PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI

Page created by Herman Brooks
 
CONTINUE READING
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
remote sensing
Article
Probing the Fault Complexity of the 2017 Ms 7.0 Jiuzhaigou
Earthquake Based on the InSAR Data
Xiongwei Tang 1,2 , Rumeng Guo 1,2,3, *, Jianqiao Xu 1 , Heping Sun 1,2 , Xiaodong Chen 1 and Jiangcun Zhou 1

 1 State Key Laboratory of Geodesy and Earth’s Dynamic, Innovation Academy for Precision Measurement
 Science and Technology, Chinese Academy of Sciences, Wuhan 430077, China;
 tangxiongwei18@mails.ucas.edu.cn (X.T.); xujq@asch.whigg.ac.cn (J.X.); heping@asch.whigg.ac.cn (H.S.);
 chenxd@whigg.ac.cn (X.C.); zjc@asch.whigg.ac.cn (J.Z.)
 2 College of Earth and Planetary Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
 3 Earth System Science Programme, The Chinese University of Hong Kong, Shatin, Hong Kong, China
 * Correspondence: rumengguo@cuhk.edu.hk

 Abstract: On 8 August 2017, a surface wave magnitude (Ms) 7.0 earthquake occurred at the buried
 faults extending to the north of the Huya fault. Based on the coseismic deformation field obtained
 from interferometric synthetic aperture radar (InSAR) data and a series of finite fault model tests,
 we propose a brand-new two-fault model composed of a main fault and a secondary fault as the
 optimal model for the Jiuzhaigou earthquake, in which the secondary fault is at a wide obtuse angle to
 the northern end of the main fault plane. Results show that the dislocation distribution is dominated
 by sinistral slip, with a significant shallow slip deficit. The main fault consists of two asperities
 bounded by an aftershock gap, which may represent a barrier. In addition, most aftershocks are
 located in stress shadows and appear a complementary pattern with the coseismic high-slip regions.
 
  We propose that the aftershocks are attributable to the background tectonic stress, which may be
 related to the velocity-strengthening zones.
Citation: Tang, X.; Guo, R.; Xu, J.;
Sun, H.; Chen, X.; Zhou, J. Probing
the Fault Complexity of the 2017 Ms
 Keywords: Jiuzhaigou earthquake; fault geometry parameters; coseismic slip distribution; InSAR
7.0 Jiuzhaigou Earthquake Based on
the InSAR Data. Remote Sens. 2021, 13,
1573. https://doi.org/10.3390/
rs13081573 1. Introduction
 On 8 August 2017, a surface wave magnitude (Ms) 7.0 earthquake occurred in Ji-
Academic Editor: Zhong Lu uzhaigou County, Aba Prefecture, Sichuan Province, China, with its epicenter at 103.82◦ E,
 33.20◦ N, and a focal depth of about 20 km (Institute of Geophysics, China Earthquake
Received: 12 April 2021
 Administration, CEA-IGP). As of 13 August 2017, the earthquake had caused 25 deaths
Accepted: 15 April 2021
 and severe damage to both the Jiuzhaigou scenic area and more than 70,000 structures [1].
Published: 19 April 2021
 The Jiuzhaigou earthquake occurred near the northeastern boundary of the Bayan Har
 block, indicating that the Bayan Har block is still active [2]. Because the southeasterly
Publisher’s Note: MDPI stays neutral
 movement of the Bayan Har block is blocked by the North China block and the South China
with regard to jurisdictional claims in
 block, there are many broom-like branches on the eastern end of the East Kunlun fault,
published maps and institutional affil-
 which intersect with the Minjiang fault and the Huya fault, resulting in a complex seismic-
iations.
 tectonic environment in this region [3]. The NWW-trending Tazang fault, the northern
 segment of the NNW-trending Huya fault, and other nearby branches accommodate most
 of the sinistral strike-slip activity of the East Kunlun fault. The nearly NS-trending Min-
 jiang fault and the southern segment of the Huya fault that are transverse compressional
Copyright: © 2021 by the authors.
 structures absorb the sinistral strike-slip motion of the remaining part of the East Kunlun
Licensee MDPI, Basel, Switzerland.
 fault, which causes the Minshan uplift on the eastern margin of the Tibetan Plateau [4,5].
This article is an open access article
 The complicated and intense tectonic activities of these faults make the eastern margin
distributed under the terms and
 of the Bayan Har block one of the most seismically active regions in China, with many
conditions of the Creative Commons
 strong earthquakes having occurred throughout its history. Examples include the 1654
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
 Tianshui earthquake (Ms = 8), the 1879 Wudu earthquake (Ms = 7), the 1933 Diexi earth-
4.0/).
 quake (Ms = 7.5), the 1976 Songpan earthquakes (Ms = 7.2, 6.7, and 7.2), the 2008 Wenchuan

