InSAR-derived Crustal Deformation and Reverse Fault Motion of the 2017 Iran-Iraq Earthquake in the Northwestern Part of the Zagros Orogenic Belt

Similar documents
DETECTION OF CRUSTAL DEFORMATION OF THE NORTHERN PAKISTAN EARTHQUAKE BY SATELLITE DATA. Submitted by Japan **

INGV. Giuseppe Pezzo. Istituto Nazionale di Geofisica e Vulcanologia, CNT, Roma. Sessione 1.1: Terremoti e le loro faglie

Field Survey of Non-tectonic Surface Displacements Caused by the 2016 Kumamoto Earthquake Around Aso Valley

The March 11, 2011, Tohoku-oki earthquake (Japan): surface displacement and source modelling

22 Bulletin of the Geospatial Information Authority of Japan, Vol.59 December, 2011 Fig. 1 Crustal deformation (horizontal) associated with the 2011 o

Fault model of the 2007 Noto Hanto earthquake estimated from PALSAR radar interferometry and GPS data

Magnitude 7.3 IRAQ. Early reports indicate that 140 have been killed with over 800 injuries reported. Sunday, November 12, 2017 at 18:18:17 UTC

to: Interseismic strain accumulation and the earthquake potential on the southern San

27th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

27th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

Episodic growth of fault-related fold in northern Japan observed by SAR interferometry

NOTES AND CORRESPONDENCE Segmented Faulting Process of Chelungpu Thrust: Implication of SAR Interferograms

Chapter 2. Earthquake and Damage

3D temporal evolution of displacements recorded on Mt. Etna from the 2007 to 2010 through the SISTEM method

Geometry of co-seismic surface ruptures and tectonic meaning of the 23 October 2011 Mw 7.1 Van earthquake (East Anatolian Region, Turkey)

Locations and types of ruptures involved in the 2008 Sichuan earthquake inferred from SAR image matching

Open Access FULL PAPER. Yukitoshi Fukahata * and Manabu Hashimoto

Line of Sight Displacement from ALOS-2 Interferometry: Mw 7.8 Gorkha Earthquake and Mw 7.3 Aftershock

Coulomb stress changes due to Queensland earthquakes and the implications for seismic risk assessment

Basics of the modelling of the ground deformations produced by an earthquake. EO Summer School 2014 Frascati August 13 Pierre Briole

INVESTIGATION OF EARTHQUAKE-CYCLE DEFORMATION IN TIBET FROM ALOS PALSAR DATA PI 168 Roland Bürgmann 1, Mong-Han Huang 1, Isabelle Ryder 2, and Eric Fi

Widespread Ground Motion Distribution Caused by Rupture Directivity during the 2015 Gorkha, Nepal Earthquake

Case study of Japan: Reference Frames in Practice

Earthquakes and Seismotectonics Chapter 5

Slow Deformation of Mt. Baekdu Stratovolcano Observed by Satellite Radar Interferometry

Haiti Earthquake (12-Jan-2010) co-seismic motion using ALOS PALSAR

Magnitude 7.1 NEAR THE EAST COAST OF HONSHU, JAPAN

Ground displacement in a fault zone in the presence of asperities

Source rupture process of the 2003 Tokachi-oki earthquake determined by joint inversion of teleseismic body wave and strong ground motion data

ON NEAR-FIELD GROUND MOTIONS OF NORMAL AND REVERSE FAULTS FROM VIEWPOINT OF DYNAMIC RUPTURE MODEL

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

EARTHQUAKE SOURCE PARAMETERS FOR SUBDUCTION ZONE EVENTS CAUSING TSUNAMIS IN AND AROUND THE PHILIPPINES

Earthquakes in Barcelonnette!

