ArticlePDF Available

Source modelling and strong ground motion simulations for the 24 January 2020, Mw 6.8 Elazig earthquake, Turkey

Authors:

Abstract and Figures

On 24 January 2020 an Mw 6.8 earthquake occurred at 20:55 local time (17:55 UTC) in eastern Turkey, close to the town of Sivrice in the Elazığ province, causing widespread considerable seismic damage in buildings. In this study, we analyse the main features of the rupture process and the seismic ground shaking during the Elazığ earthquake. We first use Interferometric Synthetic Aperture Radar (InSAR) interferograms (Sentinel-1 satellites) to constrain the fault geometry and the coseismic slip distribution of the causative fault segment. Then, we utilize this information to analyse the ground motion characteristics of the mainshock in terms of peak ground acceleration (PGA), peak ground velocity (PGV) and spectral accelerations. The absence of seismic registrations in near-field for this earthquake imposes major constraints on the computation of seismic ground motion estimations in the study area. To do this, we have used a stochastic finite-fault simulation method to generate high-frequency ground motions synthetics for the Mw 6.8 Elazığ 2020 earthquake. Finally, we evaluate the potential state of stress of the unruptured portions of the causative fault segment as well as of adjacent segments, using the Coulomb stress failure function variations. Modelling of geodetic data shows that the 2020 Elazığ earthquake ruptured two major slip patches (for a total length of about 40 km) located along the Pütürge segment of the well-known left-lateral strike-slip East Anatolian Fault Zone (EAFZ), with up to 2.3 m of slip and an estimated geodetic moment of 1.70 × 1019 Nm (equivalent to a Mw 6.8). The position of the hypocenter supports the evidence of marked WSW rupture directivity during the mainshock. In terms of ground motion characteristics, we observe that the high-frequency stochastic ground motion simulations have a good capability to reproduce the source complexity and capture the ground motion attenuation decay as a function of distance, up to the 200 km. We also demonstrate that the design spectra corresponding to 475 years return period, provided by the new Turkish building code is not exceeded by the simulated seismograms in the epicentral area where there are no strong motion stations and no recordings available. Finally, based on the Coulomb stress distribution computation, we find that the Elazığ mainshock increased the stress level of the westernmost part of the Pütürge fault and of the adjacent Palu segment and as a result of an off-fault lobe.
Content may be subject to copyright.
Geophys. J. Int. (2020) 223, 1054–1068 doi: 10.1093/gji/ggaa350
Advance Access publication 2020 July 21
GJI Seismology
Source modelling and strong ground motion simulations for the 24
January 2020, Mw6.8 Elazı˘
g earthquake, Turkey
Daniele Cheloni and Aybige Akinci
Istituto Nazionale di Geofisica e Vulcanologia, Via di Vigna Murata 605,00143 Roma, Italy. E-mail: daniele.cheloni@ingv.it
Accepted 2020 July 18. Received 2020 June 23; in original form 2020 May 4
SUMMARY
On 24 January 2020 an Mw6.8 earthquake occurred at 20:55 local time (17:55 UTC) in eastern
Turkey, close to the town of Sivrice in the Elazı˘
g province, causing widespread considerable
seismic damage in buildings. In this study, we analyse the main features of the rupture process
and the seismic ground shaking during the Elazı˘
g earthquake. We first use Interferometric
Synthetic Aperture Radar (InSAR) interferograms (Sentinel-1 satellites) to constrain the fault
geometry and the coseismic slip distribution of the causative fault segment. Then, we utilize
this information to analyse the ground motion characteristics of the main shock in terms of
peak ground acceleration (PGA), peak ground velocity (PGV) and spectral accelerations. The
absence of seismic registrations in near-field for this earthquake imposes major constraints on
the computation of seismic ground motion estimations in the study area. To do this, we have
used a stochastic finite-fault simulation method to generate high-frequency ground motions
synthetics for the Mw6.8 Elazı˘
g 2020 earthquake. Finally, we evaluate the potential state of
stress of the unruptured portions of the causative fault segment as well as of adjacent segments,
using the Coulomb stress failure function variations. Modelling of geodetic data shows that
the 2020 Elazı˘
g earthquake ruptured two major slip patches (for a total length of about 40 km)
located along the P ¨
ut¨
urge segment of the well-known left-lateral strike-slip East Anatolian Fault
Zone (EAFZ), with up to 2.3 m of slip and an estimated geodetic moment of 1.70 ×1019
Nm (equivalent to a Mw6.8). The position of the hypocentre supports the evidence of marked
WSW rupture directivity during the main shock. In terms of ground motion characteristics, we
observe that the high-frequency stochastic ground motion simulations have a good capability to
reproduce the source complexity and capture the ground motion attenuation decay as a function
of distance, up to the 200 km. We also demonstrate that the design spectra corresponding to
475 yr return period, provided by the new Turkish building code is not exceeded by the
simulated seismograms in the epicentral area where there are no strong motion stations and no
recordings available. Finally, based on the Coulomb stress distribution computation, we find
that the Elazı˘
g main shock increased the stress level of the westernmost part of the P¨
ut¨
urge
fault and of the adjacent Palu segment and as a result of an off-fault lobe.
Key words: Radar interferometry; Europe; Numerical modelling; Earthquake ground mo-
tions; Earthquake hazards; Earthquake source observations; Seismic attenuation.
1 INTRODUCTION
An earthquake of Mw6.8 occurred in the Elazı ˘
gregionofeastern
Turkey on 24 January 2020 at 20:55 local time (17:55 UTC), causing
loss of life and severe damage in the epicentral area. According to
the information provided by the Earthquake Department of the Dis-
aster and Emergency Management Presidency, AFAD, there were
46 reported fatalities and over 1600 injuries in Elazı˘
g, Malatya
and Diyarbakir. There are an estimated 10000 people homeless at
this time, sheltering in containers, tents and public refuge sites in
schools, sports facilities and dorms. The earthquake was reported
to be on a segment of the 580-km-long left-lateral continental
strike-slip East Anatolian Fault Zone (EAFZ), which is one of the
two major active strike-slip fault systems in Turkey, other being
the 1500-km-long right-lateral strike-slip North Anatolian Fault
Zone (NAFZ, inset Fig. 1). The earthquake epicentre is provided by
the different national and international Institutes (including AFAD,
KOERI, USGS, INGV, GCMT, CPPT, ERD, IPGP, GFZ and EMSC)
with rather different locations (in this work we use both the hypocen-
tre information of the main shock based on solution from AFAD and
from the Kandilli Observatory and Earthquake Research Institute,
KOERI, respectively); however, the magnitude has been assessed
1054 C
The Author(s) 2020. Published by Oxford University Press on behalf of The Royal Astronomical Society.
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1055
Figure 1. Seismotectonic framework of the study area. The solid lines represent the main fault segments of the East Anatolian Fault Zone (EAFZ) with labelled
name: (1) Amanos, (2) Pazarcik, (3) Erkenek, (4) P¨
ut¨
urge, (5) Palu and (6) Ilica fault segments, respectively (after Duman & Emre 2013); the red line is the
fault segment activated during the 2020 Elazı˘
g seismic sequence; the dashed lines are the related fault jogs. Seismicity: the green circles are the aftershocks
of the first 3 months after the main shock (available at https://deprem.afad.gov.tr); the red and white star represent the location of the main shock provided by
AFAD and KOERI, respectively, and its moment tensor solution (U.S. Geological Survey 2020); the grey and yellow stars are the location of the source of the
major (M>6.6) historical and instrumental earthquakes, respectively, in the region with labelled events date (Ambraseys 1989; Ambraseys & Jackson 1998;
Kondorskaya & Ulomov 1999) and their moment tensor solutions (Taymaz et al. 1991). The white arrows represent the direction of plate motion. The dashed
box highlights the area of Fig. 4. The main map is displayed in an oblique Mercator projection with the equator azimuth parallel to the trend of the EAFZ.
The inset shows the main fault systems in and around Turkey (modified from Cetin et al. 2003): NAFZ is the North Anatolian Fault Zone; EAFZ is the East
Anatolian Fault Zone. Arrows indicate relative plate motions. The dashed box is the area of the main figure.
as Mw6.7 or 6.8. Its epicentre was located in the Elazı˘
g province,
at a distance of about 10–20 km WSW of the Lake Hazar (Fig. 1).
Its focal depth ranged from 8 to 23 km. The focal mechanism so-
lution (Fig. 1) indicated that the earthquake is in agreement with
the activation of a ENE–WSW striking left-lateral strike-slip fault.
The affected area is still experiencing repeated aftershocks, with
over 1200 events of magnitude greater than 2.0 within 4 months; 31
aftershocks have been equal to or larger than M4.0. The largestafter-
shocks occurred on 25 January, 19 March and 5 June 2020, and their
magnitudes have been assessed as Mw5.1, 5.0 and 5.0, respectively
(earthquakes solutions available at https://deprem.afad.gov.tr). Af-
tershocks spread along a 50–60 km stretch of the EAFZ between
Sivrice to P¨
ut¨
urge, both ENE and WSW of the hypocentre, spread-
ing outward from the 30 km section of fault that ruptured on 24
January 2020 (Fig. 1).
In this study, great effort has been directed towards understand-
ing the characteristics of source and ground motion associated with
the Elazı˘
g seismic sequence. In this respect, we first use Sentinel-
1 Synthetic Aperture Radar (SAR) data to investigate the ground
displacement field and to infer, by using elastic dislocation mod-
elling, the fault geometry and slip distribution of the causative fault
segment. Then, we performed simulations aiming at reproducing
the high-frequency portion of the strong ground motion records
obtained during 2020 Elazı˘
g earthquake using the stochastic finite-
fault simulation method based on a dynamic corner frequency ap-
proach (Motazedian & Atkinson 2005; Boore 2009). The earth-
quake’s complex nature and the sparse strong motion network in
the epicentral area makes difficult to define the characteristics of
ground shaking particularly in the near-source region. The estima-
tion of ground motions becomes therefore necessary and essential
through the earthquake scenarios and several physics-based deter-
ministic, stochastic and hybrid methods (Boore 2003; Motazedian
& Atkinson 2005; Graves & Pitarka 2010;Maiet al. 2010; Irikura
& Miyake 2011;Menaet al. 2012; Akinci et al. 2017; Pitarka et al.
2017,2019).
The strong ground motion records of Elazı˘
g earthquake were
available and provided recently by AFAD that manage the national
strong motion network in Turkey. The ground motion parameters
in terms of PGA and PGV values and recorded seismograms at 43
strong motion stations within a distance of 200 km from the epi-
centre are used in this study (Table 1). The observed peak ground
accelerations (PGA) were 293 and 239 cm s–2, with the highest
maxima recorded by the two strong motion stations closest to the
fault (within 10 km of the rupture). These stations are located
near the epicentre (Sivrice and P¨
ut¨
urge stations, located 3km
to the north and 6 km to the southeast of the epicentre, respec-
tively), as well as locations that were in the heavily damaged area.
Therefore, the simulated ground motions calculated at the 1066 vir-
tual stations distributed on a regular grid with 5 km spacing are
validated with the observed ground motion parameters in terms
of PGA and PGV values, and then compared with ground mo-
tion prediction equations (GMPEs) for epicentral distances up to
200 km.
Finally, because the stress changes in the region due to this
earthquake may interact with other fault segments of the EAFZ
in the area, in this study, we investigate the static Coulomb
stress changes induced by the main shock, by examining both
the contribution to the observed aftershock triggering during
the 2020 Elazı˘
g sequence and to the adjacent fault segments
loading.
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1056 D. Cheloni and A. Akinci
Tab le 1. Peak ground accelerations and velocities from the processed recordings of the 2020 Elazı˘
g earthquake, Turkey, on the strong motion stations up to
150 km distances and local site conditions is defined in Eurocode 8 (EC 8: Seismic Design of Buildings).
Station
code Station name
Station
Lat.
Station
Lon.
