Canadian Journal of Earth Sciences

Spatio-temporal behaviour of continental transform faults: Implications from the late Quaternary slip history of the ,

Journal: Canadian Journal of Earth Sciences

Manuscript ID cjes-2018-0308.R1

Manuscript Type: Article

Date Submitted by the 04-Apr-2019 Author:

Complete List of Authors: Zabcı, Cengiz; Teknik Üniversitesi, Jeoloji Mühendisliği Bölümü

North Anatolian Fault, transform faults, slip rate, strain transfer, off-fault Keyword: deformationDraft Is the invited manuscript for Understanding tectonic processes and their consequences: A tribute to consideration in a Special A.M. Celal Sengor Issue? :

https://mc06.manuscriptcentral.com/cjes-pubs Page 1 of 59 Canadian Journal of Earth Sciences

1 Spatio-temporal behaviour of continental transform faults: 2 Implications from the late Quaternary slip history of the North 3 Anatolian Fault, Turkey 4 5 Cengiz Zabcı 6 7 İstanbul Teknik Üniversitesi, Ayazağa Yerleşkesi, Maden Fakültesi, Jeoloji Mühendisliği 8 Bölümü 34469 Maslak, İstanbul 9 Tel: +90 2122856162, e-mail: [email protected] 10 11 12 ABSTRACT

13 The slip history of the North Anatolian Fault (NAF) is constrained by displacement and age

14 data for the last 550 ka. First, I classified all available geological estimates, being member of

15 three groups: Model I for the eastern, Model II for the central and Model III for the western 16 segments where the North Anatolian ShearDraft Zone gradually widens from east to west. The 17 short-term uniform slip solutions yield similar results, 17.5 +4/-3.5 mm/a, 18.9 +3.7/-3.3

18 mm/a and 16.9 +1.2/-1.1 mm/a from east to the west. Although these model rates do not show

19 any significant spatial variations among themselves, the correlation with geodetic estimates,

20 ranging between 15 mm/a and 28 mm/a for different sections of the NAF, displays significant

21 discrepancies especially for the central and western segments of the fault. Discrepancies

22 suggest that the majority of strain is accumulated along the NAF, but some portion of it is

23 distributed along secondary structures of the North Anatolian Shear Zone. The deformation

24 rate is constant at least for the last 195 ka, whereas the limited number of data show strain

25 transfer from northern to the southern strand between 195 and 320 ka BP in the Marmara

26 Region when the incremental slip rate decreases to 13.2 +3.1/-2.9 mm/a for the northern

27 strand of the NAF. Considering the possible uncertainties of incremental displacements and

28 their timings, it is clear that more studies on slip rate are needed at different sites, including

29 major structural elements of the North Anatolian Shear Zone. Although most of the strain is

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 2 of 59

30 localized along the main displacement zone, the NAF, secondary structures are still capable of

31 generating earthquakes that can hardly reach Mw 7.

32 Keywords: North Anatolian Fault, transform faults, slip rate, strain transfer, off-fault

33 deformation

34 INTRODUCTION

35 Transform faults are subjected to great interest especially since the idea of “continuous

36 network of mobile belts”, first introduced by Wilson (1965). Oceanic transform faults are

37 known to have narrow deformation zones, whereas their continental counterparts usually

38 display a scattered seismicity and/or distribution of multiple parallel/sub-parallel tectonic

39 structures along their modern and ancient examples (e.g., Bilham and King 1989; Lacombe et

40 al. 1993; Şengör 2014; Şengör et al. 2019).Draft Thus, it is a challenging issue to understand the

41 spatial and temporal behaviour of continental transform faults considering their complex

42 structures. In this manner, the geological and geodetic studies together can provide invaluable

43 data, especially by identifying the shear zones and quantifying their rate of deformation

44 (Chery and Vernant 2006; Meade et al. 2013; Segall 2002; Segall and Davis 1997; Thatcher

45 2009). Long-term (>10 ka) geological rates are calculated as average values representing

46 multiple seismic cycles, whereas the geodetic values are like snapshots within a single cycle

47 considering the length of continuous monitoring only for the last few decades. Therefore, it

48 may be admitted to say that geological rates provide a temporal picture of fault zones, while

49 the geodetic results give more detailed spatial information.

50 Taking into account the limits of field measurements and approaches in geodesy and geology,

51 it is not surprising to have discrepancies between the results of these two methods (e.g., Chery

52 and Vernant 2006; Dixon et al. 2003; Dolan and Meade 2017). Although it is possible to have

53 errors both in geological and geodetic measurements, any variations in the rate of fault slip

https://mc06.manuscriptcentral.com/cjes-pubs Page 3 of 59 Canadian Journal of Earth Sciences

54 can be explained by the following processes: (a) mechanical layering of the earth’s crust, such

55 as the interaction of the elastic upper crust and the viscoelastic rheology of the lower crust and

56 upper mantle (Chery and Vernant, 2006; Chuang and Johnson 2011; Dixon et al. 2003; Meade

57 and Hager 2004) with post-seismic stress release in the non-seismogenic layer as response to

58 the major events (Kenner and Simons 2005) and/or hardening or weakening of ductile shear

59 zones (Chery and Vernant, 2006; Dolan et al. 2007; Oskin et al. 2008), (b) structural

60 processes, such as the off-fault permanent deformation, including cleavage development,

61 granular flow, folding, secondary faulting, pressure solution, and microcracking

62 (Bartholomew et al. 2014; Dolan and Haravitch 2014; Duebendorfer et al. 1998; Gold et al.

63 2015; Herbert et al. 2014; McClymont et al. 2009; Petersen et al. 2011; Roten et al. 2017;

64 Shelef and Oskin 2010; Titus et al. 2011), structural transformation of horizontal motion into

65 local shortening or extension with existenceDraft of thrust or normal faults along strike-slip fault

66 systems (Daeron et al. 2004; Kirby et al. 2007; Mériaux et al. 2005; Sançar et al. 2018; Xu et

67 al. 2005; Zabcı et al., 2015; Zhang et al. 2007) and/or co-dependent slip along conjugate

68 faults (Hubert-Ferrari et al. 2003; Peltzer et al. 2001) and parallel – sub-parallel fault strands

69 (Bennett et al. 2004; Dolan et al. 2007; Gaudemer et al. 1995; Ye and Liu 2017), and (c) the

70 effect of external processes, such as climatically-driven changes in crustal water content

71 (Chery and Vernant 2006) and/or changing surface loads such as those from glaciers or large

72 lakes (Grollimund and Zoback 2001; Hampel and Hetzel 2006; Hampel et al. 2007; Hampel

73 et al. 2009; Turpeinen et al. 2008). In consideration with all these processes, integrated-

74 analyses of geological and geodetic rates provide a better understanding on the spatial and

75 temporal behaviour of the continental transform faults.

76 The North Anatolian Shear Zone (NASZ) is one of the well-known continental transform

77 fault systems, which connects the East Anatolian High Plateau in the east and the Aegean

78 Taphrogen in the west (Ambraseys, 1970; Barka, 1992; Ketin 1948; Şengör, 1979; Şengör et

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 4 of 59

79 al. 2005). The North Anatolian Fault, the most prominent member of the NASZ, shows a

80 remarkable seismic activity between 1939 and 1999 when the westward migrating earthquake

81 sequence generated surface ruptures of more than 1000 km, leaving unbroken only the Main

82 Marmara Fault (MMF) to the west and the Yedisu Fault (YF) to the east (Barka 1996; Barka

83 et al. 2002; Ketin 1969; Stein et al. 1997). This well-exposed tectonic feature draws a great

84 attention of multi-disciplinary studies including many geodetic (e.g., Çakır et al., 2014;

85 Hussain et al., 2018; McClusky et al. 2003; McClusky et al. 2000; Reilinger et al. 2006) and

86 geological (e.g., Hubert-Ferrari et al. 2002; Kozacı et al. 2009; Kozacı et al. 2007) slip rate

87 estimates, which all provide invaluable data to increase our knowledge on many aspects of the

88 deformation rate along this large continental fault.

89 In this study, my aim is to analyse the Draftspatial and temporal distribution of geological slip rate 90 studies in order to understand the ongoing processes of the discrepancy between geodetic and

91 geological slip rates and late Quaternary slip history of the NASZ. My attempt does not target

92 to include the re-interpretation or re-model any of geodetic data, but try to fully review the

93 available slip rates of various time intervals. I think that the systematic comparison between

94 the geodetic and geological rates of slip may provide clues on the tectonic processes

95 governing the spatio-temporal behaviour of the NASZ and similar continental transform

96 systems.

97 TECTONIC FRAME OF ANATOLIA AND THE NORTH ANATOLIAN FAULT

98 The interaction of three major plates of Eurasia, Africa, Arabia and the smaller Anatolian

99 Scholle generates a complex tectonic setting, which is mainly controlled by the Bitlis –

100 Zagros subduction-collision to the east and/or the Hellenic subduction and rollback to the

101 west, in the eastern Mediterranean (Fig. 1). In this tectonic environment, the driving

102 mechanisms of westward extrusion of Anatolia with respect to Eurasia are (a) tectonic escape

103 system caused by the post-collisional convergence of Eurasia and Arabia creating forces at its

https://mc06.manuscriptcentral.com/cjes-pubs Page 5 of 59 Canadian Journal of Earth Sciences

104 boundaries, supplied with the extra forces applied to Anatolia from beneath or force from the

105 gravitational potential of the East Anatolia High Plateau (McKenzie 1972; Özeren and Holt

106 2010; Şengör et al. 1985), (b) slab pull of the Hellenic subduction (Armijo et al., 1999;

107 Chorowicz et al. 1999; Philippon et al. 2014; Reilinger et al. 2006), (c) asthenospheric flow

108 dragging the circular motion of lithosphere from the Levant in the east to Anatolia and

109 Aegean in the west (Le Pichon & Kreemer, 2010), (d) mantle upwelling underneath Afar and

110 the large-scale flow associated with the whole mantle, Tethyan convection cell (Faccenna et

111 al. 2013), or (e) combinations of these mechanisms (e.g., Paul et al. (2014)’s suggestion on

112 the effect of the large-scale mantle flow with the regional-scale flow driven by the suction

113 from the Hellenic slab). This westward relative motion is mainly accommodated along two

114 major tectonic structures, the NASZ and the East Anatolian Shear Zone (EASZ), respectively

115 forming the northern and eastern boundariesDraft of the Anatolian Scholle (Dewey and Şengör

116 1979; Şengör 1979; Şengör 1980; Şengör et al. 1985).

117 Figure 1.

118 The NASZ connects the East Anatolian High Plateau with the Aegean Taphrogen along a

119 course roughly parallel to the Black Sea coast of about 1600 km length (Barka 1992; Şengör

120 and Zabcı 2019; Şengör et al. 2005). This dextral shear zone (The yellow shaded region in

121 Fig. 1) is confined to the Tethyside accretionary complexes in northern Turkey that in general

122 widens from east to west reaching its maximum width of about 100 km in the Marmara Lobe

123 (See fig. 4 in Şengör et al., 2005). The deformation zone is significantly confined to a narrow

124 zone, hardly exceeding 10 km at its easternmost sections, especially around Erzincan in the

125 east and some 150-200 km to the west of it, only making an exception and becomes wider

126 again near Karlıova (Karaoğlu et al. 2017; Sançar et al. 2018; Şengör and Zabcı 2019; Şengör

127 et al. 2005; Zabcı et al. 2015). To the west of Niksar, the width of the shear zone increases, by

128 generating a broad and internally deforming wedge-shaped region, the Amasya Shear Zone

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 6 of 59

129 (ASZ), which is mainly characterised by many parallel/sub-parallel elongated basins, and

130 narrow uplifts and widely distributed seismicity, especially around its main structural

131 elements Esençay – Suluova and Ezinepazarı – Sungurlu faults (Erturaç 2009; Erturaç and

132 Tüysüz 2012; Erturaç et al. this issue). Further to the west, the deformation zone significantly

133 widens in the Marmara Region by the branching into two main strands and multiple

134 secondary faults, creating a complex structural pattern (Barka and Kadinsky-Cade 1988;

135 Görür and Elbek 2013; Le Pichon et al. 2014; Le Pichon et al. 2001; Şengör et al. 2005;

136 Şengör et al. 2014; Yaltırak 2002). This westward increasing width of the NASZ is also well

137 supported with large-scale morphological features, especially in the courses of the Anatolian

138 rivers. In the east, the Elmalı/Periçayı and Karasu rivers, tributaries of Fırat (Euphrates), are

139 displaced along a narrow corridor, but further to the west, the Yeşilırmak (Iris) and Kızılırmak

140 (Halys) rivers display a broader zone ofDraft dextral deflection. In the Marmara Region, Sakarya

141 and Susurluk rivers together display a broader deflection, because of a distributed

142 deformation of about 100 km-width (Şengör and Zabcı 2019; Şengör et al. 2005).

143 The most prominent member (or the main displacement zone, where the most of the strain has

144 been localised) of the NASZ, the North Anatolian Fault (NAF), has a simple geometry

145 between Erzincan in the east and Niksar in the west, except for a 20° restraining bend to the

146 northwest of the Erzincan Basin (Barka and Kadinsky-Cade 1988; Tatar 1978). However, to

147 the south of Niksar, the main trunk of the fault jumps to the north, making a 10 km-wide

148 releasing step-over, while two offshoots, the Esençay – Suluova Fault (EsF and SuF) and the

149 Ezinepazarı – Sungurlu Fault (Erturaç 2009; Erturaç and Tüysüz 2012, Erturaç et al. this

150 issue), bifurcate from the main course of the fault (Fig. 2). Further to the west, the NAF splits

151 into several parallel/subparallel fault branches, which are generally classified as being

152 member of the northern or the southern strands in the Marmara Region. The geodetic studies

153 and distribution of earthquakes simply show that most of the strain accumulation is localised

https://mc06.manuscriptcentral.com/cjes-pubs Page 7 of 59 Canadian Journal of Earth Sciences

154 along the northern strand (Ergintav et al. 2014; Reilinger et al. 2006; Şengör et al., 2005),

155 whereas the less active southern strand is usually expressed by multiple and discrete

156 structures (fig. 5 in Şengör et al. 2005). The NAF is suggested to be originated approximately

157 between 13 and 11 Ma in the east and propagated slowly (with the velocity of 11 cm/a) to the

158 west, reaching the Marmara Sea a few hundred thousand years ago, which has already been

159 deformed by the secondary structures of the NASZ since the initiation (or even the pre-

160 initiation stage) of the main displacement zone in the east (Şengör and Zabcı 2019; Şengör et

161 al. 2005). An alternative model based on early Pliocene offset of folded structures also

162 suggests a westward propagation, but in which the NAF reached the Marmara Region 5 Ma

163 ago (Armijo et al. 1999).

164 The NAF showed a remarkable seismicityDraft in the 20th century, especially with the westward 165 migrating earthquake sequence between the 1939 and 1999 (Akyüz et al. 2002; Ambraseys

166 1970; Barka 1996; Barka et al. 2002; Barka and Kadinsky-Cade 1988; Blumenthal 1945;

167 Blumenthal et al. 1943; Emre et al. 2003; Hartleb et al. 2002; Ketin 1969; Konca et al. 2010;

168 Kondo et al. 2005; Pamir and Ketin 1941; Pucci et al. 2006; Stein et al. 1997; Uçarkuş et al.

169 2011). These events with the addition of off-sequence, 1912a Ganos and 1912b Saros (Aksoy

170 2009; Aksoy et al. 2010a; Altunel et al. 2004), 1949 Elmalı (Barka and Kadinsky-Cade 1988;

171 Tutkun and Hancock 1990; Zabcı et al. 2015), 1992 Erzincan (Barka and Eyidoǧan 1993),

172 and 2014 Samothraki – Gökçeada (Konca et al. 2018), earthquakes have left unbroken only

173 the Main Marmara Fault (MMF) to the west and the Yedisu Fault (YF) for more than 200

174 years along the main strand of the NAF (Fig. 2).

175 Figure 2.

176 There is also a great interest for understanding the spatial and temporal behaviour of this

177 remarkable fault by measuring its deformation rate using geodetic (Fig. 3) and geological

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 8 of 59

178 (Fig. 4) methods. The increasing number of geodetic measurements, mostly by using the GPS

179 (or the Global Navigation Satellite System (GNSS) by including all navigation satellite

180 networks) and Interferometric Synthetic-Aperture Radar (InSAR) provide a very strong

181 spatial control on the distribution of deformation for most parts of the fault (Fig. 3).

182 Based on a dense GPS network, the results of elastic-block model, which cover the entire

183 fault zone, yield rates ranging between 28±0.3 mm/a and 24.2±0.2 mm/a (see the grey

184 horizontal bars in Fig. 3b) that are generally faster along the western sections of the fault

185 (Reilinger et al. 2006). On the other hand, a local GPS-based block model suggests slower

186 rates of 22.5±0.4 mm/a and 22.8±0.4 mm/a (see the green horizontal bars in Fig. 3b) for the

187 central and eastern sections of the fault between Bayramören in the west and Erzincan in the 188 east (Aktuğ et al. 2015). Another localDraft elastic block model for the most eastern sections of the 189 fault claims considerably slower rates of 11.8±0.3 mm/a and 12.1±0.3 mm/a (see the orange

190 horizontal bars in Fig. 3b) to the east and west of Erzincan, respectively (Aktuğ et al. 2013a).

191 GPS-based velocity profiles yield more scattered rates, which generally tend to decrease

192 eastward along the fault. In the Marmara Region, four profiles across the northern strand

193 suggest rates of 20±1 mm/a, 15±2 mm/a and 25±2 mm/a (see the blue vertical bars in Fig.

