Você está na página 1de 22

222

ARTICLE
The geometry of the North Anatolian transform fault in the
Sea of Marmara and its temporal evolution: implications for
the development of intracontinental transform faults1
A.M. Celâl Şengör, Céline Grall, Caner İmren, Xavier Le Pichon, Naci Görür, Pierre Henry,
Hayrullah Karabulut, and Muzaffer Siyako

Abstract: The North Anatolian Fault is a 1200 km long strike-slip fault system connecting the East Anatolian convergent area
with the Hellenic subduction zone and, as such, represents an intracontinental transform fault. It began forming some 13–11 Ma ago
within a keirogen, called the North Anatolian Shear Zone, which becomes wider from east to west. Its width is maximum at the
latitude of the Sea of Marmara, where it is 100 km. The Marmara Basin is unique in containing part of an active strike-slip fault
system in a submarine environment in which there has been active sedimentation in a Paratethyan context where stratigraphic
resolution is higher than elsewhere in the Mediterranean. It is also surrounded by a long-civilised rim where historical records
reach well into the second half of the first millennium BCE (before common era). In this study, we have used 210 multichannel
seismic reflexion profiles, adding up to 6210 km profile length and high-resolution bathymetry and chirp profiles reported in the
literature to map all the faults that are younger than the Oligocene. Within these faults, we have distinguished those that cut the
surface and those that do not. Among the ones that do not cut the surface, we have further created a timetable of fault generation
based on seismic sequence recognition. The results are surprising in that faults of all orientations contain subsets that are active
and others that are inactive. This suggests that as the shear zone evolves, faults of all orientations become activated and
deactivated in a manner that now seems almost haphazard, but a tendency is noticed to confine the overall movement to a zone
that becomes narrower with time since the inception of the shear zone, i.e., the whole keirogen, at its full width. In basins, basin
margins move outward with time, whereas highs maintain their faults free of sediment cover, making their dating difficult, but
small perched basins on top of them in places make relative dating possible. In addition, these basins permit comparison of
geological history of the highs with those of the neighbouring basins. The two westerly deeps within the Sea of Marmara seem
inherited structures from the earlier Rhodope–Pontide fragment/Sakarya continent collision, but were much accentuated by the
rise of the intervening highs during the shear evolution. When it is assumed that below 10 km depth the faults that now
constitute the Marmara fault family might have widths approaching 4 km, the resulting picture resembles a large version of an
amphibolite-grade shear zone fabric, an inference in agreement with the scale-independent structure of shear zones. We think
that the North Anatolian Fault at depth has such a fabric not only on a meso, but also on a macro scale. Detection of such broad,
vertical shear zones in Precambrian terrains may be one way to get a handle on relative plate motion directions during those
remote times.

Résumé : La faille nord-anatolienne est constituée d’un système de failles décrochantes d’une longueur de 1200 km qui relie le
secteur convergent anatolien à la zone de subduction hellénique elle représente ainsi, une faille transformante intracontinen-
tale. Elle a commencé à se former il y a environ 13 à 11 Ma dans une ceinture déformée dominée par des déplacements de
décrochement (« keirogène »), nommé la zone de cisaillement nord-anatolienne; elle s’élargit d’est en ouest. Elle atteint sa
largeur maximale de 100 km à la latitude de la mer de Marmara. Le bassin de Marmara est unique en ce qu’il contient une partie
d’un système actif de failles de décrochement dans un environnement sous-marin où il y a eu de la sédimentation active dans un
contexte de mer Paratéthys et où la résolution stratigraphique est plus élevée qu’ailleurs en Méditerranée. La faille est aussi
entourée par une bordure où, grâce aux civilisations de longue date, il existe des documents historiques datant de la deuxième
moitié du premier millénaire avant notre ère. Dans la présente étude, nous avons utilisé les données obtenues par 210 profils de
réflexion sismique multicanale, ce qui donne un total de 6210 km de profils; nous avons aussi utilisé de la bathymétrie à haute
résolution et des profils de fluctuation de longueur d’onde disponibles dans la littérature pour cartographier toutes les failles
plus récentes que l’Oligocène. Pour ces failles, nous distinguons celles qui recoupent la surface et celles qui ne la recoupent pas.

Received 31 August 2013. Accepted 1 January 2014.


Paper handled by Editor Ali Polat. / Article confié au rédacteur en chef, Ali Polat.
A.M.C. Şengör. İstanbul Teknik Üniversitesi, Maden Fakültesi, Jeoloji Bölümü, Ayazağa 34469 İstanbul, Turkey; İstanbul Teknik Üniversitesi, Avrasya
Yerbilimleri Enstitüsü, Ayazağa 34469 İstanbul, Turkey.
C. Grall and P. Henry. CEREGE, CNRS-Aix Marseille Université, Marseille, France.
C. İmren. İstanbul Teknik Üniversitesi, Maden Fakültesi, Jeofizik Bölümü, Ayazağa 34469 İstanbul, Turkey.
X. Le Pichon. Professeur Honoraire au Collège de France, Europôle de l’Arbois, Bât. Trocadéro, BP 80 13 545, Aix-en-Provence cedex 04, France.
N. Görür. İstanbul Teknik Üniversitesi, Maden Fakültesi, Jeoloji Bölümü, Ayazağa 34469 İstanbul, Turkey.
H. Karabulut. Kandilli Rasathanesi, Boğaziçi Üniversitesi, Kandilli, İstanbul, Turkey.
M. Siyako. Türkiye Petrolleri Anonim Ortaklığı, Arama Grubu, Ankara, Turkey.
Corresponding author: A.M.C. Şengör (e-mail: sengor@itu.edu.tr).
1This article is one of a selection of papers published in this special issue in honour of John Tuzo Wilson, a man who moved mountains, and his

contributions to earth sciences.

Can. J. Earth Sci. 51: 222–242 (2014) dx.doi.org/10.1139/cjes-2013-0160 Published at www.nrcresearchpress.com/cjes on 4 February 2014.
Şengör et al. 223

Pour ces dernières, nous avons créé un tableau temporel de génération de failles basée sur la reconnaissance de la séquence
sismique. Les résultats sont surprenants; en effet, peu importe l’orientation des failles, elles contiennent toutes des sous-
ensembles actifs et des sous-ensembles non actifs. Cela suggère qu’à mesure de l’évolution de la zone de cisaillement, les failles
de toutes directions sont activées et désactivées d’une manière qui semble presque aléatoire. Cependant, nous notons une
tendance à restreindre le mouvement général à une zone qui se rétrécit avec le temps depuis de début de la zone de cisaillement,
c.-à-d. tout le « keirogène », à sa pleine largeur. Dans les bassins, avec le temps, les bordures se déplacent vers l’extérieur alors que
les zones surélevées gardent leurs failles libres de couverture sédimentaire, ce qui complique la datation. Toutefois, de petits
bassins à leur sommet permettent d’obtenir une datation relative. De plus, ces bassins permettent de comparer l’histoire
géologique des zones surélevées à celle des bassins avoisinants. Deux creux, situés à l’ouest dans la mer de Marmara, semblent
être des structures héritées de la collision entre le fragment Rhodope–Pontide et le continent Sakarya. Cependant, ils ont été
grandement accentués par le soulèvement des zones surélevées durant l’évolution du cisaillement. Lorsque nous partons de
l’hypothèse qu’à une profondeur supérieure à 10 km, les failles qui forment actuellement la famille de failles de Marmara
pourraient avoir des largeurs approchant les 4 km, l’image qui en ressort ressemble à une version agrandie d’une fabrique de
zone de failles au faciès des amphibolites. Cette inférence concorde avec une structure de zones de cisaillement, indépendante
de l’échelle. Nous croyons que la faille nord-anatolienne constitue un système qui a une telle fabrique en profondeur, non
seulement à échelle moyenne mais aussi à grande échelle. La détection de telles larges zones verticales de cisaillement dans des
terrains précambriens pourrait être une manière de comprendre les directions relatives des plaques à ces temps anciens.
[Traduction par la Rédaction]

In pietate memoriam J. Tuzo Wilson ment zone(s), constituting the North Anatolian Fault. To study
the manner of that localisation, the Sea of Marmara appears to
Introduction be an ideal location, because a large number of both multi-
In 1965, J. Tuzo Wilson defined a new kind of fault that could not channel seismic and chirp profiles, shallow gravity cores (down
be accounted for by Anderson’s theory of faulting in a continuous to 30 m from the seafloor, some of which are now dated and
medium (Anderson 1951; Wilson’s 1965 paper was also the founding enabled assessments to be made of the sedimentological events
paper of plate tectonics). Wilson pointed out that the new class of recognised in them: Beck et al. 2007; Eriş et al. 2012; Çağatay
faults served to connect two plate boundaries (the word “plate” is et al., in press), a number of deeper hydrocarbon exploration
Wilson’s) with one another by transforming the motion along them wells, recent seismic studies including the water column itself,
into another kind of motion (Fig. 1). He therefore called them “trans- plus a complete, detailed bathymetric coverage are available
form faults” (Wilson 1965). Because plate boundaries are so cleanly (for references prior to 2005, see Şengör et al. 2005; Armijo et al.
delineated in the oceans, there is seldom any doubt as to the nature 2005; Carton 2005; Gazioğlu et al. 2005; Bécel 2006; Cormier
of faulting in them (e.g., Şengör, in press), but the scattering of de- et al. 2006; Seeber et al. 2006; Carton et al. 2007; Dolu et al.
formation in the continents into areas of complex strain, in places as 2007; Géli et al. 2008; Laigle et al. 2008; Kuşçu 2009; Bécel et al.
wide as 3000 km (as in Asia now), makes it difficult to decide to what 2010; Özeren et al. 2010; Şengör et al. 2011; Gasperini et al. 2012;
kind of fault class some of the large strike-slip faults moving in such Shillington et al. 2012; Sorlien et al. 2012; Zitter et al. 2012; Le
regions should be assigned. Pichon et al. 2013).
In the case of the North Anatolian Fault (Ketin 1948, 1969, 1976; Moreover, the Sea of Marmara and its surroundings are frequently
Pavoni 1962; McKenzie 1972; Şengör 1979; Barka 1992, 1996; Şengör visited by earthquakes of devastating effects with tragic conse-
et al. 1985, 2005), it is clear that it connects the East Anatolian zone quences, and the long historical records kept since at least the sec-
of convergence (Şengör and Kidd 1979; Dewey et al. 1986; Şengör ond half of the first millennium BCE (before common era) (Fig. 4) give
et al. 2003, 2008) with the Hellenic Trench (Dewey and Şengör us an unparalleled knowledge of their repeat times and natures. That
1979; McKenzie and Jackson 1983; Le Pichon et al. 1993; Şengör 2011) yet another imminent large earthquake is now threatening İstanbul,
and as such qualifies as a transform fault (Şengör 1979; Le Pichon one of the megalopolises of the world and a cultural treasure house
and Kreemer 2010). It consists of a numerous family of faults going of mankind (it is the oldest continuously inhabited human dwelling
from the well-defined Karlıova Basin in the East (Şengör 1979; in the world, since ⬃250 000 BCE, when the cave dwellers discovered
Şaroğlu 1985; Şengör et al. 1985) to the more diffuse transtensional its unique advantages as a location: see Howell et al. 2010) has fo-
region in the Northern Aegean (Fig. 2). Throughout its length, its cussed the attention of geologists, geographers, and geophysicists on
various branches have been mapped in the field (see especially this Sea after the disastrous 1999 earthquakes in İzmit and Düzce in
Herece and Akay 2003), and it is generally agreed that the broad northwestern Turkey (see Şengör et al. 2005). The fact that the Mar-
region of deformation, a veritable keirogen2, called the “North mara area has been a part alternately of the Tethyan and Paratethyan
Anatolian Shear Zone” (Fig. 3; Şengör et al. 2005; Le Pichon et al. realms since the beginning of the Miocene has given it a stratigraphy
2013), in which the fault now moves, came into existence some- of a resolution higher than the normal Tethyan regions, because of
time during the later medial Miocene between the Serravallian the complete and abrupt change of biotas during every transition (so
and the Tortonian (see especially Şengör et al. 2005; Görür and far used only for the youngest sequences and for the frame of the Sea,
Elbek 2014; between the Sarmatian and the early Pannonian in except where well information has been released by the petroleum
the Paratethyan scale and between MN7/8 and MN10 in the industry). All of these factors make the Marmara region an ideal site
European mammal stages; much of the deposits along the to study the evolution of a strike-slip system below it.
North Anatolian Shear Zone have been dated using these two We have critically reviewed the maps of the Marmara fault seg-
latter standards). What so far has been unclear is how the broad ments (e.g., Parke et al. 1999, 2002; Le Pichon et al. 2003; Rangin et al.
keirogen became localised into the narrow, single-strand (dou- 2004; Grall et al. 2012; Sorlien et al. 2012) by using that database
ble strand west of Yeniçağa near Bolu: Fig. 2) main displace- (Fig. 5A) consisting of 210 multichannel seismic reflexion profiles,

2A keirogen (from ancient Greek ␬␧␫´ ␳␻ for “shear”) is a broad strike-slip shear zone in which more than one major strike-slip fault and associated structures
take up the shear, and it is the strike-slip counterpart of the shortening orogens and extending taphrogens. It was first defined in Şengör and Natal’in (1996,
p. 639, endnote 8).