PGA-NS
(gal)
PGA-EW
(gal)
PGA-Z
(gal)
PGV-NS
(ms’1)
PGV-EW
(ms’1)
Distance
Rjb
VS30 (ms1) Site
EC8
2308 Sivrice 38.45 39.31 235.78 292.80 178.57 27.81 45.34 2.63 450-B
4404 P¨
ut¨
urge 38.20 38.87 193.59 228.44 110.62 24.82 28.71 6.16 1380-A
2301 Merkez 38.67 39.19 118.92 142.61 66.24 12.35 8.75 21.95 407-B
0204 Gerger 38.03 39.03 94.24 110.40 59.83 17.10 9.72 29.09 555-B
0212 Sincik 38.03 38.62 43.52 38.52 31.83 7.13 4.80 28.64 -
2302 Maden 38.39 39.68 25.55 31.36 22.77 3.61 2.29 33.92 907-A
2104 Ergani 38.26 39.76 26.75 25.61 24.11 3.12 3.87 45.17 -
4401 Merkez 38.35 38.34 73.23 87.63 37.35 7.74 6.91 39.96 481-B
0205 Kahta 37.79 38.62 25.49 41.01 26.02 6.81 7.38 52.34 660-B
0207 C¸ elikhan 38.03 38.25 32.94 30.50 18.18 5.06 4.27 63.36 660-B
2304 Kovancılar 38.72 39.86 8.823 13.74 5.81 2.05 2.13 67.57 489.B
4412 Yazıhan 38.60 38.18 21.46 18.97 14.44 4.41 4.43 62.90 -
4407 Arguvan 38.78 38.26 31.04 23.27 20.13 3.00 4.55 70.09 735-B
2307 Palu 38.70 39.93 12.81 20.82 11.88 3.74 3.00 67.57 329-C
2105 Dicle 38.36 40.07 10.37 11.09 8.91 1.23 1.70 68.69 -
6201 Merkez 39.07 39.53 11.99 9.70 9.91 1.26 1.21 70.39 -
0210 Merkez 37.77 38.29 23.04 27.43 18.04 8.27 5.48 70.09 -
4406 Akcada˘
g 38.34 37.97 23.60 24.02 14.06 2.07 4.14 71.16 815-A
0201 Merkez 37.76 38.27 35.9 44.61 35.14 10.08 8.10 71.15 391-B
0209 Samsat 37.58 38.48 70.84 58.44 24.71 4.45 4.80 77.42 -
2305 Beyhan 38.73 40.13 3.64 4.78 4.00 0.99 1.21 80.54 907-A
4408 Do˘
gans¸ehir 38.10 37.89 11.35 16.24 15.55 3.00 2.86 81.16 654-B
2306 Karakocan 38.96 40.04 4.356 5.41 2.89 1.12 1.52 86.63 663-B
4405 Hekimhan 38.81 37.94 11.56 11.59 6.67 4.44 1.68 94.12 579-B
2101 Ba˘
glar 37.93 40.20 24.64 26.38 13.64 2.23 2.75 97.07 519-B
2409 Kemaliye 39.28 38.49 13.71 20.43 7.04 2.79 1.73 109.83 875-A
0213 Tut 37.80 37.93 34.95 30.86 14.43 6.04 6.86 91.63 -
6304 Bozova 37.37 38.51 49.02 77.71 29.94 3.60 3.17 100.24 376-B
2415 ˙
Ilic¸ 39.46 38.55 11.30 12.26 8.99 3.13 1.92 124.97 444-B
4410 Kuluncak 38.87 37.68 13.30 15.15 6.64 2.69 1.51 116.20 -
2408 Kemah 39.60 39.03 12.08 16.46 12.08 2.69 2.17 126.10 416-B
2106 Lice 38.46 40.65 8.71 10.57 8.71 1.08 1.30 118.55 -
4409 Darende 38.56 37.49 7.91 6.95 6.68 1.03 1.24 117.11 -
0208 Golbasi 37.79 37.65 18.06 12.42 7.62 3.38 4.31 113.26 469-B
6302 Virans¸ehır 37.23 39.75 16.93 14.05 9.39 1.74 1.81 136.26 936-A
6202 P¨
ul¨
um¨
ur 39.49 39.90 8.26 6.54 3.66 1.46 0.99 125.04 -
2412 C¸a˘
glayan 39.59 39.69 2.56 2.43 1.70 0.52 0.74 129.21 955-A
1213 Adakli 39.23 40.48 8.26 6.54 3.66 2.08 1.14 149.50 -
2 TECTONIC SETTING,
SEISMOTECTONIC AND SEISMICITY
Two major faults contribute to the majority of the seismic hazard in
the region, the North Anatolian Fault Zone (NAFZ) and the Eastern
Anatolian Fault Zone (EAFZ, Fig. 1). Many large earthquakes have
occurred along these faults (Ambraseys 1989) as a result of the
ongoing movement between the Eurasian, African, Arabian and
Anatolian plates. The NAFZ with right-lateral faulting is extending
from Istanbul in the west to Karlıova in the east. During the twentieth
century this fault zone has produced many large earthquakes with
surface rupturing and with a westward migrating sequence (Barka
& Kandinsky-Cade 1988). Around the Karlıova region, NAFZ joins
the SW-trending EAFZ.
The EAFZ forms a 580 km left-lateral strike-slip transform
boundary between the northward moving Arabian Plate and west-
ward moving Anatolian block (Fig. 1), resulting in a left-lateral
slip rate of 10 ±1mmyr
–1 on the EAFZ (e.g. McClusky et al.
2000; Reilinger et al. 2006; Aktug et al. 2016), but its faulting
is less continuous and less localized than that of the NAFZ (Am-
braseys 2009). In fact, the EAFZ constitutes a complex left-lateral
strike-slip fault zone and it is divided into a number of fault seg-
ments (Fig. 1), as suggested by different authors based on variations
in trends, location of geometric discontinuities, extent of surface
ruptures, stepovers and bend structures along the EAFZ, mapping
of active faults, seismological and palaeoseismological data (e.g.
Hempton et al. 1981; Barka & Kadinsky-Cade 1988;Sarogluet al.
1992; Herece 2008; Duman & Emre 2013). According to Duman &
Emre (2013), from SW to NE, the fault segments of the main EAFZ
fault strand are: the Amanos, Pazarcık, Erkenek, P¨
ut¨
urge, Palu, Ilıca
and Karlıova segments (Fig. 1). The length of these fault segments
varies from 31 to 112 km, while their strikes vary from N35Eto
N75E (Duman & Emre 2013).
Since this zone is tectonically very active, a series of large sur-
face rupturing earthquakes occurred along the EAFZ main strand
during the last centuries (Fig. 1). From NE to SW the following
large earthquakes have occurred: the 1866 earthquake of Ms7.0
that can be correlated to the Karlıova segment (Ambraseys & Jack-
son 1998); the 1971 Bing¨
ol earthquake of Ms6.8 occurred between
Karlıova and Bing¨
ol (McKenzie 1972); the 1874 earthquake of Ms
7.1 occurred on the Palu segment (Ambraseys & Jackson 1998;
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1057
Figure 2. (a) Data, (b) model and (c) residual sampled points from the
unwrapped interferogram showing the coseismic displacement field from
the Sentinel-1 ascending track 116. The pink star indicates the main shock
epicentre provided by AFAD.
Figure 3. (a) Data, (b) model and (c) residual sampled points from the
unwrapped interferogram showing the coseismic displacement field from
the Sentinel-1 descending track 123. The pink star indicates the main shock
epicentre provided by AFAD.
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1058 D. Cheloni and A. Akinci
Figure 4. (a) Coseismic slip distribution on the P¨
ut¨
urge segment of the EAFZ. The contouring (in cm) indicates the major retrieved coseismic patches of slip.
The pink and white stars are the AFAD and KOERI locations of the main shock, respectively; the white blue stars are the major aftershocks with M>5. The
red beach ball indicates the mechanisms of the main shock, while the green beach balls are the M>4 aftershocks that occurred in the first 3 months from
the main shock. Green circles are aftershocks between 23 January and 23 April. The red box represents our best-fitting uniform slip solution. (b) Estimated
slip distribution as a function of depth (symbols as in the top panel). The white blue and green stars are, respectively, the major aftershocks with M>5and
M>4. Note that many hypocentres were located at a fixed depth of 7 km from the automatic AFAD location.
Cetin et al. 2003); the 1875 (Ms6.7) that might have been gen-
erated along the easternmost termination of the P ¨
ut¨
urge segment
(Ambraseys 1989; Cetin et al. 2003) or within the complex re-
leasing bend geometry that is inferred to exist within Lake Hazar
(Duman & Emre 2013); the 1905 (Ms6.8) earthquake occurred on
the Yarpuzlu restraining bend which is located at the western tip of
the P¨
ut¨
urge segment (Ambraseys 1989); the 1893 earthquake of Ms
7.2 occurred on the Erkenek segment (Ambraseys & Jackson 1998);
the 1513 earthquake of Ms7.4 has been attributed to the Pazarcık
segment (Herece 2008); the 1822 earthquake of M7.5 that might
have been generated on the Amonos fault segment (Ambraseys &
Jackson 1998;Seyreket al. 2007).
Prior to the 24 January 2020 Elazı˘
g earthquake sequence, taking
account of the time elapsed from the last event, the slip rate, seis-
mological and palaeoseismological data, some authors have iden-
tified some important seismic gaps along the main strand of the
EAFZ, the 80-km-long Pazarcık (Nalbant et al. 2002; Karaba-
cak et al. 2011), the 100-km-long Amanos and the 95-km-
long P¨
ut¨
urge segments (Duman & Emre 2013; Aktug et al. 2016,
Fig. 1). According to these authors these fault segments have
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1059
Figure 5. Slip distribution model on the P¨
ut¨
urge fault segment estimated in this study which is used in stochastic ground motion simulations. The pink and
white stars represent the location of the main shock provided by AFAD and KOERI, respectively; the white blue and green stars are, respectively, the major
aftershocks with M>5andM>4 projected on the fault surface (within 3 km). The grey arrows indicate the estimated slip directions for each subfault.
Tab le 2. Model parameters related to source, path and site terms that are considered for the ground motion simulations for the Mw6.8
Elazı˘
g earthquake.
Parameters Values Reference
Fault (strike and dip) 242–75This study
Fault dimension 51 ×24 km2This study
Moment magnitude 6.8 This study
Depth of the top of fault plane 0.0 km This study
Subfault dimension 1.5 ×1.5 km2This study
Stress drop 9 This study
Crustal shear wave velocity (β)3.5kms
–1 G¨
ok et al. (2007)
Crustal density 2800 kg m–3
Rupture velocity 0.8×β
Pulsing area percentage 50 per cent Boore (2009)
Kappa parameter 0.035 and 0.04 s Boore & Joyner (1997)
NEHRP generic rock and generic soil site BSSC 2001
Distance dependent of duration 0.0 (0–10 km) Atkinson & Boore (1995)
0.1 (R>10 km)
Attenuation model, Q(f) 100f0.43 Akinci et al. (2014)
Geometrical spreading Coef. r1.0 r100 km Akinci & Antonioli (2013)
r0.5 r100 km
Window function Saragoni Hart Boore (1983)
Local amplification NEHRP sites Boore & Joyner (1997)
therefore the potential to produce destructive earthquakes in the
future.
3 SENTINEL-1 SAR DATA
We use SAR data acquired by the Sentinel-1 satellites in TOPS (Ter-
rain Observation by Progressive Scans) mode, exploiting two as-
cending and two descending interferograms to measure the ground
displacement due to the 24 January, 2020, Mw6.8 Elazı˘
g earthquake.
The epicentral area is in fact covered by four Sentinel-1 tracks: the
ascending tracks 116 and 043 (Figs 2, S1 and S2), and the descend-
ing tracks 123 and 021 (Figs 3, S3 and S4). The ascending track
116 and the descending track 021 had the last pre-earthquake image
acquisitions on 21 January, while the ascending track 043 and the
descending track 123 on 22 January. Subsequent acquisitions were
made at 6-d interval. To reduce the contribution from potential post-
seismic deformation, we form the coseismic interferograms using
images acquired closest to the earthquake (the first post-seismic ac-
quisitions along the ascending track 116 and the descending track
021 were on 27 January; while for the other tracks were on 28
January).
We process the data using the Sentinel Application Platform
SNAP software that is provided freely by the European Space
Agency. We use precise orbit data and Shutter Radar Topography
Mission 1 arcsec data (Jarvis et al. 2008) for the image removal
of flat-earth phase and topographic phase. The resulting interfero-
grams were then filtered by applying the Goldstein filter (Goldstein
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1060 D. Cheloni and A. Akinci
Figure 6. Residuals between observed and simulated PGA and PGVs calcu-
lated for sixteen drop parameters, ranging from 5 to 20 MPa, that minimize
the misfit between observed and simulated ground motion.