Displacement field and slip distribution of the 2005 Kashmir earthquake from SAR imagery

Case Study of Japan: Crustal deformation monitoring with GNSS and InSAR

FOCAL MECHANISM DETERMINATION USING WAVEFORM DATA FROM A BROADBAND STATION IN THE PHILIPPINES

Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction

Seismic Activity near the Sunda and Andaman Trenches in the Sumatra Subduction Zone

Coulomb stress change for the normal-fault aftershocks triggered near the Japan Trench by the 2011 M w 9.0 Tohoku-Oki earthquake

Multi-planar structures in the aftershock distribution of the Mid Niigata prefecture Earthquake in 2004

Ground surface deformation of L Aquila. earthquake revealed by InSAR time series

Measuring Earthquake-Induced Deformation in the South of Halabjah (Sarpol-e-Zahab) Using Sentinel-1 Data on November 12, 2017

FOCAL MECHANISM DETERMINATION OF LOCAL EARTHQUAKES IN MALAY PENINSULA

Study on the feature of surface rupture zone of the west of Kunlunshan pass earthquake ( M S 811) with high spatial resolution satellite images

Journal of Geophysical Research - Solid Earth

Regional Geodesy. Shimon Wdowinski. MARGINS-RCL Workshop Lithospheric Rupture in the Gulf of California Salton Trough Region. University of Miami

SOURCE PROCESS OF THE 2003 PUERTO PLATA EARTHQUAKE USING TELESEISMIC DATA AND STRONG GROUND MOTION SIMULATION

COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES

ESTIMATES OF HORIZONTAL DISPLACEMENTS ASSOCIATED WITH THE 1999 TAIWAN EARTHQUAKE

Buried Strike Slip Faults: The 1994 and 2004 Al Hoceima, Morocco Earthquakes.

Co-seismic slip from the July 30, 1995, M w 8.1 Antofagasta, Chile, earthquake as constrained by InSAR and GPS observations

Journal of Geophysical Research Letters Supporting Information for

RELOCATION OF LARGE EARTHQUAKES ALONG THE PHILIPPINE FAULT ZONE AND THEIR FAULT PLANES

Separating Tectonic, Magmatic, Hydrological, and Landslide Signals in GPS Measurements near Lake Tahoe, Nevada-California

Crustal deformation by the Southeast-off Kii Peninsula Earthquake

Slip distributions of the 1944 Tonankai and 1946 Nankai earthquakes including the horizontal movement effect on tsunami generation

Burst overlapping of ALOS-2 PALSAR-2 ScanSAR-ScanSAR interferometry

Development of a Predictive Simulation System for Crustal Activities in and around Japan - II

Synthetic Seismicity Models of Multiple Interacting Faults

revised October 30, 2001 Carlos Mendoza

Magnitude 7.0 PAPUA, INDONESIA

Earth Science, (Tarbuck/Lutgens) Chapter 10: Mountain Building

Fault Length and Direction of Rupture Propagation for the 1993 Kushiro-Oki Earthquake as Derived from Strong Motion Duration

Conference Proceedings Paper Measuring Earthquake Induced Deformation in South of Halabjah (Sarpol-e-Zahab) Using Sentinel1 Data on November 12, 2017

Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law

RELOCATION OF LARGE EARTHQUAKES ALONG THE SUMATRAN FAULT AND THEIR FAULT PLANES

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Monitoring long-term ground movements and Deep Seated Gravitational

Geodesy (InSAR, GPS, Gravity) and Big Earthquakes

The Mw 6.2 Leonidio, southern Greece earthquake of January 6, 2008: Preliminary identification of the fault plane.