Published by NRC Research Press


224 Can. J. Earth Sci. Vol. 51, 2014

Fig. 1. Wilson’s illustration of the kinds of transform faults (reprinted by permission from Macmillan Publishers Ltd: Nature, Wilson 1965,
fig. 3, copyright 1965). For “arc” read “trench” to conform to present common usage: (A) a, ridge–ridge; b, ridge–arc facing away from the
ridge; c, ridge–arc facing toward the ridge (this is, in principle, what the North Anatolian Fault is); d, arc–arc with arcs facing away from one
another; e, arc–arc with arcs facing the same way, f; arc–arc with arcs facing toward each other; (B) d, transform point at which extension
changes to shortening.

Fig. 2. The North Anatolian Fault, forming in reality a family of faults that constitute the North Anatolian Keirogen (Şengör et al. 2005).
Not all the faults shown on this map are now active, but all have been active sometime in the last 11 Ma. Most are potential earthquake
generators. Fault traces delineated by heavier lines represent the most active parts of the keirogen, constituting the strands collectively
known as the North Anatolian Fault (NAF). Note that the keirogen is entirely confined to the area underlain by Tethyside accretionary
complexes shown in grey (slightly modified from Şengör et al. (2005, fig. 2)). A, Ankara; B, Bursa; b, Bolu; E, Erzincan; İ, İstanbul;
İ, İznik (lake); K, Karlıova; OF, Ovacık Fault; SF, Sungurlu Fault.

Published by NRC Research Press


Şengör et al. 225

Fig. 3. The North Anatolian Shear Zone (NASZ: delimited by discontinuous lines) and the courses of the major rivers traversing it.
Key to abbreviations, from east to west (black letters): E, Elmalı/Peri (tributary of the Murat before the construction of the Keban Dam);
Ka, Karasu (Elmalı/Peri + Karasu = Fırat (Euphrates) without Murat (outside this map)); Y, Yeşilırmak; K, Kızılırmak; D, Delice; F, Filyos (/Yenice/
Araç/Soğanlı (formerly Uluçay)/Gerede Suyu); S, Sakarya; Su, Susurluk. Grey letters in outline show locations of some cities and tectonic
features: A, Ankara; B, Bursa; b, Bolu; E, Erzincan; İ, İstanbul; İ, İznik (lake); K, Karlıova; OF, Ovacık Fault; SF, Sungurlu Fault. Notice that
significant abrupt deflections of river courses are confined to the area of the NASZ. Following Şengör et al. (2005), the faults shown to be parts
of the NASZ in the southern part of the Tokat Lobe are here left out of the NASZ because of their as yet uncertain relationship to the NASZ
and to the geometry of the major river courses (slightly modified from Şengör et al. (2005, fig. 6)).

Fig. 4. Historical earthquake epicentres in and around the Sea of Marmara, compiled by the ETH Geodesy and Geodynamics Laboratory (GGL),
Zurich. Most of these earthquakes presumably occurred on the faults shown in Fig. 5.

adding up to 6210 km profile length, so far unpublished except some 2005 and 2011 plus the fault-plane solutions of all the earthquakes
segments of a number of the profiles that had been published by Ateş with magnitudes >4, with a view to seeing which faults may have
et al. (2003) and high-resolution bathymetry and chirp profiles re- localised motion, and, where information is available, what kind,
ported in the literature to map all the faults that appear to cut the since 1960. In the sections that follow, we describe the faults illus-
Miocene to Recent sedimentary rocks in the Sea of Marmara. We trated in Fig. 5B and give examples of various faulting events estab-
then distinguished those that cut only a portion of the sedimentary lished using seismic sequence stratigraphy.
sequence and tentatively established a crude sequence of faulting Stratigraphic sequences that are related to 100 ka glacial/inter-
events, according to the state of knowledge about the seismic stra- glacial cyclicity over the last 450–540 ka are here depicted follow-
tigraphy of the Sea of Marmara (the crudeness being a function of the ing the recent publications by Grall et al. (2012, 2013) and Sorlien
local paucity of direct lithostratigraphic and biostratigraphic con- et al. (2012). These authors show repetitions of onlap geometries
trol). Finally, we have correlated the faults that cut the present-day (laterally correlated with condensed sections, Grall et al. 2013),
seafloor with the microearthquake epicentres between the years and forset/toplap geometries (Sorlien et al. 2012) and related them

Published by NRC Research Press


226 Can. J. Earth Sci. Vol. 51, 2014

Fig. 5. (A) The locations of the multichannel seismic profiles used in this study. (B) The faults mapped in this study. The black segments
are active, and the grey segments are presumably inactive. Legend: MTA lines, Maden Tetkik ve Arama Genel Müdürlüğü (General Directorate
of Mineral Exploration and Research in Turkey) lines; TPAO lines, Türkiye petrolleri Anonim Ortaklığı (Turkish Petroleum Company) lines.
CB, Central Basin; CH, Central High; ÇıB, Çınarcık Basin; ET, Erdek Tombolo; GG, Gulf of Gemlik; K, Kapıdağ Peninsula; KB, Kumburgaz Basin;
MI, Marmara Island; NıB: North İmralı Basin; TB, Tekirdağ Basin; WH, Western High.

to variations of sedimentation rates during low-stand/high-stand shows it to be active localising almost exclusively right-lateral
transitions and low-stand deltas, respectively. These authors use earthquakes (Figs. 6B, 6C; Örgülü and Aktar 2001; Özalaybey et al.
the following colour code associated to Marine Isotopic Stage 2002, 2003; Pinar et al. 2003). In Fig. 6, the focal mechanism solu-
(MIS) transitions: Red MIS-5 (140 ka), Blue MIS-7 (240 ka), Yellow tion 44 is not readily interpretable in terms of simple shear zone
MIS-9 (340 ka), Purple MIS-11 (445 ka) and Green MIS-13 (570 ka) geometry, but it is still compatible with a tension gash orienta-
transitions (Grall et al. 2013). We adopt this colour code, denoting tion, albeit with a tiny sinistral component. We note that the term
by these colours the sequence on top of the sequence boundary. “tension gash” is here used to denote a structure orientation as seen
in micro- and meso-scales, where the structure was originally liquid-filled.
The eastern sector The structures we describe are not necessarily “tensional” but
Figure 6 illustrates the faults we have mapped, and Fig. 5A definitely “extensional”, with variable amounts of strike-slip as-
shows the locations of the multichannel seismic profiles that sociated with them (from 0% to 50% in some rare cases). As we
were used to identify them. Figure 6 shows the eastern sector of have not read the seismograms ourselves, we cannot further com-
our study area, consisting essentially of the Çınarcık Basin (Le ment on that solution. If it is a tension gash orientation normal
Pichon et al. 2001; Ateş et al. 2003; Carton 2005; Şengör et al. 2005; fault only, then it is also compatible with pure strike-slip along
Laigle et al. 2008; Uçarkuş 2010) and regions to its north (Gökçeoğlu the northern margin of the Çınarcık Basin.
et al. 2009; Özeren et al. 2010; for the neotectonics of the land area The entire sedimentary fill seen in this basin is of Pleistocene
see Şengör 2011) and south (including the İmralı Basin: Okay et al. age. In gravity core 27 (for location, see Fig. 7B), a radiocarbon age
2000; Sorlien et al. 2012). Çınarcık Basin is a critical area because of 16.1 to 16.5 cal. kiloyears BP was obtained at a depth of 25 m. In
the exact manner of the westward prolongation of the faults from core 25, closer to the main bounding fault to the north, two ages
it to the Central High remains unclear. of 14.3–16 and 15.0–16.5 cal. kiloyears BP were obtained from
In this sector we have chosen profiles 9 (Fig. 7) and 13 (Fig. 8) to depths of 28 and 32 m, respectively (Beck et al. 2007). They indi-
illustrate the relative age sequence of the faults (for locations of cate a mean sedimentation rate of 1.5–2 mm/year during the
the profiles, see Fig. 5A). Figure 7 shows that along the profile 9, Holocene within this basin. These ages make the entire sedimen-
the main marginal fault cuts to the surface. The seismicity also tary section, as seen in Fig. 7, Pleistocene in age (Şengör et al.

Published by NRC Research Press


Şengör et al. 227

Fig. 6. Maps of the eastern sector of our study area: (A) location map of the profiles selected to characterise the faults in this sector;
(B) distribution of seismicity during the 2005–2011 interval in the eastern sector; (C) distribution of seismicity compared with the identified
faults belonging to the North Anatolian Shear Zone. Refer to Fig. 5 for the legend of the faults. GG, Gulf of Gemlik; Lİ, Lake İznik; SA, Samanlı
Mountains. Öz, from Özalaybey et al. (2003); Öa, from Örgülü and Aktar (2001).

Published by NRC Research Press


228 Can. J. Earth Sci. Vol. 51, 2014

Fig. 7. (A) Line 9 showing a cross section across the Çınarcık Basin, including its broad southern shelf that contains a subsidiary basin on a
southerly tilted normal fault block. (B) Our interpretation of the structure of the Çınarcık Basin along line 9 (vertical exaggeration (V/H) here
is 4×). In this profile, three onlap surfaces (pinch-outs, indicated by black arrows) are observed at both basin edges as well as on the 2 km
wide faulted anticline rising within the basin. These pinch-outs are thus most likely related to variations in sedimentation rates rather than
to changes in fault activity. The onlap surfaces are assigned to the Red, Blue, and Yellow sequence boundaries (130, 250, and 330 ka,
respectively). The thickness of 100 ka stratigraphic sequences in this case ranges between 400 and 150 ms twt (corresponding roughly to 300
and 110 m), which is consistent with the mean sedimentation rates in this area (Seeber et al. 2006); as well as the sequence jump correlated by
Sorlien et al. (2012, fig. 8). (C) Same as Fig. 8B, but with no vertical exaggeration to show the “real” geological structure. We have done this
only for this profile to give an idea about the difference between the vertically exaggerated and non-exaggerated sections for those not used
to looking at seismic profiles.