0.0
0.1
0.2
0.3
0.4
0.5
P
G
A
g
0.0
0.1
0.2
0.3
0.4
0.5
P
G
V
m
/
s
38˚00'
38˚00'
38˚30'
38˚30'
39˚00'
39˚00'
39˚30'
39˚30'
40˚00'
40˚00'
38˚00' 38˚00'
38˚30' 38˚30'
39˚00' 39˚00'
0.1
0.1
0.2
0.3
0.4
0.5
38˚00'
38˚00'
38˚30'
38˚30'
39˚00'
39˚00'
39˚30'
39˚30'
40˚00'
40˚00'
38˚00' 38˚00'
38˚30' 38˚30'
39˚00' 39˚00'
0 50
km
2308
4404
0204
0212
2302
4401
0205
0207
4212
2304
0210
0201
230
4
2301
2104
Malatya
Elazig
38˚00'
38˚00'
38˚30'
38˚30'
39˚00'
39˚00'
39˚30'
39˚30'
40˚00'
40˚00'
38˚00' 38˚00'
38˚30' 38˚30'
39˚00' 39˚00'
0.1
0.1
0.1
0.2
0.3
0.4
0.5
38˚00'
38˚00'
38˚30'
38˚30'
39˚00'
39˚00'
39˚30'
39˚30'
40˚00'
40˚00'
38˚00' 38˚00'
38˚30' 38˚30'
39˚00' 39˚00'
0 50
km
2308
4404
0204
0212
2302
4401
0205
0207
4212
2304
0210
0201
230
4
2301
2104
Malatya
Elazig
Figure 7. Spatial distributions of synthetics using the spectral parameters
giveninTable2in terms of (a) PGA and (b) PGV values, calculated at rock
site (BC type soil classification, VS30 =760 m s1), respectively. The region
is divided into regular grid spacing of 5 km as indicated by small-black dots
shown in figure so that the simulations are performed for the 1066 virtual
stations. The grey box represents our extended fault plane solution. The pink
star indicates the main shock epicentre provided by AFAD.
&Werner1998) to reduce interferometric phase noise. Finally, we
unwrapped the phase using Statistical-Cost Network-Flow Algo-
rithm for Phase Unwrapping (SNAPHU) (Chen & Zebker 2001)and
the interferograms were geocoded to obtain the ground deforma-
tion maps. Because of the intrinsic ambiguity of phase unwrapping,
similarly to Wang & Burgmann (2020), we flatten the unwrapped
interferograms by fitting a polynomial function to phase at pixels
away from the epicentre, where the expected ground deformation is
small. Unfortunately, there are no permanent GNSS (Global Nav-
igation Satellite System) stations in the near-field of the epicentre
for a direct comparison with InSAR Line-Of-Sight (LOS) displace-
ments. In fact, the nearest GNSS station is located about 35 km
from the epicentre, in Elazı˘
g city, and has a significant static off-
set of only 3 cm towards SW (available in an open file report
at https://deprem.afad.gov.tr/depremdokumanlari/1831). Neverthe-
less, the InSAR displacements are in good agreement with Elazı˘
g
GNSS measurements projected onto the LOS.
The ground deformation retrieved from the ascending and de-
scending unwrapped interferograms are characterized by two ENE–
WSW striking deformation lobes located on both sides of the EAFZ
alignment, with maximum LOS displacement of about 25–30 cm
(Figs 2and 3). The observed differences in the ascending and
descending LOS maps reveal a combination of mainly horizontal
movements consistently with the strike-slip left-lateral mechanisms
of the EAFZ.
4 FAULT GEOMETRY AND COSEISMIC
SLIP
In order to image the fault geometry and slip distribution of the
2020 Elazı˘
g main shock, we performed fault slip modelling using
rectangular dislocations embedded in an elastic, homogeneous and
isotropic half-space (Okada 1985), following a standard two-steps
procedure (e.g. Cheloni et al. 2010,2019): (1) we inverted the LOS
displacements to retrieve the fault geometry and then (2) the best-
fitting uniform-slip fault parameters are used as apriorifor the
estimation of the coseismic slip distribution. Before modelling, the
InSAR interferograms were down-sampled using a resolution-based
down-sampling scheme (Lohman & Simons 2005,FigsS1,S2,S3
and S4).
In the first step, we carried out a non-linear optimization of the
fault geometry by using a simulated annealing algorithm (Corana
et al. 1987). The best-fitting uniform slip model is described by
a 242ENE–WSW striking and 75NW dipping strike-slip (rake
about –7.5) 32.1 km ×8.5 km fault plane passing through the
hypocentral location (Fig. 4, red box) and in good agreement with
focal solutions. The average uniform slip is 1.3 m, which using a
value of 30 GPA for rigidity, yields an estimated seismic moment
of 1.04 ×1019 Nm, equivalent to a Mw6.7 earthquake (the results
of the uniform-slip model are displayed in Figs S5–S8).
In the second step, in order to estimate slip distribution on the
fault plane, we extended the uniform slip fault to capture the area
affected by aftershocks and subdivided the fault into small patches
of constant size (1.5 km ×1.5 km). We apply positivity constraints
and regularize the linear inversion by applying spatial smoothing
(Fig. S9). Additional terms consisting of a linear ramp for each
InSAR interferograms are also included in the inversion and relative
weights were applied to properly combine the different data sets
(Fig. S10).
The best-fitting slip distribution on the extended fault plane
(51 km ×24 km) agrees with the distribution of aftershocks (Figs 4
and 5). The fit to the data significantly improves passing from the
uniform slip to a variable slip model: from 2.75 to 1.64 cm and from
2.47 to 1.78 cm for the LOS displacements on ascending track 116
and descending track 123, respectively, and from 3.22 to 1.65 cm
and from 3.14 to 1.29 cm for the LOS displacements on ascending
track 043 and descending track 021, respectively (Figs 2,3,S11
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1061
Figure 8. Horizontal component acceleration, and velocity time history plots at three closest stations to the fault rupture. Recorded (left-hand panels, black
and red colour for two horizontal components EW and NS, respectively) and simulated (right panels) at the closest Sivrice, P¨
ut¨
urge, and Gerger sites to the
earthquake rupture for the 2020 Elazı˘
gMw6.8 earthquake.
and S12). The coseismic slip distribution model shows two major
asperities with peak slip of about 2–2.3 m, located WSW respect
to the epicentre, and mostly confined within the first 8 km, that
is roughly contained in the uniform slip fault (red box in Fig. 4),
for a length of about 40 km, thus leaving unbroken the WSW part
of the P¨
ut¨
urge fault segment. In addition, our variable slip model
shows some slip also to the ENE of the epicentre, in an area where a
number of M>4 earthquakes occurred. The resulting total seismic
moment (1.70 ×1019 Nm) agrees with an Mw6.8 earthquake. The
rake angles of the major coseismic fault patches are consistent with
a predominantly pure left-lateral strike-slip faulting mechanisms.
The two main asperities characterizing our slip model agree with
the up-dip rupture episodes and with the unilaterally WSW rupture
directivity, as retrieved by USGS finite fault analysis (U.S. Geolog-
ical Survey 2020) and by the recent study of Melgar et al. (2020).
Our retrieved slip located ENE of the epicentre, implies instead
bilateral rupture propagation.
The inversion was repeated using separately the ascending and
the descending tracks, respectively. The resulting coseismic slip
distribution is showed in Fig. S13. They are quite similar, showing
two major coseismic slip patches located WSW of the hypocentre,
suggesting these patches to be a robust feature of the retrieved slip
distribution. Notwithstanding, there are some differences, in fact,
the descending tracks-only inversion (Fig. S13b) appears to require
a more eastward position of the smaller patch of slip located ENE
of the epicentre as retrieved in the joint inversion.
5 STOCHASTIC MODELLING OF
HIGH-FREQUENCY GROUND MOTION
In this section, we attempt to simulate the high-frequency ground
motions for the Mw6.8 Elazı˘
g earthquake using a well-known
stochastic finite-fault simulation method (Motazedian & Atkinson
2005). To do so we gather several parameters related to the source,
the path and the site that are essential and requested by the adopted
approach have been chosen among those several existing models
published for the eastern Turkey.
5.1 Finite-fault source model
The necessary source parameters are defined in terms of the fault
geometry, rupture velocity, stress drop and seismic moment. The slip
distribution along the fault plane is another crucial input parameter
particularly for the ground motion simulation in the source area.
The slip model determined in this study is considered as an input
for our ground motion calculations. The fault plane geometry was
determined from the geodetic inversion with strike 242and dip
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1062 D. Cheloni and A. Akinci
Figure 9. Comparison of observed and simulated Fourier amplitude spectra of acceleration (cm s1) at six selected stations. Simulated spectra (solid blue
lines) and observed spectra for the two horizontal EW and NS components (solid red and black lines, respectively).
75and with the top of the fault plane at 0 km depth (Fig. 5). Two
patches of slip are examined over the fault plane: the larger one is
located at the centre and has the highest slip of about 2.3 m and the
other is located at southwestern part of the fault and has about 2 m
of slip. The dimensions of the modelled left-lateral strike-slip fault
plane are 50 km ×25 km and our rupture model includes 34 ×16
subfaults.
There have been several efforts to determine the fault rupture and
mechanism of the 2020 Elazı˘
g earthquake, established on various
data set (teleseismic, GPS and InSAR) and alternative inversion
methods (e.g. Melgar et al. 2020). Our inversion results agree with
these models defining the location of the slip asperities on the
fault plane and the quantitative description of the slip during the
earthquake rupture. The parameters of the finite-fault source model
used in our ground motion simulations are listed in Table 2.
5.2 High-frequency seismic wave attenuation model
The seismic wave propagation and the seismic attenuation is an
important topic and it is essential for the prediction of earthquake
ground motion in seismic hazard analysis. Determining the mech-
anism of attenuation helps to understand the regional differences
observed on the ground motion. There have been several studies
to determine the attenuation characteristics in eastern Turkey using
various database and methods (Mitchell et al. 1997; Mitchell &
Cong 1998;Zoret al. 2007; Pasyanos et al. 2009; Sertcelik 2012;
Akinci et al. 2014).
Recently Akinci et al. (2014) provided a complete description of
the characteristics of the source and the attenuation of the ground
motion around the Lake Van region (eastern Turkey) using a large
set of broadband data of the main shock and aftershocks of the 23
October 2011 Mw7.1 Van earthquake. They observed strong crustal
attenuation, Q(f)=100f0.43 together with the geometrical spreading,
g(r), occurring at a hypocentral distance of 40 km which it changes
from a body-wave-like function g (r)r1.0 to a functional form
g(r)r0.5 expected for surface waves. These results are not very
different from those observed by Zor et al. (2007)usingtheLgwaves
through the two stations method and back projection tomography;
QLg0 was around 100 with its frequency dependence n=0.4–0.6.
Sertcelik (2012) has also estimated the seismic wave attenuation
using the coda waves as a function of the lapse time and frequency
over the EAFZ. Although those studies result with a similar Qvalue
(100) over the Eastern Anatolia, in our study, we decide to use
most recently characterized seismic attenuation parameters as given
by Akinci et al. (2014)(Table2).
The geometrical spreading values governed by the crustal struc-
ture, capable of producing post-critical reflections from mid crustal
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1063
Figure 10. Comparisons between observed and simulated (a) peak ground
acceleration, PGA and (b) peak ground velocity, PGV. The simulated values
are obtained using a dynamic corner frequency approach. The PGA and
PGV simulated values were estimated for an earthquake of Mw6.8 using
the parameters given in Table 2and the proposed fault geometry and slip
distribution obtained in this study, for the 1066 virtual stations (light blue
triangles). The black crosses represent the observed data of Elazı˘
g earth-
quake recorded in two horizontal components (EW light grey and NS black
coloured crosses) at 43 stations while the red triangles represent those from
simulated seismograms at the same stations on a rock site, Vs30 =760 ms1.
Ground Motion Predictions Equations (GMPEs), obtained using Boore et al.
(2014) (short dashed line); Bindi et al. (2014) (long dashed line) and Akkar
& Cagnan (2010) (thick solid line) and the ±1σthe total standard deviation
of AC10 (thin solid lines) for a rock site of Vs30 =760 m s1are also shown.
and Moho discontinuities are also appropriated as given by Akinci
& Antonioli (2013). The stress drop parameter σ which governs
the levels of the acceleration spectrum at high frequencies is esti-
mated between 8 and 20 MPa by Akinci et al. (2006), Malagnini
et al. (2010), Akinci & Antonioli (2013) and Akinci et al. (2014)
for the 1999 Kocaeli, Mw7.2 earthquake with strike-slip faulting
and for the 2011 Van Lake Mw7.1 earthquake with reverse faulting.
Since the stress drop parameter is not estimated specifically for the
Mw6.8 Elazı˘
g earthquake, we calculate the residuals between the
observed PGA and PGV values and those from simulations over
16 different stress drop parameters, from 5 to 20 MPa. Finally, we
select the σ that minimizes the misfit between the observed and
simulated PGA and PGV data.