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

SUPPLEMENTARY INFORMATION

Bonn, Germany MOUTAZ DALATI. General Organization for Remote Sensing ( GORS ), Syria Advisor to the General Director of GORS,

High-resolution temporal imaging of. Howard Zebker

Coseismic surface-ruptures and crustal deformations of the 2008 Wenchuan earthquake Mw7.9, China

Tsunami Simulation of 2009 Dusky Sound Earthquake in New Zealand

The Size and Duration of the Sumatra-Andaman Earthquake from Far-Field Static Offsets

DETECTION OF CRUSTAL MOVEMENTS IN THE 2011 TOHOKU, JAPAN EARTHQUAKE USING HIGH-RESOLUTION SAR DATA

S. Toda, S. Okada, D. Ishimura, and Y. Niwa International Research Institute of Disaster Science, Tohoku University, Japan

Materials and Methods The deformation within the process zone of a propagating fault can be modeled using an elastic approximation.

Coseismic slip distribution of the 2005 off Miyagi earthquake (M7.2) estimated by inversion of teleseismic and regional seismograms

Occurrence of quasi-periodic slow-slip off the east coast of the Boso peninsula, Central Japan

Tandem afterslip on connected fault planes following the 2008 Nima-Gaize (Tibet) earthquake

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

ALOS PI Symposium 2009, 9-13 Nov 2009 Hawaii MOTION MONITORING FOR ETNA USING ALOS PALSAR TIME SERIES

LETTER Earth Planets Space, 58, 63 67, 2006

Preliminary slip model of M9 Tohoku earthquake from strongmotion stations in Japan - an extreme application of ISOLA code.

Coseismic displacements and slip distribution from GPS and leveling observations for the 2006 Peinan earthquake (M w 6.1) in southeastern Taiwan

Use a highlighter to mark the most important parts, or the parts. you want to remember in the background information.

14 S. 11/12/96 Mw S. 6/23/01 Mw S 20 S 22 S. Peru. 7/30/95 Mw S. Chile. Argentina. 26 S 10 cm 76 W 74 W 72 W 70 W 68 W

Source Characteristics of Large Outer Rise Earthquakes in the Pacific Plate

SUPPLEMENTARY INFORMATION

Coseismic slip model of the 2007 August Pisco earthquake (Peru) as constrained by Wide Swath radar observations

DEFORMATION KINEMATICS OF TIBETAN PLATEAU DETERMINED FROM GPS OBSERVATIONS

Kinematics of the Southern California Fault System Constrained by GPS Measurements

Estimation of deep fault geometry of the Nagamachi-Rifu fault from seismic array observations

Depth (Km) + u ( ξ,t) u = v pl. η= Pa s. Distance from Nankai Trough (Km) u(ξ,τ) dξdτ. w(x,t) = G L (x,t τ;ξ,0) t + u(ξ,t) u(ξ,t) = v pl

Case Study 1: 2014 Chiang Rai Sequence

TEMPORAL VARIABILITY OF ICE FLOW ON HOFSJÖKULL, ICELAND, OBSERVED BY ERS SAR INTERFEROMETRY

Transcription:

