine samples
require a correction for the reservoir age of the ocean where they were formed, because the
14
ocean is a large carbon reservoir, the residence time of C is long compared to the

290
atmosphere. This effect together with upwelling of older carbon from the deep ocean give an
apparent age of marine samples several hundred years older than contemporaneous
atmospheric samples (reservoir correction Ra 400 years) (Reimer and McCormac, 2002).
Moreover, the reservoir effect can varies locally (ǻR ), in several hundred years or more
(Stuiver et al., 1986).

The radiocarbon calibrate age are estimate generally by two method, the intercept
method where the mean of the radiocarbon date intercepts the calibration curve (figure 1),
and the probabilistic method (the most used and most correct from the probabilistic point of
14
view) (Telford et al., 2003). Which consists in projecting the measured C age with their
normal distribute errors (Gaussian distribution in the y-axis) on the calibration curve (together
with its uncertainty) (figure 1). This method gives the probability of each calendar age to be
the age of the sample. The complex shape of the probability density function avoids a single
value summarizing it. The only correct data is the range of values where the calibrated date is
located (figure.1). In the Marmara Sea, we used a range calibration of 95% (2V).

Intercept method
The most probable age
940r20 cal yr BP

Figure 1: Correction of radiocarbon age (sample belonging to core MD01-2425 at 190m depth)
with software OxCal v.4.1 (Bronk Ramsey, 2001). To the left of the graph, two Gaussian distribution
of the radiocarbon data can be observed. The firs one (light pink) corresponds to the age no
corrected by the local variations of the reservoir effect (ǻR). The second one (dark pink) is the

The results are a range of values comprise between 1103 and 784 cal yr BP (2V). Additionally, in
corrected one. The calibration curve (Marine09) and its uncertainty are represented in blue colour.

this figure we represented the intercept method, with a most probable age of 940r20 cal yr BP.

Despite of the correct calibrated age is constituted by a range of values. Frequently,


especially for the construction of an age model, the probability distribution is simplified by a
single value. The choice can be made in a range of several statistical values: mode, median,

291
weighted average (Telford et al., 2004b). The most used and simplest is the median age.
However, the value of a more robust statistical point of view is the weighted average. It takes
into account the probability of the every calibrated age to be the most probable age. In the
Marmara Sea it was calculated for each radiocarbon date with the software Bcal (Table III.1).
In the figure 2 we plot the probability distribution of radiocarbon date obtained for one
sample taken at 190 cm depth. The age calculate by the method of weighted average is 951
cal yr BP (represented by an orange cross), whereas the age median is 943 cal yr BP
(represented by a blue cross), they differ by 8 years. In others samples of the Marmara Sea
this difference can be more important and can reach 80 yrs.

2V

Figure 2: Correction of
radiocarbon age (sample
belonging to core MD01-2425 at
Age range calibrated (95%) 190m depth) made with the
1094 to 793 cal yr BP software Bcal (Buck et al.,
1999).The result is a probability
distribution of a radiocarbon data.
The median age is represented by a
blue arrow and the weighted
average age by an orange arrow.

2. Age-depth curve

To calculate the age of undated events is necessary construct an age-depth curve.


Three basic types of curves are the most frequently used: linear interpolation, spline
interpolation, and polynomial line-fitting (Bennet, 1994). Different software as psimpoll
(Bennet 1994), Cagedepth.fun (Heegaard 2003) and Clam (Blaauw, 2010) allow the
construction of age-depth curves

 Linear interpolation: This is the most basic and frequently used age-depth curve.
Reported radiocarbon ages are connected by straight lines. It is a superficially crude
approach, but does provide a reasonable estimation for the ages and sedimentation

292
rate. However, it takes no account of the errors on the radiocarbon ages and involves
abrupt changes in sedimentation rate at the depths corresponding to the dated points,
which is far from the reality of a sedimentary basins infill (Bennet, 1994).

 Spline interpolation: is a polynomial curve, usually cubic (4-unit). The curve is fitted
between each pair of points, but whose coefficients are deunitined slightly nonlocally:
some information is used from other points than the pair under immediate
consideration (Bennet, 1994). This non-locality is intended to make the fitted curve
smooth overall, and not change gradient abruptly at each data point (Bennet, 1994). It
not involve abrupt changes in sedimentation rate. However, this method frequently
produces curves with sedimentation rates negative.

 Polynomial line-fitting: the polynomial curve have the follow form: y = a + bx + cx2
+ dx3 …., where x = depth and y = age. This curve does not necessarily have to pass
through all the points. It is calculated by minimizing the distance to each point by least
squares. The curve is much more "flexible" when the number of units used is higher.
The aim is to find a polynomial equation as simple as possible but with a reasonable
fit. As the spline interpolation, this curve does not involve abrupt changes in
sedimentation rate (Bennet, 1994).

293
Annex D

Pictures of the MD01-2477 core

294
Section Section Section Section Section Section Section
I II III IV V VI VII
0

50
Lenght cm

100

150
Depth cm 150 300 450 500 650 800 950
Annex E

Pictures of the MD01-2478 core

297
Section Section
I II
0

150
Depth cm 150 300
Annex F

Pictures of the MD01-2481 core

299
Section Section Section Section Section Section
I II III IV V VI

150
Depth cm 150 300 450 600 750 900
Section Section Section Section Section
VII VIII IX X XI
0

50
Length cm

100

150
Depth cm 1050 1200 1350 1500 1557

Вам также может понравиться