Remote Sens. 2021, 13, 1573. https://doi.org/10.3390/rs13081573 https://www.mdpi.com/journal/remotesensing
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, x FOR PEER REVIEW 2 of 14

Remote Sens. 2021, 13, 1573 strong earthquakes having occurred throughout its history. Examples include the2 1654 of 12
 Tianshui earthquake (Ms = 8), the 1879 Wudu earthquake (Ms = 7), the 1933 Diexi earth-
 quake (Ms = 7.5), the 1976 Songpan earthquakes (Ms = 7.2, 6.7, and 7.2), the 2008 Wen-
 chuan earthquake
 earthquake (Ms = 8),(Msand= 8),
 theand
 2013the 2013 Lushan
 Lushan earthquakeearthquake
 (Ms = 7)(Ms = 7) 1).
 (Figure (Figure 1). The
 The shorter
 shorter duration, continuous earthquakes, and smaller strain of release
 duration, continuous earthquakes, and smaller strain of release compared to previous compared to pre-
 phases of earthquake clustering show that the Bayan Har block is experiencing a peak ina
 vious phases of earthquake clustering show that the Bayan Har block is experiencing
 peak in activity
 seismic seismic [4].
 activity [4]. Therefore,
 Therefore, further
 further study of study of its seismogenic
 its seismogenic environment environment
 is urgently is
 urgentlyThe
 needed. needed.
 2017 The 2017 Jiuzhaigou
 Jiuzhaigou earthquake earthquake
 provides provides a great opportunity
 a great opportunity to under-
 to understand the
 stand theloading
 tectonic tectonicmechanisms
 loading mechanisms in thisDifferent
 in this region. region. Different
 researchresearch institutions
 institutions have
 have issued
 aissued
 seriesaofseries of reports
 reports on themechanisms
 on the focal focal mechanisms of the Jiuzhaigou
 of the Jiuzhaigou earthquake.
 earthquake. All of All
 them of
 them shown
 have have shown thatearthquake
 that this this earthquake
 was a was a sinistral
 sinistral strike-slip
 strike-slip eventawith
 event with highadip
 high dip
 angle
 angle
 and and a moment
 a moment magnitude
 magnitude (Mw) (Mw)
 of 6.5,of 6.5,
 but but are
 there there are significant
 significant discrepancies
 discrepancies in thein the
 fault
 fault geometry
 geometry and depths
 and focal focal depths
 (Table(Table
 1). 1).