InSAR-derived Crustal Deformation and Reverse Fault Motion of the 2017 Iran-Iraq Earthquake in the Northwestern Part of the Zagros Orogenic Belt Tomokazu Kobayashi, Yu Morishita, Hiroshi Yarai and Satoshi Fujiwara (Published online: 5 March 2018) Abstract An inland earthquake with a moment magnitude of 7.3 occurred in western Iran / eastern Iraq on November 12, 2017. Applying an interferometric SAR (InSAR) analysis using ALOS-2 SAR data to the earthquake, we detected the crustal deformation in the northwestern part of the Zagros orogenic belt. We successfully obtained quasi-up-down and quasi-east-west displacement components by combining two sets of InSAR data. The results show that uplift occurred in the western part of the source region with ~90 cm at most, while in the east subsidence occurred with ~30 cm at most. Most of the source region moved westward with ~50 cm at most. Our preferred fault model shows that the observations can be accounted for by a nearly pure reverse fault motion with a slight dextral motion on a low-angle fault plane dipping northeast. 1. Introduction An earthquake with a moment magnitude (M w ) of 7.3 (U. S. Geological Survey (USGS), 2017) struck western Iran / eastern Iraq on November 12, 2017 (UTC). The epicenter of the main shock determined by USGS (2017) is shown by the red star in Fig. 1. In this region, the collision between the Arabian and Eurasian plates has resulted in the formation of the Zagros Mountains. The red line in Fig. 1 represents the boundary between the plates (Bird, 2003). The Arabian plate moves northward relative to the Eurasian plate at a speed of ~2 cm/year (Vernant et al., 2004; Madanipour et al. 2013). The direction of the plate movement is oblique against the plate boundary. Owing to the active collision, the source region is subjected to a NS-oriented convergent stress produced by the motion of the Arabian plate (Vernant et al., 2004), which culminates in large historical earthquakes (Fig. 1). Satellite synthetic aperture radar (SAR) data can provide detailed and spatially comprehensive ground information, and an interferometric SAR (InSAR) enables ground deformation to be measured with high precision. It can provide valuable information on the spatial features of the fault rupture. The primary purpose of this paper is to quickly report on the first results of the InSAR-derived crustal deformation and a preliminary fault model. Fig. 1 Tectonic setting and topography map. The red star and beach ball stand for the epicenter location and the centroid-moment-tensor (CMT) solution, respectively (USGS, 2017). The red line indicates a plate boundary (Bird, 2003). The black arrow represents GPS-estimated movement of the Arabian plate relative to the Eurasian plate (Vernant et al., 2004; Madanipour et al. 2013). Open circles are historical earthquakes of M>5 and shallower than 50 km, listed in the USGS catalog.