2005), with the exception of the first 15 m to at most 20 m from the (Figs. 7, 5B). Farther east there is a surface-wave magnitude (Ms) 6.4
seafloor, which represent Holocene sedimentation. earthquake that gave a normal fault solution (solution labelled T in
The fault to the south of the main fault in this profile (Fig. 7) cuts Fig. 6C: Taymaz et al. 1991) where shallow transtensive faults could be
only up into a low-amplitude layer at around 1.5 s two-way travel mapped on both bathymetry and chirp sub-bottom profiles, corrob-
time (twt), but does not penetrate farther up. On the evidence of the orating the view that at least some of the faults in this set are active
seismic profile that fault seems to have died within the Pleistocene (Fig. 5B). The microearthquake activity along these southern normal
(since 540–450 ka). Neither is there any evidence of its presence in faults seems livelier than in the north since 2005, but the southern
the Chirp profile published by Zitter et al. (2012). However, there is normal faults do not generate as large earthquakes as the northern
seismicity corresponding to its location, and available fault-plane main strike-slip strand, which is what is expected from their respec-
solutions associated with that seismicity (solutions 33, 36, 39 in tive rates of motion (see Şengör et al. 2005).
Fig. 6C) all indicate dextral strike-slip; the fault-plane solution 36 Within the Çınarcık Basin, the northern master fault takes up by
indicates a very slight extensional component either on a dextral
far the largest portion of the movement, which is nearly perfectly
Riedel shear (R) or on a sinistral anti-Riedel (R=) shear, both entirely
parallel with it as seen both from the earthquake fault plane solu-
compatible with the main shear zone (the Y-shear direction) being a
tions and from the global positioning system (GPS) data (see Şengör
right-lateral one (see Fig. 9 for the terminology we use for various
et al. 2005 for a summary of the GPS vectors here), and there is no
shear zone structures). These earthquakes indicate a non-negligible
strain being taken up on the fault that on the seismic profile 9 does indication of major oblique-slip, contrary to what is implied by
not seem to reach the surface. The earthquakes’ magnitudes are Seeber et al. (2006), who overestimated the oblique-slip component
small enough and the data of the multichannel seismic profiles are at the crustal level, as they did not take into account the compaction,
coarse enough that the sea bottom’s either not being breached or which in a mature basin, could result in a seafloor subsidence 2–3
appearing not to have been breached is entirely understandable. times larger than the subsidence of the basement. We have not taken
That the fault does not show up even on the Marnaut 54 Chirp profile into account in our study Hergert and Heidbach’s (2011) modelling of
(Zitter et al. 2012) may suggest that in levels shallower than the green the mechanical behaviour of the lithosphere in the Marmara Region
sequence the fault may branch out and even merge with the main which requires derivation of stresses. Such a derivation is hardly
boundary fault and reach the surface via that route. possible (except in a model), given the uncertainties concerning the
A south-dipping fold and trust structure along the southern mar- properties of the real lithosphere and the stresses acting in it, but if
gin of the Çınarcık Basin separated two small basinal structures one wished to invoke that model, the northern margin of the

Published by NRC Research Press


Şengör et al. 229

Fig. 8. (A) Line 13 showing another cross section farther west than line 9 across the Çınarcık Basin showing a sub-basin. (B) Our interpretation
of the structure of the Çınarcık Basin along line 13 (vertical exaggeration here is 4×); the sequences appear less obvious (except for the pinch-
out on the Red Horizon), but horizons have been correlated to line 9, and this correlation looks in agreement with the interpretation
of Sorlien et al. (2012, fig. 8). To the south, across the İmralı Basin, a thick coherent section of sedimentary horizons (reaching >3 s twt) is
imaged. As in the Çınarcık Basin, pinch-outs can be recognised, but the most remarkable feature is probably the toplap geometry (indicated
by white arrow) at the base of the Blue (250 ka) horizon (also drawn on the fig. 4 in Sorlien et al. 2012). We assigned the same sequence
boundary as Sorlien et al. (2012) in this area, and we note that onlap geometry on the sequence boundaries are also observed in this basin.

Çınarcık Basin would still require <4.5 mm/year normal component Sorlien et al. (2012) have presented a meticulous and exhaustive
of motion within its constraints, which we find reassuring. analysis of the seismic stratigraphy of the İmralı Basin fill, with a
Farther south, west of the Armutlu Peninsula, there is the view to establishing a chronostratigraphy that they then used to
southerly-to-southeasterly-tilted İmralı Basin (Okay et al. 2000), also argue that faulting continued at a constant rate since about half a
of Pleistocene age. Microearthquake activity (Figs. 6B, 6C) is much million years ago. This is based on the date of what they call the Red
subdued here, only a few events falling directly on the profile. This horizon, which has been correlated to other places in most of the Sea
shows that even the “inactive” faults probably move, albeit rarely, of Marmara and has been attributed an age of 109 ka based on ex-
during small magnitude earthquakes, but contribute essentially trapolation from Marion–Dufresne cores (Sorlien et al. 2012; Grall
nothing to the further straining of the area. Along its easterly strike et al. 2013). Lower reflectors have been correlated to low-stand deltas
continuation, on the Samanlı Mountains (for their detailed geomor- and age attributed with the assumption that the sediment input into
phology, see Bilgin 1967), there is a livelier earthquake activity. The the İmralı Basin has been of the same order for each glacial cycle.
fault plane solutions available from there readily fall into two However, this reasoning also led to the conclusion that the low-stand
groups: a western group (fault plane solutions 16, 45, 47, 49, 51 in delta observed in stratigraphic succession does not systematically
Fig. 6C) and an eastern group (solutions 53 and 54). It seems that the correspond to successive glacial cycles. In the absence of indepen-
western group faults stretch the ground in a NNE–SSW direction and dent age constraints, ambiguities in the age attributions thus cannot
are related to those that created the İmralı Basin and the ones farther be resolved. In detail, Sorlien et al. (2012) age model taken as face
east cause extension in a NE–SW direction and are associated with value does suggest some slowing down of the subsidence over the
the fault family that caused the Ms = 6.4 Çınarcık earthquake farther last 130–330 ka once compaction is taken into account. In both pro-
north. Within the İmralı Basin, the main faults strike WNW–ESE and files it is difficult to ascertain whether the basin fill of the Çınarcık
are related to the faults that cause the western group of earthquakes and the İmralı Basin is conformable or unconformable with their
in the neighbouring Samanlı Mountains. But they are probably not stratigraphic basement. Because at least one fault is truncated by the
mainly responsible for the origin of the İmralı Basin. Otherwise, the bottom sequence (Fig. 8B) and some reflectors below the lowest se-
neighbouring Samanlı Mountains with similar and active faults quence seem to have a geometry in places not exactly parallel with
would not have stood as high as they do now. the basement layering, we assume an unconformity bounding the
Figure 8A shows profile 13, which is farther west than the pre- basin at the bottom, which is, however, not easy to defend conclu-
vious one (Fig. 7A), and Fig. 8B shows our interpretation of it. The sively on the basis of the available observations.
recognisable sequences indicate a much narrower basin of a sim- Interestingly, once one goes southward over the Samanlı Mountains
ilar age to the one we saw farther east. Although this sub-basin is on the Armutlu Peninsula (SA in Fig. 6A), the microseismicity
bounded by faults that cut to the seafloor, there is one between again becomes lively in the Gulf of Gemlik (GG in Fig. 5A). A major
them that is much older, possibly pre-Pleistocene. The southerly fault strand, which we here term the Gemlik–Bandırma Fault,
tilted basin here appears completely inactive, as none of the faults goes from there to the Erdek tombolo (ET in Fig. 5) and then turns
that appear related to its formation cut to the surface. The spread southwestward in the county Gönen into the Biga Peninsula
of microseismicity (Figs. 6B, 6C) corroborates this picture, show- (Fig. 5). Young scarps all along the fault suggest recent activity and
ing only a few scattered epicentres. at least two estimated Ms = 7 earthquakes (128 and 543 CE) seem to

Published by NRC Research Press


230 Can. J. Earth Sci. Vol. 51, 2014

Fig. 9. Development of a shear zone (from Şengör, in press). Brown areas are structures associated with extension and (or) transtension.
In transtension, all extensional structures become “flatter”, and all shortening structures become “steeper” with respect to the Y shear.
However, this is only an instantaneous picture, because as the shear zone evolves, all structures will rotate and strains will increase, as we
now see in the Sea of Marmara case. Key to lettering: F is fold axial trace; P is P shear; R is a Riedel shear; R= is an anti-Riedel shear; T is a
tension gash (in meso and macro examples, depending on the complexity of the system, the stress along the “tension gash orientation” line
may not necessarily be purely tensional, but there must be a major component of extension); X is an X shear; ␸ is the angle of internal
friction of the material being sheared.

have taken place on it during historical times (Fig. 4). Kuşçu et al. broad southern shelf of the Sea of Marmara (Fig. 5A). Therefore,
(2009) mapped a group of active faults in the Gulf of Gemlik and we go along the recent trend in following Dewey and Şengör (1979)
microearthquakes align along smaller faults on the isthmus sep- and Şengör (1979) and stick to two major strands, but we never-
arating Lake İznik and the Gulf of Gemlik, all betraying a lively theless point out that the Gemlik–Bandırma fault does maintain
activity although not accomplishing much strain (no earthquakes an activity, although much subdued.
above magnitude 4), except obviously at intervals of a minimum
or a millennium-and-a-half. Although Barka and Kadinsky-Cade The central sector
(1988) called it the “middle branch of the North Anatolian Fault”,
it is much less significant than the two major strands originally This sector (Fig. 10), as herein used, corresponds mostly to what
identified by Dewey and Şengör (1979) and Şengör (1979) and is not is called the Central High in the literature (it is the high region
much different from the many deactivated faults populating the that completely surrounds the Kumburgaz Basin, KB; Le Pichon

Published by NRC Research Press


Şengör et al. 231

Fig. 10. Maps of the central sector of our study area: (A) location map of the profiles selected to characterise the faults in this sector;
(B) distribution of seismicity during the 2005–2011 interval in the central sector; (C) distribution of seismicity compared with the identified
faults belonging to the North Anatolian Shear Zone.

et al. 2001; Şengör et al. 2005) but that also includes the Kumbur- Profile 23 (Figs. 11A, 11B) is located in the eastern and central part of
gaz Basin and almost the whole of the Central Basin (Le Pichon the basin and shows that both its bounding faults there cut to the
et al. 2001; Şengör et al. 2005). The presence of the Kumburgaz surface (Fig. 11B). The basin also has other faults in it, but they seem
Basin atop the Central High provides a sediment receptacle and to have died shortly after the basin initially formed. We also see that
thus a welcome means to date the faults cutting the Central High the basin was initially narrower. However, the faults that bound the
relative to one another (see especially Sorlien et al. 2012, fig. 6). basin on this profile are not parts of a single fault family: the northern

Published by NRC Research Press


232 Can. J. Earth Sci. Vol. 51, 2014

Fig. 11. (A) Line 23 showing a cross section across the eastern moiety of the Kumburgaz Basin, showing the well-localised basin and its temporal
enlargement outward. For details of the seismic stratigraphy, see the caption of Fig. 10B. (B) Our interpretation of the geology along line 23.