194 3b), respectively from western to the eastern sections (Ergintav et al. 2014). The profile with

195 the lowest velocity, 15±2 mm/a, crosses the Çınarcık Basin (profile B of Ergintav et al. 2014),

196 where the fault makes a releasing bend and some portion of the horizontal motion is

197 transformed into the vertical deformation according to tilted deposits (Seeber et al. 2006).

198 Furthermore, one of the profiles of Ergintav et al. (2014) displays signs of creeping for the

199 central Marmara Sea, which is opposed by studies showing multiple evidence for

200 palaeoearthquakes (Yakupoğlu et al. 2019) and evidence for a locked state or very moderate

201 surface creeping according to sea floor acoustic ranging (Sakic et al. 2016) at the Kumburgaz

202 Basin, the presence of fresh fault scarps (Uçarkuş, 2019), and a partial creep by using

https://mc06.manuscriptcentral.com/cjes-pubs Page 9 of 59 Canadian Journal of Earth Sciences

203 submarine geodetic measurements at the Western High (Yamamoto et al. 2019). Thus, this

204 velocity profile is excluded in this study. Another GPS velocity profile yields a similar slip

205 rate result, 25.6 ± 4.4 mm/a with a higher uncertainty (see the purple vertical bar in Fig. 3b) to

206 the east of the Marmara Sea (Vernant 2015). Results of the closely spaced GPS stations

207 indicate an eastward decreasing in deformation rate from 24.0 ± 2.9 to 16.2 ± 2.3 mm/yr (see

208 the yellow vertical bars in Fig. 3b) between Niksar in the west and Erzincan in the east (Tatar

209 et al. 2012).

210 Figure 3.

211 There is also an increasing trend in using InSAR for monitoring the post-seismic or inter-

212 seismic activity of the NAF almost for the last two decades (Fig. 3c). The first pioneering

213 study of measuring the interseismic strainDraft accumulation proposes a right-lateral slip rate of

214 24.5 ± 7.5 mm/a (green vertical bar in Fig. 3c) for the eastern segments of the fault

215 immediately west of Erzincan (Wright et al. 2001). Later studies on the same fault segments

216 display results between 20 ± 3 mm/a (see the blue vertical bar in Fig. 3c) and 23 ± 3 mm/a

217 (see the thick and black vertical bar in Fig. 3b) with significantly lower (better) uncertainties

218 for the same region (Walters et al. 2014; Walters et al. 2011). The value of 20 ± 3 mm/a (see

219 the yellow bars in Fig. 3c) are also suggested by another study with a larger geographical

220 coverage with only one exception to the west of Niksar where the rate increases to 25 ± 3

221 mm/a (Çakır et al. 2014). The easternmost parts of the fault are also suggested to deform with

222 a similar rate of about 20 mm/a (brown vertical bar in Fig. 3c) (Cavalié and Jónsson 2014).

223 InSAR velocity profiles constructed for almost the entire NAF again propose a westward

224 decreasing in rates from 28.5 ± 6.5 mm/a to 24 ± 3 mm/a (see the dark grey vertical bars in

225 Fig. 3c) between İzmit and west of Erzincan (Hussain et al. 2018). In addition, there are two

226 more discrete InSAR-based slip rate estimates, one suggesting 23.5 ± 3.5 mm/a (see the

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 10 of 59

227 orange vertical bar in Fig. 3c) for the Ganos Fault (Motagh et al. 2007) and another giving 20

228 mm/a (see the dark green vertical bar in Fig. 3c) for the central part (Peyret et al. 2013).

229 METHODS AND MATERIALS

230 In this study, my first purpose is to classify the available geological slip rate estimates (Fig. 4

231 and Table 1), which are sparsely located along different sections of the NAF, in order to make

232 a better correlation with geodetic velocity rates. Secondly, I used a Monte-Carlo approach to

233 derive the fault slip history by using the method suggested by Gold and Cowgill (2011). I

234 briefly explain these steps below.

235 Figure 4.

236 Collecting and Organising Data Draft 237 First, I collected most of published estimates of geological slip rates in order to prepare a

238 database as full as possible. Then, I organised each entry to be enlisted in a geographical

239 order. I refer all palaeoseismology- and morphochronology-based slip rate estimates as

240 ‘geological’ throughout the text, but each of them is identified of being palaeoseismological

241 or morhphochronological in Table 1. If these estimates had been calculated without taking

242 consideration of any uncertainty either in their displacement or in time span, I recalculated

243 geological slip rates by simply adding the missing uncertainties. For example, I added 10% of

244 uncertainty both to the displacement and their suggested timing for the streams of different

245 orders along the Ganos Fault (Aksoy 2009; Aksoy et al. 2010b). In case of the depocenter

246 displacement of the Çınarcık Basin (Kurt et al. 2013), I re-interpolated the supplementary

247 information of Kurt et al. (2013) in order to precisely measure the offset of the depocenter

248 between each identified seismic horizons. Moreover, I used the uncertainty age limits of Grall

249 et al. (2013) for the timing of these seismic horizons that correspond to Marine Isotopic

250 Stages (MIS). In terms of palaeoseismological studies, I used the incremental offsets of each

https://mc06.manuscriptcentral.com/cjes-pubs Page 11 of 59 Canadian Journal of Earth Sciences

251 earthquakes and ages of events predating these displaced features for the Kavakköy

252 (Rockwell et al. 2009), Güzelköy (Meghraoui et al. 2012) and Demirtepe (Kondo et al. 2010)

253 trench sites. All calculations of slip rate estimates are evaluated by using the probability

254 density functions according to Zechar and Frankel (2009) that include 2-sigma error intervals.

255 Table 1 summarizes all organised geological slip rate studies and my recalculations.

256 Classification of Slip Rate Sites

257 Geodetic rates, especially deduced from elastic block models, display the total strain

258 accumulation on boundaries of each identified block. In order to make an adequate correlation

259 between geodetic and geological values, I do not take into account any geological slip rate site

260 in regions of distributed strain where the deformation is expressed with closely spaced

261 multiple parallel/sub-parallel fault strandsDraft and/or any other active structures. Thus, I excluded

262 sites at Saros Gulf Fault (Gasperini et al. 2011), Düzce Fault (Pucci et al. 2008), Mudurnu

263 Valley Fault (Erturaç et al. 2019), and Elmalı Valley Fault (Zabcı et al. 2015), which are

264 characterised by parallel/subparallel structural elements of the deformation zone (Fig. 5).

265 Moreover, I also did not include any site from the southern strand of the NAF (Gemlik Bay;

266 Gasperini et al. 2011; Vardar et al. 2014) in the Marmara Region to avoid miscorrelation.

267 There are only two more exclusions, which are not either from the southern strand or the

268 region of distributed strain; İzmit Gulf (Gasperini et al. 2011; Polonia et al. 2004) and Alıç

269 (Sugai et al. 1999) sites, because calculated slip rates at these sites are suggested to be

270 underestimated as the channel offset in the İzmit Gulf (Dolan 2009) and the co-seismic slip of

271 the 1943 earthquake in the Alıç Site (Emre et al. 2005) are claimed to be mismeasured.

272 Figure 5.

273 In the second step, I grouped the rest of the geological slip rate estimates to be the member of

274 one of three groups along the NAF (Table 2). First group inludes Ayanoğlu (Zabcı 2012) and

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 12 of 59

275 Koçyatağı (Hubert-Ferrari et al. 2002; Zabcı 2012) sites, which are located along the fault

276 section between Yedisu in the east and Niksar in the west. This part of the NAF is known to

277 have a simple geometry where the deformation is highly localised along a single branch of

278 faulting. The second group, including the Destek (Hubert-Ferrari et al. 2002), Karapürçek

279 (Zabcı 2012), Tahtaköprü (Kozacı et al. 2009), Eksik (Kozacı et al. 2007), Üçoluk (Hubert-

280 Ferrari et al. 2002; Zabcı 2012), Berçin (Hubert-Ferrari et al. 2002), Demirtepe (Kondo et al.

281 2010) and Gerede (Hubert-Ferrari et al. 2002) sites, covers a region between Niksar in the

282 east and Bolu in the west. The third group includes the geological slip rate sites in the

283 Marmara Region between Akyazı and Saros (Aksoy 2009; Aksoy et al. 2010b; Dikbaş et al.

284 2018; Grall et al. 2013; Kurt et al. 2013; Meghraoui et al. 2012; Rockwell et al. 2009), where

285 the NASZ reaches its maximum width. In summary, the final grouping is done according to

286 the width of the NASZ, which generallyDraft becomes wider from east to west (please see the

287 section on the ‘Tectonic Frame of Anatolia and the North Anatolian Fault’ for the relevant

288 discussions).

289 Slip History Analyses

290 One of the main purposes of this study is to understand whether the NAF has experienced any

291 secular variations in its slip rate history. I used the Monte Carlo-based approach of Gold and

292 Cowgill (2011) for estimating the distribution of geologically reasonable fault-slip histories

293 that fit the displacement and age data from a population of offset landforms with known ages.

294 In this approach, first displacement (d) and time (t) envelopes are constructed from primary

295 age and offset determinations. Secondly, bounds of the d–t envelopes are trimmed in

296 assumption of the no-negative slope rule (the sense of motion cannot be reversed). For slip

297 history iteration, one point is selected in each trimmed d–t envelope using a uniformly

298 distributed pseudorandom number generator so that calculations can be reproduced and tested.

299 The line that connects the resulting set of points is one of the slip history candidates. Then,

https://mc06.manuscriptcentral.com/cjes-pubs Page 13 of 59 Canadian Journal of Earth Sciences

300 the calculated slip history line is tested to control if it has a non-negative slope. For a valid

301 slip history, only line segments with slope ≥ 0 are accepted. In the final stage, the iteration of

302 choosing random points and slope checking are repeated with a certain number of times. In all

303 relevant calculations, the iteration number is selected to be 1000 in this study (For details

304 please see supplementary material).

305 I separately analysed all selected geological slip rate estimates in three identified groups

306 (Table 2; they will be referred as models in the following sections). In the Model I, I ran the

307 model twice first for all slip rate estimates of the group and then a second time only for

308 morphochronology- or palaeoseismology-based studies (Fig. 6 and Table 3). I repeated the

309 same procedure also for the Model II (Fig. 7 and Table 3). I made three distinct calculations 310 for the Model III; one to include all geologicalDraft estimates (Fig. 8a), a second one excluding the 311 estimates that were calculated by using the displaced submarine features (Fig. 8b), and a third

312 one only including morphochronological and palaeoseismological sites (Fig. 8c). While

313 incorporating the slip data from the Çınarcık Basin, I used incremental offsets between each

314 seismic horizon (Fig. 9 and Table 3) and the time difference between each MISs, which are

315 assigned to determined seismic horizons (Table 3).

316 RESULTS: ANALYSES OF GEOLOGICAL SLIP RATES

317 Model I – Yedisu – Niksar section

318 There are three geological slip rate estimates, only two of which are morphochronologically

319 measured for the NAF’s section between Yedisu in the east and Niksar in the west. This part

320 of the NASZ is suggested to have the narrowest width where most of the strain is localised

321 along its main displacement zone, NAF.

322 The Model I, accounting all geological slip rate estimates (Hubert-Ferrari et al. 2002; Zabcı

323 2012), yield a uniform slip rate solution of 16.8 +4/-3.5 mm/a (2-sigma) for the last 12 ka

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 14 of 59

324 (Fig. 6A.1 and Table 3). This uniform rate is calculated as 17.5 +4/-3.5 mm/a for the last 4 ka

325 (Fig. 6B.1 and Table 3), when I include only two available morphochronology-based

326 estimates (Zabcı 2012) for this portion of the fault. The slip history results for both solutions

327 display a similar pattern that does not show any significant variation in slip history (Figs.

328 6A.2 and 6B.2).

329 Figure. 6

330 Model II – Niksar -Bolu section

331 To the west of Niksar, the NASZ starts to have a wider zone relative to its eastern part.

332 Although, the main strain is accumulated along the NAF itself, smaller portion of deformation 333 is shared between the splays from the principalDraft displacement zone of the NASZ and the 334 discrete structures of the Amasya Shear Zone. My model includes a total of nine geological

335 slip rate estimates for this section of the NAF (Table 2). Three of these rates are estimated by

336 using the palaeoclimatological (Last Glacial Maximum ~12 ka) correlation to the systematic

337 alluvial fan formations at Destek, Berçin and Gerede sites (Hubert-Ferrari et al. 2002),

338 whereas others are based on morphochronology and palaeoseismology studies and covers a

339 time span of about last 5 ka (Hubert-Ferrari et al. 2002; Kondo et al. 2010; Kozacı et al. 2009;

340 Kozacı et al. 2007; Zabcı 2012).

341 The uniform slip solutions yield 18.9 +3.7/-3.3 mm/a (2-sigma) and 17.4 +2.7/-2.4 mm/a (2-

342 sigma) for the last 5 ka and the 12 ka, respectively (Figs. 7A.1, 7B.1 and Table 3). In terms of

343 slip history solutions, there is almost a constant slope between about 2.5 ka and 12 ka ago

344 (Fig. 7A.2). However, there is a slight change of the slope between 1 ka and 2 ka BP, mostly

345 because of the limited time and slip coverage of the Demirtepe trench site (Fig. 7B.2).

346 Figure 7.

https://mc06.manuscriptcentral.com/cjes-pubs Page 15 of 59 Canadian Journal of Earth Sciences

347 Model III- Akyazı – Saros section

348 The Model III covers a region where the NASZ reaches to the widest territory along its

349 course. This is not only characterised by the number of discrete structures, but also the NAF,

350 itself, which bifurcates into two main strands to the west of Bolu. The southern strand is made

351 of multiple parallel/sub-parallel fault segments, which almost all lay to the south of the

352 Marmara Sea, whereas the northern strand is characterised as a more localised structure

353 (except its local complexities). All geological slip rate estimates are compiled for the more

354 active northern strand of the NAF and integrated into the Model III.

355 Available geological slip rate estimates are grouped into three according to the source of the

356 age data; (a) morphochronology and palaeoseismology-based results for the last 3 ka (Aksoy

357 2009; Aksoy et al. 2010b; Dikbaş et al.Draft 2018; Meghraoui et al. 2012; Rockwell et al. 2009),

358 (b) estimates calculated by the correlation of palaeoclimatological history and stream

359 incisions for the last 20 ka (Aksoy 2009; Aksoy et al. 2010b), (c) and rates obtained by the

360 correlation of displaced submarine formations and stratigraphic age model linked to Marine

361 Isotopic Stages for the last 550 ka (Grall et al. 2013; Kurt et al. 2013).

362 The uniform slip rate solution suggests a rate of 16.9 +1.2/-1.1 mm/a (2-sigma) for the last 3

363 ka (Fig. 8C.1 and Table 3), while the slip history solution has a general constant trend except

364 a fluctuation at about 1 ka BP. The second dataset of offset streams yield slightly a faster rate

365 of 17.7 +1.9/-1.7 mm/a (2-sigma) for the last 20 ka (Fig. 8B.1 and Table 3). In this time span,

366 the slip history solution also points a local change in the slope at about 12 to 14 ka BP (within

367 the 3rd order offset streams along the Ganos Fault in Fig. 8B.2). The very long-term uniform

368 slip solution gives a rate of 16.7 +1/-0.9 mm/a (2-sigma) similar to the rate for the last 3 ka

369 (Fig. 8A.1). However, the slip history solution displays a differentiation in the slope,

370 especially for two incremental intervals 195 – 320 ka and 320 – 540 ka (Fig. 8A.2). This

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 16 of 59

371 peculiarity of the slip rate behaviour is also clear in the temporal distribution of the displaced

372 landforms, where the d-t envelope for the Çınarcık Basin (H3/H2) offset significantly stays

373 out of the uniform solution line. Thus, I constructed individual uniform solutions for

374 incremental steps between Çınarcık Basin H2/H1 – H3/2 and Çınarcık Basin H3/H2 – H5/H4,

375 which yield a strong variation in the deformation rate (13.2 +3.1/-2.9 mm/a and 22.9 +3.1/-2.6

376 mm/a (2-sigma), respectively; Fig. 8A.2 and Table 3).

377 Figure 8.

378 Figure 9.

379 DISCUSSIONS: SPATIAL AND TEMPORAL BEHAVIOUR OF THE NORTH 380 ANATOLIAN FAULT 381 Spatial Behaviour of the North AnatolianDraft Fault: Implications from geological slip rate 382 estimates

383 The geodetic slip rate estimates have a general decreasing trend from west to east, especially

384 in case of velocity profiles (Figs. 3 and 10b). However, unclassified geological estimates

385 display great varieties almost all along the NAF without any regular pattern. Geological sites

386 are also scattered and outnumbered in the eastern parts, whereas they are relatively more

387 abundant at the central and western sections (Figs. 4 and 10b). Thus, I used results of model

388 rates, which are calculated by using only the carefully selected morphochronology- and

389 palaeoseismology-based estimates (Model IB, Model IIB, and Model IIIC; Fig. 10c).

390 The probability distributions of Model IB (17.5 +4/-3.5 mm/a, 2-sigma) and Model IIB (18.9

391 +4.3/-4, 2-sigma) is relatively larger than the Model IIIC (16.9 +1.2/-1.1 mm/a), because they

392 either include limited number of observations or contain high uncertainties in measurements

393 of displacements and/or ages (Table 1 and 2). Nevertheless, there is no significant spatial slip

394 rate variation between Yedisu in the east and Bolu in the west for about 800 km along the

395 NAF (Fig. 10b). The Model IIIC only corresponds to the lower limits of the Model IB and

https://mc06.manuscriptcentral.com/cjes-pubs Page 17 of 59 Canadian Journal of Earth Sciences

396 Model IIB, which can be interpreted as a very slight decrease in the deformation rate along

397 the northern strand of the NAF in the Marmara Region. This decrease is understandable,

398 considering the increasing width of the NASZ, especially in its western parts, where some of

399 the total strain are distributed between the northern strand and discrete faults of the southern

400 strand.