2. SAR Data Analysis A Japanese L-band synthetic aperture radar satellite, known as the Advanced Land Observing Satellite 2 (ALOS-2), made emergency observations for the purpose of detecting crustal deformation using InSAR. The images used in the study are listed in Table 1. We analyzed the ScanSAR-ScanSAR interferometry. The ScanSAR data we processed were produced by a full-aperture method in which range/azimuth compression was performed for the data whose gaps between neighboring bursts were filled with zeroes. The original interferograms were contaminated by a long-wavelength phase change noise of unknown causes, maybe due in part to the ionospheric effect. To pick out the crustal deformation only, we modeled the noises by estimating a phase curve with a second-order polynomial from far-field data, and subtracted the estimated phase curve from the original interferograms. Observation Date Flight Direction Beam Direction Incidence Angle Observation Mode Perpendicular Baseline Table 1 Analyzed ALOS-2 images. Aug. 09, 2016 Nov. 14, 2017 Ascending Right Oct. 04, 2017 Nov. 15, 2017 Descending Right 47 47 ScanSAR- ScanSAR ScanSAR- ScanSAR 70 m +160 m Figure # Fig. 2 Fig. 3 We processed the ALOS-2 data with GSISAR software (Fujiwara and Tobita, 1999; Tobita et al., 1999; Fujiwara et al., 1999; Tobita, 2003), and used ASTER GDEM for the InSAR analyses. 3. Ground displacement detected by InSAR 3.1 Line-of-sight component Figs. 2a and 3a show interferograms obtained from ascending and descending orbit data, respectively. The coherence is high for the most part, probably due to the less-vegetated environment. An intensive deformation is located around 20 km NNW of Sarpol-e Zahab city. Two clear roundish fringes appear in both images. In Fig. 2a, slant range shortening signals are distributed in both lobes. The maximum line-of-sight (LOS) displacements are ~90 cm shortening for the southwestern lobe at most and ~15 cm shortening for the northeastern lobe at most. On the other hand, in Fig. 3a, slant range shortening and lengthening signals appear in the southwest and the northeast, respectively. The maximum LOS displacements are ~50 cm shortening for the southwestern lobe at most and ~35 cm lengthening for the northeastern lobe at most. 3.2 Quasi-up-down and quasi east-west components We have both descending and ascending data that are right-looking, in which the microwaves are emitted from two opposite directions from the eastern sky and the western sky. This enables the derived LOS displacements to be converted into two components which are the quasi-up-down (QUD) and quasi-east-west (QEW) components (Fig. 4) (Fujiwara et al., 2000). Figs 5a and 5b show the estimated QUD and QEW components, respectively. In Fig. 5a, the red and blue colors mean uplift and subsidence, respectively. We can identify a large uplift is distributed in the southwest of the epicenter with ~90 cm at most. In contrast, in and around the epicentral area, the ground has subsided with ~30 cm. In Fig. 5b, the red and green colors mean eastward and westward movements, respectively. There appear two main westward movement areas in the southwest with ~50 cm at most and in the northeast with ~35 cm at most. These spatial patterns are in agreement with crustal deformation produced by a reverse slip motion on a NW-SE striking fault plane. 3.3 Local Ground Displacement Here we stress that we can identify many localized LOS displacements in the InSAR images, apart from the fault-slip-related crustal deformation. Figs. 2b and 2c show enlarged views of Fig. 2a. Note that we filtered out the long-wavelength signals corresponding to the fault-related deformation to pick out localized displacements. We can identify small and local but clear phase changes in the high-pass filtered image. Also, applying the high-pass filter to the InSAR data of descending orbit in the same manner,