fault is along the strike of the northern boundary fault of the epicentres on it. Yet it sits perfectly on strike with, and seems to be
Çınarcık Basin coming from the east, here forming a part of the main the northwesterly continuation of, a major inactive fault extend-
zone of displacement along the northern branch of the North Ana- ing down into the Gulf of Gemlik.
tolian Fault (Şengör et al. 2005). The southern fault, however, has a Oblique (N120°-N140°) transtensional blind and deep-seated faults,
WNW strike and seems to be a somewhat rotated, extensional Riedel identified in the western part of the Sea of Marmara, are part of a
shear so identified simply on account of its orientation. Immediately northeast dipping extensional fault system named the South Tran-
to its south there is another fault similar to it that also bounds the stensional Zone (STZ) by Grall et al. (2012). Many of these faults ap-
basin. Both of these faults cut to the surface, and both are active as pear still active, although we do not know what portion of the total
shown by the microseismicity they localise. To their east, the south- shear they accommodate. In particular, the western part of this fault
ern basin margin seems unfaulted as can be seen on fig. 6 by Sorlien system is suspected to play a rôle in the landslide activity in the
et al. (2012). Sorlien et al. (2012) carry what they call the Red-1 reflex- Ganos area (in our western sector: Zitter et al. 2012). They are the
ion into the Kumburgaz Basin where it is clearly cut twice, whereas dominantly strike-slip transtensional parts of a system that includes
in the İmralı Basin it is not, where it records the lower boundary of a transpression with dominant thrusting along the western foot of the
non-faulted floor of a sedimentary infill. Therefore, although a Ganos Mountain (Fig. 5; see also Fig. 16). The long fault mentioned in
throughgoing main strike-slip fault appears to have already formed
the previous paragraph is the partly active eastward termination of
in the Kumburgaz Basin, the auxiliary faults, such as the two Riedels
this transtensional system providing a connection between the main
just mentioned, appear not to have not terminated their activity.
northern branch of the North Anatolian Fault and the fault system to
They probably do not localise major shocks (no earthquakes above
the south recently named the Southern Marmara Fault by Le Pichon
magnitude 4) and therefore do not take up much displacement, but
et al. (2013). ⬃N100° striking, mainly wrench faults occur also in the
they nevertheless remain active and it is hard to tell up to what size
earthquakes they might one day localise. eastern part of the Sea. These faults, which may be blind, are proba-
Profile 27 (Fig. 12A and 12B) shows a much wider basin, of almost bly a part of the Central Fault System (CFS, Grall et al. 2012). Other,
bovine-head shape when considered together with the narrower shorter, faults close to and parallel with the inactive strand are, by
fault trough underlying it, and numerous faults cut its most recent contrast, still active. A small family of short faults with the same
sequence. Although the major displacement is along the northern orientation to the east of the meandering submarine canyon seen in
master fault bounding all the sequences and reaching the seafloor, it Fig. 10C are all active and clearly localise microearthquakes. By con-
seems that during the deposition of the youngest sequence, the basin trast, some abutting against the bounding faults of the spindle-
had already become much wider. The numerous faults to the south, shaped basin within the Central Basin appear dead, perhaps because
sub-parallel with the two previously mentioned Riedel shears, are all they have been recently decapitated by the main northern branch of
either active or terminated their activity very recently during the the North Anatolian Fault. All of the faults belonging to Grall et al.’s
deposition of the top sequence. Many of these faults seem to be short, (2012) Central Fault System belong to the tension-gash-orientation
therefore insignificant, in terms of the displacement they can accom- (refer to Fig. 9 to see what that orientation means) normal faults
plish, but they nevertheless are mostly active. related to the North Anatolian Shear Zone.
When we look farther south, an active fault that strikes about Curiously, the longer E-W orientated faults along Y-shear direc-
N45°W, thus in an extensional orientation in the overall North tions (cf. Figure 9), some turning into Riedel and P-shear orientations
Anatolian Shear Zone, is seen to be active with microearthquake at their ends, seem in some segments active and in others inactive.

Published by NRC Research Press


Şengör et al. 233

Fig. 12. (A) Line 27 showing another cross section farther west than line 23 across the western moiety of the Kumburgaz Basin, showing a broad,
very young, bovine-head basin that is nested on the older, narrower, fault-bounded basin. (B) Our interpretation of the geology along line 27. On
both profiles, 23 and 27 pinch-outs of horizons on the Red (130 ka) and Blue (250 ka), as well as on the Violet (430 ka), horizons are indicated by
black arrows. The depths of these horizons along the Northern Main Fault are similar to the ones proposed by Sorlien et al. (2012, figs. 6, 11), and
imply a mean sedimentation rate of around 1–1.5 mm/year over 100 ka period. The well-localised basin below the Blue horizon, which was widened
slowly outward through time, i.e., the blue sequence (250 ka) time, by contrast, looks like the period when the basin widening set on drastically.
This change is concomitant to the deactivation/slowdown of the faults running across the Kumburgaz Basin. Parts of these faults are presumably
still active, as some en-échelon discrete structures can be recognised along them on the seafloor (en-échelon structures mentioned in Fig. 11).

Farther west, within the Central Basin, the microearthquake activ- the main displacement zone of the northern branch of the North
ity shows that the smaller spindle-shaped basin within the Central Anatolian Fault here. Some of the faults that cut the high are active,
Basin is where most of the present activiy is located, but the north- others dead and yet both dead and active faults seem to belong to the
eastern border of the larger basin is also active (Figs. 10B, 10C). The same set of steep thrusts. This is another region where we see active
active, spindle-shaped basin within the larger Central Basin is in fact faults in a thrust orientation. In fact, the entire geometry of the
where the Holocene subsidence rate is the highest in the Sea of Western High, with its flanking basins, gives the impression of a
Marmara (reaching >7 mm/year, Grall et al. 2012). This centralised shortening structure that originally consisted of a broad synclinal
subsidence is in stark contrast to the situation encountered in the basin, in the middle of which an anticline later rose. This is also
Çınarcık and the Kumburgaz basins, where the main displacement supported by the topography of the sediment/acoustic basement in-
zone is along the northern master faults of the basins. terface, which reveals a deep trough which extends from the Central
So far, we had encountered no active faults in a shortening Basin towards the east and to the Tekirdağ Basin towards the West,
orientation. We encounter such a fault for the first time here, such as Bayrakçı et al. (2013) recently depicted. It seems that the
along the main displacement zone of the North Anatolian Fault in throughgoing faulting chose the anticline to follow and in so doing
the Sea of Marmara just east of the Central Basin where the main sliced it with small and steep thrusts.
strand is connected to the small spindle-shaped basinette in the We have no data as to the nature of the sequences seen in the
middle of the Central Basin (Fig. 10B). seismic profiles across the Western High making up its founda-
Within the southern Marmara Shelf, the available seismic pro- tion. It is, however, very likely that they consist of the Ganos
files show that all the faults terminated their major movement at Flysch wedge that delimits the Thrace Basin to the south and the
the end of the early Pliocene (3.5 Ma), but there is sparse, broadly Mio-Pliocene gently dipping sedimentary rocks covering the
spread microseismicity (Le Pichon et al. 2013). flysch wedge as one sees on the Gallipoli Peninsula, where even
deeper portions of the accretionary wedge crop out (e.g., Şentürk
The western sector and Okay 1984 report blueschists that crop out from under the
The western sector (Fig. 13) houses what is termed the Western flysch and the surrounding younger Cainozoic rocks).
High and the Tekirdağ (ancient Raedestos) Basin. We selected two The second profile, 78 (Fig. 15), goes across the Tekirdağ Basin.
profiles to show here, one across the High (profile 70: Fig. 14A and B) In principle it shows the same structures that one sees in the
and another, across the Basin (profile 78: Fig. 16A and B). Western High, except the central anticline is here asymmetric and
Profile 70 (Fig. 14A) shows a much-faulted anticlinal ridge rising in north-vergent, delimited against the northern, deeper half of the
the deepest part of a trough-shaped basin that is not seriously faulted basin by a very steep, essentially vertical fault. Okay et al. (1999)
at either end. Although what faults there are dip towards the basin, interpreted it as a south-vergent thrust, which is contrary to what
their vertical displacements are insignificant compared with the is seen along profile 78 (Fig. 15A). The southern part of the anti-
marginal faults of the Çınarcık, Kumburgaz and the Central basins. cline is not broken by faults (except two very shallow landslide
The basin in fact has the appearance of a sag in the midst of which faults) but has the form of a syncline. The sequences betray a
the peculiar anticlinal ridge of the Western High rises. It seems that faster subsidence in the basin centre than along its sides consis-
the only active faults are those that bound the Western High, the tent with the syncline interpretation.
southern with a higher rate of vertical motion as judged from the The steep fault is a continuation westwards of the main dis-
thickness of the green sequence. It is that fault which coincides with placement zone. The smaller large fault to the north achieves not

Published by NRC Research Press


234 Can. J. Earth Sci. Vol. 51, 2014

Fig. 13. Maps of the western sector of our study area: (A) location map of the profiles selected to characterise the faults in this sector;
(B) distribution of seismicity during the 2005–2011 interval in the western sector; (C) distribution of seismicity compared with the identified
faults belonging to the North Anatolian Shear Zone. For the profile 84, see Fig. 16.

much of a vertical displacement and seems a part of a series of Basin. That attempt seems to have been abandoned, however,
short, en-échelon faults that seem an abortive attempt to reach the while the youngest sequence was being deposited. There is essen-
northern end of the Ganos (MG in Fig. 13A) constraining bend tially no microseismic activity along that trend of small faults.
thrust zone (Dewey and Şengör 1979; Şengör 1979; see Fig. 16 Two other short trends of faults on the basin floor to the east of
herein) by tearing a strike-slip fault from the northern basin north the middle part of the Ganos constraining bend suggest other at-
of the Western High across the northern part of the Tekirdağ tempts at reaching it farther north than its present southernmost

Published by NRC Research Press


Şengör et al. 235

Fig. 14. (A) Line 70 showing a cross section across the Western High, showing its faulted anticline structure and the saucer-shaped basin in
which the anticline seems to have risen along its bounding faults. (B) Our interpretation of the geology along line 70. This basin displays a
thick coherent sedimentary section (which can reach 2 s twt of thickness), suggesting a long history of sedimentation. Along the basin
margins, one sees the very same seismic facies as one does within the small perched basins on top of the High (Grall et al. 2013). In particular,
one recognises the chaotic reflectors interpreted as slide deposits on the Violet horizon (430 ka) and also the bed with low-amplitude
reflectors above the Green horizon (which is widespread in the entire Sea of Marmara). These observations show that the Western High rose
in a basin that was simply an eastern continuation of the Tekirdağ Basin! The blue patchy reflector (250 ka) and the Red horizon in reverse
polarity (130 ka) have been also recognised above these characteristic layers. This profile confirms what Grall et al. (2014) pointed out
regarding the mass transport deposits on the Violet horizon, which is likely related to a regional (at least at the scale of the Western High)
process (such as the paleo-environmental changes proposed in Grall et al. 2014).