401 Figure 10.

402 Temporal Behaviour of the North Anatolian Fault

403 I analyse the temporal behaviour of the NAF for three different time periods; (a) short-term

404 (the last 10 ka) and (b) long- (10 – 100 ka) and (c) very long-term (> 100 ka);

405 Short-term (<10 ka) slip behaviour Draft 406 The short-term solutions of almost all models do not show any secular variation in their slip

407 history for almost the entire fault except a slight fluctuation at about 1 ka in Model IIIC (Fig.

408 8C.1). This apparent change in the slip history slope is most possibly due to relatively high

409 slip rate estimate of Dikbaş et al. (2018), which may be an overestimate considering the older

410 ages of Erturaç et al. (2019) for the same offset terrace formation. Using Erturaç et al.

411 (2019)’s two OSL ages of 1060±100 a and 1110±200 a to date the 18.5±0.5 m slip

412 measurement of Dikbaş et al. (2018) yields a revised slip rate of 16.7 +3.6/-2.5 mm/a, which

413 is perfectly compatible with other palaeoseismological estimates (e.g., Meghraoui et al. 2012)

414 in the Marmara Region.

415 The correlation of short-term model estimates and geodetic rates display a difference in

416 favour of geodetic values especially for the central and western sections. The block model

417 estimates significantly exceed rates of the Model IB, Model IIB and Model IIIC. Most of the

418 geodetic velocity profiles also show the same picture for the western and the central sections

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 18 of 59

419 of the fault. However, the rate of Model IB slightly correlate with geodetic estimates of

420 velocity profiles within their uncertainty limits at the eastern portion of the NAF. I consider

421 three possible cases, (a) the effect of external processes, (b) the interaction of the elastic upper

422 crust and the viscoelastic lower crust/upper mantle, and (c) structural processes, but exclude

423 the possibility of temporally changing plate rates due to the complex geodynamics of Anatolia

424 and limited information on controlling factors such as the short-term behaviour of the

425 Hellenic Trench while discussing the discrepancy between geological and geodetic rates

426 along the NAF.

427 Climatically-driven changes in crustal water content cannot be a valid explanation, because

428 there is no geological record documenting any remarkable sea level changes that can be used 429 as a proxy for the climatic history for theDraft last 5 ka (Lericolais et al. 2009; Lericolais et al. 430 2011). Furthermore, there is no indication for removal or volumetric change of large

431 lacustrine or similar bodies for the same time period. Although there are studies documenting

432 old lacustrine sediments, these are very old formations such as the 800 ka old Suluova Lake

433 deposits (Erturaç et al. this issue) or local features, such as the one at the Gölova pull-apart

434 that was occupying relatively a small volume in very near past (Hubert-Ferrari et al. 2012).

435 Briefly, there is no significant evidence of an external force to explain the discrepancy

436 between the geodetic and short-term morphochronological/palaeoseismological slip rates

437 along the NASZ.

438 The strain transient as result of the interaction between different mechanical layers of the

439 crust is suggested to be one of the possible explanations for the discrepancy between geodetic

440 and geological velocities along the NAF. This mechanical behaviour is proposed to may

441 happen with weakening or hardening-annealing of the lower crust/upper mantle, following

442 large events or cluster of earthquakes (Dolan and Meade 2017; Kozacı et al. 2009; Zabcı

443 2012). The weakening of the lower crust and the upper mantle can happen by the thermo-fluid

https://mc06.manuscriptcentral.com/cjes-pubs Page 19 of 59 Canadian Journal of Earth Sciences

444 induced structural and metamorphic reactions such as dynamic recrystallization and grain-size

445 reduction (e.g., Bürgmann and Dresen 2009; Jamtveit et al. 2018; Oskin et al. 2008), which

446 promote faster elastic strain accumulation in the upper crust. The Ms 8.0 1668 Great

447 Anatolian Earthquake that ruptured a multiple segments between Bolu in the west and west of

448 Erzincan in the east (Ambraseys and Finkel, 1988; Kondo et al. 2010; Fraser et al. 2012;

449 Hubert-Ferrari et al. 2012; Zabcı et al. 2011) and/or the significant earthquake sequence of the

450 20th century might have generated weakening in the lower crust/upper mantle and triggered

451 the strain transient for the NAF. The geodetic evidences for the post-seismic viscoelastic

452 relaxation (Ergintav et al. 2009; Hearn et al. 2009) or creeping (Çakır et al. 2005; Çakır et al.

453 2012; Çetin et al. 2014; Karabacak et al. 2011) are reported for the central-to-western parts of

454 the fault. Nevertheless, postseismic measurements along the İsmetpaşa Creep, which happens Draft 455 along of about 100 km section of the Ms 7.3 1944 Bolu-Gerede and Ms 7.4 1943 Tosya-Ladik

456 earthquake segments, show an exponential decay in creep rate that appears to have a steady

457 state for about last 30 years (Bilham et al. 2016; Çetin et al. 2014) and suggest that this

458 aseismic behaviour is mainly controlled by the fault rock rheology (Kaduri et al. 2017). An

459 analogue spring-dashboard-slider experiment also shows that high peak strain rates that

460 remain well above the geological estimates only about 15% of the earthquake cycle (Kenner

461 and Simons 2005), which means that it is normal for İzmit and the surrounding region to

462 experience a post-seismic effect, where the last major earthquake occurred within the last 20

463 years ago. But, it is unlikely to have it along the 1943 and 1944 earthquake segments, which

464 exceed the 15% of the earthquake cycle when we consider ~300 years of mean repeat time

465 (Kondo et al. 2010) for central to western parts of the fault.

466 An alternative explanation is to consider that the NAF accumulates the majority of the total

467 strain, but some portion is distributed on secondary structures of the shear zone (the NASZ).

468 Not only the existence of these active secondary structures (e.g., the Amasya Shear Zone in

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 20 of 59

469 the central and the multi-strand structure in the Marmara Region), but also the scattered

470 seismicity (Bulut, 2015; Karasözen et al. 2014; Örgülü, 2011; Şengör and Zabcı, 2019) and

471 results of a 3D mechanical model, yielding slip partitioning and internal deformation (Hergert

472 and Heidbach 2010) strongly supports this hypothesis. The off-the-main displacement zone

473 earthquakes, such as the Mw 6.0 2000 Orta (Çakır and Akoğlu 2008) and the Ms 6.9 1951

474 Kurşunlu (Ketin 1969) seismic events are good examples for the distribution of some of the

475 total strain along the secondary features (Fig. 2). Moreover, geodetic studies conducted across

476 the central NASZ suggest a slip partitioning between the NAF and the Erçenek Fault, from

477 which the NAF gets big portion of about 90% and the Erçenek Fault has 10% of the total

478 strain (Peyret et al. 2013). In Almacık Blok where it well-known that the total slip is

479 distributed between fault segments to the north and to the south (Fig. 5b), the sum of slip rate

480 estimates of ~15 mm/a for the Düzce FaultDraft (Pucci et al. 2008) and ~10 mm/a for the Mudurnu

481 Valley Fault (Erturaç et al. 2019) provide a similar right-lateral velocity with the GPS-based

482 block model estimate of Reilinger et al. (2006). There is also no significant discrepancy

483 between the geological slip rates and results of geodetic velocity profiles at the eastern part of

484 the NAF, where most of the strain have confined to the narrowest zone along its entire course

485 (Fig. 10c). In general, the discrepancy between geodetic and geological rates increases

486 parallel with the width of the NASZ from east to west. However, we still need more multi-

487 disciplinary studies such as meticulous mapping of structures (especially deformed

488 Quaternary deposits), seismic distribution of earthquakes along these secondary structures and

489 identification of more slip rate sites particularly on major splays of the NASZ in order to

490 increase our understanding on the distribution of the active deformation.

491 Long-term (10 – 100 ka) slip behaviour

492 There are only two geological slip rate studies among only one is based on the geochronology

493 of offset landforms, along the entire NASZ for the time span between 10 and 100 ka (Aksoy,

https://mc06.manuscriptcentral.com/cjes-pubs Page 21 of 59 Canadian Journal of Earth Sciences

494 2009; Aksoy et al. 2010; Pucci et al. 2008). The single long-term morphochronology-based

495 slip rate is from the Düzce Fault and covers the last 60 ka (Pucci et al. 2008). I did not include

496 this estimate in my models, because it only represents one segment of closely spaced multi-

497 branch faulting to the north and south of the Almacık Block (Fig. 5b). Any previous analysis

498 of the NAF’s slip history based on the slip rate estimates of the Düzce Fault (e.g. Dolan and

499 Meade, 2017) are inadequate, because the neighbouring parallel segment, the Mudurnu Valley

500 Fault, has been proven to have frequent surface rupturing earthquakes (Ikeda et al 1991;

501 Palyvos et al. 2007) and a slip rate of about 10 mm/yr that is higher than the previous

502 assumptions (Erturaç et al. 2019).

503 The second study is based on the dating of offset streams by correlating the major 504 palaeoclimatological changes and initiationDraft of river incisions (Aksoy, 2009; Aksoy et al. 505 2010). The general slip history solution of these offset markers does not show a remarkable

506 variation, except a slight decrease between 2nd and 3rd order streams at about 12 ka. This

507 apparent slight change in the slip rate might have occurred because of the major climatic

508 change at beginning of the Holocene or is just an artefact due to resolution of data. Therefore,

509 morphochronology-based slip rate studies, covering the same time span, are mandatory.

510 Very long-term (>100 ka) slip behaviour

511 In terms of very long-term slip history solutions, there are data only for the Marmara Region

512 all along the NASZ. Although, finite measurement of offset features yields similar slip rates

513 with the short-term geological estimates (Grall et al. 2013; Kurt et al. 2013), the incremental

514 slip history solutions suggest an anomaly for the NAF segments in the Marmara Sea (Fig. 8A

515 and Table 3). The Çınarcık Basin H2/H1 – H3/2 and Çınarcık Basin H3/H2 – H5/H4 offset

516 markers yield a strong variation in the deformation rate history (13.2 +3.1/-2.9 mm/a and 22.9

517 +3.1/-2.6 mm/a), respectively. Considering the multi-strand structure of the NAF in the

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 22 of 59

518 Marmara Region, the most possible case is the transfer of total strain between different

519 strands of the NAF. In more details, southern strand was accelerated between 195 and 320 ka,

520 while the deformation rate of the northern strand decreased to 13.2 +3.1/-2.9 mm/a.

521 Furthermore, between 320 and 550 ka, the southern strand should have significantly slower

522 deformation rate with respect to its recent state, while the northern strand was deforming with

523 a much faster rate, 22.9 +3.1/-2.6 mm/a. A similar co-dependent slip history is also suggested

524 for the San Andreas and San Jacinto faults (Bennett et al. 2004; Blisniuk et al. 2010), the San

525 Andreas Fault and Eastern California Shear Zone (Dolan et al. 2007), within the Eastern

526 California Shear Zone (Reheis and Dixon 1996), between the Eastern California Shear Zone

527 and the Walker Lane faults (Frankel et al. 2007) or even within the individual fault segments

528 of the Walker Lane System (Gold et al. 2013; Walker et al. 2005), and Wellington-Wairarapa

529 and Awatare faults (Ninis et al. 2013; RobinsonDraft et al. 2011) that have occurred in different

530 time scakes. Short term variations are mostly related to the mechanical coupling of the crust,

531 in which the hardening/weakening and the annealing processes trigger activity on

532 neighbouring fault segments (Dolan et al. 2007; Ninis et al. 2013). However, the low

533 resolution data in the Marmara Sea points a variation of strain accumulation in a scale of >105

534 a, hence, it cannot be linked with a great earthquake or cluster of events. The co-dependent

535 behaviour of neighbouring fault segments are usually observed at evolving faults systems as it

536 is observed during the initiation of the San Jacinto Fault (Bennett et al. 2004). Thus, this can

537 be a plausible mechanism for the strain transfer between the different structural elements of

538 the NASZ, considering that the northern strand is claimed to be the youngest member and

539 initiated not more than a million years ago (Le Pichon et al. 2001; Şengör et al. 2005; Şengör

540 and Zabcı 2019). However, sedimentation of about 4 km-thickness in the Çınarcık Basin,

541 which is supposed to be deposited during the last 550 ka, possibility of miscorrelating any of

542 the MIS to the identified seismic horizons and non-existent slip rate sites along the southern

https://mc06.manuscriptcentral.com/cjes-pubs Page 23 of 59 Canadian Journal of Earth Sciences

543 strand of the NAF for the same time period are the main shortcomings in this hypothesis.

544 Therefore, we strongly need additional studies both along the northern and southern strands in

545 order to have a better picture of this apparent temporal behaviour.

546 CONCLUSIONS

547 I analysed almost all geological slip rate estimates in order to have a better understanding of

548 the spatial and temporal behaviour of the NAF. The modelled uniform slip solutions yield

549 17.5 +4/-3.5 mm/a, 18.9 +3.7/-3.3 mm/a and 16.9 +1.2/-1.1 mm/a as average values for the

550 last ~5 ka for the eastern, central and western sections of the fault, respectively. Although

551 there is no significant variation of these model rates along the NAF, there are discrepancies

552 between model and geodetic rates especially for the central and western sections. I suggest 553 that this apparent behaviour happens becauseDraft of the geometry of the NASZ, which constantly 554 widens from east to west. In terms of slip history analyses, the deformation rate is constant at

555 least for the last 195 ka. Nevertheless, the limited number of data show strain transfer from

556 northern to the southern strand between 195 and 320 ka BP in the Marmara Region.

557 Considering the possible defects in determination of incremental displacements and their

558 timings, it is clear that we need more slip rate sites both on the northern and southern strands.

559 The same necessity is also crucial for the eastern segments of the NAF, where there is very

560 limited number of geological slip rate estimates.

561 The distributed strain pattern along the NASZ is similar to other continental transform faults,

562 such as the San Andreas Fault System in California. Although most of the strain is localized

563 along the main displacement zone, the NAF, secondary structures are still capable of

564 generating earthquakes hardly ever reaching up to magnitude 7. This behaviour must be taken

565 into account for seismic risk assessments.

566 ACKNOWLEDGEMENTS

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 24 of 59

567 H. Serdar Akyüz inspired me to extent my studies on the North Anatolian Fault. Many ideas

568 in this manuscript has been discussed with Semih Ergintav and Ziyadin Çakır for whom I am

569 thankful for their fruitful comments and suggestions. M. Korhan Erturaç was very kind to

570 share his fault database for the Amasya Shear Zone and information on the Suluova Basin.

571 Ryan D. Gold provided the Matlab codes of his Monte-Carlo approach for the inversion of

572 geological slip rates. I thank Pierre Henry, Christopher Sorlien and Hülya Kurt for discussions

573 on the timing of seismic horizons in the Marmara Sea. I am also indebted to Ali Polat, John

574 Dewey, Mustapha Meghraoui and an anonymous reviewer for their invaluable comments that

575 significantly improved the quality of the manuscript. Most of the figures are made by using

576 the General Mapping Tools (Wessel et al. 2013).

577 Draft

https://mc06.manuscriptcentral.com/cjes-pubs Page 25 of 59 Canadian Journal of Earth Sciences

578 REFERENCES

579 Aksoy, M.E. 2009. Active Tectonics and Paleoseismology of the Ganos Fault Segment and 580 Seismic Characteristics of the 9 August 1912 Mürefte Earthqıake of the North Anatolian 581 Fault (Western Turkey). Avrasya Yer Bilimleri Enstitüsü ve IPGS, İstanbul Teknik 582 Üniversitesi ve L'Universite de Strasbourg, İstanbul ve Strasbourg. 583 Aksoy, M.E., Meghraoui, M., Vallée, M., and Çakır, Z. 2010a. Rupture characteristics of the 584 A.D. 1912 Mürefte (Ganos) earthquake segment of the North Anatolian fault (western 585 Turkey). Geology, 38: 991-994. doi:10.1130/g31447.1. 586 Aksoy, M.E., Meghaoroui, M., Çakır, Z., Ferry, M., and Uçarkuş, G. Short-term and long-term 587 slip rate along the westernmost segment of the North Anatolian Fault using paleoseismic 588 trenching and drainage offsets. In EGU. Vienna 2010b. Geophysical Research 589 Abstracts. Vol. 12, pp. EGU2010-12447. 590 Aktuğ, B., Dikmen, Ü., Doğru, A., and Özener, H. 2013a. Seismicity and strain accumulation 591 around Karliova Triple Junction (Turkey). Journal of Geodynamics, 67: 21-29. 592 doi:http://dx.doi.org/10.1016/j.jog.2012.04.008. 593 Aktuğ, B., Doğru, A., Özener, H., and Peyret, M. 2015. Slip rates and locking depth variation 594 along central and easternmost segments of North Anatolian Fault. Geophysical Journal 595 International, 202: 2133-2149. 596 Aktuğ, B., Parmaksız, E., Kurt, M., Lenk, O., Kılıçoğlu, A., Gürdal, M.A., and Özdemir, S. 597 2013b. Deformation of Central Anatolia: GPS implications. Journal of Geodynamics, 598 67: 78-96. doi:http://dx.doi.org/10.1016/j.jog.2012.05.008. 599 Akyüz, H.S., Hartleb, R., Barka, A., Altunel,Draft E., and Sunal, G. 2002. Surface rupture and slip 600 distribution of the 12 November 1999 Düzce Earthquake (M 7.1), North Anatolian 601 Fault, Bolu, Turkey. Bulletin of the Seismological Society of America, 92: 61-66. 602 Altınok, Y., Alpar, B., Yaltırak, C., Pınar, A., and Özer, N. 2012. The earthquakes and related 603 tsunamis of October 6, 1944 and March 7, 1867; NE Aegean Sea. Natural Hazards, 60: 604 3-25. doi:10.1007/s11069-011-9949-7. 605 Altunel, E., Megharoui, M., Akyüz, H.S., and Dikbaş, A. 2004. Characteristics of the 1912 co- 606 seismic rupture along the North Anatolian Fault Zone (Turkey): implications for the 607 expected Marmara earthquake. Terra Nova, 16: 198-204. 608 Ambraseys, N.N., 1970. Some characteristic features of the Anatolian Fault Zone. 609 Tectonophysics, 9: 143-165. 610 Ambraseys, N.N., and Finkel, C.F. 1988. The Anatolian Earthquake of 17 August 1668. In 611 Historical Seismograms and Earthquakes of the World. Edited by W.H.K. Lee, H. 612 Meyers, and K. Shimazaki. Academic Press, San Diego. pp. 173–180. 613 Ambraseys, N.N., and Zatopek, A. 1969. The Mudurnu valley, west Anatolia, Turkey, 614 earthquake of 22 July 1967. Bulletin of the Seismological Society of America, 59: 521- 615 589. 616 Armijo, R., Meyer, B., Hubert-Ferrari, A., and Barka, A.A. 1999. Westward propagation of 617 North Anatolian Fault into the Northern Agean: timing and kinematcis. Geology, 27: 618 267-270. 619 Barka, A. 1996. Slip distribution along the North Anatolian Fault associated with large 620 earthquakes of the period 1939 to 1967. Bulletin of the Seismological Society of 621 America, 86: 1238-1254. 622 Barka, A., and Eyidoǧan, H. 1993. The Erzincan earthquake of 13 March 1992 in Eastern 623 Turkey. Terra Nova, 5: 190-194. doi:10.1111/j.1365-3121.1993.tb00245.x. 624 Barka, A., Akyüz, H.S., Altunel, E., Sunal, G., Çakır, Z., Dikbaş, A., Yerli, B., Armijo, R., 625 Meyer, B., De Chabalier, J.B., Rockwell, T., Dolan, J.R., Hartleb, R., Dawson, T., 626 Christofferson, S., Tucker, A.Z., Fumal, T.E., Langridge, R., Stenner, H.D., Lettis, W.,