Fig. 2 An InSAR image for the ascending orbit. (a) An InSAR image covering the source region. (b) and (c) Enlarged views. A high-pass filter was applied to remove the long-wavelength signals. Fig. 3 Same as Fig. 2 but for the descending orbit data.

Fig. 4 Geometry of 2.5D displacement analysis. Fig. 5 2.5D displacement fields. (a) A quasi-up-down component. The red and blue colors mean uplift and subsidence, respectively. (b) A quasi-east-west component. The red and green colors mean eastward and westward movements, respectively. the localized phase changes can be identified at almost the same locations, but with the opposite sign, indicating that the horizontal (EW) displacement is dominant (Figs. 3b and 3c). Most of the localized displacements are detected at mountain slopes and are consistent with downward movement along the slopes. These displacements might have been triggered by the seismic motion. 3.4 Linear Surface Ruptures The InSAR images generally show elastic deformation caused by the main earthquakes, but many other linear discontinuities showing displacement are also found (Fujiwara et al., 2016). We mapped the ruptures in the two InSAR images in Fig. 6. Most of the ruptures are likely secondary faults that are not directly related to the main earthquake but whose slip was probably triggered by the main earthquake or

aftershocks. 4. Fault model We constructed a slip distribution model by a least squares approach (e.g., Kobayashi et al. 2015). In the inversion, we used both the ascending and the descending orbit data shown in Figs. 2a and 3a. The fault geometry was assumed to be a plane fault. We set a rectangular fault of 100 km long and 80 km wide, and the fault was divided into square patches of 5 5 km in size. The fault geometry basically followed the USGS s Centroid Moment Tensor (CMT) solution (USGS, 2017). According to previous geological and seismological studies, a low-angle thrust system with a NE-dipping plane may have developed in this region (e.g., Vergés et al., 2011; Madanipour et al. 2013). Thus we followed this idea; that is, we took a nodal plane dipping to NE with a dip angle of 16. The strike was fixed so as to be almost parallel to the plate boundary (Bird, 2003). We adjusted the fault depth so that the USGS's hypocenter (19 km in depth) could be on the fault plane. Consequently, the fault top position was fixed to a depth of 3 km below the surface. We used the dislocation equations proposed by Okada (1985) to calculate the surface displacement in the LOS directions. Only the dip-slip and strike-slip components were estimated for each patch. A large number of model parameters often give rise to instability of the solution. To stabilize the solution, we imposed a spatial smoothness constraint on the slip distribution using a Laplacian operator. The relative weight of the constraints was determined by Akaike s Baysian information criterion (Akaike, 1980). Here, we assumed a rigidity of 30 GPa. The data points of the interferograms were too many to be easily assimilated in a modeling scheme. In order to reduce the number of data for the modeling analysis, we resampled the interferogram data beforehand, using a quadtree decomposition method. Essentially, we followed an algorithm presented by Jónsson et al. (2002). For a given quadrant, after removing the mean, if the residue was greater than a prescribed threshold (5 cm in our case), the quadrant was further divided into four new quadrants. This process was iterated either until each block met the specified criterion or until the quadrant reached a minimum block size (8 8 pixels in our case). Upon application of this procedure, the sizes of the respective interferogram data sets were reduced from 1,443,200 to 239 for the ascending data and to 279 for the descending data. Fig. 7 shows the estimated slip distribution, and Figs. 8 and 9 show the LOS displacements calculated from the derived model and the residuals, respectively. Our model reproduces well the LOS displacement fields for both the ascending and descending data. Fig. 10 shows the model-predicted UD, EW, and NS displacement components. Both UD and EW displacement fields are reproduced well. A major slip with a maximum slip amount of approximately 3 m occurred beneath the southeastern part of the epicenter. The slips were a nearly pure reverse fault motion, but with a slight right-lateral motion. The obtained oblique slip was consistent with the tectonic background that the Arabian plates obliquely subside against the Eurasian plate (Fig. 1). The moment is estimated to be 1.18 10 20 Nm, equivalent to M w 7.31. This is consistent with the seismic moment determined with seismic data (USGS CMT: 1.12 10 20 Nm (M w 7.3); Global CMT Catalog (2017): 1.72 10 20 Nm (M w 7.4)). 5. Concluding remarks We conducted InSAR analysis using ALOS-2 data to detect the crustal deformation associated with the 2017 Iran-Iraq earthquake. The following conclusions were obtained from the analyses: 1. Large displacement (~90 cm upward and ~50 cm westward) was detected around 20 km NNW of Sarpol-e Zahab. 2. Around the epicenter, ~30 cm downward and ~35 cm westward displacement was detected. 3. Apart from the deformation associated with the fault slip, lots of localized displacements were detected at mountain slopes, and might have been triggered by the seismic motion. 4. Our fault model shows a nearly pure reverse-fault with a slight dextral slip motion. A major slip occurred with a slip amount of ~3 m beneath the southwestern part of the epicenter. 5. The estimated moment magnitude was 7.31 (seismic moment 1.18 10 20 Nm).

Fig. 6 InSAR images for the ascending orbit (a) and the descending orbit (b). The solid lines show the identified linear surface ruptures. Fig. 7 (a) Slip distribution model. The arrows show slip vectors of the hanging wall. The contour interval is 1 m. The red star represents the hypocenter projected on the fault plane. (b) Slip distribution on map. (c) Schematic view of the source mechanism.

Fig. 8 (a) Model-calculated LOS displacement for the ascending orbit data. (b) Residual. Fig. 9 (a) Same as Fig. 7 but for the descending orbit data.