end. Although both attempts seem to have failed, the latter group 2003) must be compensated in part by shortening, and it is more
of faults are still active as indicated by a lively microseismic activ- economical to distribute the shortening and consequent crustal
ity at their eastern ends, but this is probably because of the load- thickening on many faults. However, here too the strike-slip
ing of the crust in which they reside by the Ganos Wedge (see faults display their characteristic complex strains by localising
Fig. 16 for the cross-sectional geometry of the Ganos overthrust earthquakes of all three kinds of fault types (Fig. 17)!
zone along the restraining bend). This loading and the presence of
E–W-striking faults going under the thrust wedge, including in Discussion
part the main displacement zone itself, which are probably in
part extensional because of their orientation in a zone of mainly Temporal evolution of the North Anatolian Fault in the Sea
WNW–ESE orientation, may be the reason of a very high rate of of Marmara
3He escape here (Burnard et al. 2012). When we started this study, we were expecting, as expressed in
At the latitude of the Western segment, seismic activity on the Şengör et al. (2005), an orderly evolution of a shear belt in which
South Marmara Shelf is considerably livelier than in the Central an initially broad shear zone characterised by such “secondary” or
Segment. Here we have in fact four earthquakes with magni- “auxiliary structures” as Riedel (R) and Anti-Riedel (R=) shears, X
tudes >6 (Fig. 17) on the Biga Peninsula, the western continuation of and P shears, and tension gashes (T) with shortening structures (S)
the southernmost part of the South Marmara Shelf. The microe- gradually evolving and eventually abandoning the field of deforma-
arthquake activity is also much higher than it is in the Central Sector. tion to a narrow throughgoing fault consisting of segments of Y
In this sector, the Marmara Island (the ancient Proconnesos: the shears, much as in the experiments of Riedel (1929), Tchalenko
largest island in Fig. 17) is a divider: to the east nearly all major (1970), Wilcox et al. (1973), Bartlett et al. (1981), Naylor et al. (1986), and
faults strike W to NW; to its west nearly all strike NE, indicating Ahlgren (2001), and as depicted in our Fig. 9. What we have found
that the shear zone takes a fairly sharp bend to the southwest with instead is a much more haphazard development. If one separates the
accompanying appearance of major thrust earthquakes along the fault families expected from the experiments into R, R=, X, P, and Y
fault segments. The faults also acquire here a much greater con- shears and T and S structures according to their orientations and,
tinuity than they have in the east, as one would expect from thrust wherever available, from earthquake first-motion solutions and
faults as opposed to normal faults which are numerous east of the stratigraphic vertical separations and note their sequence of occur-
Marmara Island. Interestingly, west of the Marmara Island all the rence in the analogue experiments as R=, R, P, X, and, if the material
major strands are active, presumably because here the 2.5 cm/year being sheared is suitable to form them, first S and then T structures.
motion of the Marmara Block delimited by the northern and the Nature shows us that in the Sea of Marmara almost all of these
southern strands of the North Anatolian Fault (Le Pichon et al. structure types are active now. We have actually seen no X shears

Published by NRC Research Press


236 Can. J. Earth Sci. Vol. 51, 2014

Fig. 15. (A) Line 78 showing another cross section farther west than line 70 across the Tekirdağ Basin, showing the central anticline covered
with young sedimentary layers. (B) Our interpretation of the geology along line 78. Note the fault that bounds the anticline to the north. It
now seems vertical, but the internal structure of the anticline implies that it might have originated as a very steeply south-dipping thrust
fault, contrary to the interpretation by Okay et al. (1999). 1.5 s of coherent sedimentary section is imaged across the Tekirdağ Basin, on which
onlap sequences (indicated by black arrows) occur periodically, as in the Central Basin (Grall et al. 2012), and in the Çınarcık (Fig. 8A). The first
Red onlap surface (dated at ⬃130 ka) is at around 300 ms twt (roughly 230 m) below the seafloor within the depocentre. This age estimation is
thus consistent with the mean sedimentation rates (1.5–2 mm/year) derived from the last marine/lacustrine transition (Mercier de Lepinay,
personal communication). The transparent lens in green (430–540 ka interval) in the basin, as well as in the bounded anticline, looks very
similar to the one encountered on the Western High and elsewhere in the Sea of Marmara (Grall et al. 2013). The Red and the Blue horizons
have been correlated from the basin to the southern anticline across the faults. The Red horizon is tentatively proposed on the base of
sediments affected by gravity-gliding, as it has been shown that main slides occur preferentially during glacial/interglacial transitions (Grall
et al. 2014). Nonetheless, we cannot exclude that the prominent reflector on the base of this sliding mass, which apparently functions as a
décollement, could have no stratigraphic meaning. Besides, it is also probably interesting to note that across this basin, as across other deep
basins, no mass transport deposits have yet been recognised on top of the Violet horizon, despite the fact that they have been recognised in
the ponded basins of the Western High (Grall et al. 2013), as well as on the margins of the basin within which the Western High seems to rise (Fig. 13B).

that can be confidently identified and some of the fault segments by downslope creep on steep (3° to 10°) slopes in and around the main
bending into a P-shear orientation are all now dead. What is even trough of the Sea of Marmara. Interestingly, with the exception of
more peculiar is the observation that some faults may have both the slope north of the Armutlu Peninsula and to the south of the
dead and active segments along them (within the time frame of a Central Basin, all of the slopes on which such creep folds have
part of the Pleistocene, mostly even late Pleistocene–Holocene)! An formed and in places continue to form are on the slopes of structures
extreme example has been recently documented by Le Pichon et al. that are in a shortening orientation and have been so since the shear
(2013), who showed that what they called a South Marmara Fault began. The Çınarcık exception is easy to explain in terms of the
under the late Pliocene beds in the South Marmara Shelf appears
active normal faulting going on at the foot of the Armutlu Peninsula.
entirely inactive there at least for the last 2.5 or 2.6 Ma, yet both at its
The slope south of the Central Basin may have been accentuated by
eastern and western ends it is today active! Kuşçu et al. (2009) showed
that the faults in the Gulf of Gemlik are active and the prolongation the faults that there exist along the South Marmara Shelf (see Le
of the South Marmara Fault into the Aegean in the west just localised Pichon et al. 2013), but the data there are not adequate to address this
a moment magnitude (Mw) 6.2 dextral slip earthquake on 8 January question. Shillington et al.’s (2012) work is a very nice possible exam-
2013 (Le Pichon et al. 2013)! This suggests that some dead fault strands ple of structures forming concurrently with growing folds on a large
may be connected to active ones by as yet unrecognised intermediate scale, perhaps not dissimilar to the situation as documented by the
segments. progressive growth in time (late Cretaceous to Eocene) of the Hathira
An extremely interesting observation has been reported recently (Makhtesh–Hagadol or Kurnub) anticline in southern Israel (Eyal
by Shillington et al. (2012): they showed the presence of folds forming 2011).

Published by NRC Research Press


Şengör et al. 237

Fig. 16. Profile 84 across the Mt. Ganos shelf and slope showing the underthrusting of the Tekirdağ Basin floor under the rising pillar of
Mr. Ganos and the development of a small marginal fold-thrust belt. This underthrusting activity is probably the reason of the
microseismicity seen along the small faults at high angles to the overthrust margin breaking up the basin floor (Fig. 13C).

Although most of the elements of the shear zone in the Sea of Marmara area (Akbayram and Okay 2012; Şengör et al. 2005) that the
Marmara are active now, the strain they accomplish is extremely Sea of Marmara has a total history of at least three million years.
variable. It seems that almost all the motion of the Eurasia/Marmara The difference between the experiments and Nature as it appears
Block (the area delimited by the northern and the southern strands to us in the neotectonics of the Marmara region may be twofold: one
of the North Anatolian Fault: Le Pichon et al. 2003) is taken up by the is the time scale. What happens in Nature in thousands to hundreds
Y shears making up what Le Pichon et al. (2001) and Şengör et al. of thousands, and even a couple of million years, happens in exper-
(2005) termed the main displacement zone forming, from east to iments in a matter of minutes or at most hours. We may be seeing a
west, the main northern bounding faults of the Çınarcık and the process of slow demise of the auxiliaries in Nature, which to us seem
Kumburgaz basins, the fault cutting through the Central Basin and a time of activity in concert. In reality, the auxiliaries may be slowly
bounding the small spindle-shaped basin in the middle, and the fault passing away as the main displacement zone gathers the action onto
delimiting the Western High and the Tekirdağ Basin to their south. itself. The fact that nearly all the main displacement today is concen-
All other active faults seem to cut to the seafloor and nucleate micro- trated on the almost completely throughgoing main displacement
seismic shocks with few exceptions, such as the normal faults south zone may be a result of its gradual climb to prominence, a climb that
of the Çınarcık Basin that nucleated an Ms = 6.4 earthquake and has not yet reached the summit.
possibly another one in 1894 that greatly damaged İstanbul. They The second difference pertains to the extreme heterogeneity of
contribute very little to the overall strain, yet Le Pichon et al. (2003) the natural material which the shear zone deforms as opposed to the
documented that the small strain was by no means negligible and it nearly homogeneous materials used in the experiments. No experi-
is detectable by GPS observations. Sorlien et al.’s (2012) thesis that the ment can hope to reproduce the multifarious nature of the materials
faults have moved with a constant rate during the last half a million making up the area in and around the Sea of Marmara. Entire do-
years, which implies in view of the ⬃55 km total offset for the entire mains consisting of very different types of materials dominate the

Published by NRC Research Press


238 Can. J. Earth Sci. Vol. 51, 2014

Fig. 17. The locations, magnitudes, and the first-motion solutions of the major shocks known from the southern part of the western sector
studied in this paper. Fault plane solutions are from McKenzie (1972), Ekström and England (1989), and Taymaz et al. (1991).

Marmara area and the developing shear zone no doubt exploited Anatolian Shear Zone is 75–80 km along its entire length (Şengör
every bimaterial surface to break through on all sorts of scales. That et al. 2005), we subtracted the offset of some 26 km seen along the
is mainly why we have been unable to use with confidence the kind Pamukova strand leading to the main southern strand (Şengör et al.
of modelling represented by Hergert and Heidbach’s (2011) paper, 2005) from that to obtain the 55 km left-lateral offset on the entire
where rock properties in the sedimentary carapace are accepted to be Sea of Marmara fault family to reverse the present right-lateral shear
distributed homogeneously — which is manifestly not the case, if for for the purposes of Fig. 18. This seems consistent with the 52 km
no other reason than the depth differences among basin basements, offset reported by Akbayram and Okay (2012) from the northern
even within a single basin — and the great variations in the rock strand east of the Sea of Marmara. The timing of the offset reported
properties in the basement are not even considered. Şengör et al. by Akbayram and Okay (2012) is problematic, however, because
(2005) pointed out that the North Anatolian Fault as a whole seems to some of it could have formed just after the medial Eocene collision
have nucleated along the northern “wall” of the Tethyside accretion- along the Intra-Pontide suture in a completely different system of
ary complexes against the more indurated basement of the Rhodo- strike-slip faults (Akbayram et al. 2013). It is clear, however, that at
pe–Pontide Fragment (see Fig. 2), very much corroborating the least the Marmara Fault family north of the South Marmara Shelf
theoretical model of Ben-Zion and Andrews (1998) and Ben-Zion takes up a displacement which is concentrated farther east on the
(2001). There is little doubt that the sensitivity of bimaterial surfaces
northern strand of the North Anatolian Fault, and that offset is an
to preferentially produce rupture under stress is also present at
absolute maximum of 57 km. Figure 18C shows what is seen when we
much smaller scales between the tectonic units made up of different
superimpose Figs. 18A and 18B keeping the northern boundary fixed.
materials under and around the Sea of Marmara and between differ-
The highs do not rotate as we expected, but acquire a lumpier appearance! In
ent rock packages within those tectonic units.
fact, as a result, they present an optical illusion as if they rotated into
The origin of the Marmara Deeps a more E–W orientation when the displacement is reversed.
Both the Western and the Central highs have east-northeasterly This to us suggests that the highs, or their much more subdued
trends. Figure 14A shows the Western High having an overall anticli- forerunners, were probably there before the shear began and were
nal structure albeit sliced by numerous faults. We had originally similar to gentle axial culminations along outer arc highs in present-
hypothesised that the anticline was a function of the shear evolution day forearcs, in this case possibly either accentuated or even initially
and asked ourselves that if we reverse the shear along the North caused by the collision. When the shear began they were shortened
Anatolian Shear Zone, what orientation the high would acquire. Our into S structures and became elongated in an ENE orientation. The
expectation was that it would rotate into a more northeasterly direc- Western High was probably only a gentle “suggestion” with a broad
tion, and we expected a similar result from the Central High. Fig- basin, a simple “undulation”, of the basin floor, which later became
ure 18A shows the present-day fault geometry within the Sea of accentuated as the shear evolved. The fig. 6 in Sorlien et al. (2012) very
Marmara and the outlines of the two highs in blue. Figure 18B shows nicely shows the anticlinal structure of the Central High and the
the situation if a homogeneous strain is imposed on the system to smaller wavelength folds right across it involving even the youngest
reverse a 55 km offset. Because the overall offset along the North sediments, although there is a clear decrease in folding amplitude up