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 26 of 59

627 Bachhuber, J., and Page, W.D. 2002. The surface rupture and slip distribution of the 17 628 August 1999 İzmit earthquake (M 7.4), North Anatolian Fault. Bulletin of the 629 Seismological Society of America, 92: 43-60. 630 Barka, A.A. 1992. The North Anatolian fault zone. Annales Tectonicae, 6: 164-195. 631 Barka, A.A., and Kadinsky-Cade, K. 1988. Strike-Slip Fault Geometry in Turkey and Its 632 Influence on Earthquake Activity. Tectonics, 7: 663-684. 633 Bartholomew, T.D., Little, T.A., Clark, K.J., Van Dissen, R., and Barnes, P.M. 2014. 634 Kinematics and paleoseismology of the Vernon Fault, Marlborough Fault System, New 635 Zealand: Implications for contractional fault bend deformation, earthquake triggering, 636 and the record of Hikurangi subduction earthquakes. Tectonics, 33: 2014TC003543. 637 doi:10.1002/2014TC003543. 638 Bennett, R.A., Friedrich, A.M., and Furlong, K.P. 2004. Codependent histories of the San 639 Andreas and San Jacinto fault zones from inversion of fault displacement rates. 640 Geology, 32: 961-964. doi:10.1130/g20806.1. 641 Bilham, R., and King, G. 1989. The morphology of strike-slip faults: Examples from the San 642 Andreas Fault, California. Journal of Geophysical Research: Solid Earth, 94: 10204- 643 10216. doi:10.1029/JB094iB08p10204. 644 Bilham, R., Ozener, H., Mencin, D., Dogru, A., Ergintav, S., Cakir, Z., Aytun, A., Aktug, B., 645 Yilmaz, O., Johnson, W., and Mattioli, G. 2016. Surface creep on the North Anatolian 646 Fault at Ismetpasa, Turkey, 1944–2016. Journal of Geophysical Research: Solid Earth, 647 121: 7409–7431. doi:10.1002/2016JB013394. 648 Blisniuk, K., Rockwell, T., Owen, L.A., Oskin, M., Lippincott, C., Caffee, M.W., and Dortch, 649 J., 2010, Late Quaternary slip Draftrate gradient defined using high-resolution topography 650 and 10Be dating of offset landforms on the southern San Jacinto Fault zone, California: 651 Journal of Geophysical Research: Solid Earth, v. 115, p. B08401, 652 doi:10.1029/2009JB006346. 653 Blumenthal, M.M. 1945. Ladik deprem hatti (Samsun ili); la ligne sismique de Ladik, Vilayet 654 de Samsun. Maden Tetkik ve Arama Dergisi, 33: 153-174. 655 Blumenthal, M.M., Nafiz Pamir, H., and Hakki Akyol, I. 1943. Simal Anadolu zelzele sahasinin 656 jeolojisi ve 1942 yili sonunda buralarda yapilan makro-sismik muesahedeler 657 (Osmancik- Erbaa); Zur Geologie der Landstrecken der Erdbeben von Ende 1942 in 658 Nord-Anatolien und dortselbst ausgefuehrte makroseismische Beobachtungen. Maden 659 Tetkik ve Arama Dergisi, 29: 33-58. 660 Bulut, F. 2015. Different phases of the earthquake cycle captured by seismicity along the North 661 Anatolian Fault. Geophysical Research Letters, 42: 2219–2227. 662 doi:10.1002/2015GL063721. 663 Bürgmann, R., and Dresen, G. 2008. Rheology of the Lower Crust and Upper Mantle: Evidence 664 from Rock Mechanics, Geodesy, and Field Observations. Annual Review of Earth and 665 Planetary Sciences, 36: 531–567. Annual Reviews. 666 doi:10.1146/annurev.earth.36.031207.124326. 667 Çakır, Z., and Akoğlu, A.M. 2008. Synthetic aperture radar interferometry observations of the 668 M = 6.0 Orta earthquake of 6 June 2000 (NW Turkey): Reactivation of a listric fault. 669 Geochemistry, Geophysics, Geosystems, 9. doi:10.1029/2008GC002031. 670 Çakır, Z., Akoglu, A.M., Belabbes, S., Ergintav, S., and Meghraoui, M. 2005. Creeping along 671 the Ismetpasa section of the North Anatolian fault (Western Turkey): Rate and extent 672 from InSAR. Earth and Planetary Science Letters, 238: 225-234. 673 doi:10.1016/j.epsl.2005.06.044. 674 Çakır, Z., Ergintav, S., Akoğlu, A.M., Çakmak, R., Tatar, O., and Meghraoui, M. 2014. InSAR 675 velocity field across the North Anatolian Fault (eastern Turkey): Implications for the

https://mc06.manuscriptcentral.com/cjes-pubs Page 27 of 59 Canadian Journal of Earth Sciences

676 loading and release of interseismic strain accumulation. Journal of Geophysical 677 Research: Solid Earth, 119: 2014JB011360. doi:10.1002/2014JB011360. 678 Çakır, Z., Ergintav, S., Özener, H., Dogan, U., Akoglu, A.M., Meghraoui, M., and Reilinger, 679 R. 2012. Onset of aseismic creep on major strike-slip faults. Geology, 40: 1115-1118. 680 Available from http://geology.gsapubs.org/content/40/12/1115.abstract [accessed. 681 Cavalié, O., and Jónsson, S. 2014. Block-like plate movements in eastern Anatolia observed by 682 InSAR. Geophysical Research Letters, 41: 26-31. doi:10.1002/2013GL058170. 683 Çetin, E., Çakır, Z., Meghraoui, M., Ergintav, S., and Akoğlu, A.M. 2014. Extent and 684 distribution of aseismic slip on the Ismetpaşa segment of the North Anatolian Fault 685 (Turkey) from Persistent Scatterer InSAR. Geochemistry, Geophysics, Geosystems, 15: 686 2883-2894. doi:10.1002/2014GC005307. 687 Chery, J., and Vernant, P. 2006. Lithospheric elasticity promotes episodic fault activity. Earth 688 and Planetary Science Letters, 243: 211-217. doi:10.1016/j.epsl.2005.12.014. 689 Chorowicz, J., Dhont, D., and Gündogdu, N. 1999. Neotectonics in the eastern North Anatolian 690 fault region (Turkey) advocates crustal extension: mapping from SAR ERS imagery and 691 Digital Elevation Model [doi: DOI: 10.1016/S0191-8141(99)00022-X]. Journal of 692 Structural Geology, 21: 511-532. 693 Chuang, R.Y., and Johnson, K.M. 2011. Reconciling geologic and geodetic model fault slip- 694 rate discrepancies in Southern California: Consideration of nonsteady mantle flow and 695 lower crustal fault creep. Geology, 39: 627-630. doi:10.1130/g32120.1. 696 Daeron, M., Benedetti, L., Tapponnier, P., Sursock, A., and Finkel, R.C. 2004. Constraints on 697 the post ~25-ka slip rate of the Yammo˚neh fault (Lebanon) using in situ cosmogenic 698 36Cl dating of offset limestone-clastDraft fans. Earth and Planetary Science Letters, 227: 699 105-119. doi:10.1016/j.epsl.2004.07.014. 700 Dewey, J.F., and Şengör, A.M.C. 1979. Aegean and surrounding regions: Complex multiplate 701 and continuum tectonics in a convergent zone. Geological Society of America Bulletin, 702 90: 84-92. doi:10.1130/0016-7606(1979). 703 Dikbaş, A., Akyüz, H.S., Meghraoui, M., Ferry, M., Altunel, E., Zabcı, C., Langridge, R., and 704 Yalçıner, C.Ç. 2018. Paleoseismic history and slip rate along the Sapanca-Akyazı 705 segment of the 1999 İzmit earthquake rupture (Mw = 7.4) of the North Anatolian Fault 706 (Turkey). Tectonophysics, 738-739: 92-111. 707 doi:https://doi.org/10.1016/j.tecto.2018.04.019. 708 Dixon, T.H., Norabuena, E., and Hotaling, L. 2003a. Paleoseismology and Global Positioning 709 System: Earthquake-cycle effects and geodetic versus geologic fault slip rates in the 710 Eastern California shear zone. Geology, 31: 55-58. 711 Dolan, J. Possible Transient Strain Accumulation along the North Anatolian Fault (NAF): 712 Precursor or Artifact? In 1509 Marmara Earthquake Symposium Program. Istanbul 713 2009. Istanbul Teknik Universitesi. 714 Dolan, J.F., and Haravitch, B.D. 2014. How well do surface slip measurements track slip at 715 depth in large strike-slip earthquakes? The importance of fault structural maturity in 716 controlling on-fault slip versus off-fault surface deformation. Earth and Planetary 717 Science Letters, 388: 38-47. doi:http://dx.doi.org/10.1016/j.epsl.2013.11.043. 718 Dolan, J.F., and Meade, B.J. 2017. A Comparison of Geodetic and Geologic Rates Prior to 719 Large Strike-Slip Earthquakes: A Diversity of Earthquake-Cycle Behaviors? 720 Geochemistry, Geophysics, Geosystems, 18: 4426-4436. doi:10.1002/2017GC007014. 721 Dolan, J.F., Bowman, D.D., and Sammis, C.G. 2007. Long-range and long-term fault 722 interactions in Southern California. Geology Boulder, 35: 858-855. 723 Duebendorfer, E.M., Vermilye, J., Geiser, P.A., and Davis, T.L. 1998. Evidence for aseismic 724 deformation in the western Transverse Ranges, southern California: Implications for 725 seismic risk assessment. Geology, 26: 271-274.

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 28 of 59

726 Emre, Ö., Awata, Y., and Duman, T.Y. (eds). 2003. 17 Ağustos 1999 İzmit Depremi Yüzey 727 Kırığı/Surface Rupture Associated with the 17 August 1999 İzmit Earthquake. MTA 728 Genel Müdürlüğü Special Publication 1, Ankara. 729 Emre, Ö., Kondo, H., Yıldırım, C., and Özaksoy, V. Fault geometry and slip distribution of the 730 1943 Tosya Earthquake rupture, North Anatolian Fault, Turkey. In AGU Faal Meeting 731 Abstracts. San Francisco 2005, p. B1291. 732 Ergintav, S., McClusky, S., Hearn, E., Reilinger, R., Cakmak, R., Herring, T., Ozener, H., Lenk, 733 O., and Tari, E. 2009. Seven years of postseismic deformation following the 1999, M = 734 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence. J. Geophys. Res., 114: 1- 735 19. Available from: http://www.agu.org/journals/jb/jb0907/2008JB006021/ [accessed. 736 Ergintav, S., Reilinger, R.E., Çakmak, R., Floyd, M., Cakir, Z., Doğan, U., King, R.W., 737 McClusky, S., and Özener, H. 2014. Istanbul's earthquake hot spots: Geodetic 738 constraints on strain accumulation along faults in the Marmara seismic gap. 739 Geophysical Research Letters, 41: 2014GL060985. doi:10.1002/2014GL060985. 740 Erturaç, M.K. 2009. Amasya ve Çevresinin Morfotektonik Evrimi. PhD, Avrasya Yer Bilimleri 741 Enstitüsü, İstanbul Teknik Üniversitesi, İstanbul. 742 Erturaç, M.K., and Tüysüz, O. 2012. Kinematics and basin formation along the Ezinepazarı- 743 Sungurlu Fault Zone, NE Anatolia, Turkey. Turkish Journal of Earth Sciences, 21: 497- 744 520. doi:10.3906/yer-0910-27. 745 Erturaç, M.K., Erdal, O., Sunal, G., Tüysüz, O., and Şen, Ş. this issue. Quaternary evolution of 746 the Suluova Basin: Implications on tectonics and palaeoenvironments of the central 747 North Anatolian Fault. Canadian Journal of Earth Science. 748 Erturaç, M.K., Şahiner, E., Zabcı, C., Okur,Draft H., Polymeris, G.S., Meriç, N., and İkiel, C. 2019. 749 Fluvial response to rising levels of the Black Sea and to climate changes during the 750 Holocene: Luminescence geochronology of the Sakarya terraces. The Holocene,: 751 0959683619831428. SAGE Publications Ltd. doi:10.1177/0959683619831428. 752 Erturaç, M.K., Zabcı, C., Şahiner, E., and Sunal, G. 2019. Holocene slip rate of the Mudurnu 753 Valley Segment: Implications for the strain partitioning of the North Anatolian Shear 754 Zone in the eastern Marmara Region, Turkey. In Geophysical Research Abstracts. 755 European Geosciences Union, Vienna. pp. EGU2019-5390. 756 Faccenna, C., Becker, T.W., Jolivet, L., and Keskin, M. 2013. Mantle convection in the Middle 757 East: Reconciling Afar upwelling, Arabia indentation and Aegean trench rollback. Earth 758 and Planetary Science Letters, 375: 254-269. 759 doi:http://dx.doi.org/10.1016/j.epsl.2013.05.043. 760 Frankel, K.L., Dolan, J.F., Finkel, R.C., Owen, L.A., and Hoeft, J.S. 2007. Spatial variations in 761 slip rate along the Death Valley-Fish Lake Valley fault system determined from LiDAR 762 topographic data and cosmogenic 10Be geochronology. Geophysical Research Letters, 763 34: L18303. doi:10.1029/2007GL030549. 764 Fraser, J.G., Hubert Ferrari, A., Verbeeck, K., Garcia-Moreno, D., Avşar, U., Maricq, N., 765 Coudijzer, A., Vlamynck, N., and Venneste, K. 2012. A 3000-year record of surface- 766 rupturing earthquakes at Günalan: variable fault-rupture lengths along the 1939 767 Erzincan earthquake-rupture segment of the North Anatolian Fault, Turkey. Annals of 768 Geophysics, 55: 895–927. doi:10.4401/ag-4884. 769 Gasperini, L., Polonia, A., Çağatay, M.N., Bortoluzzi, G., and Ferrante, V. 2011. Geological 770 slip rates along the North Anatolian Fault in the Marmara region. Tectonics, 30: 771 TC6001. doi:10.1029/2011tc002906. 772 Gaudemer, Y., Tapponnier, P., Meyer, B., Peltzer, G., Shunmin, G., Zhitai, C., Huagung, D., 773 and Cifuentes, I. 1995. Partitioning of crustal slip between linked, active faults in the 774 eastern Qilian Shan, and evidence for a major seismic gap, the ‘Tianzhu gap’, on the

https://mc06.manuscriptcentral.com/cjes-pubs Page 29 of 59 Canadian Journal of Earth Sciences