Fig. 10 Model-predicted displacements. (a) Up-Down component. (b) East-West component. (c) North-South component. Acknowledgments ALOS-2 data were provided by the Earthquake Working Group under a cooperative research contract with the Japan Aerospace Exploration Agency (JAXA). The ownership of ALOS-2 data belong to JAXA. We used Generic Mapping Tools (GMT) provided by Wessel and Smith (1998) to construct the figures for the fault model. We used ASTER GDEM for the InSAR analyses. ASTER GDEM is a product of the National Aeronautics and Space Administration (NASA) and the Ministry of Economy, Trade and Industry (METI). References Akaike, H (1980): Likelihood and the Bayes procedure. In: Bernardo. JM, DeGroot. MH, Lindley. DV, Smith.AFM, (eds) Bayesian Statistics. University Press, Valencia. Bird, P. (2003): An updated digital model of plate boundaries, Geochemistry Geophysics Geosystems, 4(3), 1027. Fujiwara, S and M. Tobita (1999): SAR interferometry techniques for precise surface change detection, J. Geod. Soc. Japan, 45, 283-295 (in Japanese with English abstract). Fujiwara, S., M. Tobita, M. Murakami, H. Nakagawa and P. A. Rosen (1999): Baseline determination and correction of atmospheric delay induced by topography of SAR interferometry for precise surface change detection, J. Geod. Soc. Japan, 45, 315-325 (in Japanese with English abstract). Fujiwara, S., T. Nishimura, M. Murakami, H. Nakagawa, M. Tobita and P. A. Rosen (2000): 2.5-D surface deformation of a M6.1 earthquake near Mt Iwate detected by SAR interferometry, Geophys. Res. Lett., 27, 2049-2052. Fujiwara, S., H. Yarai, T. Kobayashi, Y. Morishita, T. Nakano, B. Miyahara, H. Nakai, Y. Miura, H. Ueshiba, Y. Kakiage and H. Une (2016): Small-displacement linear surface ruptures of the 2016 Kumamoto earthquake sequence detected by ALOS-2 SAR interferometry, Earth Planets Space, 68: 160. Global Centroid Moment Tensor Catalog (2017): 201711121818A IRAN-IRAQ BORDER REGION, http://www.globalcmt.org (accessed Dec.7, 2017). Jónsson, S., H. Zebker, P. Segall and F. Amelung (2002) : Fault slip distribution of the 1999 Mw 7.1 Hector Mine, California, earthquake, estimated from satellite radar and GPS measurements, Bull. Seism. Soc. Amer., 92, 1377-1389. Kobayashi, T., Y. Morishita and H. Yarai (2015): Detailed crustal deformation and fault rupture of the 2015 Gorkha earthquake, Nepal, revealed from ScanSAR-based interferograms of ALOS-2, Earth Planets Space, 67:201. Madanipour, S., Ehlers, T., Yassaghi, A., Rezaeian, M., Enkelmann E and Bahroudi, A (2013): Synchronous deformation on orogenic plateau margins: Insights from the Arabia Eurasia collision, Tectonophysics, 608, 440-451. http://dx.doi.org/10.1016/j.tecto.2013.09.003 (accessed Dec.7, 2017). Okada, Y (1985): Surface deformation due to shear and tensile faults in a halfspace, Bull Seismol Soc Am, 75, 1135-1154.

Tobita, M., S. Fujiwara, M. Murakami, H. Nakagawa and P. A. Rosen (1999): Accurate offset estimation between two SLC images for SAR interferometry, J. Geod. Soc. Japan, 45, 297-314 (in Japanese with English abstract). Tobita, M (2003): Development of SAR interferometry analysis and its application to crustal deformation study, J. Geod. Soc. Japan, 49, 1-23 (in Japanese with English abstract). U. S. Geological Survey (2017): M7.3 30km S of Halabjar, Iraq. https://earthquake.usgs.gov/earthquakes/eventpage/us 2 000bmcg#executive (accessed Dec.7, 2017). Vergés, J., Saura, E., Casciello, E., Fernàndez, M., Villaseñor, A., Jiménez-Munt, I., García-Castellanos, D (2011): Crustal-scale cross-sections across the NW Zagros belt: implications for the Arabian margin reconstruction, Geological Magazine, 148, 739-761. Vernant, Ph., F. Nilforoushan, D. Hatzfeld, M.R. Abbassi, C. Vigny, F. Masson, H. Nankali, J. Martinod, M. Ghafory-Ashtiany, R. Bayer, F. Tavakoli and J. Chéry (2004) : Present-Day Crustal Deformation and Plate Kinematics in the Middle East Constrained by GPS Measurements in Iran and Northern Oman, Geophys. J., Int., 157, 381-398. Wessel, P and Smith, WH (1998): New, improved version of Generic Mapping Tools released, Eos Trans AGU, 79, 579.