Results presented in Fig. 6demonstrate how the choice of the
stress drop parameter affects the misfit. The lowest bias determined
by averaging the residuals over the 43 stations indicates the best
parameter that could be considered for the simulations. The chosen
stress drop parameter for our ground motion estimations that pro-
duces a good fit to the observed data for the PGA and PGV is equal
to σ =9MPa.
6 HIGH FREQUANCY GROUND
MOTION SIMULATIONS FOR THE Mw
6.8 ELAZIG EARTHQUAKE
6.1 Spatial distribution of simulated ground motions
In order to investigate the spatial variation of ground motion caused
by the Mw6.8 Elazı˘
g earthquake, we calculate ground motion pa-
rameters together with the synthetic time histories at 43 recording
stations, and at 1066 virtual stations on a regular grid with 5 km
spacing, covering a regionbetween 38.25–39.5E and 38.0–38.75N
up to 200 km distances. The high frequency seismograms are gen-
erated using a stochastic finite-fault model approach, based on a
dynamic corner frequency (Motazedian & Atkinson 2005; Boore
2009) and considering the spectral parameters as given in Table 2.
Spatial distribution of the estimated ground motion parameters in
terms of PGA and PGV values within the study area is shown in
Figs 7(a) and (b), respectively.
The site-amplification effect is considered as uniform for the
whole region and referred to the BC type generic rock site condition
(Vs30 =760 m/s, the shear wave velocity averaged over the top 30 m
of the soil, BSSC 2001). So that, the spatial distributions of PGA
and PGV values mainly reflects the source effects. We observe that
the largest ground shaking is concentrated along the rupture fault
plane, where PGA and PGV values raising up to 0.5 g and 40 cm/s,
respectively. Particularly, we observe the strongest ground shaking
within the surface projection of the fault around the location of the
two slip asperities; one asperity is located at the centre of the fault
plane and is larger and much stronger than the second asperity that
is located in the southern section of the fault plane. Finally, while
the near-field results are governed by the source effects, such as the
distribution of asperities on the fault plane, intermediate distances
are controlled mainly by the propagation and the attenuation of
seismic waves with distance. However, since the stochastic approach
adopted in our study is not very sensitive to large-scale source
related directivity effects, the simulated ground motion may not be
suitable for producing the expected ground motion variability in the
near-fault region.
Some of the simulated synthetic waveforms together with the reg-
istered seismograms are shown in Fig. 8at three near-fault stations
(Sivirce, P¨
ut¨
urge and Gerger). The soil conditions are provided for
Sivrice, P¨
ut¨
urge and Gerger stations in the AFAD website. Station
Sivrice (2304) located on the soil type with Vs30 450 m s1is the
closest station to the fault rupture, with Joyner and Boore distance
to the fault surface of 3 km. The simulated PGA and PGV values are
328 cm s–2 and31cms
–1, computed considering the soft soil site am-
plifications, while the observed PGA values are 238 and 293 cm s–2
and the observed PGV values are 27 and 45 cm s–1 for the two
horizontal components, respectively. Simulations at P ¨
ut¨
urge station
(2305) located on the rock site with a Vs30 1380 m s1and at 6 km
distance from the fault surface, result in a simulated PGA value of
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1064 D. Cheloni and A. Akinci
Figure 11. The spectral accelerations from the recorded (thin lines) and simulated (thick lines) seismograms for Sivrice, P¨
ut¨
urge and one for a virtual station
(red colour) located close to the nucleation point where maximum slip is released over the fault rupture, and comparison with the code-based spectrum from
old DBYBHY2007 and new TBDY2018 building codes, respectively.
268 cm s–2 while the observed values are 228 and 193 cm s–2 .The
simulated PGV value is instead 25 cm s–1, while the observed values
are 29 and 25 cm s–1 for the two horizontal components (NS and
EW, respectively). The site condition at the Gerger (0204) station is
given as Vs30 =550 ms1; our simulated ground motion parameters
at this site are 87.86 cm s–2 for PGA and 10.60 cm s–1 for PGV val-
ues, respectively, while the recorded values are 110.1–94.24 cm s–2
and 17.1–9.71 cm s–1 for two horizontal components, respectively.
Therefore, the PGA and PGV parameters are very well reproduced
by the stochastic modelling approach for short distances.
In order to show the effectiveness of our simulations to reproduce
observations in frequency domain, in Fig. 9we compare the Fourier
amplitude spectra with the recorded ones at six selected strong
ground motion stations. As shown in Fig. 9, simulations provide
modest estimates of the general shape and amplitudes of the spectra
for almost all of the stations. The misfit at P¨
ut¨
urge and Gerger sta-
tions, at the higher frequencies could be attributed to inaccurate site
amplification function and κparameter adopted for those stations in
our study. The synthetic spectra calculated using a high-frequency
attenuation parameter κ=0.035 s and the frequency-dependent site
amplification of the BC type site classification may be insufficient
to capture all the features of the spectral variation and real transfer
function driven by the detailed velocity-depth profile (Vs30).
6.2 Comparison of observed and simulated ground
motions with selected GMPEs
We validate our simulated ground motion parameters against the
43 observed PGA and PGV values up to 200 km distances from
the recordings of the Elazı˘
g earthquake (Table 1) and compare
our simulations with the three selected GMPEs derived for the
active shallow crustal regions. These include the GMPEs developed
(1) within the context of the Next Generation Attenuation (NGA)
models given by Boore et al. (2014) (hereafter, BSSA14); (2) from
European and the Middle East strong motion database of Bindi
et al. (2014) (hereafter, BIN14) and (3) from national strong motion
database and events as the Turkish attenuation model of Akkar &
Cagnan (2010, hereafter AC10).
In Fig. 10 the two horizontal components of the PGAs recorded
by the total of 43 strong motion stations of Turkish network are
plotted up to 200 km as a function of the distance, and compared
with the values calculated from the two GMPEs, BSSA14, BIN14
and the AC10, together with our simulated PGAs. All GMPEs are
derived for strike-slip faulting style and rock conditions, Vs30 =760
ms
1. Simulations are also performed for the BC type site class,
Vs30 =760 m s1and κ=0.035 s for all the sites (Boore & Joyner
1997; BSSC 2001). As can be seen from Fig. 10(a) the observed
PGAs are mostly overestimated both by BSSA14 and BIN14 at all
distances, while the AC10 model offers better fit to the observed
and the simulated data at the rock sites. At distances greater than
100 km, the recorded PGAs decay with distance is larger than that
predicted by the GMPEs.
The three GMPE models are in good agreement with the simu-
lated and recorded PGV data at short distances although the BIN14
slightly overestimates both the recorded and the simulated data.
Moreover, the observed PGVs are much scattered than the PGAs
and makes ambitious to confirm the better fit with the predictions
particularly at larger distances. It is interesting to note that the
simulated PGA and PGV values determined using region specific
parameters and regional strong motion data are in good agreement
both with those estimated from the empirical Turkish GMPE model
of AC10 and the simulated data at all distances.
In Fig. 11, we compare the spectral accelerations from the sim-
ulated and registered seismograms for Sivrice, P¨
ut¨
urge and from a
virtual station close to the nucleation point (where maximum slip is
released over the fault rupture), with the code-based design spectra
(for Z3, Vs30 >200 m s–1 ,andZC,V
s30 360–760 m s–1 , site classifi-
cation) from both old DBYBHY2007 and new TBDY2018 (T¨
urkiye
Bina Deprem Y¨
onetmeli˘
gi 2018) building codes, respectively. As
it is seen in Fig. 11, response spectra of P¨
ut¨
urge and Sivrice sta-
tion records are well-below the 475 yr design spectra from the new
Turkish building codes (TBDY-2018), differently from the old code
(DBYBHY-2007). The spectral acceleration of the virtual station
(0533) over the fault rupture (where there are no strong motion reg-
istration/recordings available) is also covered by the spectra design
by the new TBDY-2018. It can be observed from Fig. 11 that the
Sivrice’s response spectra has long periods and higher amplitudes
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1065
Figure 12. Slip distribution and Coulomb failure stress variation produced by the 24 January main shock for left-lateral ENE–WSW striking faults, in
agreement with fault orientations of the EAFZ in this area. (a) Slip distribution: the green circles represent the aftershocks of the Elazig seismic sequence;
pink star is the main shock epicentre provided by AFAD, while grey stars are previous large (M>6.6) historical and instrumental events along the EAFZ.
Solid lines represent the main fault segments, while dashed lines the fault jogs. Abbreviations: Y-RDB, Yarpuzlu restraining double bend; H-RB, lake Hazard
releasing bend. The black box represents our extended fault plane solution. The main map is displayed in an oblique Mercator projection with the equator
azimuth parallel to the trend of the EAFZ as in Fig. 1, (b) Change of Coulomb stress (friction coefficient =0.4) computed at a reference depth of 10 km. The
contouring (black solid lines, in cm) indicated the major coseismic slip distribution of the Elazı˘
g earthquake. The ellipse corresponds to the slip deficit area as
suggested in this study. Other symbols are as in panel (a).
due to its site characteristic when compared to that from the rock
motions recorded at P¨
ut¨
urge.
7 VARIATIONS OF STATIC COULOMB
STRESS IN THE STUDY AREA
It is well established that the coseismic slip causes some varia-
tions in static stress that may trigger subsequent earthquakes as well
as patches of aseismic slip on unbroken portions of the causative
fault itself or/and on adjacent faults (e.g. Lin & Stein 2004). In
this context it is important for hazard assessment in the study area
to evaluate the potential state of stress of the unruptured portions
of the P¨
ut¨
urge segment as well as of adjacent segments follow-
ing the 24 January 2020, Mw6.8 Elazı˘
g main shock. To inves-
tigate such stress variations, we calculated the Coulomb stress
change induced by the main shock using our preferred slip dis-
tribution (Fig. 12a) on fault segments having the same mechanism
and geometry as the main event, and assuming an effective fric-
tion coefficient of 0.4, as is commonly used in stress interaction
studies (e.g. Freed 2005; Akinci & Antonioli 2013). As expected,
we find increased Coulomb stress when projected on ENE–WSW
striking left-lateral slip faults at both terminations of the mod-
elled fault plane; (1) ENE near Lake Hazar and (2) WSW of the
hypocentre, respectively, in areas where also a large number of
aftershocks occurred following the 24 January Elazı˘
g earthquake
(Fig. 12b).
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1066 D. Cheloni and A. Akinci
As regard the first stressed area, this zone was likely ruptured
during the 1875 earthquake. In fact, the 1875 event has been gen-
erated along the easternmost termination of the P ¨
ut¨
urge segment
(Ambraseys 1989) or within the complex releasing bend geometry
that is inferred to exist within Lake Hazar (Duman & Emre 2013,
H-RB in Fig. 12). Differently, the second stressed area, WSW of the
hypocentre, might be a portion of the P¨
ut¨
urge fault segment that has
not yet been broken by any seismic event in the last centuries (red
ellipse in Fig. 12b). In fact, there is a debate on the exact location
of the 1905 (Ms6.8) earthquake, that is the last relevant seismic
event possibly located around the western end of the P¨
ut¨
urge seg-
ment (Ambraseys 1989). This event might have happened either
on the Yarpuzlu restraining bend (Y-RDB in Fig. 12) or along the
western part of the P¨
ut¨
urge fault segment itself. Depending of the
real location of the 1905 earthquake, the first hypothesis allows the
possibility that the western part of the P¨
ut¨
urge segment may be an
unbroken area that may rupture in a future earthquake(s).
8 CONCLUSIONS
The 24 January 2020 Mw6.8 Elazı˘
g earthquake in Eastern Turkey
was caused by the rupture of a segment of the EAFZ, the (75)
NNW dipping left-lateral strike-slip P¨
ut¨
urge segment, which has
not ruptured in the recent past and that was considered as a seismic
gap prior of the 2020 seismic sequence (Duman & Emre 2013).
The slip distribution obtained from the geodetic inversion shows
two major asperities with peak slip of 2.3 m, located WSW from
the epicentre, thus implying a marked westward directivity for the
Elazı˘
g main shock. The seismic moment release calculated with
the geodetic data is 1.70 ×1019 Nm (equivalent to a Mw6.8),
in agreement with magnitude estimates provided by different na-
tional and international Institutes. The slip distribution along the
90-km-long causative P¨
ut¨
urge fault segment implies also that the
western part of the seismogenic fault remained unruptured and was
positively stressed by an increase of Coulomb stress.