775 western Haiyuan Fault, Gansu (China). Geophysical Journal International, 120: 599- 776 645. 777 Gold, R.D., and Cowgill, E. 2011. Deriving fault-slip histories to test for secular variation in 778 slip, with examples from the Kunlun and Awatere faults. Earth and Planetary Science 779 Letters, 301: 52-64. doi:10.1016/j.epsl.2010.10.011. 780 Gold, R., dePolo, C., Briggs, R., Crone, A., and Gosse, J., 2013, Late Quaternary Slip‐Rate 781 Variations along the Warm Springs Valley Fault System, Northern Walker Lane, 782 California–Nevada Border: Bulletin of the Seismological Society of America, v. 103, 783 p. 542–558. 784 Gold, R.D., Reitman, N.G., Briggs, R.W., Barnhart, W.D., Hayes, G.P., and Wilson, E. 2015. 785 On- and off-fault deformation associated with the September 2013 Mw 7.7 Balochistan 786 earthquake: Implications for geologic slip rate measurements. Tectonophysics, 660: 65- 787 78. doi:http://dx.doi.org/10.1016/j.tecto.2015.08.019. 788 Görür, N., and Elbek, Ş. 2013. Tectonic events responsible for shaping the and 789 its surrounding region. Geodinamica Acta, 26: 1-11. 790 doi:10.1080/09853111.2013.859346. 791 Grall, C., Henry, P., Thomas, Y., Westbrook, G.K., Çağatay, M.N., Marsset, B., Saritas, H., 792 Çifçi, G., and Géli, L. 2013. Slip rate estimation along the western segment of the Main 793 Marmara Fault over the last 405–490 ka by correlating mass transport deposits. 794 Tectonics, 32: 2012TC003255. doi:10.1002/2012TC003255. 795 Grollimund, B., and Zoback, M.D. 2001. Did deglaciation trigger intraplate seismicity in the 796 New Madrid seismic zone? Geology, 29: 175-178. 797 Gürsoy, H., Tatar, O., Akpınar, Z., Polat,Draft A., Mesci, L., and Tunçer, D. 2013. New observations 798 on the 1939 Erzincan Earthquake surface rupture on the Kelkit Valley segment of the 799 North Anatolian Fault Zone, Turkey. Journal of Geodynamics, 65: 259-271. 800 doi:http://dx.doi.org/10.1016/j.jog.2012.06.002. 801 Hampel, A., and Hetzel, R. 2006. Response of normal faults to glacial-interglacial fluctuations 802 of ice and water masses on Earth's surface. J. Geophys. Res., 111: B06406. 803 doi:10.1029/2005jb004124. 804 Hampel, A., Hetzel, R., and Densmore, A.L. 2007. Postglacial slip-rate increase on the Teton 805 normal fault, northern Basin and Range Province, caused by melting of the Yellowstone 806 ice cap and deglaciation of the Teton Range? Geology Boulder, 35: 1107-1110. 807 Hampel, A., Hetzel, R., Maniatis, G., and Karow, T. 2009. Three-dimensional numerical 808 modeling of slip rate variations on normal and thrust fault arrays during ice cap growth 809 and melting. J. Geophys. Res., 114: B08406. doi:10.1029/2008jb006113. 810 Hartleb, R., Dolan, J.F., Akyüz, H.S., Dawson, T.E., Tucker, A.Z., Yerli, B., Rockwell, T.K., 811 Toroman, E., Çakır, Z., Dikbaş, A., and Altunel, E. 2002. Surface rupture and slip 812 distributions along the Karadere segment of the 17 August 1999 İzmit and the western 813 section of the 12 November Düzce, Turkey, earthquakes. Bulletin of the Seismological 814 Society of America, 92: 67-78. 815 Hearn, E.H., McClusky, S., Ergintav, S., and Reilinger, R.E. 2009. Izmit earthquake 816 postseismic deformation and dynamics of the North Anatolian Fault Zone. J. Geophys. 817 Res., 114: 1-21. Available from: 818 http://www.agu.org/journals/jb/jb0908/2008JB006026/ [accessed. 819 Herbert, J.W., Cooke, M.L., Oskin, M., and Difo, O. 2014. How much can off-fault deformation 820 contribute to the slip rate discrepancy within the eastern California shear zone? 821 Geology, 42: 71-74. 822 Hergert, T., and Heidbach, O. 2010. Slip-rate variability and distributed deformation in the 823 Marmara Sea fault system [10.1038/ngeo739]. Nature Geoscience, 3: 132-135. 824 Available from http://dx.doi.org/10.1038/ngeo739 [accessed.

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 30 of 59

825 Hubert-Ferrari, A., Armijo, R., King, G., Meyer, B., and Barka, A. 2002. Morphology, 826 displacement, and slip rates along the North Anatolian Fault, Turkey. Journal of 827 Geophysical Research: Solid Earth, 107: 2235. doi:10.1029/2001JB000393. 828 Hubert-Ferrari, A., King, G., Manighetti, I., Armijo, R., Meyer, B., and Tapponnier, P. 2003. 829 Long-term elasticity in the continental lithosphere; modelling the Aden Ridge 830 propagation and the Anatolian extrusion process. Geophysical Journal International, 831 153: 111-132. doi:10.1046/j.1365-246X.2003.01872.x. 832 Hubert Ferrari, A., Avşar, U., El Ouahabi, M., Lepoint, G., Martinez, P., and Fagel, N. 2012. 833 Paleoseismic record obtained by coring a sag-pond along the North Anatolian Fault 834 (Turkey). Annals of Geophysics, 55: 929–953. 835 Hussain, E., Hooper, A., Wright, T.J., Walters, R.J., and Bekaert, D.P.S. 2016. Interseismic 836 strain accumulation across the central North Anatolian Fault from iteratively unwrapped 837 InSAR measurements. Journal of Geophysical Research: Solid Earth, 121: 9000-9019. 838 doi:10.1002/2016JB013108. 839 Hussain, E., Wright, T.J., Walters, R.J., Bekaert, D.P.S., Lloyd, R., and Hooper, A. 2018. 840 Constant strain accumulation rate between major earthquakes on the North Anatolian 841 Fault. Nature Communications, 9: 1392. doi:10.1038/s41467-018-03739-2. 842 Ikeda, Y., Suzuki, Y., Herece, E., Şaroğlu, F., Işıkara, A.M., and Honkura, Y. 1991. Geological 843 evidence for the last two faulting events on the North Anatolian fault zone in the 844 Mudurnu Valley, western Turkey. Tectonophysics, 193: 335–345. 845 Jamtveit, B., Ben-Zion, Y., Renard, F., and Austrheim, H. 2018. Earthquake-induced 846 transformation of the lower crust. Nature, 556: 487–491. doi:10.1038/s41586-018- 847 0045-y. Draft 848 Kaduri, M., Gratier, J.-P., Renard, F., Çakir, Z., and Lasserre, C. 2017. The implications of fault 849 zone transformation on aseismic creep: Example of the North Anatolian Fault, Turkey. 850 Journal of Geophysical Research: Solid Earth, 122: 4208–4236. 851 doi:10.1002/2016JB013803. 852 Karabacak, V., Altunel, E., and Çakır, Z. 2011. Monitoring aseismic surface creep along the 853 North Anatolian Fault (Turkey) using ground-based LIDAR. Earth and Planetary 854 Science Letters, 304: 64–70. doi:10.1016/j.epsl.2011.01.017. 855 Karaoğlu, Ö., Selçuk, A.S., and Gudmundsson, A. 2017. Tectonic controls on the Karlıova 856 triple junction (Turkey): Implications for tectonic inversion and the initiation of 857 volcanism. Tectonophysics, 694: 368-384. 858 doi:http://dx.doi.org/10.1016/j.tecto.2016.11.018. 859 Karasözen, E., Özacar, A.A., Biryol, C.B., and Beck, S.L. 2013. Seismicity, focal mechanisms 860 and active stress field around the central segment of the North Anatolian Fault in 861 Turkey. Geophysical Journal International, 196: 405–421. doi:10.1093/gji/ggt367. 862 Kenner, S.J., and Simons, M. 2005. Temporal clustering of major earthquakes along individual 863 faults due to post-seismic reloading [10.1111/j.1365-246X.2005.02460.x]. Geophysical 864 Journal International, 160: 179-194. 865 Ketin, I., and Rösli, F. 1953. Makroseismische über das nordwest-anatolische Beben vom 18 866 März 1953. Eclogae Geologicae Helvetiae, 46: 187-208. 867 Ketin, İ. 1948. Über die tektonisch-mechanischen Folgerungen aus den großen anatolischen 868 Erdbeben des letzten Dezenniums. Geologische Rundschau, 36: 77-83. 869 doi:10.1007/BF01791916. 870 Ketin, İ. 1969. Kuzey Anadolu Fayı Hakkında. Maden Tetkik ve Arama Dergisi, 72: 1-27. 871 Kirby, E., Harkins, N., Wang, E., Shi, X., Fan, C., and Burbank, D. 2007. Slip rate gradients 872 along the eastern Kunlun fault. Tectonics, 26: 1-16. 873 Koçyiğit, A., Rojay, B., Cihan, M., and Özacar, A. 2001. The June 6, 2000, Orta (Çankırı, 874 Turkey) Earthquake: Sourced from a New Antithetic Sinistral Strike-slip Structure of

https://mc06.manuscriptcentral.com/cjes-pubs Page 31 of 59 Canadian Journal of Earth Sciences

875 the North Anatolian Fault System, the Dodurga Fault Zone. Turkish Journal of Earth 876 Sciences, 10: 69-82. 877 Konca, A.O., Leprince, S., Avouac, J.-P., and Helmberger, D.V. 2010. Rupture Process of the 878 1999 Mw 7.1 Duzce Earthquake from Joint Analysis of SPOT, GPS, InSAR, Strong- 879 Motion, and Teleseismic Data: A Supershear Rupture with Variable Rupture Velocity. 880 Bulletin of the Seismological Society of America, 100: 267-288. 881 doi:10.1785/0120090072. 882 Konca, A.O., Cetin, S., Karabulut, H., Reilinger, R., Dogan, U., Ergintav, S., Cakir, Z., and 883 Tari, E. 2018. The 2014, Mw 6.9 North Aegean earthquake: seismic and geodetic 884 evidence for coseismic slip on persistent asperities. Geophysical Journal International, 885 213: 1113-1120. doi:10.1093/gji/ggy049. 886 Kondo, H., Özaksoy, V., and Yıldirim, C. 2010. Slip history of the 1944 Bolu-Gerede 887 earthquake rupture along the North Anatolian fault system: Implications for recurrence 888 behavior of multisegment earthquakes. Journal of Geophysical Research: Solid Earth, 889 115: B04316. doi:10.1029/2009JB006413. 890 Kondo, H., Awata, Y., Emre, O., Dogan, A., Ozalp, S., Tokay, F., Yildirim, C., Yoshioka, T., 891 and Okumura, K. 2005. Slip distribution, fault geometry, and fault segmentation of the 892 1944 Bolu-Gerede earthquake rupture, North Anatolian Fault, Turkey. Bulletin of the 893 Seismological Society of America, 95: 1234-1249. 894 Kozacı, Ö., Dolan, J.F., and Finkel, R.C. 2009. A late Holocene slip rate for the central North 895 Anatolian fault, at Tahtaköprü, Turkey, from cosmogenic 10Be geochronology: 896 Implications for fault loading and strain release rates. Journal of Geophysical Research: 897 Solid Earth, 114: B01405. doi:10.1029/2008JB005760.Draft 898 Kozacı, Ö., Dolan, J., Finkel, R., and Hartleb, R. 2007. Late Holocene slip rate for the North 899 Anatolian fault, Turkey, from cosmogenic 36Cl geochronology: Implications for the 900 constancy of fault loading and strain release rates. Geology, 35: 867-870. 901 Kurt, H., Sorlien, C.C., Seeber, L., Steckler, M.S., Shillington, D.J., Cifci, G., Cormier, M.H., 902 Dessa, J.X., Atgin, O., Dondurur, D., Demirbag, E., Okay, S., Imren, C., Gurcay, S., 903 and Carton, H. 2013. Steady late quaternary slip rate on the Cinarcik section of the North 904 Anatolian fault near Istanbul, Turkey. Geophysical Research Letters, 40: 4555-4559. 905 doi:10.1002/grl.50882. 906 Lacombe, O., Angelier, J., Byrne, D., and Dupin, J.M. 1993. Eocene-Oligocene tectonics and 907 kinematics of the Rhine-Saone Continental Transform Zone (eastern France). 908 Tectonics, 12: 874-888. doi:10.1029/93TC00233. 909 Le Pichon, X., İmren, C., Rangin, C., Şengör, A.M.C., and Siyako, M. 2014. The South 910 Marmara Fault. International Journal of Earth Sciences, 103: 219-231. 911 doi:10.1007/s00531-013-0950-0. 912 Le Pichon, X., Şengör, A.M.C., Demirbağ, E., Rangin, C., İmren, C., Armijo, R., Görür, N., 913 Çağatay, N., Mercier de Lapinay, B., Meyer, B., Saatçılar, R., and Tok, B. 2001. The 914 active Main Marmara Fault. Earth and Planetary Science Letters, 192: 595-616. 915 Lericolais, G., Bulois, C., Gillet, H., and Guichard, F. 2009. High frequency sea level 916 fluctuations recorded in the Black Sea since the LGM. Global and Planetary Change, 917 66: 65-75. doi:10.1016/j.gloplacha.2008.03.010. 918 Lericolais, G., Guichard, F., Morigi, C., Popescu, I., Bulois, C., Gillet, H., and Ryan, W.B.F. 919 2011. Assessment of Black Sea water-level fluctuations since the Last Glacial 920 Maximum. In Geology and Geoarchaeology of the Black Sea region: Beyond the Flood 921 Hypothesis. Edited by I.V. Buynevich and V. Yanko-Hombach and A.S. Gilbert and 922 R.H. Martin. GSA, Colorado. pp. 33-50.

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 32 of 59

923 McClusky, S., Reilinger, R., Mahmoud, S., Ben Sari, D., and Tealeb, A. 2003. GPS constraints 924 on Africa (Nubia) and Arabia plate motions. Geophysical Journal International, 155: 925 126-138. 926 McClusky, S., Balassanian, S., Barka, A., Demir, C., Ergintav, S., Georgiev, I., Gurkan, O., 927 Hamburger, M., Hurst, K., Kahle, H., Kastens, K., Kekelidze, G., King, R., Kotzev, V., 928 Lenk, O., Mahmoud, S., Mishin, A., Nadariya, M., Ouzounis, A., Paradissis, D., Peter, 929 Y., Prilepin, M., Reilinger, R., Sanli, I., Seeger, H., Tealeb, A., Toksöz, M., and Veis, 930 G. 2000. Global Positioning System constraints on plate kinematics and dynamics in 931 the eastern Mediterranean and Caucasus. J. Geophys. Res., 105: 5695-5719. 932 McClymont, A.F., Villamor, P., and Green, A.G. 2009. Assessing the contribution of off-fault 933 deformation to slip-rate estimates within the Taupo Rift, New Zealand, using 3-D 934 ground-penetrating radar surveying and trenching. Terra Nova, 21: 446-451. 935 doi:10.1111/j.1365-3121.2009.00901.x. 936 McKenzie, D. 1972. Active Tectonic of the Mediterranean Region. The Geophysical Journal of 937 the Royal Astronomical Society, 30: 109-185. 938 Meade, B.J., and Hager, B.H. 2004. Viscoelastic deformation for a clustered earthquake cycle. 939 Geophys. Res. Lett., 31: L10610. doi:10.1029/2004gl019643. 940 Meade, B.J., Klinger, Y., and Hetland, E.A. 2013. Inference of Multiple Earthquake‐Cycle 941 Relaxation Timescales from Irregular Geodetic Sampling of Interseismic Deformation. 942 Bulletin of the Seismological Society of America, 103: 2824-2835. 943 doi:10.1785/0120130006. 944 Meghraoui, M., Aksoy, M.E., Akyüz, H.S., Ferry, M., Dikbaş, A., and Altunel, E. 2012. 945 Paleoseismology of the North DraftAnatolian Fault at Güzelköy (Ganos segment, Turkey): 946 Size and recurrence time of earthquake ruptures west of the Sea of Marmara. 947 Geochemistry, Geophysics, Geosystems, 13: Q04005. doi:10.1029/2011GC003960. 948 Mériaux, A.S., Tapponnier, P., Ryerson, F.J., Xiwei, X., King, G., Van der Woerd, J., Finkel, 949 R.C., Haibing, L., Caffee, M.W., Zhiqin, X., and Wenbin, C. 2005. The Aksay segment 950 of the northern Altyn Tagh fault: Tectonic geomorphology, landscape evolution, and 951 Holocene slip rate. Journal of Geophysical Research: Solid Earth, 110: B04404. 952 doi:10.1029/2004JB003210. 953 Motagh, M., Hoffmann, J., Kampes, B., Baes, M., and Zschau, J. 2007. Strain accumulation 954 across the Gazikoy–Saros segment of the North Anatolian Fault inferred from Persistent 955 Scatterer Interferometry and GPS measurements. Earth and Planetary Science Letters, 956 255: 432-444. doi:http://dx.doi.org/10.1016/j.epsl.2007.01.003. 957 Örgülü, G. 2011. Seismicity and source parameters for small-scale earthquakes along the splays 958 of the North Anatolian Fault (NAF) in the Marmara Sea. Geophysical Journal 959 International, 184: 385–404. doi:10.1111/j.1365-246X.2010.04844.x. 960 Oskin, M., Perg, L., Shelef, E., Strane, M., Gurney, E., Singer, B., and Zhang, X. 2008. Elevated 961 shear zone loading rate during an earthquake cluster in eastern California. Geology, 36: 962 507-510. doi:10.1130/g24814a.1. 963 Özener, H., Arpat, E., Ergintav, S., Doğru, A., Çakmak, R., Turgut, B., and Doğan, U. 2010. 964 Kinematics of the eastern part of the North Anatolian Fault Zone. Journal of 965 Geodynamics, 49: 141-150. doi:10.1016/j.jog.2010.01.003. 966 Özeren, M.S., and Holt, W.E. 2010. The dynamics of the eastern Mediterranean and eastern 967 Turkey. Geophysical Journal International, 183: 1165-1184. doi:10.1111/j.1365- 968 246X.2010.04819.x. 969 Palyvos, N., Pantosti, D., Zabcı, C., and D’Addezio, G. 2007. Paleoseismological evidence of 970 recent earthquakes on the 1967 Mudurnu Valley earthquake segment of the North 971 Anatolian fault zone. Bulletin of the Seismological Society of America, 97: 1646–1661. 972 doi:10.1785/0120060049.

https://mc06.manuscriptcentral.com/cjes-pubs Page 33 of 59 Canadian Journal of Earth Sciences