Published by NRC Research Press


Şengör et al. 239

Fig. 18. (A) A simplified map of the faults reported in this paper. The two light blue patches are the Central and the Western highs. (B) Same as A,
but with a left-lateral shear imposed on all the elements to reverse the right-lateral shear of some 55 km along the Sea of Marmara. (C) A
comparison of the present geometry of the faults belonging to the northern part of the north Anatolian Shear Zone in the Sea of Marmara, with
their presumed geometry before a homogeneous right-lateral shear of some 55 km took place. Note how little many of the elements rotate. No
wonder so many of the R shears are so long-lived. Note also that the Central and the Western highs are compressed into a more pronounced NE–SW
orientation as the shear evolves. The compression appears as a rotation of the entire structure, which, in reality, it is not.

section (the reverse of what one sees in the so-called “generative branch of the North Anatolian Fault (we here mean the last, clean,
folds”; our sorts of folds may thus be called “degenerative”) indicat- throughgoing fault), at least west of the Çınarcık Basin, is only some
ing a waning of the shortening commensurate with the concentra- 4 km, falls out of the distributed shear interpretation. If the shear has
tion of the shear along the northern branch of the North Anatolian been really as distributed as the Fig. 18 approximates, then the off-
Fault. set along the level of the Central Basin/Western High comes to at
The Çınarcık Basin is clearly not a part of this system, as indeed it most 5 km. and no more. Therefore, the fault families we have
does not sit in front of the Thracian forearc basin, but in front of the mapped in the Sea of Marmara must have initially taken up the
more solid and homogeneous İstanbul Zone (Şengör 2011, fig. 1) and entire displacement that falls to the share of the present-day North-
in a totally different orientation, viz WNW. We think that the inter- ern Branch altogether. That many of them are still active (although
pretation of it by Uçarkuş (2010) as a fault wedge basin sensu Crowell probably dying) corroborates this view. That the South Marmara
(1974) is so far the best explanation for the origin and the present Shelf fault families are almost all dead (with the exception of the present
structure of the basin, so we do not further elaborate on it here, microseismic activity), but that more activity is seen in the northern
except to say that it is a completely different sort of basin from the auxiliary structures, on both seismic and stratigraphic (Sorlien et al.
other, more westerly, basins of the Sea of Marmara. 2012) evidence, may be an indication of the fact that the displacement is
gradually gathering itself to the northern branch alone.
The offset along the northern branch of the
North Anatolian Fault Conclusions
Figure 18C further shows that the statement by Le Pichon et al. The present-day fault families younger than about the medial Mio-
(2001) that the offset along the very young (ca. 200 ka) northern cene in the Sea of Marmara are all parts of the North Anatolian Shear

Published by NRC Research Press


240 Can. J. Earth Sci. Vol. 51, 2014

Fig. 19. The presumed geometry of the fault network depicted in Fig. 18A at a depth of about 10 km below the floor of the Sea of Marmara.
All the faults are assumed to have widened into shear zones with a uniform width of 4 km. This is probably wrong, as not all faults have
the same displacement, but it gives an idea as to what the North Anatolian Fault family under the Sea of Marmara may look like close to
mid-crustal depths. The inset marked “b” is a hand specimen showing the mesostructure of the Ailao shear zone at amphibolite facies
metamorphic conditions. Its similarity to our presumed North Anatolian fault structure at depth is remarkable.

Zone, and they all seem to have shared among themselves a total lateral, and the similarity of the hand sample to the map pattern
offset of some 55 km along this zone since the later medial Miocene. displayed by the presumed depth-map of the Sea of Marmara fault
Contrary to expectation from analogue claycake and sandbox exper- family is staggering: it is one of the best examples of Pumpelly’s rule
iments, a regular development from distributed structures to a sin- that small structures mimic big structures and a further demonstra-
gle, throughgoing strand cannot now be established, but this may be
tion of Tchalenko’s (1970) point that shear structures are indepen-
more a function of the vastly different timescales between the exper-
dent of scale. We thus think that at a depth below 10 km, the Sea of
iments and Nature herself than a fundamental difference in the
evolutionary sequence. That the Southern Shelf fault family is now Marmara basement consists of sheared rocks of the kind seen below
dead and that no major earthquake seems to be nucleated on the the Ailoa Shan strike-slip fault, which, by an interesting coincidence,
auxiliaries (except in the southern part of the Çınarcık Basin) may is also nucleated on a suture zone bearing steep serpentinite screens
indicate a gradual localisation of the offset along the single strand of (Şengör and Natal’in 1996; Burchfiel and Chen 2012, chap. 15, espe-
the Northern Brach of the North Anatolian Fault. cially fig. 15-1). In deeply eroded Precambrian terrains, the pres-
This study gives support to the older ideas that intracontinental ence of such broad, vertical, anastomosing ductile shear zones are
transform faults begin their lives as broad shear zones and be- most likely expressions of large strike-slip faults at the surface,
come gradually converted into single- (or at most a few-) strand many of which were probably transform faults. As transform
structures. What such a broad shear zone at depth looks like may
faults are dependable indicators of relative plate motion direc-
be displayed by the 15 km wide mylonite zone seen along the giant
Irtysh shear zone (displacement ≈2000 km) in Central Asia (Şengör tions, this is one way to get a handle on such old plate motion
and Natal’in 1996). The numerous active fault zones we see at the directions.
surface no doubt broaden with depth (earthquake hypocentres
are hardly deeper than some 10 km under the Sea of Marmara; Acknowledgements
below that the deformation is most likely ductile), and given their This paper was written upon the invitation of Professor Ali Polat
density at the surface, they most likely coalesce into a very broad to whom we are grateful for the honour of being included in a
ductile shear zone. Le Pichon et al.’s (2003) study has corroborated volume in memory of Tuzo Wilson. Professor Namık Çağatay was
the presence of such a general, distributed deformation south of a great source of knowledge and critique in all matters strati-
the Northern Branch of the North Anatolian Fault. graphic. MARSITE (New Directions in Seismic Hazard Assessment
Figure 19 shows a situation where the faults depicted on Fig. 18A through Focused Earth Observation in the MARmara SuperSITE)
have been widened to shear zones of some 4 km width, the pre-
EU FP7 Project number 308417) funding made substantial parts of
sumed width of the San Andreas Fault below 10 km (e.g., McBride and
Brown 1986). In these regions of amphibolite-grade metamorphism, this study possible. All marine work in the Sea of Marmara since
what a shear zone looks like is shown in the inset from the Diancan 2000 forming the basis of this study was undertaken with the
Shang late Cainozoic metamorphic shear zone Ailao Shan along the active and cheerful support of the Turkish Navy. Detailed and
Red River strike-slip fault in Southeast Asia (copied from Cao et al. constructive reviews by Professors Kevin Burke and Mark Zoback
2011). We have flipped Cao et al.’s figure to make the shear right- substantially improved the original presentation.