We simulate the high-frequency ground motion and seismograms
at 1066 virtual stations ranging between 0.1 and 100 km distances in
the epicentral area to have detailed point of view on ground motion
intensity distribution particularly close to the fault rupture where
the maximum damaged observed. In the near-fault area we observe
that our simulations have a good capability to detect near source
effects and to reproduce the source complexity. The general good
consistency found between synthetic and observed ground motion
both in time and frequency domain, suggests the importance of
the retrieving specific regional seismic parameters. We remark that
ground motion parameters decay faster than the empirical ground
motion equations except that of Turkish GMPEs of AC10 both at
moderate and particularly at larger distance (around 100 km) this
feature is captured by our simulated data. Although the adopted
stochastic approach does not fully produce source directivity ef-
fects, such as coherent pulses in near-fault ground motion, it can be
easily and quickly implemented for both region-specific and path-
specific applications. We also demonstrate that the design spectra
corresponding to 475 yr return period, provided by the new Turkish
building code is not exceeded by the simulated seismograms in the
epicentral area where there are no strong motion stations and no
recordings available.
Finally, our prefer red fault model, computed stress redistribution,
and location of historical earthquakes along the EAFZ highlights
some interesting features that are relevant to seismic hazard assess-
ment in the region. In fact, to the WSW of the 2020 Elazı˘
gseismic
sequence, our fault modelling and stress calculation suggests the
presence of a stressed and possibly unbroken area of the P¨
ut¨
urge
segment that should be considered for future hazard assessment.
For this reason, our results suggest that the occurrence of future
significant earthquakes to the WSW of P¨
ut¨
urge city cannot be ruled
out, and therefore a significant seismic hazard still remains in the
area.
ACKNOWLEDGEMENTS
We would like to thank the Editor, Dr Kosuke Heki, the reviewer Ar-
ben Pitarka of Lawrence Livermore National Laboratory, CA, and
the reviewer Farnaz Kamranzad of University of Tehran, for their
constructive suggestions, which helped to improve the manuscript.
Most of the figures have been created using the Generic Mapping
Tools version 4.2.1 (www.soest.hawwai.edu/gmt) and the software
of Seismic Analysis Code (SAC) is used for many of the calculations
throughout several set of macros. We use Copernicus Sentienel-1
InSAR imagery (https://scihub.copernicus.eu/). Sentinel-1 data are
copyright of Copernicus (2020). We thank everyone at the Earth-
quake Department of the Disaster and Emergency Management
Presidency, AFAD for making the strong motion data available
(https://tadas.afad.gov.tr/).
REFERENCES
Akinci, A, Malagnini, L., Herrmann, R.B. & Kalafat, D., 2014. High-
frequency attenuation in the Lake Van Region, Eastern Turkey, Bull.
seism. Soc. Am., 104(3), doi:10.1785/0120130102.
Akinci, A. & Antonioli, A., 2013. Observations and stochastic modelling
of strong ground motions for the 2011 October 23 Mw 7.1 Van, Turkey,
earthquake, Geophys. J. Int., 192, doi:10.1093/gji.ggs075.
Akinci, A., Aochi, H., Herrero, A., Pischiutta, M. & Karanikas, D., 2017.
Physics-based broadband ground-motion simulations for probable Mw
7.0 earthquakes in the Marmara Sea Region (Turkey), Bull. Seismol.
Soc. Am, 107(3), doi:10.1785/0120160096.
Akinci, A., Malagnini, L., Herrmann, R.B., Gok, R. & Sorensen, M.B.,
2006. Ground motion scaling in Marmara region, Turkey, Geophys. J.
Int., 166, 635–651.
Akkar, S. & Cagnan, Z., 2010. A local ground-motion predictive model for
Turkey and its comparison with other regional and global ground-motion
models, Bull. seism. Soc. Am., 100, 2978–2995.
Aktug, B., Ozener, H., Dogru, A., Sanbucu, A., Turgut, B., Halicioglu, K.,
Yilmaz, O. & Havazli, E., 2016. Slip rates and seismic potential on the
East Anatolian Fault System using an improved GPS velocity field, J.
Geod., 94–95, 1–12.
Ambraseys, N., 2009. Earthquakes in the Mediterraneanand Middle East: A
Multidisciplinary Study of Seismicity up to 1900. Cambridge Univ. Press,
968pp.
Ambraseys, N.N., 1989. Temporary seismic quiescence: SE Turkey,
Geophys. J. Int., 96, 311–331.
Ambraseys, N.N. & Jackson, J.A., 1998. Faulting associated with historical
and recent earthquakes in the Eastern Mediterranean region, Geophys. J.
Int., 133, 390–406.
Atkinson, G.M. & Boore, D.M., 1995. Ground-motion relations for eastern
North America, Bull. seism. Soc. Am., 85, 17–30.
Barka, A.A. & Kandinsky-Cade, K., 1988. Strike-slip fault geometry in
Turkey and its influence in earthquake activity, Tectonics, 7, 663–684.
Bindi, D., Massa, M., Luzi, L., Ameri, G., Pacor, F., Puglia, R. & Augliera, P.,
2014. Pan-European ground motion prediction equations for the average
horizontal component of PGA, PGV and 5%-damped PSA at spectral
periods of up to 3.0 s using the RESORCE dataset, Bull. Earthq. Eng.,
12(1), 391–340
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
The 24 January 2020, Mw 6.8 Elazi˘
g earthquake 1067
Boore, D.M., 1983. Stochastic simulation of high-frequency ground motions
based on seismological models of the radiated spectra, Bull. seism. Soc.
Am., 73, 1865–1894.
Boore, D.M., 2003. Simulation of ground motion using the stochastic
method, Pure appl. Geophys., 160, 635–676.
Boore, D.M., 2009. Comparing stochastic point-source and finite-source
ground-motion simulations: SMSIM and EXSIM, Bull. seism. Soc. Am.,
99, doi:10.1785/0120090056.
Boore, D.M. & Joyner, W.B., 1997. Site amplifications for generic rock sites,
Bull. seism. Soc. Am., 87, 327–341.
Boore, D.M., Stewart, P.J., Seyhan, E. & Atkinson, G.M., 2014. NGAWest2
equations for predicting PGA, PGV, and 5% damped PSA for shallow
crustal earthquakes, Earthq. Spectra, 30(3), 1057–1085.
BSSC (Building Seismic Safety Council), 2001. NEHRP recommended pro-
visions for seismic regulations for new buildings, and other structures,
2000 Edition. Part 1: Provisions, Building Seismic Safety Council for the
Federal Emergency Management Agency (Report FEMA368), Washing-
ton, DC, USA.
Cetin, H., Guneyli, H. & Mayer, K., 2003. Paleoseismology of the Palu-Lake
Hazar segment of the East Anatolian Fault Zone, Turkey, Tectonophysics,
374, 163–197.
Cheloni, D. et al., 2010. Coseismic and initial post-seismic slip of the 2009
Mw 6.3 L’Aquila earthquake, Italy, from GPS measurements, Geophys. J.
Int., 181(3), 1539–1546.
Cheloni, D. et al., 2019. Heterogeneous Behavior of the Campotosto Normal
Fault (Central Italy) Imaged by InSAR GPS and Strong-Motion Data:
Insights from the 18 January 2017 Events, Remote Sens., 11, 1482.
Chen, C.W. & Zebker, H.A. 2001. Two-dimensional phase unwrapping with
use of statistical models for cost functions in nonlinear optimization, J.
Opt. Soc. Am., 18(2), 338–351.
Corana, A., Marchesi, M., Martini, C. & Ridella, S., 1987. Minimizing mul-
timodal functions of continuous variables with the “Simulated Annealing”
algorithm, ACM Trans. Math. Softw., 13, 262–280.
Duman, T.Y. & Emre, ¨
O., 2013. The East Anatolian Fault: geometry, seg-
mentation and jog characteristics, Geol. Soc., Lond., Spec. Publ., 372,
495–529.
Freed, T.G., 2005. Earthquake triggering by static, dynamic, and
post-seismic stress transfer, Annu. Rev. Earth Planet Sci., 33,
doi:10.1146/annurev.earth.33.092203.122505.
Goldstein, R.M. & Werner, C.L., 1998. Radar interferograms filtering for
geophysical applications, Geophys. Res. Lett., 25, 4035–4038.
Graves, R. & Pitarka, A., 2010. Refinements to the Graves and Pitarka (2010)
broadband ground motion simulation method, Seismol. Res. Lett., 86(1),
75–80.
G¨
ok, R., Pasyanos, M.E. & Zor, E., 2007. Lithospheric structure of the
continent collision zone: Eastern Turkey, Geophys. J. Int., 169, 3, 1079–
1088.
Hempton, M.R., Dewey, J.F. & Saroglu, F., 1981. The East Anatolian trans-
form fault: along strike variations in geometry and behavior, EOS, Tran.
Am. Geophys. Un., EOS, 62, 393.
Herece, E., 2008. Dogu Anadolu Fayi (DAF) Atlasi. General Directorate of
Mineral Research and Exploration. Special Publications, Ankara, Serial
Number, 14, 359.
Irikura, K. & Miyake, H., 2011. Recipe for predicting strong ground motion
from crustal earthquake scenarios, Pure appl. Geophys., 168, 85–104.
Jarvis, A., Reuter, H.I., Nelson, A. & Guevara, E., 2008. Hole-filled SRTM
for the globe, version4, CGIAR-CSI SRTM 90m Database, available at ht
tp://srtm.csi.cgiar.org.
Karabacak, V., Onder, Y., Altunel, E., Yalciner, C.C., Akyuz, H.S. & Kiyak,
N.G., 2011. Dogu Anadolu Fay Zonunun guney bati uzaniminin paleo-
sismolojisi ve ilk kayma hizi. Proceeding of the Aktif Tektonik Arastirma
Grubu Onbesinci Calistayi (ATAG-15), 19–22 Ekim 2011, Cukurova Uni-
versitesi, Karatas-Adana, 17.
Kondorskaya, N.V. & Ulomov, V.I., 1999. Special catalogue of earthquakes
of the Northern Eurasia (SECNE). http://www.seismo.ethz.ch/static/gsha
p/neurasia/nordasiacat.txt.
Lin, J. & Stein, R.S., 2004. Stress triggering in thrusts and subduction
earthquakes, and stress interaction between the southern San Andreas
and nearby thrust and strike-slip faults, J. geophys. Res., 109(B02303),
doi:10.1029/2003JB002607.
Lohman, R.B. & Simons, M., 2005. Some thoughts on the use of In-
SAR data to constrain models of surface deformation: noise struc-
ture and data downsampling, Geochem. Geophys. Geosyst., 6(Q01007),
doi:10.1029/2004GC000841.
Mai, P.M., Imperatori, W. & Olsen, K.B., 2010. Hybrid broadband ground
motion simulations: combining long-period deterministic synthetics with
high-frequency multiple S-to-S backscattering, Bull. seism. Soc. Am.,
100(5A), 2124–2142.
Malagnini, L., Nielsen, S., Mayeda, K. & Boschi, E., 2010. Energy
radiation from intermediate to large magnitude earthquakes: impli-
cations for dynamic fault weakening, J. geophys. Res., 115(B6),
doi:10.1029/2009JB006786.
McClusky, S. et al., 2000. GPS constraints on plate motions and deformation
in the Eastern Mediterranean: implications for plate dynamics, J. geophys.
Res., 105, 5695–5719.
McKenzie, D.P., 1972. Active tectonics of the Mediterranean region,
Geophys.J.R.astr.Soc.,30, 109–185.
Melgar, D., et al.,, 2020. Rupture kinematics of January 24, 2020
Mw 6.7 Do˘
ganyol-Sivrice, Turkey, earthquake on the East Anatolian
Fault Zone imaged by space geodesy, Geophys. J. Int. , ggaa345,
doi:10.1093/gji/ggaa345 .
Mena, B., Dalguer, L.A. & Mai, P.M., 2012. Pseudodynamic source
characterization for strike slip faulting including stress hetero-
geneity and super-shear ruptures, Bull. seism. Soc. Am., 102(4),
1654–1680.
Mitchell, B. & Cong, L., 1998. Lg coda Q and its relation to the structure
and evolution of continents: a global perspective, Pure appl. Geophys.,
153, 655–663.
Mitchell, B.J., Pan, Y., Xie, J. & Cong, L., 1997. Lg coda Q variation across
Eurasia and its relation to crustal evolution, J. geophys. Res., 102, 22767–
22779.
Motazedian, D. & Atkinson, G.M., 2005. Stochastic finite-fault 800 mod-
elling based on a dynamic 801 corner frequency, Bull. seism. Soc. Am.,
95, 995–1010.
Nalbat, S.S., McCluskey, J., Steacy, S. & Barka, A.A., 2002. Stress accu-
mulation and increased seismic risk in eastern Turkey, Earth. Planet. Sci.
Lett., 195, 291–298.