973 Pamir, H., and Ketin, İ. 1941. Das anatolische Erdbeben Ende 1939 [10.1007/BF01799760]. 974 Geologische Rundschau, 32: 279-287. 975 Paul, A., Karabulut, H., Mutlu, A.K., and Salaün, G. 2014. A comprehensive and densely 976 sampled map of shear-wave anisotropy in the Aegean-Anatolia region. Earth and 977 Planetary Science Letters, 389: 14-22. 978 Peltzer, G., Crampé, F., Hensley, S., and Rosen, P. 2001. Transient strain accumulation and 979 fault interaction in the Eastern California shear zone. Geology, 29: 975-978. 980 Petersen, M.D., Dawson, T.E., Chen, R., Cao, T., Wills, C.J., Schwartz, D.P., and Frankel, A.D. 981 2011. Fault Displacement Hazard for Strike-Slip Faults. Bulletin of the Seismological 982 Society of America, 101: 805–825. doi:10.1785/0120100035. 983 Peyret, M., Masson, F., Yavasoglu, H., Ergintav, S., and Reilinger, R. 2013. Present-day strain 984 distribution across a segment of the central bend of the North Anatolian Fault Zone from 985 a Persistent-Scatterers InSAR analysis of the ERS and Envisat archives. Geophysical 986 Journal International, 192: 929-945. 987 Philippon, M., Brun, J.-P., Gueydan, F., and Sokoutis, D. 2014. The interaction between 988 Aegean back-arc extension and Anatolia escape since Middle Miocene. 989 Tectonophysics, 631: 176–188. doi:http://dx.doi.org/10.1016/j.tecto.2014.04.039. 990 Polonia, A., Gasperini, L., Amorosi, A., Bonatti, E., Bortoluzzi, G., Cagatay, N., Capotondi, 991 L., Cormier, M.H., Gorur, N., McHugh, C., and Seeber, L. 2004. Holocene slip rate of 992 the North Anatolian Fault beneath the Sea of Marmara. Earth and Planetary Science 993 Letters, 227: 411-426. 994 Pucci, S. 2006. The Düzce segment of the North Anatolian Fault Zone (Turkey): Understanding 995 its seismogenic behavior throughDraft earthquake geology, tectonic geomorphology and 996 paleoseismology. PhD, Universita degli Studi de Perugia. 997 Pucci, S., De Martini, P.M., and Pantosti, D. 2008. Preliminary slip rate estimates for the Düzce 998 segment of the North Anatolian Fault Zone from offset geomorphic markers. 999 Geomorphology, 97: 538-554. 1000 Pucci, S., Palyvos, N., Zabcı, C., Pantosti, D., and Barchi, M. 2006. Coseismic ruptures and 1001 tectonic landforms along the Düzce segment of the North Anatolian Fault Zone (Ms 7.1, 1002 November 1999). Journal of Geophysical Research, 111: B06312-B06312. 1003 Reilinger, R., McClusky, S., Vernant, P., Lawrence, S., Ergintav, S., Cakmak, R., Ozener, H., 1004 Kadirov, F., Guliev, I., Stepanyan, R., Nadariya, M., Hahubia, G., Mahmoud, S., Sakr, 1005 K., ArRajehi, A., Paradissis, D., Al-Aydrus, A., Prilepin, M., Guseva, T., Evren, E., 1006 Dmitrotsa, A., Filikov, S.V., Gomez, F., Al-Ghazzi, R., and Karam, G. 2006. GPS 1007 constraints on continental deformation in the Africa-Arabia-Eurasia continental 1008 collision zone and implications for the dynamics of plate interactions. Journal of 1009 Geophysical Research: Solid Earth, 111: B05411. doi:10.1029/2005JB004051. 1010 Reheis, M.C., and Dixon, T.H. 1996. Kinematics of the Eastern California shear zone: Evidence 1011 for slip transfer from Owens and Saline Valley fault zones to Fish Lake Valley fault 1012 zone. Geology, 24: 339–342. doi:10.1130/0091- 1013 7613(1996)024<0339:KOTECS>2.3.CO;2. 1014 Rockwell, T., Regona, D., Seitz, G., Langridge, R., Aksoy, M.E., Uçarkuş, G., Ferry, M., 1015 Meltzner, A.J., Meghraoui, M., Satir, D., Barka, A., and Akbalık, B. 2009. 1016 Paleoseismology of the North Anatolian Fault near the Marmara Sea: Implications for 1017 fault segmentation and seismic hazard. In Paleoseismology: Historical and Prehistorical 1018 Records of Earthquake Ground ffects for Seismic Hazard Assessment. Edited by K. 1019 Reicherter and A.M. Michetti and P.G. Silva. The Geological Society, London. pp. 31- 1020 54.

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 34 of 59

1021 Roten, D., Olsen, K.B., and Day, S.M. 2017. Off-fault deformations and shallow slip deficit 1022 from dynamic rupture simulations with fault zone plasticity. Geophysical Research 1023 Letters, 44: 7733-7742. doi:10.1002/2017GL074323. 1024 Sakic, P., Piété, H., Ballu, V., Royer, J.Y., Kopp, H., Lange, D., Petersen, F., Özeren, M.S., 1025 Ergintav, S., Geli, L., Henry, P., and Deschamps, A. 2016. No significant steady state 1026 surface creep along the North Anatolian Fault offshore Istanbul: Results of 6 months of 1027 seafloor acoustic ranging. Geophysical Research Letters, 43: 6817–6825. 1028 doi:10.1002/2016GL069600. 1029 Sançar, T., Akyüz, H.S., Schreurs, G., and Zabcı, C. 2018. Mechanics of plio-quaternary 1030 faulting around the Karliova triple junction: implications for the deformation of Eastern 1031 part of the Anatolian Scholle. Geodinamica Acta, 30: 287-305. 1032 doi:10.1080/09853111.2018.1533736. 1033 Seeber, L., Cormier, M.-H., McHugh, C., Emre, O., Polonia, A., and Sorlien, C. 2006. Rapid 1034 subsidence and sedimentation from oblique slip near a bend on the North Anatolian 1035 transform fault in the Marmara Sea, Turkey. Geology, 34: 933-936. 1036 Segall, P. 2002. Integrating Geologic and Geodetic Estimates of Slip Rate on the San Andreas 1037 Fault System. International Geology Review, 44: 62-82. doi:10.2747/0020- 1038 6814.44.1.62. 1039 Segall, P., and Davis, J.L. 1997. GPS Applications For Geodynamics and Earthquake Studies. 1040 Annual Review of Earth and Planetary Sciences, 25: 301-336. 1041 doi:doi:10.1146/annurev.earth.25.1.301. 1042 Şengör, A.M.C. 1979. The North Anatolian transform fault; its age, offset and tectonic 1043 significance. Journal of the GeologicalDraft Society of London, 136, Part 3: 269-282. 1044 Şengör, A.M.C. 1980. Türkiye'nin neotektoniğinin esasları (Principles of the Neotectonism of 1045 Turkey). Türkiye Jeoloji Kurumu Konferans Serisi, Ankara. pp. 40. 1046 Şengör, A.M.C. 2014. Transform Fault. In Encyclopedia of Marine Geosciences. Edited by J. 1047 Harff and M. Meschede and S. Petersen and J. Thiede. Springer Netherlands, Dordrecht. 1048 pp. 1-22. 1049 Şengör, A.M.C. 2017. Diversion of river courses across major strike-slip faults and keirogens. 1050 In Active Global Seismology. Edited by İ. Çemen and Y. Yılmaz. John Wiley & Sons, 1051 Inc., New Jersey. pp. 93-101. 1052 Şengör, A.M.C., and Zabcı, C. 2019. The North Anatolian Fault and the North Anatolian Shear 1053 Zone. In Landscapes and landforms of Turkey. Edited by C. Kuzucuoğlu and A. Çiner 1054 and N. Kazancı. Springer. 1055 Şengör, A.M.C., Görür, N., and Şaroğlu, F. 1985. Strike-slip faulting and related basin 1056 formation in zones of tectonic escape: Turkey as a case study. In Strike-slip 1057 Deformation, Basin Formation, and Sedimentation, Soc. Econ. Paleontol. Spec. Publ. 1058 Edited by K.T. Biddle and N. Christie-Blick. Society of Economic Paleontologists and 1059 Mineralogists, Oklahoma. pp. 227-264. 1060 Şengör, A.M.C., Zabcı, C., and Natal'in, B.A. 2019. Chapter 9 - Continental Transform Faults: 1061 Congruence and Incongruence With Normal Plate Kinematics. In Transform Plate 1062 Boundaries and Fracture Zones. Edited by J.C. Duarte. Elsevier. pp. 169-247. 1063 Şengör, A.M.C., Tüysüz, O., İmren, C., Sakınç, M., Eyidoğan, H., Görür, N., Le Pichon, X., 1064 and Rangin, C. 2005. The North Anatolian Fault: A New Look. Annual Review of Earth 1065 and Planetary Sciences, 33: 37-112. 1066 Şengör, A.M.C., Grall, C., İmren, C., Le Pichon, X., Görür, N., Henry, P., Karabulut, H., and 1067 Siyako, M. 2014. The geometry of the North Anatolian transform fault in the Sea of 1068 Marmara and its temporal evolution: implications for the development of 1069 intracontinental transform faults1. Canadian Journal of Earth Sciences, 51: 222-242. 1070 doi:10.1139/cjes-2013-0160.

https://mc06.manuscriptcentral.com/cjes-pubs Page 35 of 59 Canadian Journal of Earth Sciences

1071 Shelef, E., and Oskin, M. 2010. Deformation processes adjacent to active faults: Examples from 1072 eastern California. Journal of Geophysical Research: Solid Earth, 115: B05308. 1073 doi:10.1029/2009JB006289. 1074 Stein, R.S., Barka, A.A., and Dieterich, J.H. 1997. Progressive failure on the North Anatolian 1075 fault since 1939 by earthquake stress triggering. Geophysical Journal International, 128: 1076 594-604. 1077 Sugai, T., Emre, O., Duman, T.Y., Yoshioka, T., and Kuscu, I. 1999. Geologic evidence for 1078 five large earthquakes on the North Anatolian Fault at Ilgaz, during the last 2000 years; 1079 a result of GSJ-MTA international cooperative research. In The Peleoseismology 1080 Workshop. Edited by K. Satake and D.P. Schwartz. USGS Open-File Report, Tsukuba, 1081 Japan. pp. 66-72. 1082 Tatar, O., Poyraz, F., Gürsoy, H., Cakir, Z., Ergintav, S., Akpınar, Z., Koçbulut, F., Sezen, F., 1083 Türk, T., Hastaoğlu, K.Ö., Polat, A., Mesci, B.L., Gürsoy, Ö., Ayazlı, İ.E., Çakmak, R., 1084 Belgen, A., and Yavaşoğlu, H. 2012. Crustal deformation and kinematics of the Eastern 1085 Part of the North Anatolian Fault Zone (Turkey) from GPS measurements. 1086 Tectonophysics, 518–521: 55-62. doi:10.1016/j.tecto.2011.11.010. 1087 Tatar, Y. 1978. Tectonic investigations on the North Anatolian Fault Zone between Erzincan 1088 and Refahiye. Publ. Inst Earth Sci. Hacettepe Univ., 4: 136-201. 1089 Thatcher, W. 2009. How the Continents Deform: The Evidence From Tectonic Geodesy*. 1090 Annual Review of Earth and Planetary Sciences, 37: 237-262. 1091 doi:doi:10.1146/annurev.earth.031208.100035. 1092 Titus, S.J., Dyson, M., DeMets, C., Tikoff, B., Rolandone, F., and B√ºrgmann, R. 2011. 1093 Geologic versus geodetic deformationDraft adjacent to the San Andreas fault, central 1094 California. Geological Society of America Bulletin, 123: 794-820. 1095 doi:10.1130/b30150.1. 1096 Turpeinen, H., Hampel, A., Karow, T., and Maniatis, G. 2008. Effect of ice sheet growth and 1097 melting on the slip evolution of thrust faults. Earth and Planetary Science Letters, 269: 1098 230-241. doi:10.1016/j.epsl.2008.02.017. 1099 Tutkun, S.Z., and Hancock, P.L. 1990. Tectonic landforms expressing strain at the Karlıova 1100 continental triple junction (E Turkey). Annales Tectonicae, IV: 182–195. 1101 Uçarkuş, G. 2019. A synthesis on fault geometry, earthquake rupture and fault creep along the 1102 North Anatolian Fault in the Sea of Marmara. European Geosciences Union, Vienna. 1103 pp. EGU2019-14947–1. Available from 1104 https://meetingorganizer.copernicus.org/EGU2019/EGU2019-14947-1.pdf. 1105 Uçarkuş, G., Çakır, Z., and Armijo, R. 2011. Western termination of the Mw 7.4, 1999 İzmit 1106 Earthquake Rupture: Implications for the Expected Large Earthquake in the Sea of 1107 Marmara. Turkish Journal of Earth Sciences, 20: 379-394. 1108 Ustaömer, T., Gökaşan, E., Tur, H., Görüm, T., Batuk, F., Kalafat, D., Alp, H., Ecevitoğlu, B., 1109 and Birkan, H. 2008. Faulting, mass-wasting and deposition in an active dextral shear 1110 zone, the Gulf of Saros and the NE Aegean Sea, NW Turkey. Geo-Marine Letters, 28: 1111 171-193. doi:10.1007/s00367-007-0099-6. 1112 Vardar, D., Öztürk, K., Yaltırak, C., Alpar, B., and Tur, H. 2014. Late Pleistocene–Holocene 1113 evolution of the southern Marmara shelf and sub-basins: middle strand of the North 1114 Anatolian fault, southern Marmara Sea, Turkey. Marine Geophysical Research, 35: 69- 1115 85. doi:10.1007/s11001-013-9210-8. 1116 Vernant, P. 2015. What can we learn from 20 years of interseismic GPS measurements across 1117 strike-slip faults? Tectonophysics, 644–645: 22-39. 1118 doi:http://dx.doi.org/10.1016/j.tecto.2015.01.013.

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 36 of 59

1119 Walker, J.D., Kirby, E., and Andrew, J.E. 2005. Strain transfer and partitioning between the 1120 Panamint Valley, Searles Valley, and Ash Hill fault zones, California. Geosphere, 1: 1121 111–118. doi:10.1130/GES00014.1. 1122 Walters, R.J., Parsons, B., and Wright, T.J. 2014. Constraining crustal velocity fields with 1123 InSAR for Eastern Turkey: Limits to the block-like behavior of Eastern Anatolia. 1124 Journal of Geophysical Research: Solid Earth, 119: 5215-5234. 1125 doi:10.1002/2013JB010909. 1126 Walters, R.J., Holley, R.J., Parsons, B., and Wright, T.J. 2011. Interseismic strain accumulation 1127 across the North Anatolian Fault from Envisat InSAR measurements. Geophysical 1128 Research Letters, 38: L05303. doi:10.1029/2010GL046443. 1129 Wessel, P., Smith, W.H.F., Scharroo, R., Luis, J., and Wobbe, F. 2013. Generic Mapping Tools: 1130 Improved Version Released. Eos, Transactions American Geophysical Union, 94: 409- 1131 410. doi:10.1002/2013EO450001. 1132 Wilson, J.T. 1965. A New Class of Faults and their Bearing on Continental Drift 1133 [10.1038/207343a0]. Nature, 207: 343-347. 1134 Wright, T., Parsons, B., and Fielding, E. 2001. Measurement of interseismic strain 1135 accumulation across the North Anatolian Fault by satellite radar interferometry. 1136 Geophysical Research Letters, 28: 2117-2120. doi:10.1029/2000GL012850. 1137 Xu, X.W., Tapponnier, P., Van Der Woerd, J., Ryerson, F.J., Wang, F., Zheng, R.Z., Chen, 1138 W.B., Ma, W.T., Yu, G.H., Chen, G.H., and Meriaux, A.S. 2005. Late quaternary 1139 sinistral slip rate along the Altyn Tagh Fault and its structural transformation model. 1140 Science in China Series D-Earth Sciences, 48: 384-397. 1141 Yakupoğlu, N., Uçarkuş, G., Kadir DraftEriş, K., Henry, P., and Namık Çağatay, M. 2019. 1142 Sedimentological and geochemical evidence for seismoturbidite generation in the 1143 Kumburgaz Basin, Sea of Marmara: Implications for earthquake recurrence along the 1144 Central High Segment of the North Anatolian Fault. Sedimentary Geology, 380: 31–44. 1145 doi:https://doi.org/10.1016/j.sedgeo.2018.11.002. 1146 Yaltırak, C. 2002. Tectonic evolution of the Marmara Sea and its surroundings. Marine 1147 Geology, 190: 493-529. doi:Doi: 10.1016/s0025-3227(02)00360-2. 1148 Yamamoto, R., Kido, M., Ohta, Y., Takahashi, N., Yamamoto, Y., Pinar, A., Kalafat, D., 1149 Özener, H., and Kaneda, Y. 2019. Seafloor Geodesy Revealed Partial Creep of the North 1150 Anatolian Fault Submerged in the Sea of Marmara. Geophysical Research Letters, 46: 1151 1268–1275. doi:10.1029/2018GL080984. 1152 Yavaşoğlu, H., Tarı, E., Tüysüz, O., Çakır, Z., and Ergintav, S. 2011. Determining and 1153 modeling tectonic movements along the central part of the North Anatolian Fault 1154 (Turkey) using geodetic measurements. Journal of Geodynamics, 51: 339-343. 1155 doi:10.1016/j.jog.2010.07.003. 1156 Ye, J., and Liu, M. 2017. How fault evolution changes strain partitioning and fault slip rates in 1157 Southern California: Results from geodynamic modeling. Journal of Geophysical 1158 Research: Solid Earth, 122: 6893-6909. doi:10.1002/2017JB014325. 1159 Zabcı, C. 2012. Kuzey Anadolu Fayı’nın Ilgaz (Çankırı) - Karlıova (Bingöl) arasında kalan 1160 kesiminin morfokronoloji tabanlı son beşbin yıllık kayma hızı tarihçesi ve 1161 depremselliği. PhD, Avrasya Yer Bilimleri Enstitüsü, İstanbul Teknik Üniversitesi, 1162 İstanbul. 1163 Zabcı, C., Akyüz, H.S., Karabacak, V., Sançar, T., Altunel, E., Gürsoy, H., and Tatar, O. 2011. 1164 Palaeoearthquakes on the Kelkit Valley Segment of the North Anatolian Fault, Turkey: 1165 Implications for the Surface Rupture of the Historical 17 August 1668 Earthquake. 1166 Turkish Journal of Earth Sciences, 20: 411–427.