 Figure 1. Tectonic setting and seismicity around the 2017 Jiuzhaigou earthquake. (a) Tectonic setting of the eastern margin
 Figure 1. Tectonic setting and seismicity around the 2017 Jiuzhaigou earthquake. (a) Tectonic setting of the eastern margin
 of the Tibetan Plateau. The solid blue and purple lines indicate the primary and secondary block boundaries, respectively.
 of the Tibetan Plateau. The solid blue and purple lines indicate the primary and secondary block boundaries, respectively.
 The color-coded circles represent historical earthquakes (Ms ≥ 5) from the Data Sharing Infrastructure of National Earth-
 The
 quakecolor-coded
 Data Centercircles
 (NEDC)represent historical earthquakes
 (http://data.earthquake.cn, (Ms ≥on5)14
 accessed from the2021).
 April DataThe
 Sharing
 black Infrastructure
 beach balls show of National
 the focal
 Earthquake
 mechanisms of the historical earthquakes (Ms ≥ 6.5) in this region from the Global Centroid Moment Tensor,show
 Data Center (NEDC) (http://data.earthquake.cn, accessed on 14 April 2021). The black beach balls the
 Harvard
 focal mechanisms of the historical earthquakes (Ms ≥ 6.5) in this region from the Global Centroid Moment
 University Global Moment Tensor Solution (GCMT). The red beach ball represents the focal mechanism of the Jiuzhaigou Tensor, Harvard
 University
 earthquakeGlobal Moment
 from the Tensor
 Institute Solution (GCMT).
 of Geophysics, The red beach
 China Earthquake ball represents
 Administration the focal(http://www.cea-igp.ac.cn/en/,
 (CEA-IGP) mechanism of the Jiuzhaigou
 accessed on 14 April 2021). The inset shows the large-scale tectonic environment. The red rectangle is the region shown in
 earthquake from the Institute of Geophysics, China Earthquake Administration (CEA-IGP) (http://www.cea-igp.ac.cn/en/,
 (b). Note: QT—Qiangtang block; BH—Bayan Har block; QL—Qilian block; SC—South China; NC—North
 accessed on 14 April 2021). The inset shows the large-scale tectonic environment. The red rectangle is the region China;
 shownTB—
 in
 Tibet; TR—Tarim Basin. (b) Fault geometry and aftershocks of the Jiuzhaigou earthquake. The
 (b). Note: QT—Qiangtang block; BH—Bayan Har block; QL—Qilian block; SC—South China; NC—North China; TB—Tibet;blue dots are the relocated
 aftershocks within a month from Fang et al. [6], and the red rectangle is the projection of our optimal fault model on the
 TR—Tarim Basin. (b) Fault geometry and aftershocks of the Jiuzhaigou earthquake. The blue dots are the relocated
 surface. The gray, black, red, and purple beach balls show the focal mechanisms determined by the GCMT, United States
 aftershocks within a month from Fang et al. [6], and the red rectangle is the projection of our optimal fault model on the
 Geological Survey (USGS) (https://www.usgs.gov/, accessed on 14 April 2021), CEA-IGP, and Institute of Earthquake
 surface. The gray,
 Forecasting, Chinablack, red, and
 Earthquake purple beach balls
 Administration show (http://www.ief.ac.cn/,
 (CEA-IEF) the focal mechanisms determined
 accessed onby14the GCMT,
 April 2021),United States
 respectively.
 Geological Survey (USGS) (https://www.usgs.gov/, accessed on 14 April 2021), CEA-IGP, and Institute of Earthquake
 Forecasting, China Earthquake Administration
 No significant (CEA-IEF) (http://www.ief.ac.cn/,
 coseismic surface rupture was accessed on 14 April
 observed 2021),
 for the respectively.earth-
 Jiuzhaigou
 quake, making it challenging to determine the seismogenic fault [3,5]. Therefore, surface
 Table 1. Focal mechanisms of the 2017 Jiuzhaigou earthquake determined by different institutions.
 deformation as a direct response to the earthquake provides critical information for stud-
 ying the seismogenic
 Nodalfault
 Plane and
 I rupture model.
 Nodal PlaneAs II one of the important means for obtain-
 ing surface deformation, interferometric synthetic aperture Magnitude (Mw)
 radar (InSAR)Depth (km)ad-
 has the
 Strike, Dip, Rake Strike, Dip, Rake
 vantages of a wide◦geographical span, high precision, and high spatial resolution [7]. It is
 GCMT 242 , 77◦ , −168◦ 150◦ , 78◦ , −13◦
 widely used to decipher earthquake cycle processes, including6.5 14.9
 the interseismic locking [8],
 USGS 246◦ , 57◦ , −173◦ 153◦ , 84◦ , −33◦ 6.5 13.5
 coseismic
 CEA-IGP rupture [9,10], and
 328◦ , 48◦ , −11◦ postseismic afterslip
 65◦ , 82◦ , 137◦ and viscoelastic
 6.5 relaxation [11].
 20.0 Re-
 CEA-IEF 59◦ , 77◦ , 164◦ 152◦ , 75◦ , 13◦ 6.5 5.0
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 3 of 12