Okada, Y., 1985. Surface deformation due to shear and tensile faults in a
half-space, Bull. seism. Soc. Am., 75, 1135–1154.
Pasyanos, M.E., Matzel, E.M., Walter, W.R. & Rodgers, A.J., 2009. Broad-
band Lg attenuation modeling in the Middle East, Geophys. J. Int., 177,
1166–1176.
Pitarka, A., Graves, R., Irikura, K., Miyake, H. & Rodgers, A., 2017. Per-
formance of Irikura recipe rupture model generator in earthquake ground
motion simulations with graves and Pitarka hybrid approach, Pure appl.
Geophys., 174(9), doi:10.1007/s00024-017-1504-3.
Pitarka, A., Graves, R., Irikura, K., Miyakoshi, K. & Rodgers, A., 2019.
Kinematic rupture modeling of ground motion from the M7 Kumamoto,
Japan, Earthquake, Pure appl. Geophys., 177, 2199–2221.
Reilinger, R. et al., 2006. GPS constraints on continental deformation in
the Africa-Arabia-Eurasia continental collision zone and implications
for the dynamics of plate interactions, J. geophys. Res., 111(B05411),
doi:10.1029/2005JB004051.
Saroglu, F., Emre, O. & Kuscu, O., 1992. The East Anatolian fault zone of
Tur key , Ann. Tecton., 6, 99–125.
Sertcelik, F., 2012. Estimation of coda wave attenuation in the east Anatolia
fault zone, Turkey, Pure appl. Geophys., 169(7), 1189–1204.
Seyrek, A., Demir, T., Pringle, M.S., Yurtem, S., Westway, R.W.C., Beck,
A. & Rowbotham, G., 2007. Kinematics of the Amos Fault, southern
Turkey, from Ar/Ar dating of offset Pleistocene basalt plates, in Tectonics
of Strike-Slip Restraining and releasing Bends, Vol . 290, pp. 255–284,
eds Cunningham, W.D. & Mann, P., Geological Society London Special
Publications.
Taymaz, T., Eyidogan, H. & Jackson, J., 1991. Source parameters of large
earthquakes in the East Anatolian Fault Zone (Turkey), Geophys. J. Int.,
106, 537–550.
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
1068 D. Cheloni and A. Akinci
T¨
urkiye Bina Deprem Y¨
onetmeli˘
gi (TBDY), 2018. T.C. Ba s¸bakanlık Afet
ve Acil Durum Y¨
onetimi Bas¸kanlı˘
gı, Deprem Dairesi Bas¸kanlı ˘
gı, http:
//www.deprem.afad.gov.tr, Ankara.
U.S. Geological Survey, 2020. Earthquake Hazards Program, available
at: https://earthquake.usgs.gov/earthquakes/eventpage/us60007ewc/; last
accessed on 20 April 2020.
Wang, K. & Burgmann, R., 2020. Co- and postseismic deformation due
to the 2019 Ridgecrest earthquake sequence constrained by Sentinel-
1 and COSMO-SkyMed Data, Seismol. Soc. Lett., 91(4), 1998-2009,
doi:10.1785/0220190299.
Zor, E., Sandvol, E., Xie, J., Turkelli, N., Mitchell, B., Gasanov, A.H. &
Yetirmishli, G., 2007. Crustal attenuation within the Turkish Plateau and
surrounding regions, Bull. seism. Soc. Am., 97, 151–161.
SUPPORTING INFORMATION
Supplementary data are available at GJI online.
Figure S1. Downsampling of the Sentinel-1 ascending track 116
LOS coseismic displacements using a resolution-based downsam-
pling scheme: (a) input full unwrapping interferogram; (b) resam-
pled interferogram. Squares indicate areas within which the LOS
displacements are averaged. (c) LOS profile across sections A–A’
showing the full unwrapped interferogram (coloured circles) versus
the resampled one (open circles). The pink star is the main shock
epicentre provided by AFAD.
Figure S2. Downsampling of the Sentinel-1 ascending track 043
LOS coseismic displacements using a resolution-based downsam-
pling scheme: (a) input full unwrapping interferogram; (b) resam-
pled interferogram. Squares indicate areas within which the LOS
displacements are averaged. (c) LOS profile across sections A–A’
showing the full unwrapped interferogram (coloured circles) versus
the resampled one (open circles). The pink star is the main shock
epicentre provided by AFAD.
Figure S3. Downsampling of the Sentinel-1 descending track 123
LOS coseismic displacements using a resolution-based downsam-
pling scheme: (a) input full unwrapping interferogram; (b) resam-
pled interferogram. Squares indicate areas within which the LOS
displacements are averaged. (c) LOS profile across sections A–A’
showing the full unwrapped interferogram (coloured circles) versus
the resampled one (open circles). The pink star is the main shock
epicentre provided by AFAD.
Figure S4. Downsampling of the Sentinel-1 descending track 021
LOS coseismic displacements using a resolution-based downsam-
pling scheme: (a) input full unwrapping interferogram; (b) resam-
pled interferogram. Squares indicate areas within which the LOS
displacements are averaged. (c) LOS profile across sections A–A’
showing the full unwrapped interferogram (coloured circles) versus
the resampled one (open circles). The pink star is the main shock
epicentre provided by AFAD.
Figure S5. (a) Data, (b) model and (c) residuals sampled points
from the unwrapped ascending track 116 interferogram. The violet
box represents our best-fitting uniform-slip solution. The pink star
is the main shock epicentre provided by AFAD.
Figure S6. (a) Data, (b) model and (c) residuals sampled points
from the unwrapped ascending track 043 interferogram. The violet
box represents our best-fitting uniform-slip solution. The pink star
is the main shock epicentre provided by AFAD.
Figure S7. (a) Data, (b) model and (c) residuals sampled points from
the unwrapped descending track 123 interferogram. The violet box
represents our best-fitting uniform-slip solution. The pink star is the
main shock epicentre provided by AFAD.
Figure S8. (a) Data, (b) model and (c) residuals sampled points from
the unwrapped descending track 021 interferogram. The violet box
represents our best-fitting uniform-slip solution. The pink star is the
main shock epicentre provided by AFAD.
Figure S9. Trade-off curve between solution roughness and
weighted misfit. The red circle indicates the chosen scalar smooth-
ing factor.
Figure S10. Relationship between RMS data reduction and relative
weighting factors. (a) The blue circles represent the RMS reduction
of the descending track 123, while the red ones are relevant to the
ascending track 116. (b) The green circles represent RMS reduction
of the ascending track 043 relative to both the descending track 123
(blue circles) and the ascending track 116 (red circles). (c) The light
blue circles show the RMS reduction of the descending track 021
relative to all the other data sets.
Figure S11. (a) Data, (b) model and (c) residuals sampled points
from the unwrapped ascending track 043 interferogram. The pink
star is the main shock epicentre provided by AFAD.
Figure S12. (a) Data, (b) model and (c) residuals sampled points
from the unwrapped descending track 021 interferogram. The pink
star is the main shock epicentre provided by AFAD.
Figure S13. Coseismic slip distributions on the causative fault seg-
ment from the inversion of (a) only the ascending tracks, (b) only
the descending tracks, and (c) the full data sets. The pink star is the
main shock epicentre provided by AFAD.
Please note: Oxford University Press is not responsible for the con-
tent or functionality of any supporting materials supplied by the
authors. Any queries (other than missing material) should be di-
rected to the corresponding author for the paper.
Downloaded from https://academic.oup.com/gji/article/223/2/1054/5874258 by INGV user on 27 September 2020
... The maximum and minimum horizontal principal stresses are depicted by red and blue arrows, respectively. The relocated aftershocks [7] Several studies [7,[10][11][12][13][14][15][16][17] have investigated the rupture kinematics of 2020 Mw 6.8 Elaziğ earthquake through joint or individual inversions of geodetic (i.e., InSAR and GNSS) and seismic (i.e., strong motion and teleseismic) observations. These kinematic models generally reveal a similar heterogeneous slip distribution with a shallow slip deficit and predominant southwestward rupture propagation. ...
... These kinematic models generally reveal a similar heterogeneous slip distribution with a shallow slip deficit and predominant southwestward rupture propagation. However, they differ significantly in their representations of the causative fault geometry, including single planar fault, e.g., [7,11,12,[15][16][17], two planar faults, e.g., [10,14], and a curved fault [13]. In fact, the EAF is known for its complex geometry, comprising multiple main segments characterized by bends, pull-apart basins, and compressional structures, e.g., [2]. ...
... Following their results, we constructed a realistic curved fault geometry with two bends along the strike direction ( Figure 2). The fault has an average strike of N239°E and dips northwest at 75°, consistent with previous studies, e.g., [12,17]. The fault plane was discretized into 35 along-strike and 14 downdip subfaults, each measuring 1.5 by 1.5 km. ...
Article
Full-text available
Dynamic rupture simulations of earthquakes offer crucial insights into the physical mechanisms of driving fault slip and seismic hazards. By incorporating non-planar fault models that accurately represent subsurface structures, this study provides a realistic depiction of the rupture processes of the 2020 Mw 6.8 Elazığ, Türkiye earthquake, influenced by geometric complexities. Initially, we determined its coseismic slip on the non-planar fault using near-field strong motion and InSAR observations. Subsequently, we established the heterogeneous initial stress on the fault plane based on the coseismic slip and integrated it into the dynamic rupture modeling to assess physics-based ground motion and seismic hazards. The numerical simulations utilized the curved grid finite-difference method (CGFDM), which effectively models rupture dynamics with heterogeneities in fault geometry, initial stress, and other factors. Our synthetic surface deformation and seismograms align well with the observational data obtained from InSAR and seismic instruments. We observed localized occurrences of supershear rupture during fault propagation. Furthermore, the intensity distribution we simulated closely aligns with the actual observations. These findings highlight the critical role of source heterogeneity in seismic hazard assessment, advancing our understanding of fault dynamics and enhancing predictive capabilities.
... Several studies have utilized ground motion simulations to reconstruct historical earthquakes and validate their results against empirical data. For instance, Tanırcan and Yelkenci-Necmioğlu (2020) Askan et al. (2013), Karimzadeh and Askan (2018), Cheloni and Akinci (2020), Arslan Kelam et al. (2022) have employed stochastic simulation techniques to address the scarcity of recorded accelerograms for large-magnitude events by generating ground motions simulations (Ugurhan and Askan, 2010;Ozmen et al., 2020;Askan et al., 2013;Karimzadeh and Askan, 2018;Cheloni and Akinci, 2020;Arslan Kelam et al., 2022). They modeled several major historical earthquakes in Türkiye, including the 1999 Düzce (Mw 7.1), 1939 and 1992 Erzincan (Mw 7.8, 6.6 respectively) as well as more recent events such as the 2020 Elazığ (Mw 6.8) and 2023 Kahramanmaraş (Mw 7.8) earthquakes. ...
... Several studies have utilized ground motion simulations to reconstruct historical earthquakes and validate their results against empirical data. For instance, Tanırcan and Yelkenci-Necmioğlu (2020) Askan et al. (2013), Karimzadeh and Askan (2018), Cheloni and Akinci (2020), Arslan Kelam et al. (2022) have employed stochastic simulation techniques to address the scarcity of recorded accelerograms for large-magnitude events by generating ground motions simulations (Ugurhan and Askan, 2010;Ozmen et al., 2020;Askan et al., 2013;Karimzadeh and Askan, 2018;Cheloni and Akinci, 2020;Arslan Kelam et al., 2022). They modeled several major historical earthquakes in Türkiye, including the 1999 Düzce (Mw 7.1), 1939 and 1992 Erzincan (Mw 7.8, 6.6 respectively) as well as more recent events such as the 2020 Elazığ (Mw 6.8) and 2023 Kahramanmaraş (Mw 7.8) earthquakes. ...
Article
Full-text available
This study evaluates the Soviet-era ground motion prediction equation (named as A&K-1979) and introduces an Artificial Neural Network (ANN)-based GMM specifically designed for Azerbaijan to improve prediction accuracy. Ground motion models (GMMs) are essential for predicting earthquake-induced ground motions, aiding seismic hazard assessments and engineering designs. While traditional linear empirical models have been widely used, they often struggle to capture complex nonlinear ground motion behaviors. The performance of A&K-1979 was assessed using a strong-motion dataset comprising 500 records collected between 2022 and 2024. Two variants of A&K-1979 were tested: A&K-1979-1 for PGA ≥160 cm/s² and A&K-1979-2 for PGA <160 cm/s². An ANN-based GMM was developed using earthquake magnitude and hypocentral distance as inputs, followed by three hidden layers (32-32-16 neurons) with the Rectified Linear Unit (ReLU) activation function. The model was validated with a separate dataset of 268 records, evaluated using metrics such as bias, standard deviation of residuals (σ), mean absolute error (MAE), root mean squared error (RMSE), and R². The A&K-1979 model exhibited notable prediction biases: A&K-1979-1 overestimated PGA values, while A&K-1979-2 underestimated them. The ANN-based GMM achieved improved performance metrics, with a bias of -0.0076, σ of 0.5971, MAE of 0.4416, RMSE of 0.5972, and an R² of 0.4601. The improved accuracy of the ANN-based GMM highlights its potential as a valuable tool for seismic hazard assessments in Azerbaijan. By providing enhanced prediction capabilities, the ANN model demonstrates greater reliability and practical value than A&K-1979, reinforcing the need for updated predictive models in the region and supporting its use in preliminary hazard analysis.