Published by NRC Research Press


Şengör et al. 241

References the three-dimensional architecture of the Çinarcik Basin along the North
Anatolian Fault. Journal of Geophysical Research, 112: B06101. doi:10.1029/
Ahlgren, S.G. 2001. The nucleation and evolution of Riedel shear-zones as defor-
2006JB004548.
mation bands in porous sandstone. Journal of Structural Geology, 23: 1203–
Cormier, M.-H., Seeber, L., McHugh, C.M.G., Polonia, A., Çağatay, N., Emre, Ö.,
1214. doi:10.1016/S0191-8141(00)00183-8.
Gasperini, L., Görur̈, N., Bortoluzzi, G., Bonatti, E., Ryan, W.B.F., and
Akbayram, K., and Okay, A.I. 2012. The cumulative offset of North Anatolian
Newman, K.R. 2006. North Anatolian Fault in the Gulf of Izmit (Turkey): rapid
Fault in the Marmara region, northwest Turkey. EGU General Assembly,
vertical motion in response to minor bends of a nonvertical continental trans-
Vienna, Austria, Abstract pp. 240-2012.
form: Journal of Geophysical Research, 111: B04102. doi:10.1029/2005JB003633.
Akbayram, K., Şengör, A.M.C., and Özcan, E. 2013. The evolution of the Intra -
Crowell, J.C. 1974. Origin of late Cenozoic basins in southern California. In
Pontide suture: implications of the discovery of late Cretaceous – early Ter-
Tectonics and Sedimentation. Edited by W.R. Dickinson. Special Publication -
tiary melanges. Geological Society of America Abstracts with Programs,
Society of Economic Paleontologists and Mineralogists, 22, pp. 190–204.
45(7): 672.
Dewey, J.F., and Şengör, A.M.C. 1979. Aegean and surrounding regions: Complex
Anderson, E.M. 1951. The dynamics of faulting and dyke formation with applica-
multiplate and continuum tectonics in a convergent zone. Geological Society
tions to Britain, 2nd edition. Oliver and Boyd, Edinburgh and London, 206 pp.
of America Bulletin, 90: 84–92. doi:10.1130/0016-7606(1979)90<84:AASRCM>2.
Armijo, R., Pondard, N., Meyer, B., Uçarkuş, G., Mercier de Lépinay, B.,
0.CO;2.
Malavieille, J., Dominguez, S., Gustcher, M.-A., Schmidt, S., Beck, C.,
Dewey, J.F., Hempton, M.R., Kidd, W.S.F., Şaroğlu, F., and Şengör, A.M.C. 1986.
Çagatay, N., Çakir, Z., İmren, C., Eriş, K., Natalin, B., Özalaybey, S., Tolun, L.,
Shortening of continental lithosphere: the neotectonics of Eastern Anatolia -
Lefèvre, I., Seeber, L., Gasperini, L., Rangin, C., Emre, O., and Sarikavak, K.
a young collision zone. In Collision Tectonics. Edited by M.P. Coward and
2005. Submarine fault scarps in the Sea of Marmara pullapart (North Anato-
A.C. Ries. Geological Society of London Special Publication 19 (R.M. Shackleton
lian Fault): implications for seismic hazard in Istanbul. Geochemistry, Geo-
Volume), pp. 3–36.
physics, Geosystems 6: Q06009. doi:10.1029/2004GC000896.
Dolu, E., Gökaşan, E., Meriç, E., Ergin, M., Görüm, T., Tur, H., Ecevitoğlu, B.,
Ateş, A., Kayıran, T., and Sincer, İ. , 2003. Structural interpretation of the Mar-
Avşar, N., Görmüş, M., Batuk, F., Tok, B., and Çetin, O. 2007. Quaternary
mara region, NW Turkey, from aeromagnetic, seismic and gravity data. Tec-
evolution of the Gulf of Izmit (NW Turkey): A sedimentary basin under con-
tonophysics, 367: 41–99. doi:10.1016/S0040-1951(03)00044-1.
trol of the North Anatolian Fault Zone. Geo-Marine Letters, 27: 355–381.
Barka, A. 1992. The North Anatolian Fault zone. Annales Tectonicae, 6: 164–195. doi:10.1007/s00367-007-0057-3.
Barka, A. 1996. Slip distribution along the North Anatolian Fault associated with Ekström, G., and England, P. 1989. Seismic strain rates in regions of distributed
the large earthquakes of the period 1939 to 1967. Bulletin of the Seismologi- continental deformation. Journal of Geophysical Research, 94(B8): 10231–
cal Society of America, 86(5): 1238–1254. 10257. doi:10.1029/JB094iB08p10231.
Barka, A.A., and Kadinsky-Cade, K. 1988. Strike-slip fault geometry in Turkey and Eriş, K.K., Çağatay, N., Beck, C., and Mercier de Lepinay, B. 2012. Late-Pleistocene
its influence on earthquake activity. Tectonics, 7(3): 663–684. doi:10.1029/ to Holocene sedimentary fills of the Çınarcık Basin of the Sea of Marmara.
TC007i003p00663. Sedimentary Geology, 281: 151–165. doi:10.1016/j.sedgeo.2012.09.001.
Bartlett, W.L., Friedman, M., and Logan, J.M. 1981. Experimental folding and Eyal, Y. 2011. The Syrian Arc Fold System: age and rate of folding. Geophysical
faulting of rocks under confining pressure. Part IX. Wrench faults in lime- Research Abstracts, 13: EGU2011-7401, 2011 EGU General Assembly 2011.
stone layers. Tectonophysics, 79: 255–277. doi:10.1016/0040-1951(81)90116-5.
Gasperini, L., Polonia, A., Del Bianco, F., Favali, P., Marinaro, G., and Etiope, G.
Bayrakçı, G., Laigle, M., Bécel, A., Hirn, A., Taymaz, T., and Yolsal-Çevikbilen, S. , and 2012. Cold seeps, active faults and the earthquake cycle along the North
the SEISMARMARA Team. 2013. 3-D basement tomography of the northern Anatolian Fault system in the Sea of Marmara (NW Turkey). Bollettino di
Marmara trough by a dense OBS network at the nodes of a grid of controlled Geofisica Teorica ed Applicata, 53: 371–384.
source profiles along the North Anatolian Fault. Geophysical Journal Inter-
Gazioğlu, C., Yuc̈el,Z. Y.Doğan, E. 2005. Morphological features of major subma-
national, 194: 1335–1357. doi:10.1093/gji/ggt211.
rine landslides of Marmara Sea using multibeam data. Journal of Coastal
Bécel, A. 2006. Structure Sismique de la Faille Nord Anatolienne en Mer de Research, 214: 664–673. doi:10.2112/03-0060.1.
Marmara: École Doctorale des Sciences de la Terre, Institut de Physique du
Géli, L., Henry, P., Zitter, T., Dupré, S., Tryon, M., Çağatay, M.N.,
Globe de Paris, 286 pp.
Mercier de Lépinay, B., Le Pichon, X., Şengör, A.M.C., Görür, N., Natalin, B.,
Bécel, A., Laigle, M., de Voogd, B., Hirn, A., Taymaz, T., Yolsal-Çevikbilen, S., and
Uçarkuş, G., Özeren, S., Volker, D., Gasperini, L., Burnard, P., and
Shimamura, H. 2010. North Marmara Trough architecture of basin infill,
Bourlange, S. , and the Marnaut Scientific Party. 2008. Gas emissions and
basement and faults, from PSDM reflection and OBS refraction seismics.
active tectonics within the submerged section of the North Anatolian Fault
Tectonophysics, 490: 1–14. doi:10.1016/j.tecto.2010.04.004.
zone in the Sea of Marmara. Earth and Planetary Science Letters, 274: 34–39.
Beck, C., Mercier de Lépinay, B., Schneider, J.-L., Cremer, M., Çağatay, N.,
doi:10.1016/j.epsl.2008.06.047.
Wendenbaum, E., Boutareaud, S., Ménot, G., Schmidt, S., Weber, O., Eris, K.,
Armijo, R., Meyer, B., Pondard, N., Gutscher, M.-A., MARMARACORE Cruise Party, Gökçeoğlu, C., Tunusluoğlu, M.C., Görüm, T., Tur, H., Gökaşan, E., Tekkeli, A.B.,
Turion, J.-L., Labeyrie, L., Cortijo, E., Gallet, Y., Bouqueral, H., Görur̈, N., Gervais, A., Batuk, F., and Alp, H. 2009. Description of dynamics of the Tuzla landslide
Castera, M.-H., Londeix, L., de Resseguier, A., and Jaouen, A. 2007. Late Quaternary and its implications for further landslides in the northern slope and shelf of
co-seismic sedimentation in the Sea of Marmara’s deep basins. Sedimentary the Çınarcık basin (Marmara Sea, Turkey). Engineering Geology, 106: 133–
Geology, 199: 65–89. doi:10.1016/j.sedgeo.2005.12.031. 153. doi:10.1016/j.enggeo.2009.02.007.
Ben-Zion, Y. 2001. Dynamic ruptures in recent models of earthquake faults. Görür, N., and Elbek, Ş. 2014. Tectonic events responsible for shaping the Sea of
Journal of the Mechanics and Physics of Solids, 49: 2209–2244. doi:10.1016/ Marmara and its surrounding region. Geodinamica Acta, http://dx.doi.org/
S0022-5096(01)00036-9. 10.1080/09853111.2013.859346
Ben-Zion, Y., and Andrews, D.J. 1998. Properties and implications of dynamic Grall, C., Henry, P., Tezcan, D., Géli, L.M., de Lepinay, B., Rudkiewicz, J.-L.,
rupture along a material interface. Bulletin of the Seismological Society of Zitter, T., and Harmegnies, F. 2012. Heat flow in the Sea of Marmara Central
America, 88: 1085–1094. Basin: Possible implications for the tectonic evolution of the North Anatolian
Bilgin, T. 1967. Samanlı Dağları—Coğrafî Etüd. İstanbul Üniversitesi Edebiyat Fault. Geology, 40(1): 3. doi:10.1130/g32192.1.
Fakültesi yayınıNo.: 1294, İstanbul Üniversitesi Coğrafya Enstitüsü YayınıNo. Grall, C., Henry, P., Thomas, Y., Westbrook, G.K., Cagatay, M.N., Marsset, B.,
50, VIII+196 pp. Saritas, H., Cifci, G., and Géli, L. 2013. Slip rate estimation along the western
Burchfiel, B.C., and Chen, Z.L. 2012. Tectonics of the Southeastern Tibetan Pla- segment of the Main Marmara Fault over the last 405-490 ka by correlating
teau and its Adjacent Foreland. Geological society of America Memoir 210, Mass Transport Deposits. Tectonics, 32: 1587–1601. doi:10.1002/2012TC003255.
[iii]+231 pp. Grall, C., Henry, P., Westbrook, G.K., Cagatay, M.N., Thomas, Y., Marsset, B.,
Burnard, P., Bourlange, S., Henry, P., Géli, L., Tyron, M.D., Natal’in, B.A., Bornschneck, D., Saritas, H., Cifci, G., and Géli, L. 2014. Mass-Transport De-
Şengör, A.M.C., Özeren, M.S., and Çağatay, M.N. 2012. Constraints on fluid posits cyclicity related to glacial cycles and marine-lacustrine transitions on
origins and migration velocities along the Marmara Main Fault (Sea of Mar- the Sea of Marmara (Turkey) over the last 500 ka. Advances in Natural and
mara, Turkey) using helium isotopes. Earth and Planetary Science Letters, Technological Hazards Research. doi:10.1007/978-3-319-00972-8_53.
341: 68–78. Herece, E., and Akay, E. 2003. Kuzey Anadolu Fayı(KAF) Atlası/Atlas of North
Çağatay, N., Wulf, S., Guichard, F., Özmaral, A., Sancar, Ü., Akçer-Ön, S., Henry, P. Anatolian Fault (NAF). Maden Tetkik ve Arama Genel Müdürlüğü, Özel
and Gasperini, L. Tephra record from the Sea of Marmara for the last 70 ka Yayınlar Serisi 2, Ankara, [IV]+61 pp.+13 appendices as separate maps of
and its paleoceanographic implications. Quaternary Science Reviews. In 100 000 scale.
press. Hergert, T., and Heidbach, O. 2011. Geomechanical model of the Marmara Sea
Cao, S., Neubauer, F., Liu, J., Genser, J., and Leiss, B. 2011. Exhumation of the region - II. 3-D contemporary background stress field. Geophysical Journal
Diancang Shan metamorphic complex along the Ailao Shan-Red River belt, International, 185: 1090–1102. doi:10.1111/j.1365-246X.2011.04992.x.
southwestern Yunnan, China: Evidence from 40Ar/39Ar thermochronology. Howell, F.C., Arsebük, G., Kuhn, S.L., Özbaşaran, M., and Stiner, M.C. (Editors).
Journal of Asian Earth Sciences, 42: 525–550. doi:10.1016/j.jseaes.2011.04.017. 2010. Culture and biology at a Crossroads: The Middle Pleistocene Record of
Carton, H. 2005. Études Tectoniques en Mediterranée Orientale par Analyse de Yarımburgaz Cave (Thrace, Turkey). Ege Yayınları, İstanbul, xvi+[ii]+329 pp.+
Données de Sismique Réflexion, Mer de Marmara (Bassin de Çınarcık) et 11+27 plates.
Marge du Liban: École Doctorale des Sciences de la Terre, Institut de Physique Ketin, İ. 1948. Über die tektonisch-mechanischen Folgerungen aus den großen
du Globe de Paris, Laboratoire de Géosciences Marines, 297 pp. Anatolischen Erdbeben des letzten Dezenniums. Geol. Rundschau, 36: 77–83.
Carton, H., Singh, S.C., Hirn, A., Bazin, S., de Voogd, B., Vigner, A., Richolleau, A., Ketin, İ. 1969. Über die nordanatolische Horizontalverschiebung. Bulletin of the
Çetin, S., Ocakoğlu, N., Karakoç, F., and Sevilgen, V. 2007. Seismic imaging of Mineral Research and Exploration Institute of Turkey, 72: 1–28.