 No significant coseismic surface rupture was observed for the Jiuzhaigou earthquake,
 making it challenging to determine the seismogenic fault [3,5]. Therefore, surface defor-
 mation as a direct response to the earthquake provides critical information for studying
 the seismogenic fault and rupture model. As one of the important means for obtaining
 surface deformation, interferometric synthetic aperture radar (InSAR) has the advantages
 of a wide geographical span, high precision, and high spatial resolution [7]. It is widely
 used to decipher earthquake cycle processes, including the interseismic locking [8], co-
 seismic rupture [9,10], and postseismic afterslip and viscoelastic relaxation [11]. Recently,
 InSAR has been applied to monitor induced earthquakes, which is helpful to scientifically
 guide industrial production [12,13]. After the Jiuzhaigou earthquake, many studies on
 surface deformation and seismotectonics have been carried out using InSAR technology.
 Shan et al. [14] used Sentinel-1A data to acquire the coseismic surface deformation fields
 and inverted the rupture model. Hong et al. [15] derived the dislocation distribution from
 InSAR and Global Positioning System (GPS) data. Zhao et al. [16] retrieved the coseis-
 mic deformation field from Sentinel-1A InSAR data. Although a lot of research has been
 performed on the seismogenic fault of the Jiuzhaigou earthquake, a unified knowledge
 has not been achieved. From previous studies, we know that different one-fault models
 have a wide range of dips (50◦ –84◦ ) and a small range of strikes (152◦ –155◦ ) [1,5,9,14–20].
 The two-fault models using the aftershock gap as the boundary have a small difference
 in fault geometry on the southern fault (strike: 145◦ –148◦ , dip: 84◦ –88◦ ), but a significant
 difference on the northern fault (strike: 151◦ –171◦ , dip: 77◦ –88◦ ) [2,21,22]. Sun et al. [10]
 proposed a more complex three-fault model to resolve the crustal deformation, including
 a fault branch forming an obtuse angle with the main subvertical fault at the northern
 end [23,24]. Thus, it is necessary to further discuss the seismogenic fault and rupture
 model of the Jiuzhaigou earthquake. Accurate coseismic slip distribution could provide
 a reasonable driving source for the postseismic viscoelastic relaxation analysis, Coulomb
 stress change assessment, and ground motion simulation. Moreover, it is of great practical
 significance for improving our knowledge of the background tectonics of the source region
 and scientifically guiding earthquake relief efforts.
 In this study, four finite fault models are tested, and a brand-new two-fault model
 composed of a main fault and a secondary fault is finally determined as the optimal model
 of the 2017 Jiuzhaigou earthquake, elucidating a complex fault system. In addition, we
 further discuss the possible reasons for the aftershock distribution and the occurrence of
 the aftershock gap.

 2. Data and Methods
 2.1. Data
 We derived the coseismic deformation field in the line-of-sight direction based on the
 Sentinel-1 synthetic aperture radar (SAR) images following the strategies of Ji et al. [3].
 Interferometric processing was carried out using the GAMMA software. Two images,
 including a master image (20170730) and a slave image (20170811), from the ascending
 Path128 were used to obtain the interferogram based on the two-pass InSAR method [25].
 The shuttle radar topography mission (SRTM) digital elevation model (DEM) with 30 m
 resolution was used to remove the topographic phase. In order to suppress the noise, the
 interferogram was multi-looked by a factor of 8 in range and 2 in azimuth and filtered
 twice using the weighted power spectrum algorithm with windows of 128 × 128 and then
 32 × 32 pixels [26]. The minimum cost flow algorithm was used to phase unwrapping.
 The residual orbital phase in the interferogram was removed by a nonlinear least-squares
 adjustment [27]. To minimize the atmospheric signal caused by the layered atmosphere,
 the atmospheric delay model was established based on the existing digital elevation model
 and removed from the interferogram [27].
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 4 of 12

 2.2. Methods
 According to the Okada elastic dislocation theory [28], there is a linear relationship
 between the coseismic surface displacement and the fault dislocation distribution. Ideally,
 we would directly use the least-squares method to invert the fault dislocation distribution.
 However, due to practical limitations, to describe the fault more objectively, we need
 to discretize it into many meshes. Moreover, the distribution of the surface observation
 is usually heterogeneous, and the Green function matrix is often rank-deficient or ill-
 conditional. Generally, additional constraints are needed to ensure the stability of the
 inversion, and a dislocation smoothing constraint is usually used:

 F (s) = |Gs − d| 2 + α2 |Lτ | 2 → min,

 where G, s, and d represent the Green function, the slip vector of each sub-fault, and the
 displacement vector, respectively, and ||Gs − d||2 is the variance between the observations
 and predictions. The second term is the smoothing constraint, where L represents the
 finite difference approximation of the Laplacian operator multiplied by a weighting factor
 proportional to the slip amplitude, and τ is the shear stress drop that is linearly related to
 the slip distribution. The parameter α is the smoothing factor, which controls the trade-off
 between the model roughness and misfit. Generally, the “inflection point” is taken as the
 optimal value. We used the steepest descent method (SDM) [29–32] to estimate the slip
 model of the Jiuzhaigou earthquake. Compared with other inversion methods, the SDM
 has the advantage of providing stable inversion results with high efficiency. The Green
 function used for the inversion was based on the Crust1.0 layered model [33].