... Turkey has been a seismically active region with a long history. Specifically for Eastern Turkey (Region of this study), a number of destructive earthquakes occurred and many scholars employed stochastic finite fault method to simulate these earthquakes (Karimzadeh and Askan 2018;Askan et al. 2013;Ugurhan and Askan 2010;Zengin and Cakti 2014;Cheloni and Akinci 2020;Can et al. 2021;Kelam et al. 2022). Karimzadeh and Askan (2018) simulated the 1939 M s 7.8 Erzincan earthquake using stochastic finite fault method with available regional seismic and geologic parameters, and obtained the expected intensity map of the 1939 Erzincan earthquake. ...
... Zengin and Cakti (2014) used the stochastic finite fault method to simulate the 2011 M w 7.1 Van earthquake, and evaluated the effect of two different slip models on ground motion intensity. Cheloni and Akinci (2020) simulated 2020 M w 6.8 Elazığ earthquake using the stochastic finite fault method, and observed that it is able to reproduce source complexity and capture seismic motion attenuation with distance. In addition, seismic hazard analysis with buildings damage evaluations also adopted stochastic finite fault simulation, for example, Kelam et al. (2022) analyzed the seismic hazard in Gaziantep, Turkey and performed stochastic finite fault simulations with scenario events and estimated regional buildings damage in central districts of Gaziantep. ...
Article
Full-text available
On February 6, 2023, an Mw 7.8 earthquake occurred in southern Turkey, and only nine hours later, an Mw 7.5 earthquake occurred 95 km north of the first earthquake epicenter. This study employed stochastic finite fault method to simulate the ground motions from the earthquake doublet. The input parameters of source, path, site are mostly determined by regression of station records. The simulated ground motions are validated by comparing with eight station records, and results show that simulated PGA, waveform, PSA curve, duration match with those from station records with minor discrepancies. In addition, goodness-of-fit evaluation is also performed. Regional building damage estimation results show that severely damaged and collapsed buildings increased from 28 to 42% after the second earthquake, and 1/4 buildings damage state experienced one-level jump, which indicates that the second earthquake might significantly intensify buildings damage and should be carefully evaluated within an earthquake doublet context. The stochastic finite fault simulation in this study could provide a basis for future studies on the Turkey earthquake doublet, and the regional buildings damage estimation could be helpful for improvement of earthquake rescue and disaster mitigation policies.
... The magnitude-dependent apparent attenuation contains a reference effective distance term (R ref ) computed from Eq. 10 with R rup = 1 km. As for the geometrical attenuation (given in Eq. 8), Cheloni and Akinci (2020) selects transition distance (R t ) as 100 km. For distances less than 100 km, the geometrical spreading is assumed to be controlled by direct wave radiation in a layered medium with a rate of (b 1 = -1.0), ...
Article
Full-text available
We developed a data-driven and simulation-based ground motion model (GMM) for Eastern Türkiye as a follow up of the Sandıkkaya et al. (Bull Earthq Eng, 2023) paper, which proposes a regional GMM for Western Türkiye. This way, we complete the region-specific ground-motion models for the entire country. As in the case of Sandıkkaya et al. (Bull Earthq Eng, 2023) paper, we investigated the influence of regional stress parameter, geometrical spreading,and anelastic attenuation in Eastern Türkiye for ground-motion prediction. The proposed model was developed for estimating the geometric means of PGA, PGV and 5%-damped horizontal spectral acceleration up to 10 s. The model is usable for earthquake scenarios of magnitudes between 3.5 and 8, and the path effects are consistent for distances up to 400 km. The magnitude-dependent partially non-ergodic sigma is computed as part of the model. In the paper, we also addressed the differences in locality residuals and anelastic attenuation rates for each event cluster that are determined from the seismotectonic features of the region, event sequences and residual distributions.
... In particular, before the February 6th, 2023 Kahramanmaras earthquake sequence, multiple studies identified important seismic gaps along both the main and the northern strand of the EAFZ (e.g., Duman and Emre 2013;Aktug et al. 2016;Bulut et al. 2012) such as the Pazarcik, the Amanos and the Pütürge segments of the main strand, and the Çardak-Sürgü-Savrun segments of the northern strand. Following the Mw6.8 Doğanyol-Elazığ Earthquake sequence in 2020, affecting a part of the Pütürge segment, numerous investigations have extensively examined the true seismic potential of the EAFZ (e.g., Cheloni and Akinci 2020;Ragon et al. 2021;Tatar et al. 2020). ...
Article
Full-text available
On February 6th, 2023, two severe earthquakes struck southeastern Türkiye near the Syrian border. The first earthquake, Mw7.8, occurred at 04:17 local time in the East Anatolian Fault Zone near the city of Gaziantep. The second earthquake, Mw7.5, occurred approximately 9 h later at 13:24 local time near Elbistan County, in Kahramanmaraş province. These seismic events ruptured multiple segments of the East Anatolian Fault Zone (EAFZ), with rupture lengths exceeding 300 km, and deformation exceeding 5 m on both sides of the faults. In this study, we aim to analyze characteristics of the strong ground motion induced by the mainshocks, focusing on ground motion intensity measures such as the peak ground acceleration (PGA), the peak ground velocity (PGV), and the pseudo-acceleration response spectra (PSA). The first earthquake produced extremely high PGA values in both horizontal (> 2 g) and vertical (> 1 g) components. At near field distances, large PGVs are measured (> 180 cm/s) with more than 30 impulsive motions which may indicate source-related effects. Large spectral demands are also recorded for both earthquakes, partially underestimated by Ground Motion Models (GMMs), especially in the near-field. Specifically, we compare the PSA for horizontal directions with the design spectra provided by both the new and previous Turkish building codes. We also present building and ground damage observations that provide insights into the observed ground motions in the heavily damaged areas.
... Figure 1 shows the epicenters of these four earthquakes. The first, Sivrice-Elazığ, struck the eastern part of Türkiye on January 24, 2020, with a moment magnitude of M w = 6.8, causing a maximum peak ground acceleration (PGA) of 0.298 g (AFAD 2020a; Cheloni and Akinci 2020;Cetin et al. 2021;Mertol et al. 2021a and2021b;Bayık et al. 2022;Dedeoğlu et al. 2023). The earthquake's focal depth was 8.1 km. ...
Article
Full-text available
Türkiye is located in an earthquake-prone region where almost all of its population resides in risky areas. In the past 100 years, there has been a strong earthquake every two years and a major one every 3 years. This study investigates the impact of four recent earthquakes, that occurred between 2020 and 2023, on reinforced concrete (RC) buildings. The first, Sivrice-Elazığ, struck the eastern part of Türkiye on January 24, 2020, with a moment magnitude of Mw = 6.8. The second, the Aegean Sea, hit the western part of the country on October 30, 2020, with an Mw of 6.6. The third and fourth are the February 6, 2023 dual Kahramanmaraş earthquakes with Mws of 7.7 and 7.6, which struck the eastern part of Türkiye approximately 9 h apart. Immediately following these earthquakes, a technical team investigated each of the damaged areas. This study summarizes their findings on RC buildings. It was discovered that the majority of the collapsed or severely damaged RC buildings were constructed before 2000. The main reasons for this included technological limitations, specifically on producing high-quality concrete, as well as a lack of public policies and enforced laws in the construction sector to maintain an acceptable international standard. Furthermore, the damage patterns of buildings from these four earthquakes indicated poor workmanship, low material quality, improper structural framing, a common appearance of soft and weak stories, the inadequate use of shear walls, and defective reinforcement configuration. The significance of soil studies and the enforcement of building inspections are also discussed, along with the earthquake codes. The study concludes that the maximum peak ground accelerations from the dual Kahramanmaraş earthquakes were almost triple the code-prescribed values. Therefore, it is recommended that the current mapped spectral acceleration values be revised and that buildings constructed before 2000 should be prioritized while determining their structural performances.
Article
Near-fault fling-step ground motions (NFFS-GMs) are known to cause significant permanent ground displacements, resulting in greater structural damage for long-period flexible structures compared to far-field ground motions. Furthermore, even during the same seismic event, the pulse parameters—such as permanent ground displacements Dsite and pulse period Tp—can vary considerably. Despite their importance, research on stochastic NFFS-GMs remains limited. To address this gap, this paper proposes a method for synthesizing the timefrequency non-stationary stochastic near-fault fling-step ground motion for a specific seismic scenario. Firstly, we employ the discrete wavelet transform (DWT) method, utilizing five mother wavelet functions (MWFs) to analyze 210 Chi-Chi ground motions. This analysis identifies 41 valid NFFS-GMs. The effectiveness of the identification method is validated by comparing the displacement time histories of the original ground motion. Pulse parameters are subsequently derived using the fling-step (FS) pulse model proposed by Abrahamson, in conjunction with the nonlinear least-squares method. A regression model correlating pulse parameters with the seismological parameter of the fault distance R is then developed through Pearson correlation analysis and the nonlinear least-squares method. The residuals of the regression model σln Dsite and Tp are treated as random variables, and their probability distributions are determined. After that, a new stochastic pulse model is introduced to simulate low-frequency ground motions, while a time-frequency non-stationary model is used to simulate high-frequency ground motions. These components are synthesized in the frequency domain to obtain the time-frequency non-stationary stochastic near-fault fling-step ground motion (TFNS-SNFFS-GM) via inverse Fourier transform. Finally, the effectiveness of the proposed method is confirmed by comparing the response spectrum of the synthesized ground motion with that of actual NFFS-GMs.
Article
The five Mw≥7.8 continental transform earthquakes since 2000 all nucleated on branch faults. This includes the 2001 Mw 7.8 Kokoxili, 2002 Mw 7.9 Denali, 2008 Mw 7.9 Wenchuan, 2016 Mw 7.8 Kaikōura, and 2023 Mw 7.8 Pazarcık events. A branch or splay is typically an immature fault that connects to the transform at an oblique angle and can have a different rake and dip than the transform. The branch faults ruptured for at least 25 km before they joined the transforms, which then ruptured an additional 250–450 km, in all but one case (Pazarcık) unilaterally. Branch fault nucleation is also likely for the 1939 M 7.8 Erzincan earthquake, possible for the 1906 Mw∼7.8 and 1857 Mw∼7.9 San Andreas earthquakes, but not for the 1990 Mw 7.7 Luzon, 2013 Mw 7.7 Balochistan, and 2023 Mw 7.7 Elbistan events. Here, we argue that because fault continuity and cataclastite within the fault damage zone develop through cumulative fault slip, mature transforms are pathways for dynamic rupture. Once a rupture enters the transform from the branch fault, flash shear heating causes pore fluid pressurization and sudden weakening in the cataclastite, resulting in very low dynamic friction. But the static friction on transforms is high, and so they are usually far from failure, which could be why they tend to be aseismic between, or at least for centuries after, great events. This could explain why the largest continental transform earthquakes either begin on a branch fault or nucleate along the transform at locations where the damage zone is absent or the fault continuity is disrupted by bends or echelons, as in the 1999 Mw 7.6 İzmit earthquake. Recognition of branch fault nucleation could be used to strengthen earthquake early warning in regions such as California, New Zealand, and Türkiye with transform faults.
Article
When conducting coseismic slip distribution inversion with interferometric synthetic aperture radar (InSAR) data, there is no universal method to objectively determine the appropriate size of InSAR data. Currently, little is also known about the computing efficiency of variance component estimation implemented in the inversion. Therefore, we develop a variance component adaptive estimation algorithm to determine the optimal sampling number of InSAR data for the slip distribution inversion. We derived more concise variation formulae than conventional simplified formulae for the variance component estimation. Based on multiple sampling data sets with different sampling numbers, the proposed algorithm determines the optimal sampling number by the changing behaviors of variance component estimates themselves. In three simulation cases, four evaluation indicators at low levels corresponding to the obtained optimal sampling number validate the feasibility and effectiveness of the proposed algorithm. Compared with the conventional slip distribution inversion strategy with the standard downsampling algorithm, the simulation cases and practical applications of five earthquakes suggest that the developed algorithm is more flexible and robust to yield appropriate size of InSAR data, thus provide a reasonable estimate of slip distribution. Computation time analyses indicate that the computational advantage of variation formulae is dependent of the ratio of the number of data to the number of fault patches and can be effectively suitable for cases with the ratio smaller than five, facilitating the rapid estimation of coseismic slip distribution inversion.