https://mc06.manuscriptcentral.com/cjes-pubs Page 37 of 59 Canadian Journal of Earth Sciences

1167 Zabcı, C., Akyüz, H.S., and Sançar, T. 2017. Palaeoseismic history of the eastern part of the 1168 North Anatolian Fault (Erzincan, Turkey): Implications for the seismicity of the Yedisu 1169 seismic gap. Journal of Seismology, 21: 1407-1425. doi:10.1007/s10950-017-9673-1. 1170 Zabcı, C., Sançar, T., Akyüz, H.S., and Kıyak, N.G. 2015. Spatial slip behavior of large strike- 1171 slip fault belts: Implications for the Holocene slip rates of the eastern termination of the 1172 North Anatolian Fault, Turkey. Journal of Geophysical Research: Solid Earth, 120: 1173 8591-8609. doi:10.1002/2015JB011874. 1174 Zechar, J.D., and Frankel, K.L. 2009. Incorporating and reporting uncertainties in fault slip 1175 rates. Journal of Geophysical Research: Solid Earth, 114: B12407. 1176 doi:10.1029/2009JB006325. 1177 Zhang, P.-Z., Molnar, P., and Xu, X. 2007. Late Quaternary and present-day rates of slip along 1178 the Altyn Tagh Fault, northern margin of the Tibetan Plateau. Tectonics, 26: TC5010. 1179 doi:10.1029/2006TC002014. 1180

Draft

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 38 of 59

FIGURE CAPTIONS

Figure 1. Neotectonic structures of the eastern Mediterranean, which are late medial Miocene (13 Ma) or younger (Şengör and Zabcı 2019; Şengör et al. 2019). All red lines show strike-slip faults, whereas the green is for extensional and the purple colour is for compressional structures. The yellow shaded region marks the North Anatolian Keirogen (NAK) and the North Anatolian Shear Zone (NASZ), which constantly widen from east to west (Şengör and Zabcı 2019; Şengör et al. 2005). The yellow rectangle shows the boundaries of the Figure 2, which includes the most prominent member of the NASZ, the North Anatolian Fault. Key to lettering; H – Hellenic Trench, C – Cyprus Trench, EASZ – East Anatolian Shear Zone, EAHP – East Anatolian High Plateau

Figure 2. The North Anatolian Fault and the distribution of the surface ruptures of the last century earthquakes. The fault map represents surface ruptures of the 20th century earthquakes, including 1912a and 1912b Ganos (Aksoy 2009; Aksoy et al. 2010a; Altunel et al. 2004), 1939 Erzincan (Barka 1996; Gürsoy et al. 2013; Ketin 1969; Pamir and Ketin 1941; Tatar 1978), 1942 Niksar (Barka 1996; Blumenthal et al. 1943; Ketin 1969), 1943 Tosya–Lâdik (Barka 1996; Blumenthal 1945), 1944 Bolu–Gerede (Barka 1996; Kondo et al. 2005), 1944 Edremit (Altınok et al. 2012), 1949 Elmalı (Barka and Kadinsky-Cade 1988; Sançar et al. 2018; Zabcı et al. 2015), 1951 Kurşunlu (Ketin 1969), 1953 Yenice – Gönen (Ketin and Rösli 1953), 1957 Abant (Ketin 1969), 1964 ManyasDraft , 1967 Mudurnu Valley (Ambraseys and Zatopek 1969), 1992 Erzincan (Barka and Eyidoǧan 1993), 1999a İzmit (Barka et al. 2002; Emre et al. 2003; Hartleb et al. 2002; Uçarkuş et al. 2011), 1999b Düzce (Akyüz et al. 2002; Konca et al. 2010; Pucci 2006; Pucci et al. 2006), 2000 Orta (Çakır and Akoğlu 2008; Koçyiğit et al. 2001) and 2014 Samothraki–Gökçeada (Gasperini et al. 2011; Konca et al. 2018; Ustaömer et al. 2008) earthquakes with addition of some unruptured important segments Yedisu Fault (Zabcı et al. 2017), Esençay–Suluova faults (Erturaç 2009; Erturaç and Tüysüz 2012), and the Main Marmara Fault (Le Pichon et al. 2001; Şengör et al. 2014). The rectangles show inSAR studies (Çakır et al. 2014; Cavalié and Jónsson 2014; Hussain et al. 2016; Hussain et al. 2018; Motagh et al. 2007; Peyret et al. 2013; Walters et al. 2014; Walters et al. 2011; Wright et al. 2001), whereas arrows represent GPS velocities (Aktuğ et al. 2015; Aktuğ et al. 2013b; Ergintav et al. 2014; Özener et al. 2010; Reilinger et al. 2006; Tatar et al. 2012; Yavaşoğlu et al. 2011) and stars are for geological slip rate sites. Key to lettering; K – Karlıova, Er – Erzincan, N – Niksar, A – Akyazı, B – Bayramören, YF – Yedisu Fault, EsF – Esençay Fault, SuF – Suluova Fault, E-SF – Ezinepazarı-Sungurlu Fault, MMF – Main Marmara Fault, KB – Kumburgaz Basin, SG – Saros Gulf.

Figure 3. (a) The location of geodetic studies along the North Anatolian Fault and the distribution of the (b) GPS and (c) inSAR-based geodetical slip rates along the North Anatolian Fault (for references of the fault, please see Fig. 2). On each figure, the vertical bars represent rates measured from velocity profiles with their error margins, whereas horizontal bars show block model velocities.

Figure 4. (a) The location of geological slip rate sites along the North Anatolian Fault (for references of the fault map, please see Fig. 2) and (b) the projection of each slip rate estimate along the distance of the fault. For detailed coordinates, definitions, numerical values and references please refer to Table 1.

https://mc06.manuscriptcentral.com/cjes-pubs Page 39 of 59 Canadian Journal of Earth Sciences

Figure 5. Geological slip rate sites, where the North Anatolian Fault bifurcates into parallel branches. (a) shows the geological slip rate site of Gasperini et al. (2011) (site no. 1) where there are two sub-parallel branches within the Saros Gulf. Likewise, (b) displays the geological sites to the north and south of the Almacık Block, which is bounded by the Düzce Fault to the north and the Mudurnu Valley to the south (for references of the fault map, please see Fig. 2). It is clear that these parallel/sub-parallel branches share the total strain. For details of slip rate sites, please see Fig. 4 and Table 1.

Figure 6. The slip history solution of the Model I, which is calculated for the fault section between Yedisu in the east and Niksar in the west (Fig. 10 and Table 3). This part of the NASZ is particularly known to have the narrowest width. (A.1 and A.2) represent the last 12 ka, whereas (B.1 and B.2) is calculated only with the morphochronology-based geological slip rate sites for the last 4 ka. The left column (A.1 and B.1) shows the uniform slip rate solution, while the right one (A.2 and B.2) displays the slip history solutions for two different time intervals. Site names and site numbers (in parenthesis) are indicated next to each D-t envelope. (A.1 and B.1). In A.2 and B.2, the black solid line represent the median, whereas the grey-shaded region is for 68.27% and the white region is for the 95.45% confidence bounds. The modelled median slip rate is 16.8 and 17.5 mm/a for the last 12 ka and 4 ka, respectively. The slip history solutions show no significant change in the slip history paths for two time intervals. Draft Figure 7. The slip history solution of the Model II, which is calculated for the fault section between Niksar in the east and Bolu in the west (Fig. 10 and Table 3). Here, the deformation width of the NASZ shows a significant thickening, which is well-marked with broad deflection of Yeşilırmak and Kızılırmak rivers (Şengör 2017; Şengör and Zabcı 2019; Şengör et al. 2005). (A.1 and A.2) represent the last 12 ka, whereas (B.1 and B.2) is calculated only with the morphochronology- and palaeoseismology-based geological slip rate sites for the last 5 ka. The left column (A.1 and B.1) shows the uniform slip rate solution, while the right one (A.2 and B.2) displays the slip history solutions for two different time intervals. Site names and site numbers (in parenthesis) are indicated next to each D-t envelope. (A.1 and B.1). In A.2 and B.2, the black solid line represent the median, whereas the grey-shaded region is for 68.27% and the white region is for the 95.45% confidence bounds. The modelled median slip rate is 17.4 and 18.9 mm/a for the last 12 ka and 5 ka, respectively. The slip history solutions show no significant change in the slip history paths for two time intervals. There is only a slight change of slope at 1000 years (B.2), mostly because of the limited temporal control of the Demirtepe Site.

Figure 8. The slip history solution of the Model II, which is calculated for the fault section between Akyazı in the east and Saros Gulf in the west (Fig. 10 and Table 3). The NASZ reaches its maximum thickness with respect to the distribution of the Tethyan sutures in the region (Şengör and Zabcı 2019; Şengör et al. 2005). (A.1 and A.2) represent the last 550 ka, whereas (B.1 and B.2) is for the last 20 ka and (C.1 and C.2) is calculated only with the morphochronology- and palaeoseismology-based geological slip rate sites for the last 3 ka. The left column (A.1, B.1 and C.1) shows the uniform slip rate solutions, while the right one (A.2, B.2 and C.2) displays the slip history solutions for three different time intervals. Site names and site numbers (in parenthesis) are indicated next to each D-t envelope. (A.1, B.1

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 40 of 59

and C.1). In A.2, B.2 and C.2, the black solid line represent the median, whereas the grey- shaded region is for 68.27% and the white region is for the 95.45% confidence bounds. The modelled and average median slip rate is 16.7, 17.7 and 16.9 mm/a for the last 550 ka, 20 ka, and 3 ka respectively. The slip history solutions show no significant change in the slip history paths for the last 20 ka. However, there are two significant breaks in the median rate slopes in my analysis for the last 3 ka and the 550 ka. For the case of the last 3 ka (C.2), the slip rate slope changes at the Nehirkent Site, where the rate could be overestimated due to misdating of the offset landform here. Nevertheless, the slope break in A.2 may show a significant deceleration between 195 and 320 ka.

Figure 9. Reconstruction of the incremental displacements of the Çınarcık Basin depocenter, where the left column represents the minimum and the right one shows the maximum offsets for each vertical thickness (isochore) maps between different seismic horizons. Isochore maps between each seismic horizon are re-interpolated by using the supplementary data of Kurt et al. (2013). Together with the colour of the particular seismic horizon, black strips show the initial and white strips represent the recent positions of each 500 m isochore contours. (A1 and A2) 1550 and 1950 m are offsets of the H2/H1 (Blue) to H1/Sea floor (Red) isochores in 130 ka; (B1 and B2) 2900 and 3550 m are offsets of the H3/H2 (Yellow) to H1/Sea floor (Red) isochores in 250 ka; (C1 and C2) 4900 and 5600 m are offsets of the H4/H3 (Violet) to H1/Sea floor (Red) isochores in 350 ka; and (D1 and D2) 7550 and 8400 m are offsets of the H5/H4 (Green)Draft to H1/Sea floor (Red) isochores in 430 ka. Figure 10. The comparison of geodetic and geological slip rates along the North Anatolian Fault. (a) shows the location of all geological and geodetic sites (for references of the fault map, please see Fig. 2). (b) includes all geological slip rate estimates, whereas (c) only includes the outcomes of three models for the three distinct sections of the fault. The model outcomes only include analyses of morphochronology- and palaeoseismology-based slip rates. The light grey shade represents 2-sigma, whereas dark grey is for 1-sigma confidence boundary. In (b) the geological slip rates display a chaotic and messy distribution, while (c) represents a more controlled comparison between the geodetic and modelled geological rates. Modelled geological rates are all exceeded by all GPS-based block model rates, whereas they are slightly slower than rates of inSAR- or GPS-based velocity profiles. The exceptions are some GPS profiles in the eastern sections and the one across the Çınarcık Basin. Orange bars are used for inSAR, blue is for GPS and black is for geological rates.

https://mc06.manuscriptcentral.com/cjes-pubs Page 41 of 59 Canadian Journal of Earth Sciences

Draft

Figure 1. Neotectonic structures of the eastern Mediterranean, which are late medial Miocene (13 Ma) or younger (Şengör and Zabcı 2019; Şengör et al. 2019). All red lines show strike-slip faults, whereas the green is for extensional and the purple colour is for compressional structures. The yellow shaded region marks the North Anatolian Keirogen (NAK) and the North Anatolian Shear Zone (NASZ), which constantly widen from east to west (Şengör and Zabcı 2019; Şengör et al. 2005). The yellow rectangle shows the boundaries of the Figure 2, which includes the most prominent member of the NASZ, the North Anatolian Fault. Key to lettering; H – Hellenic Trench, C – Cyprus Trench, EASZ – East Anatolian Shear Zone, EAHP – East Anatolian High Plateau

177x138mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 42 of 59

Figure 2. The North Anatolian Fault and the distribution of the surface ruptures of the last century earthquakes. The fault map represents surface ruptures of the 20th century earthquakes, including 1912a and 1912b Ganos (Aksoy 2009; Aksoy etDraft al. 2010a; Altunel et al. 2004), 1939 Erzincan (Barka 1996; Gürsoy et al. 2013; Ketin 1969; Pamir and Ketin 1941; Tatar 1978), 1942 Niksar (Barka 1996; Blumenthal et al. 1943; Ketin 1969), 1943 Tosya–Lâdik (Barka 1996; Blumenthal 1945), 1944 Bolu–Gerede (Barka 1996; Kondo et al. 2005), 1944 Edremit (Altınok et al. 2012), 1949 Elmalı (Barka and Kadinsky-Cade 1988; Sançar et al. 2018; Zabcı et al. 2015), 1951 Kurşunlu (Ketin 1969), 1953 Yenice – Gönen (Ketin and Rösli 1953), 1957 Abant (Ketin 1969), 1964 Manyas , 1967 Mudurnu Valley (Ambraseys and Zatopek 1969), 1992 Erzincan (Barka and Eyidoǧan 1993), 1999a İzmit (Barka et al. 2002; Emre et al. 2003; Hartleb et al. 2002; Uçarkuş et al. 2011), 1999b Düzce (Akyüz et al. 2002; Konca et al. 2010; Pucci 2006; Pucci et al. 2006), 2000 Orta (Çakır and Akoğlu 2008; Koçyiğit et al. 2001) and 2014 Samothraki–Gökçeada (Gasperini et al. 2011; Konca et al. 2018; Ustaömer et al. 2008) earthquakes with addition of some unruptured important segments Yedisu Fault (Zabcı et al. 2017), Esençay–Suluova faults (Erturaç 2009; Erturaç and Tüysüz 2012), and the Main Marmara Fault (Le Pichon et al. 2001; Şengör et al. 2014). The rectangles show inSAR studies (Çakır et al. 2014; Cavalié and Jónsson 2014; Hussain et al. 2016; Hussain et al. 2018; Motagh et al. 2007; Peyret et al. 2013; Walters et al. 2014; Walters et al. 2011; Wright et al. 2001), whereas arrows represent GPS velocities (Aktuğ et al. 2015; Aktuğ et al. 2013b; Ergintav et al. 2014; Özener et al. 2010; Reilinger et al. 2006; Tatar et al. 2012; Yavaşoğlu et al. 2011) and stars are for geological slip rate sites. Key to lettering; K – Karlıova, Er – Erzincan, N – Niksar, A – Akyazı, B – Bayramören, YF – Yedisu Fault, EsF – Esençay Fault, SuF – Suluova Fault, E-SF – Ezinepazarı-Sungurlu Fault, MMF – Main Marmara Fault, KB – Kumburgaz Basin, SG – Saros Gulf.

266x145mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Page 43 of 59 Canadian Journal of Earth Sciences

Draft

Figure 3. (a) The location of geodetic studies along the North Anatolian Fault and the distribution of the (b) GPS and (c) inSAR-based geodetical slip rates along the North Anatolian Fault (for references of the fault, please see Fig. 2). On each figure, the vertical bars represent rates measured from velocity profiles with their error margins, whereas horizontal bars show block model velocities.

191x208mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 44 of 59

Draft

Figure 4. (a) The location of geological slip rate sites along the North Anatolian Fault (for references of the fault map, please see Fig. 2) and (b) the projection of each slip rate estimate along the distance of the fault. For detailed coordinates, definitions, numerical values and references please refer to Table 1.

191x137mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Page 45 of 59 Canadian Journal of Earth Sciences

Figure 5. Geological slip rate sites, where the North Anatolian Fault bifurcates into parallel branches. (a) shows the geological slip rate site of Gasperini et al. (2011) (site no. 1) where there are two sub-parallel branches within the Saros Gulf. Likewise, (b) displays the geological sites to the north and south of the Almacık Block, which is bounded by the Düzce Fault to the north and the Mudurnu Valley to the south (for references of the fault map, please see Fig. 2). It is clear that these parallel/sub-parallel branches share the total strain. For details of slip rate sites, please see Fig. 4 and Table 1.