 3. Model Tests and Results
 We first constructed a one-fault model (Figure 2a) to retrieve the surface deformation.
 To obtain a high-resolution finite fault slip model, we divided the entire fault into 247 rect-
 angular sub-faults with a grid of 2 km × 2 km. As shown in Figure S1, we choose an optimal
 smoothing factor of 0.1 based on the trade-off curve. The optimal strike and dip are 154◦ and
 51◦ , respectively, which is similar to the results of Shan et al. [14] (strike = 153◦ , dip = 50◦ ).
 The one-fault model reproduces most of the observations, but it cannot decipher the surface
 deformation on the southeastern end of the fault satisfactorily (Table 2, Figure 3c). The slip
 distribution is mainly dominated by sinistral strike slips, with some thrust components on
 the southern segment and small normal components on the northern segment (Figure 4a).
 Unfortunately, this model has a significant shallow slip on the southern end of the fault,
 which is not consistent with the fact that there was no surface rupture.
 Table 2. Parameters of finite fault models.

 Average Error Maximum Moment Magnitude
 Faults Strike Dip
 (m) Residual (m) (Mw)
 One-fault Main fault 154◦ 51◦ 0.0211 −0.0900 6.4
 Northern fault 154◦ 66◦
 Two-fault 0.0203 −0.0684 6.4
 Southern fault 147◦ 74◦
 Northern fault 156◦ 77◦
 Three-fault Southern fault 147◦ 82◦ 0.0198 −0.0602 6.4
 Secondary fault 204◦ 80◦
 Main fault 151◦ 77◦
 New two-fault 0.0188 −0.0487 6.5
 Secondary fault 196◦ 77◦
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 5 of 12
 Remote Sens. 2021, 13, x FOR PEER REVIEW 5 of 14

 Remote Sens. 2021, 13, x FOR PEER REVIEW 6 of 14

 80°) (Figure S3), improving the data fit in the southwestern region compared to the two-
 fault model (Table 2, Figure 3h). Figure 4c shows the slip distribution, whose high-slip
 zone is located at a depth of 2–10 km on the main fault, with a peak slip of 1.30 m. The slip
 of the northern fault is composed of sinistral strike-slip and normal motion, while the
 southern fault is almost exclusively composed of sinistral strike-slip motion. We conclude
 that the average slip of the northern fault is larger than that of the southern fault, and that
 the secondary fault extends to the northeast of the main fault, which is consistent with the
 large number of aftershocks that occurred in this region.

 Table 2. Parameters of finite fault models.