Article
Full-text available
The 2019 Ridgecrest earthquake sequence ruptured a series of conjugate faults in the broad eastern California shear zone, north of the Mojave Desert in southern California. The average spacing between Global Navigation Satellite System (GNSS) stations around the earthquakes is 20–30 km, insufficient to constrain the rupture details of the earthquakes. Here, we use Sentinel-1 and COSMO-SkyMed (CSK) Synthetic Aperture Radar data to derive the high-resolution coseismic and early postseismic surface deformation related to the Ridgecrest earthquake sequence. Line of sight (LoS) Interferometric Synthetic Aperture Radar displacements derived from both Sentinel-1 and CSK data are in good agreement with GNSS measurements. The maximum coseismic displacement occurs near the Mw 7.1 epicenter, with an estimated fault offset of ∼4.5 m on a northwest-striking rupture. Pixel tracking analysis of CSK data also reveals a sharp surface offset of ∼1 m on a second northwest-striking fault strand on which the Mw 6.4 foreshock likely nucleated, which is located ∼2–3 km east of the major rupture. The lack of clear surface displacement across this fault segment during the Mw 6.4 event suggests this fault might have ruptured twice, with more pronounced and shallow slip during the Mw 7.1 mainshock. Both Sentinel-1 and CSK data reveal clear postseismic deformation following the 2019 Ridgecrest earthquake sequence. Cumulative postseismic deformation near the Mw 7.1 epicenter ∼2 months after the mainshock reaches ∼5 cm along the satellites’ LoSs. The observed postseismic deformation near the fault is indicative of both afterslip and poroelastic rebound. We provide data derived in this study in various data formats, which will be useful for the broad community studying this earthquake sequence.
Article
Full-text available
On 18 January 2017, the 2016–2017 central Italy seismic sequence reached the Campotosto area with four events with magnitude larger than 5 in three hours (major event Mw 5.5). To study the slip behavior on the causative fault/faults we followed two different methodologies: (1) we use Interferometric Synthetic Aperture Radar (InSAR) interferograms (Sentinel-1 satellites) and Global Positioning System (GPS) coseismic displacements to constrain the fault geometry and the cumulative slip distribution; (2) we invert near-source strong-motion, high-sampling-rate GPS waveforms, and high-rate GPS-derived static offsets to retrieve the rupture history of the two largest events. The geodetic inversion shows that the earthquake sequence occurred along the southern segment of the SW-dipping Mts. Laga normal fault system with an average slip of about 40 cm and an estimated cumulative geodetic moment of 9.29×1017 Nm (equivalent to a Mw 6). This latter estimate is larger than the cumulative seismic moment of all the events, with Mw > 4 which occurred in the corresponding time interval, suggesting that a fraction (35%) of the overall deformation imaged by InSAR and GPS may have been released aseismically. Geodetic and seismological data agree with the geological information pointing out the Campotosto fault segment as the causative structure of the main shocks. The position of the hypocenters supports the evidence of an up-dip and northwestward rupture directivity during the major shocks of the sequence for both static and kinematic inferred slip models. The activated two main slip patches are characterized by rise time and peak slip velocity in the ranges 0.7–1.1 s and 2.3–3.2 km/s, respectively, and by 35–50 cm of slip mainly concentrated in the shallower northern part of causative fault. Our results show that shallow slip (depth < 5 km) is required by the geodetic and seismological observations and that the inferred slip distribution is complementary with respect to the previous April 2009 seismic sequence affecting the southern half of the Campotosto fault. The recent moderate strain-release episodes (multiple M 5–5.5 earthquakes) and the paleoseismological evidence of surface-rupturing events (M 6.5) suggests therefore a heterogeneous behavior of the Campotosto fault.
Article
Full-text available
We analyzed a kinematic earthquake rupture generator that combines the randomized spatial field approach of Graves and Pitarka (Bull Seismol Soc Am 106:2136–2153, 2016) (GP2016) with the multiple asperity characterization approach of Irikura and Miyake (Pure Appl Geophys 168:85–104, 2011) (IM2011, also known as Irikura recipe). The rupture generator uses a multi-scale hybrid approach that incorporates distinct features of both original approaches, such as small-scale stochastic rupture variability and depth-dependent scaling of rupture speed and slip rate, inherited from GP2016, and specification of discrete high slip rupture patches, inherited from IM2011. The performance of the proposed method is examined in simulations of broadband ground motion from the 2016 Kumamoto, Japan earthquake, as well as comparisons with ground motion prediction equations (GMPEs). We generated rupture models with multi-scale heterogeneity, including a hybrid one in which the slip is a combination of high- slip patches and stochastic small scale variations. We find that the ground motions simulated with these rupture models match the general characteristics of the recorded near-fault motion equally well, over a broad frequency range (0–10 Hz). Additionally, the simulated ground motion is in good agreement with the predictions from Ground Motion Prediction Equations (GMPEs). Nonetheless, due to sensitivity of the ground motion to the local fault rupture characteristics, the performance among the models at near-fault sites is slightly different, with the hybrid model producing a somewhat better fit to the recorded ground velocity waveforms. Sensitivity tests of simulated near-fault ground motion to variations in the prescribed kinematic rupture parameters show that average rupture speeds higher than the default value in GP2016 (average rupture speed = 80% of local shear wave speed), as well as slip rate durations shorter than the default value in GP2016 (rise time coefficient = 1.6), generate ground motions that are higher than the recorded ones at periods longer than 1 s. We found that these two parameters also affect the along strike and updip rupture directivity effects, as illustrated in comparisons with the Kumamoto observations.
Article
Full-text available
The city of Istanbul is characterized by one of the highest levels of seismic risk in the Mediterranean region. An important source of such increased risk is the high probability of large earthquake occurrence during the coming years, which stands at about 65% likelihood owing to the existing seismic gap and the post-1999 earthquake stress transfer at the western portion of the North Anatolian fault zone. In this study, we simulated hybrid broadband time histories from selected earthquakes having magnitude Mw >7:0 in the Sea of Marmara within 10–20 km of Istanbul, the most probable scenarios for simulated generation of the devastating 1509 event in this region. Physics-based rupture scenarios, which may be an indication of potential future events, are adopted to estimate the ground-motion characteristics and its variability in the region. Two simulation techniques are used to compute a realistic time series, considering generic rock site conditions. The first is a full 3D wave propagation method used for generating low-frequency seismograms, and the second is a stochastic finite-fault model approach based on dynamic corner-frequency high-frequency seismograms. Dynamic rupture is generated and computed using a boundary integral equation method, and the propagation in the medium is realized through a finite-difference approach. The results from the two simulation techniques are then merged by performing a weighted summation at intermediate frequencies to calculate a broadband synthetic time series. The simulated hybrid broadband ground motions are validated by comparing peak ground acceleration, peak ground velocity (PGV), and spectral accelerations (5% damping) at different periods with the ground-motion prediction equations in the region. Our simulations reveal strong rupture directivity and supershear rupture effects over a large spatial extent, which generate extremely high near-fault motions exceeding the 250 cm=s PGV along the entire length of the ruptured fault.
Article
Full-text available
We analyzed the performance of the Irikura and Miyake (Pure and Applied Geophysics 168(2011):85–104, 2011) (IM2011) asperity-based kinematic rupture model generator, as implemented in the hybrid broadband ground motion simulation methodology of Graves and Pitarka (Bulletin of the Seismological Society of America 100(5A):2095–2123, 2010), for simulating ground motion from crustal earthquakes of intermediate size. The primary objective of our study is to investigate the transportability of IM2011 into the framework used by the Southern California Earthquake Center broadband simulation platform. In our analysis, we performed broadband (0–20 Hz) ground motion simulations for a suite of M6.7 crustal scenario earthquakes in a hard rock seismic velocity structure using rupture models produced with both IM2011 and the rupture generation method of Graves and Pitarka (Bulletin of the Seismological Society of America, 2016) (GP2016). The level of simulated ground motions for the two approaches compare favorably with median estimates obtained from the 2014 Next Generation Attenuation-West2 Project (NGA-West2) ground motion prediction equations (GMPEs) over the frequency band 0.1–10 Hz and for distances out to 22 km from the fault. We also found that, compared to GP2016, IM2011 generates ground motion with larger variability, particularly at near-fault distances (<12 km) and at long periods (>1 s). For this specific scenario, the largest systematic difference in ground motion level for the two approaches occurs in the period band 1–3 s where the IM2011 motions are about 20–30% lower than those for GP2016. We found that increasing the rupture speed by 20% on the asperities in IM2011 produced ground motions in the 1–3 s bandwidth that are in much closer agreement with the GMPE medians and similar to those obtained with GP2016. The potential implications of this modification for other rupture mechanisms and magnitudes are not yet fully understood, and this topic is the subject of ongoing study. We concluded that IM2011 rupture generator performs well in ground motion simulations using Graves and Pitarka hybrid method. Therefore, we recommend it to be considered for inclusion into the framework used by the Southern California Earthquake Center broadband simulation platform.
Article
A complete suite of closed analytical expressions is presented for the surface displacements, strains, and tilts due to inclined shear and tensile faults in a half-space for both point and finite rectangular sources. These expressions are particularly compact and free from field singular points which are inherent in the previously stated expressions of certain cases. The expressions derived here represent powerful tools not only for the analysis of static field changes associated with earthquake occurrence but also for the modeling of deformation fields arising from fluid-driven crack sources.
Article
Here, we present the results of a kinematic slip model of the 2020 Mw 6.7 Doğanyol-Sivrice, Turkey Earthquake, the most important event in the last 50 yr on the East Anatolian Fault Zone. Our slip model is constrained by two Sentinel-1 interferograms and by 5 three-component high-rate GNSS (Global Navigation Satellite System) recordings close to the earthquake source. We find that most of the slip occurs predominantly in three regions, two of them at between 2 and 10 km depth and a deeper slip region extending down to 20 km depth. We also relocate the first two weeks of aftershocks and find a distribution of events that agrees with these slip features. The HR-GNSS recordings suggest a predominantly unilateral rupture with the effects of a directivity pulse clearly seen in the waveforms and in the measure peak ground velocities. The slip model supports rupture propagation from northeast to southwest at a relatively slow speed of 2.2 km s−1 and a total source duration of ∼20 s. In the absence of near-source seismic stations, space geodetic data provide the best constraint on the spatial distribution of slip and on its time evolution.
Article
We report four new Ar/Ar dates and 18 new geochemical analyses of Pleistocene basalts from the Karasu Valley of southern Turkey. These rocks have become offset left-laterally by slip on the N20°E-striking Amanos Fault. The geochemical analyses help to correlate some of the less-obvious offset fragments of basalt flows, and thus to measure amounts of slip; the dates enable slip rates to be calculated. On the basis of four individual slip-rate determinations, obtained in this manner, we estimate a weighted mean slip rate for this fault of 2.89±0.05mm/a (±2σ). We have also obtained a slip rate of 2.68±0.54mm/a (±2σ) for the subparallel East Hatay Fault farther east. Summing these values gives 5.57±0.54mm/a (±2σ) as the overall left-lateral slip rate across the Dead Sea fault zone (DSFZ) in the Karasu Valley. These slip-rate estimates and other evidence from farther south on the DSFZ are consistent with a preferred Euler vector for the relative rotation of the Arabian and African plates of 0.434±0.012° Ma−1 about 31.1°N, 26.7°E. The Amanos Fault is misaligned to the tangential direction to this pole by 52° in the transpressive sense. Its geometry thus requires significant fault-normal distributed crustal shortening, taken up by crustal thickening and folding, in the adjacent Amanos Mountains. The vertical component of slip on the Amanos Fault is estimated as c. 0.15mm/a. This minor component contributes to the uplift of the Amanos Mountains, which reaches rates of c. 0.2–0.4mm/a. These slip rate estimates are considered representative of time since. 3.73±0.05Ma, when the modern geometry of strike-slip faulting developed in this region; an estimated 11km of slip on the Amanos Fault and c. 10km of slip on the East Hatay Fault have occurred since then. It is inferred that both these faults came into being, and the associated deformation in the Amanos Mountains began, at that time. Prior to that, the northern part of the Africa–Arabia plate boundary was located further east.