197x54mm (300 x 300 DPI) Draft

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 46 of 59

Draft

Figure 6. The slip history solution of the Model I, which is calculated for the fault section between Yedisu in the east and Niksar in the west (Fig. 10 and Table 3). This part of the NASZ is particularly known to have the narrowest width. (A.1 and A.2) represent the last 12 ka, whereas (B.1 and B.2) is calculated only with the morphochronology-based geological slip rate sites for the last 4 ka. The left column (A.1 and B.1) shows the uniform slip rate solution, while the right one (A.2 and B.2) displays the slip history solutions for two different time intervals. Site names and site numbers (in parenthesis) are indicated next to each D-t envelope. (A.1 and B.1). In A.2 and B.2, the black solid line represent the median, whereas the grey-shaded region is for 68.27% and the white region is for the 95.45% confidence bounds. The modelled median slip rate is 16.8 and 17.5 mm/a for the last 12 ka and 4 ka, respectively. The slip history solutions show no significant change in the slip history paths for two time intervals.

194x146mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Page 47 of 59 Canadian Journal of Earth Sciences

Draft

Figure 7. The slip history solution of the Model II, which is calculated for the fault section between Niksar in the east and Bolu in the west (Fig. 10 and Table 3). Here, the deformation width of the NASZ shows a significant thickening, which is well-marked with broad deflection of Yeşilırmak and Kızılırmak rivers (Şengör 2017; Şengör and Zabcı 2019; Şengör et al. 2005). (A.1 and A.2) represent the last 12 ka, whereas (B.1 and B.2) is calculated only with the morphochronology- and palaeoseismology-based geological slip rate sites for the last 5 ka. The left column (A.1 and B.1) shows the uniform slip rate solution, while the right one (A.2 and B.2) displays the slip history solutions for two different time intervals. Site names and site numbers (in parenthesis) are indicated next to each D-t envelope. (A.1 and B.1). In A.2 and B.2, the black solid line represent the median, whereas the grey-shaded region is for 68.27% and the white region is for the 95.45% confidence bounds. The modelled median slip rate is 17.4 and 18.9 mm/a for the last 12 ka and 5 ka, respectively. The slip history solutions show no significant change in the slip history paths for two time intervals. There is only a slight change of slope at 1000 years (B.2), mostly because of the limited temporal control of the Demirtepe Site.

187x152mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 48 of 59

Draft

Figure 8. The slip history solution of the Model II, which is calculated for the fault section between Akyazı in the east and Saros Gulf in the west (Fig. 10 and Table 3). The NASZ reaches its maximum thickness with respect to the distribution of the Tethyan sutures in the region (Şengör and Zabcı 2019; Şengör et al. 2005). (A.1 and A.2) represent the last 550 ka, whereas (B.1 and B.2) is for the last 20 ka and (C.1 and C.2) is calculated only with the morphochronology- and palaeoseismology-based geological slip rate sites for the last 3 ka. The left column (A.1, B.1 and C.1) shows the uniform slip rate solutions, while the right one (A.2, B.2 and C.2) displays the slip history solutions for three different time intervals. Site names and site numbers (in parenthesis) are indicated next to each D-t envelope. (A.1, B.1 and C.1). In A.2, B.2 and C.2, the black solid line represent the median, whereas the grey-shaded region is for 68.27% and the white region is for the 95.45% confidence bounds. The modelled and average median slip rate is 16.7, 17.7 and 16.9 mm/a for the last 550 ka, 20 ka, and 3 ka respectively. The slip history solutions show no significant change in the slip history paths for the last 20 ka. However, there are two significant breaks in the median rate slopes in my analysis for the last 3 ka and the 550 ka. For the case of the last 3 ka (C.2), the slip rate slope changes at the Nehirkent Site, where the rate could be overestimated due to misdating of the offset

https://mc06.manuscriptcentral.com/cjes-pubs Page 49 of 59 Canadian Journal of Earth Sciences

landform here. Nevertheless, the slope break in A.2 may show a significant deceleration between 195 and 320 ka.

187x224mm (300 x 300 DPI)

Draft

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 50 of 59

Draft

Figure 9. Reconstruction of the incremental displacements of the Çınarcık Basin depocenter, where the left column represents the minimum and the right one shows the maximum offsets for each vertical thickness (isochore) maps between different seismic horizons. Isochore maps between each seismic horizon are re- interpolated by using the supplementary data of Kurt et al. (2013). Together with the colour of the particular seismic horizon, black strips show the initial and white strips represent the recent positions of each 500 m isochore contours. (A1 and A2) 1550 and 1950 m are offsets of the H2/H1 (Blue) to H1/Sea floor (Red) isochores in 130 ka; (B1 and B2) 2900 and 3550 m are offsets of the H3/H2 (Yellow) to H1/Sea floor (Red) isochores in 250 ka; (C1 and C2) 4900 and 5600 m are offsets of the H4/H3 (Violet) to H1/Sea floor (Red) isochores in 350 ka; and (D1 and D2) 7550 and 8400 m are offsets of the H5/H4 (Green) to H1/Sea floor (Red) isochores in 430 ka.

203x284mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Page 51 of 59 Canadian Journal of Earth Sciences

Draft

Figure 10. The comparison of geodetic and geological slip rates along the North Anatolian Fault. (a) shows the location of all geological and geodetic sites (for references of the fault map, please see Fig. 2). (b) includes all geological slip rate estimates, whereas (c) only includes the outcomes of three models for the three distinct sections of the fault. The model outcomes only include analyses of morphochronology- and palaeoseismology-based slip rates. The light grey shade represents 2-sigma, whereas dark grey is for 1- sigma confidence boundary. In (b) the geological slip rates display a chaotic and messy distribution, while (c) represents a more controlled comparison between the geodetic and modelled geological rates. Modelled geological rates are all exceeded by all GPS-based block model rates, whereas they are slightly slower than rates of inSAR- or GPS-based velocity profiles. The exceptions are some GPS profiles in the eastern sections and the one across the Çınarcık Basin. Orange bars are used for inSAR, blue is for GPS and black is for geological rates.

191x204mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 52 of 59

TABLE CAPTIONS

Table 1. The geological slip rate sites along the North Anatolian Fault

Table 2. Models and their main parameters for the slip history analysis

Table 3. Results of slip history analyses for three models

Draft

https://mc06.manuscriptcentral.com/cjes-pubs Page 53 of 59 Canadian Journal of Earth Sciences

SITE Average slip-rate SITE LON (E) LAT (N) Offset (m) Age (a) Dating Method Type Model NO (mm/a) 1 Saros 26.13 40.41 130±10 12500±1200 10.4±2.5 Sea Level, Radiocarbon Morphochronology - 4.5 m/per Radiocarbon, historical Kavakköy 26.86 40.61 15.9 +10/-4 - event earthquakes 2 Palaeoseismology Radiocarbon, historical Kavakköy 26.86 40.61 9±1 563±5a 16±1.8a III earthquakes Offset streams along Ganos the Ganos fault: 1st 70±10b 4000±400b 17.5 +3.2/-2.9 Palaeoclimate III order Offset streams along Ganos the Ganos fault: 2nd 183±20b 10200±1000b 17.9 +2.8/-2.5 Palaeoclimate III order Offset streams along 3 Ganos the Ganos fault: 3rd 202±20b 12500±1250Draftb 16.2 +2.4/-2.1 Palaeoclimate III order Offset streams along Ganos the Ganos fault: 4th 255±25b 14500±1450b 17.6 +2.6/-2.3 Palaeoclimate III order Offset streams along Ganos the Ganos fault: 5th 320±35b 17500±1750b 18.3 +2.9/-2.6 Palaeoclimate III order

4 Yeniköy 26.99 40.65 46±1 2715±125 16.9 +0.9/-0.8 Radiocarbon Palaeoseismology III

Radiocarbon, historical Güzelköy 27.27 40.73 16.5±1.5 last 3 events 17±5 - earthquakes 5 Palaeoseismology 10.5±0.5 564±5.5 18.6±0.9 Radiocarbon, historical III Güzelköyc 27.27 40.73 16.5±1.5 1000±120 16.5 +2.7/-2.3 earthquakes III 6 Western High 27.77 40.82 7700±300 448000±41000 17.4±2.3 Seismic Stratigraphy III 7 Gemlik Bay 29.03 40.42 60±5 30000 2 Sea Level - 8 Gemlik Bay 29.03 40.42 42±4 11250±60 3.7±0.7 Radiocarbon Morphochronology - Çınarcık Basin 29.21 40.74 8000 in 430000 18.5 Seismic Stratigraphy -

9

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 54 of 59

29.16 40.74 2800±320 180000±14000 15.5 +2.3/-2.1 III 9 29.14 40.74 4200±410 300000±18000 14.0 +1.7/-1.6 III Çınarcık Basind Seismic Stratigraphy 29.11 40.74 6300±430 400000±24000 15.7 +1.5/-1.4 III 29.08 40.76 9000±490 510000±31000 17.6 +1.5/-1.4 III

10 İzmit Gulf 29.42 40.73 100±10 10500±500 9.3±1.9 Sea Level, Radiocarbon Morphochronology -

11 Nehirkent 30.41 40.71 18.5±0.5 860±110 21.9±3 OSL Palaeoseismology III 12a Amcahasanbey 31.15 40.77 105±25 6720±170 15.7±4 Radiocarbon - 12b Beykent 300±20 21700±1850 14±2.1 OSL Morphochronology - 31.18 40.77 12c Beykent 890±110 60170±6280 15.2±3.5 Radiocarbon -

13 Taşkesti 31.02 40.58 54±11 5800±100 9.2±1.9 OSL, Radiocarbon, U/Th Morphochronology -

14a Geredee 32.32 40.82 180±25 11000±1000 16.3 +2.9/-2.6 Palaeoclimate II Demirtepe 17 Radiocarbon - 15 32.33 40.82 Palaeoseismology Demirtepef 15.5±1.5 Draft909 16.7±1.9 Radiocarbon II 16 Alıç 33.51 40.98 12.5±2.5 Radiocarbon Palaeoseismology - 14b Eksike 33.66 41.02 200±20 11000±1000 18.2 +2.6/-2.3 Palaeoclimate II 17a Eksik 33.67 41.02 46±10 2250±150 20.5±5.5 Cosmogenic (Cl-36) II Morphochronology 17b Eksik 33.67 41.02 46±10 2435±485 20.5±8.5 Radiocarbon - 14c Berçine 33.89 41.07 200±20 11000±1000 18.5±3.5 Palaeoclimate II 14d Üçoluk 33.92 41.07 34±3.5 1640±40 21±2 Radiocarbon Morphochronology II 18a Üçoluk 33.93 41.07 29±9 1580 +115/-240 18.8 +6.6/-6.0 OSL Morphochronology II 18b Karapürçek 34.59 41.12 82±13 4385 +445/-505 18.8 +4.0/-3.4 OSL Morphochronology II 19 Tahtaköprü 35.05 41.10 55±10 3030±150 18.6 +3.5/-3.3 Cosmogenic (Be-10) Morphochronology II 14e Destek 36.3 40.84 200±50 11000±1000 18.1 +5.0/-4.7 Palaeoclimate Morphochronology II 14f Koçyatağı (Mihar) 39.21 39.91 185±35 11000±1000 16.8 +3.7/-3.4 Palaeoclimate Morphochronology I 18c Koçyatağı (Mihar) 39.19 39.91 29±8 1630±150 17.7 +5.4/-5.1 OSL Morphochronology I 18d Ayanoğlu 40.42 39.47 65±7 3275±310 20.2 +3.2/-2.8 OSL Morphochronology I 20a Dinarbey 40.71 39.38 57±10 5700±1300 10.0 +4.0/-2.7 OSL - Morphochronology 20b Dinarbey 40.71 39.38 27±9 2300±400 12.0 +5.4/-4.3 OSL - 20c Kaynarpınar 40.76 39.39 25±7 2000±600 12.6 +8.4/-4.8 OSL - Morphochronology 20d Kaynarpınar 40.76 39.39 36±5 2400±500 15.5 +4.6/-3.3 OSL -

https://mc06.manuscriptcentral.com/cjes-pubs Page 55 of 59 Canadian Journal of Earth Sciences

aThe slip-rate was calculated with the offset measurement and ages of last two events. The total strain was assumed to accumulate between AD 1344 (or 1354) and 1912. bI added 10% uncertanity to both offset and age measurements. cI chose two incremental offset measurements and used ages of events predating these displaced features. dIncremental displacements of the basin dypocentre were remeasured using the supplementary data of Kurt et al. (2013). In addition, age uncertainties of Grall et al. (2013) were considered for each seismic horizon. eMeasurements with their uncertainities were spatially grouped and recalculated for each individual location. fThe total slip of 15.5±1.5 for the last 3 events was accepted to occur between AD 1035 and 1944.

Draft

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 56 of 59

Reference

Gasperini et al. (2011)

Rockwell et al. (2009)

Draft Aksoy (2009); Aksoy et al. (2010)

Meghraoui et al. (2012)

Grall et al. (2013) Vardar et al. (2014) Gasperini et al. (2011)

Kurt et al. (2013)

https://mc06.manuscriptcentral.com/cjes-pubs Page 57 of 59 Canadian Journal of Earth Sciences

Kurt et al. (2013)

Polonia et al. (2004); Gasperini et al. (2011) Dikbaş et al. (2018)

Pucci et al. (2008)

Erturaç et al. (2019)

Hubert-Ferrari et al. (2002) Kondo et al. (2010) Kondo et al. (2010) Draft Sugai et al. (1999) Hubert-Ferrari et al. (2002)

Kozacı et al. (2007)

Hubert-Ferrari et al. (2002)

Zabcı (2012) Zabcı (2012) Kozacı et al. (2009) Hubert-Ferrari et al. (2002) Hubert-Ferrari et al. (2002) Zabcı (2012) Zabcı (2012)

Zabcı et al. (2015)

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 58 of 59

aThe slip-rate was calculated with the offset measurement and ages of last two events. The total strain was assumed to accumulate between AD 1344 (or 1354) and 1912. bI added 10% uncertanity to both offset and age measurements. cI chose two incremental offset measurements and used ages of events predating these displaced features. dIncremental displacements of the basin dypocentre were remeasured using the supplementary data of Kurt et al. (2013). In addition, age uncertainties of Grall et al. (2013) were considered for each seismic horizon. eMeasurements with their uncertainities were spatially grouped and recalculated for each individual location. fThe total slip of 15.5±1.5 for the last 3 events was accepted to occur between AD 1035 and 1944.

Draft

https://mc06.manuscriptcentral.com/cjes-pubs Page 59 of 59 Canadian Journal of Earth Sciences

Offset (m) Age (a) Site Reference min max min max Model I: Between Niksar and Yedisu Koçyatağı (Mihar) 21 37 1480 1780 Zabcı (2012) Ayanoğlu 47 67 2965 3585 Zabcı (2012) Koçyatağı (Mihar) 150 220 10000 12000 Hubert-Ferrari et al. (2002)

Model II: Between Bolu and Niksar Demirtepe 14 17 909 909 Kondo et al. (2010) Üçoluk 22 34 1340 1695 Zabcı (2012) Üçoluk 30 38 1600 1680 Hubert-Ferrari et al. (2002) Eksik 36 56 2100 2400 Kozacı et al. (2007) Tahtaköprü 45 65 2880 3180 Kozacı et a. (2009) Karapürçek 69 95 3880 4830 Zabcı (2012) Gerede 155 205 10000 12000 Hubert-Ferrari et al. (2002) Berçin 180 220 10000 12000 Hubert-Ferrari et al. (2002) Destek 150 250 10000 12000 Hubert-Ferrari et al. (2002)

Model III: Between Akyazı and Saros Kavakköy 8 10 558 568DraftRockwell et al. (2009) Güzelköy 10 11 558 568 Meghraoui et al. (2012) Güzelköy 15 18 880 1120 Meghraoui et al. (2012) Nehirkent 18 19 750 970 Dikbaş et al. (2018) Yeniköy 45 47 2590 2840 Aksoy (2009); Aksoy et al. (2010) Ganos 60 80 3600 4400 Aksoy (2009); Aksoy et al. (2010) Ganos 163 203 9200 11200 Aksoy (2009); Aksoy et al. (2010) Ganos 182 222 11250 13750 Aksoy (2009); Aksoy et al. (2010) Ganos 230 280 13050 15950 Aksoy (2009); Aksoy et al. (2010) Ganos 285 355 15750 19250 Aksoy (2009); Aksoy et al. (2010) Çınarcık Basin 2480 3120 166000 194000 Kurt et al. (2013) Çınarcık Basin 3790 4610 282000 318000 Kurt et al. (2013) Çınarcık Basin 5870 6730 376000 424000 Kurt et al. (2013) Western High 7400 8000 407000 489000 Grall et al. (2013) Çınarcık Basin 8510 9490 479000 542000 Kurt et al. (2013)

https://mc06.manuscriptcentral.com/cjes-pubs Canadian Journal of Earth Sciences Page 60 of 59

68.27% interval 95.45% interval Model Fault Section Time interval (ka) Median slip- rate (mm/a) (+) (-) (+) (-) Yedisu-Niksar (40.5-37°E) 4 ka to present 17.5 2.3 2 4 3.5 I Yedisu-Niksar (40.5-37°E) 12 ka to present 16.8 2.5 2.3 4.3 4 Niksar-Bolu (37-31.5°E) 5 ka to present 18.9 2 1.8 3.7 3.3 II Niksar-Bolu (37-31.5°E) 12 ka to present 17.4 1.4 1.3 2.7 2.4 Akyazı-Saros (30.5-26°E) 3 ka to present 16.9 0.6 0.6 1.2 1.1 Akyazı-Saros (30.5-26°E) 20 ka to present 17.7 1 0.9 1.9 1.7 Akyazı-Saros (30.5-26°E) 550 ka to present 16.7 0.5 0.5 1 0.9 III Akyazı-Saros (30.5-26°E) 195 to present 15.8 1.3 1.4 2.4 2.3 Akyazı-Saros (30.5-26°E) 320 to 195 ka 13.2 1.7 1.6 3.1 2.9 Akyazı-Saros (30.5-26°E) 540 to 320 ka 22.9 1.5 1.3 3.1 2.6

Draft

https://mc06.manuscriptcentral.com/cjes-pubs