 Average Error Maximum Resid- Moment Magni-
 Faults Strike Dip
 (m) ual (m) tude ( )
 One-fault Main fault 154° 51° 0.0211 −0.0900 6.4
 Northern fault 154° 66°
 Two-fault 0.0203 −0.0684 6.4
 Southern fault 147° 74°
 Northern fault 156° 77°
 Three-fault Figure 2. Projections
 Southernoffault
 the four finite147°
 fault models82°
 on the surface,
 0.0198including the−0.0602
 (a) one-fault 6.4
 Figure (b)Projections
 model, 2. of the
 two-fault model four finite
 (a northern fault
 fault and modelsfault),
 a southern on the (c) surface,
 three-faultincluding
 model, and the
 (d) (a) one-fault model,
 Secondary fault 204° 80°
 brand-new two-fault
 (b) two-faultMain model model (a main
 (a northern151°fault and a secondary
 fault and 77° fault). Other symbols are the same as
 a southern fault), (c) three-fault model, and (d) brand-new
 Figure 1b. fault
 New two-fault
 two-fault Secondary
 model (a main fault and a secondary 0.0188 −0.0487 6.5
 fault 196° 77° fault). Other symbols are the same as Figure 1b.
 In addition, given that the depth of the aftershocks is different in the north and the
 south and that there is a seismic gap [6], we also constructed a two-fault model similar to
 the previous studies (Figure 2b, [2,3,19,21,22,34]). The trade-off curve is shown in Figure
 S2a, and we set the smoothing factor as 0.15. Results demonstrate that the optimal model
 is composed of a northern fault (strike = 154°, dip = 66°) and a southern fault (strike = 147°,
 dip = 74°) (Figure S2). Our preferred strikes are similar to the model of Zheng et al. [2],
 and the optimal dip in the north is slightly smaller than that in the south [3,6,21,34]. Figure
 3f shows that the two-fault model further improves the fit of the data (Table 2). It has a
 similar slip pattern to the results of joint inversion by Zheng et al. [2], whose average slip
 on the northern fault is larger than that on the southern fault, with a peak slip of around
 0.9 m. The differences are that our results have a partial normal slip component on the
 northern fault, a smaller slip on the aftershock gap, and no significant slip near the surface
 (Figure 4b). This is consistent with the fact that the Jiuzhaigou earthquake did not result
 in surface rupture.
 Based on the coseismic deformation field and the aftershock distribution, Sun et al.
 [10] proposed a fault branch at the northern end of the main fault plane at an obtuse angle.
 The geological survey also illustrates that there is indeed a secondary fault in this area
 [35]. Therefore, we introduced a similar secondary fault to construct a three-fault model
 (Figure 2c). During the inversion, we applied the same smoothing factor as the two-fault
 model. The optimal three-fault model is composed of a northern fault (strike = 156°, dip =
 77°), a southern fault (strike = 145°, dip = 82°), and a secondary fault (strike = 204°, dip =

 Figure 3. Comparison between interferometric synthetic aperture radar (InSAR) observations and predictions from the
 Figure
 three test models. 3. Comparison
 (a,d,g) between
 Coseismic line-of-sight interferometric
 (LOS) synthetic
 deformation. Simulated aperture
 LOS radar
 deformation from(InSAR) observations and
 the (b) one-fault
 model, (e) two-fault model, and (h) three-fault model. (c,f,i) Residuals. Red star represents the epicenter
 predictions from the three test models. (a,d,g) Coseismic line-of-sight (LOS) deformation. of the 2017 Jiu- Simulated
 zhaigou earthquake from the CEA-IGP.
 LOS deformation from the (b) one-fault model, (e) two-fault model, and (h) three-fault model. (c,f,i)
 Residuals. Red star represents the epicenter of the 2017 Jiuzhaigou earthquake from the CEA-IGP.
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 6 of 12
Remote Sens. 2021, 13, x FOR PEER REVIEW 7 of 14