Published by NRC Research Press


242 Can. J. Earth Sci. Vol. 51, 2014

Ketin, İ. 1976. San Andreas ve Kuzey Anadolu Faylarıarasında bir karşılaştırma. active tectonics beneath the Sea of Marmara. Geophysical Journal Interna-
Türkiye Jeoloji Kurumu Bülteni, 19: 149–154. tional, 153: 133–145. doi:10.1046/j.1365-246X.2003.01897.x.
Kuşçu, İ. 2009. Cross-basin faulting and extinction of pull-apart basins in the Sea Rangin, C., Le Pichon, X., Demirbağ, E., and İmren, C. 2004. Strain localization in
of Marmara, NW Turkey. Turkish Journal of Earth Sciences, 18: 331–349. the Sea of Marmara: Propagation of the North Anatolian Fault in a now
Kuşçu, İ., Okamura, M., Matsuoka, H., Yamamori, K., Awata, Y., and Özalp, S. inactive pull-apart. Tectonics, 23: TC2014. doi:10.1029/2002TC001437.
2009. Recognition of active faults and stepover geometry in Femlik Bay, Sea Riedel, W. 1929. Zur Mechanik geologischer Brucherscheinungen. Centralbl. f.
of Marmara, NW Turkey. Marine Geology, 260: 90–101. doi:10.1016/j.margeo. Mineral. Geol. Pal., B, pp. 354–368.
2009.02.003. Şaroğlu, F. 1985. Doğu Anadolu’nun Neotektonik Dönemde Jeolojik ve Yapısal
Laigle, M., Bécel, A., de Voogd, B., Hirn, A., Taymaz, T., Özalaybey, S., Çetin, S., Evrimi. PhD Thesis, İstanbul Üniversitesi, Fen Bilimleri Enstitüsü,İstanbul,
Galvé, A., Karabulut, H., Lépine, J.C., Saatçılar, R., Sapin, M., Shimamura, H., 240 pp.+7 foldouts.
Murai, Y., Singh, S., Tan, O., Vigner, A., and Yolsal, S. 2008. A first deep seismic Seeber, L., Cormier, M.-H., McHugh, C., Emre, O., Polonia, A., and Sorlien, C.
survey of the Sea of Marmara: Deep basins and whole crust architecture and 2006. Rapid subsidence and sedimentation from oblique slip near a bend on
evolution. Earth and Planetary Science Letters, 270: 168–179. doi:10.1016/j. the North Anatolian transform fault in the Marmara Sea, Turkey. Geology,
34: 933–936. doi:10.1130/G22520A.1.
epsl.2008.02.031.
Şengör, A.M.C. 1979. The North Anatolian transform fault: its age, offset and
Le Pichon, X., and Kreemer, C. 2010. The Miocene-to-Present Kinematic Evolu-
tectonic significance. Journal of the Geological Society of London, 136: 269–
tion of the Eastern Mediterranean and Middle East and Its Implications for
282. doi:10.1144/gsjgs.136.3.0269.
Dynamics. Annual Review of Earth and Planetary Sciences, 38: 323–351. doi:
Şengör, A.M.C. 2011. İstanbul Boğazıniçin Boğaziçi’nde açılmıştır: in Ekinci, D.,
10.1146/annurev-earth-040809-152419.
editor, Fiziki Coğrafya Araştırmaları: sistematik ve Bölgesel (Hoşgören vol-
Le Pichon, X., Chamot-Rooke, N., Huchon, P., and Luxey, P. 1993. Implications ume). Türk Coğrafya Kurumu Yayınları, no. 6, pp. 57–102.
des nouvelles mesures de géodésie spatiale en Grèce et en Turquie sur Şengör, A.M.C. Transform faults, in Harff, J., Meschede, M., Petersen, S., Thiede, J.,
l’extrusion latérale de l’Anatolie et de l’Egée. C.R. Acad. Sci. Ser. II, 316: editors, Encyclopedia of Marine Sciences, Encyclopedia of Earth Sciences Series,
983–990. Springer Verlag, Berlin. In press.
Le Pichon, X., Şengör, A.M.C., Demirbağ, E., Rangin, C., İmren, C., Armijo, R., Şengör, A.M.C., and Kidd, W.S.F. 1979. The post-collisional tectonics of the
Görür, N., Çağatay, N., Mercier de Lepinay, B., Meyer, B., Saatçılar, R., and Turkish-Iranian Plateau and a comparison with Tibet. Tectonophysics, 55:
Tok, B. 2001. The active Main Marmara Fault. Earth and Planetary Science 361–376. doi:10.1016/0040-1951(79)90184-7.
Letters, 192(4): 595–616. doi:10.1016/S0012-821X(01)00449-6. Şengör, A.M.C., and Natal’in, B.A. 1996. Palaeotectonics of Asia: Fragments of a
Le Pichon, X., Chamot-Rooke, N., and Rangin, C. 2003. The North Anatolian Fault synthesis. In Tectonic Evolution of Asia, Rubey Colloquium. Edited by A. Yin
in the Sea of Marmara. Journal of Geophysical Research, 108(B4): 2179. doi: and M. Harrison. Cambridge University Press, Cambridge. pp. 486–640.
10.1029/2002JB001862. Şengör, A.M.C, Görür, N., and Şaroğlu, F. 1985. Strike-slip faulting and related
Le Pichon, X., İmren, C., Rangin, C., Şengör, A.M.C., and Siyako, M. 2013. The basin formation in zones of tectonic escape: Turkey as a case study. In Strike-
South Marmara Fault. International Journal of Earth Sciences, August 2013. slip Deformation, Basin Formation, and Sedimentation. Edited by K.T. Biddle
doi:10.1007/S00531-013-0950-0. and N. Christie-Blick. Society of Economic Paleontologists and Mineralogists
McBride, J.H., and Brown, L.D. 1986. Reanalysis of the cocorp deep seismic re- Special Publication 37 (in honor of J.C. Crowell), pp. 227–264.
flection profile across the San Andreas fault, Parkfield, California. Bulletin of Şengör, A.M.C., Özeren, S., Genç, T., and Zor, E. 2003. East Anatolian high pla-
the Seismological Society of America, 76(6): 1668–1686. teau as a mantle-supported, north-south shortened domal structure. Geo-
McKenzie, D.P. 1972. Active tectonics of the Mediterranean region. Geophysical physical Research Letters, 30: 8045. doi:10.1029/2003GL017858.
Journal of the Royal Astronomical Society, 30: 109–185. doi:10.1111/j.1365-246X. Şengör, A.M.C., Tüysüz, O., İmren, C., Sakınç, M., Eyidoğan, H., Görür, N.,
1972.tb02351.x. Le Pichon, X., and Rangin, C. 2005. The North Anatolian Fault. A new look:
McKenzie, D., and Jackson, J. 1983. The relationship between strain rates, crustal Annual Review of Earth and Planetary Sciences, 33: 37–112. doi:10.1146/annu
thickening, palaeomagnetism, finite strain and fault moments within a de- rev.earth.32.101802.120415.
forming zone. Earth and Planetary Science Letters, 65: 182–202. doi:10.1016/ Şengör, A.M.C., Özeren, M.S., Keskin, M., Sakınç, M., Özbakır, A.D., and Kayan, I.
0012-821X(83)90198-X. 2008. Eastern Turkish high plateau as a small Turkic-type orogen. implica-
Naylor, M.A., Mandl, G., and Supesteijn, C.H.K. 1986. Fault geometries in tions for post-collisional crust-forming processes in Turkic-type orogens.
Earth Science Reviews, 90: 1–48. doi:10.1016/j.earscirev.2008.05.002.
basement-induced wrench faulting under different initial stress states. Jour-
nal of Structural Geology, 8(7): 725–830. doi:10.1016/0191-8141(86)90021-0. Şengör, A.M.C., Uçarkuş, G., İmren, C., Rangin, C., Le Pichon, X., Özeren, S., and
Natal’in, B. 2011. Broad shear zones and narrow strike-slip faults in orogens
Okay, A.I., Demirbağ, E., Kurt, H., Okay, N., and Kuşçu, İ. 1999. An active, deep
and their role in forming the orogenic architecture: The North Anatolian
marine strike-slip basin along the North Anatolian Fault in Turkey. Tecton-
Fault as an active example, Joint Meeting GeoMunich, Fragile Earth. Geolog-
ics, 18(1): 129–147. doi:10.1029/1998TC900017.
ical Processes from Global to Local Scales, associated Hazards and Resources,
Okay, A. İ., Kaşlılar-Özcan, A., İmren, C., Boztepe-Güney, A., Demirbağ, E., and 14-2, A18, Munich, Germany, September 4–7.
Kuşçu, İ. 2000. Active faults and evolving strike-slip basins in the Marmara Şentürk, K., and Okay, A.I. 1984. Blueschists discovered east of Saros Bay in
Sea, northwest Turkey: a multichannel seismic reflection study. Tectono- Thrace. Bulletin of the Mineral Research and Exploration Institute of Turkey,
physics, 321: 189–218. doi:10.1016/S0040-1951(00)00046-9. 97/98: 72–75.
Örgülü, G., and Aktar, M. 2001. Regional moment tensor inversion for strong Shillington, D.J., Seeber, L., Sorlien, C.C., Steckler, M.S., Kurt, H., Dondurur, D.,
aftershocks of the August 17, 1999 İzmit earthquake (Mw=7.4). Geophysical Çifçi, G., İmren, C., Cormier, M.-H., McHugh, C.M.G., Gürçay, S., Poyraz, D.,
Research Letters, 28(2): 371–374. doi:10.1029/2000GL011991. Okay, S., Atgın, O., and Diebold, J.B. 2012. Evidence for widespread creep on
Özalaybey, S., Ergin, M., Aktar, M., Tapırdamaz, C., Biçmen, F., and Yörük, A. the flanks of the Sea of Marmara transform basin from marine geophysical
2002. The 1999 İzmit earthquake sequence in Turkey: seismological and data. Geology, 40: 439–442. doi:10.1130/G32652.1.
tectonic aspects. Bulletin of the Seismological Society of America, 92(1): 376– Sorlien, C.C., Akhun, S.D., Seeber, L., Steckler, M.S., Shillington, D.J., Kurt, H.,
386. doi:10.1785/0120000838. Çifçi, G., Timur Poyraz, D., Gurcay, S., Dondurur, D., İmren, C., Perincek, E.,
Özalaybey, S., Karabulut, H., Ergin, M., Aktar, M., Tapırdamaz, C., Biçmen, F., Okay, S., Küçük, H.M., and Diebold, J.B. 2012. Uniform basin growth over the
and Yörük, A. 2003. Seismogenic zones of the Sea of Marmara: recent seis- last 500ka, North Anatolian Fault, Marmara Sea, Turkey, Tectonophysics,
micity and focal mechanism solutions. EGS-AGU-EUG Joint Assembly, 5, 518 –521: 1–16.
p. 12680, Nice, France. Taymaz, T., Jackson, J.A., and McKenzie, D. 1991. Active tectonics of the north
Özeren, M.S., Çağatay, M.N., Postacıoğlu, N., Şengör, A.M.C., Görür, N., and and central Aegean Sea. Geophysical Journal International, 106: 433–490.
Eriş, K. 2010. Mathematical modeling of a potential tsunami associated with doi:10.1111/j.1365-246X.1991.tb03906.x.
a late glacial submarine landslide in the Sea of Marmara. Geo-Marine Letters, Tchalenko, J.S. 1970. Similarities between shear zones of different magnitudes.
30: 523–539. doi:10.1007/s00367-010-0191-1. Geological Society of America Bulletin, 81: 1625–1640. doi:10.1130/0016-
Parke, J.R., Minshull, T.A., Anderson, G., White, R.S., McKenzie, D., Kuşçu, İ., 7606(1970)81[1625:SBSZOD]2.0.CO;2.
Bull, J.M., Görür, N., and Şengör, A.M.C. 1999. Active faults in the Sea of Uçarkuş, G. 2010. Active faulting and earthquake scarps along the North Anato-
Marmara, western Turkey, imaged by seismic reflection profiles. Terra Nova, lian fault in the Sea of Marmara: Ph.D. Thesis, İstanbul Technical University,
11: 223–227. doi:10.1046/j.1365-3121.1999.00248.x. Eurasia Institute of Earth Sciences, xxx+178 pages.
Parke, J.R., White, R.S., McKenzie, D., Minshull, T.A., Bull, J.M., Kuşçu, İ., Wilcox, R.E., Harding, T.P., and Seely, D.R. 1973. Basic wrench tectonics. The
Görür, N., and Şengör, A.M.C. 2002. Interaction between faulting and sedi- American Association Petroleum Geologists Bulletin, 57(1): 74–96.
mentation in the Sea of Marmara, western Turkey. Journal of Geophysical Wilson, J.T. 1965. A new class of faults and their bearing on continental drift.
Research, 107(B11): 2286. doi:10.1029/2001JB000450. Nature, 207(4995): 343–347. doi:10.1038/207343a0.
Pavoni, N. 1962. Die Nordanatolische Horizontalverschiebung. Geologische Zitter, T.A.C., Grall, C., Henry, P., Özeren, S., Çağatay, M.N., Şengör, A.M.C.,
Rundschau, 51: 122–139. doi:10.1007/BF01803232. Gasperini, L., Mercier de Lépinay, B., and Géli, L. 2012. Distribution, morphol-
Pinar, A., Kuge, K., and Honkura, Y. 2003. Moment tensor inversion of recent ogy and triggers of submarine mass wasting in the Sea of Marmara. Marine
small to moderate sized earthquakes: implications for seismic hazard on Geology, 329 –331: 58–74. doi:10.1016/j.margeo.2012.09.002.

Published by NRC Research Press


Copyright of Canadian Journal of Earth Sciences is the property of Canadian Science
Publishing and its content may not be copied or emailed to multiple sites or posted to a
listserv without the copyright holder's express written permission. However, users may print,
download, or email articles for individual use.

Você também pode gostar