 Figure 4. Slip distribution with arrows delineating the average rake of each sub-fault for the three test models, whose
 Figure 4. Slip distribution with arrows delineating the average rake of each sub-fault for the three test models, whose
 magnitude is color-coded. (a) Slip distribution for the optimal one-fault model. (b) Slip distribution for the optimal two-
 magnitude is color-coded. (a) Slip distribution for the optimal one-fault model. (b) Slip distribution for the optimal two-fault
 fault model, including the (1) northern fault and (2) southern fault. (c) Slip distribution for the optimal three-fault model,
 model, including
 including the (1) northern
 the (1) secondary fault, (2)fault and (2)
 northern southern
 fault, and (3)fault. (c) Slip
 southern fault.distribution for the optimal three-fault model,
 including the (1) secondary fault, (2) northern fault, and (3) southern fault.
 Given that the fault geometry of the southern fault is similar to that of the northern
 In addition, given that the depth of the aftershocks is different in the north and the
 fault in the optimal three-fault model, we proposed a simpler two-fault model including
 south and that there is a seismic gap [6], we also constructed a two-fault model similar to the
 a main fault and a secondary fault that is at an obtuse angle to the northern end of the
 previous studies (Figure 2b, [2,3,19,21,22,34]). The trade-off curve is shown in Figure S2a,
 main fault (Figure 2d). The trade-off curve shows that the optimal smoothing factor is
 and we set the smoothing factor as 0.15. Results demonstrate that the optimal model is
 consistent with the three-fault model. Figure 5 illustrates that the optimal strikes of the
 composed of a northern fault (strike = 154◦ , dip = 66◦ ) and a southern fault (strike = 147◦ ,
 main fault (Seg1) and the secondary fault (Seg2) are 151° and 196°, respectively, and that
 dip = 74◦ ) (Figure S2). Our preferred strikes are similar to the model of Zheng et al. [2], and
 their dips are
 the optimal dipboth 77°.
 in the Interestingly,
 north this brand-new
 is slightly smaller than that two-fault
 in the southmodel has theFigure
 [3,6,21,34]. best fit
 3f
 among the four models (Table 2). The southeastern source region, which
 shows that the two-fault model further improves the fit of the data (Table 2). It has a similar is poorly recov-
 ered by the other
 slip pattern to themodels,
 resultsisofwelljointconstrained
 inversion by by Zheng
 this model
 et al.(Figure 6). Results
 [2], whose average showslipthat
 on
 18
 the released seismic moment is around 6.3 × 10 N∙m, equivalent
 the northern fault is larger than that on the southern fault, with a peak slip of around to an Mw 6.5 event,
 consistent
 0.9 m. Thewith the results
 differences areof several
 that different
 our results institutions
 have a partial (Table
 normal1).slip
 Most of the coseismic
 component on the
 slip on the
 northern main
 fault, fault occurred
 a smaller slip on thearound two shallow
 aftershock gap, and asperities, whichslip
 no significant arenear
 separated by a
 the surface
 seismic gap with few aftershocks. The rupture is dominated by sinistral
 (Figure 4b). This is consistent with the fact that the Jiuzhaigou earthquake did not result in slip, accompanied
 by a small
 surface number of normal components. The high-slip regions are distributed between
 rupture.
 2–8 kmBased ona the
 with peak slip of 1.51
 coseismic m, consistent
 deformation field with no aftershock
 and the surface rupture (FigureSun
 distribution, 7c).etInal.
 addi-
 [10]
 tion, most aftershocks are concentrated under the high-slip patches
 proposed a fault branch at the northern end of the main fault plane at an obtuse an- and appear a comple-
 mentary
 gle. The pattern
 geologicalwith the coseismic
 survey slip. Interestingly,
 also illustrates that there isthe slip on
 indeed the secondary
 a secondary faultfault is
 in this
 also
 areamainly concentrated
 [35]. Therefore, we in two patches,
 introduced but it is
 a similar separatedfault
 secondary by antoaftershock
 construct swarm. They
 a three-fault
 extend up to 112c).
 model (Figure km, with athe
 During peak slip of 0.85
 inversion, m, coinciding
 we applied the same with the depth
 smoothing of aftershocks
 factor as the two-
 (Figures 2d and 7b). Therefore, we propose that this two-fault
 fault model. The optimal three-fault model is composed of a northern fault (strike model is our = 156◦ ,
 preferred
 dip = 77 ), a southern fault (strike = 145 , dip = 82 ), and a secondary fault (strike = 204◦ ,
 model for◦ the Jiuzhaigou earthquake. ◦ ◦

 dip = 80◦ ) (Figure S3), improving the data fit in the southwestern region compared to the
 two-fault model (Table 2, Figure 3h). Figure 4c shows the slip distribution, whose high-slip
 zone is located at a depth of 2–10 km on the main fault, with a peak slip of 1.30 m. The slip
 of the northern fault is composed of sinistral strike-slip and normal motion, while the
 southern fault is almost exclusively composed of sinistral strike-slip motion. We conclude
 that the average slip of the northern fault is larger than that of the southern fault, and that
 the secondary fault extends to the northeast of the main fault, which is consistent with the
 large number of aftershocks that occurred in this region.
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 7 of 12

 Given that the fault geometry of the southern fault is similar to that of the northern
 fault in the optimal three-fault model, we proposed a simpler two-fault model including a
 main fault and a secondary fault that is at an obtuse angle to the northern end of the main
 fault (Figure 2d). The trade-off curve shows that the optimal smoothing factor is consistent
 with the three-fault model. Figure 5 illustrates that the optimal strikes of the main fault
 (Seg1) and the secondary fault (Seg2) are 151◦ and 196◦ , respectively, and that their dips
 are both 77◦ . Interestingly, this brand-new two-fault model has the best fit among the four
 models (Table 2). The southeastern source region, which is poorly recovered by the other
 models, is well constrained by this model (Figure 6). Results show that the released seismic
 moment is around 6.3 × 1018 N·m, equivalent to an Mw 6.5 event, consistent with the
 results of several different institutions (Table 1). Most of the coseismic slip on the main fault
 occurred around two shallow asperities, which are separated by a seismic gap with few
 aftershocks. The rupture is dominated by sinistral slip, accompanied by a small number of
 normal components. The high-slip regions are distributed between 2–8 km with a peak slip
 of 1.51 m, consistent with no surface rupture (Figure 7c). In addition, most aftershocks are
 concentrated under the high-slip patches and appear a complementary pattern with the
 coseismic slip. Interestingly, the slip on the secondary fault is also mainly concentrated in
 two patches, but it is separated by an aftershock swarm. They extend up to 11 km, with a
 Remote Sens. 2021, 13, x FOR PEER REVIEW 8 of 14
 peak
 Remote Sens. 2021, 13, x FOR PEER REVIEW slip of 0.85 m, coinciding with the depth of aftershocks (Figures 2d and 7b). Therefore,
 8 of 14
 we propose that this two-fault model is our preferred model for the Jiuzhaigou earthquake.

 Figure 5. Search results of our preferred model. (a) Trade-off curve between the model roughness and misfit. Dips of the
 Figure
 Figure5. Search results of our preferred model. (a)(a)
 Trade-off curve between thethemodel roughness and misfit. Dips of the (b)
 (b) main5.fault
 Search results
 (Seg1) and of(c)our preferred
 secondary model.
 fault (Seg2). Trade-off
 Strikes curve
 of the (d) between model
 main fault (Seg1) roughness
 and and
 (e) secondary misfit. Dips
 fault (Seg2).of the
 main fault (Seg1) and (c) secondary fault (Seg2). Strikes of the (d) main fault (Seg1) and (e) secondary fault
 (b) main fault (Seg1) and (c) secondary fault (Seg2). Strikes of the (d) main fault (Seg1) and (e) secondary fault (Seg2).(Seg2).

 Figure 6. Comparison between InSAR observations and predictions from our preferred model. (a) Coseismic LOS defor-
 Figure 6.
 mation. Comparison between InSAR observations and predictions
 symbolsfrom oursame
 preferred model. (a) Coseismic LOS defor-
 Figure 6.(b)Comparison
 Simulated LOS deformation.
 between InSAR (c) Residuals.
 observations Other arefrom
 and predictions the our as Figure
 preferred 3.model. (a) Coseismic LOS
 mation. (b) Simulated LOS deformation. (c) Residuals. Other symbols are the same as Figure 3.
 deformation. (b) Simulated LOS deformation. (c) Residuals. Other symbols are the same as Figure 3.
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 8 of 12
 Remote Sens. 2021, 13, x FOR PEER REVIEW 9o

 Figure 7. Slip distribution fordistribution
 Figure 7. Slip our preferred formodel. (a) Projection
 our preferred model.of(a)
 theProjection
 slip modelofon
 thethe surface.
 slip modelSlip distribution
 on the surface. of the
 (b) secondary fault, and (c) main fault. The white dots represent the relocated aftershocks. The slip distribution with ar-
 Slip distribution of the (b) secondary fault, and (c) main fault. The white dots represent the relocated
 rows delineating the average rake of each sub-fault is color-coded. Other symbols are the same as Figure 1b.
 aftershocks. The slip distribution with arrows delineating the average rake of each sub-fault is
 color-coded. Other symbols are the same as Figure 1b.
 4. Discussion
 4. Discussion
 4.1. Uncertainty Test
 4.1. Uncertainty Test
 In order to verify the effectiveness of our preferred model, we applied the bootstr
 In order to verify
 method the effectiveness
 to estimate of our preferred
 the uncertainty. model, is
 Bootstrapping wea applied
 statistical theestimation
 bootstrap method
 method to randomly
 estimate the uncertainty.
 resamples Bootstrapping
 from the original dataistoacreate
 statistical estimation
 multiple samples. method.
 That is, m grou
 It randomlyofresamples
 data are from the original
 randomly data
 selected to create
 from multiple
 n groups samples.data
 of original That foris,inversion,
 m groups of and B resu
 data are randomly
 could be selected
 obtainedfrombynrepeating
 groups ofthe
 original data forB inversion,
 experiment times, whichandallows
 B results
 thecould
 standard dev
 be obtainedtions
 by repeating the experiment
 to be inferred B times,
 [36]. Therefore, which
 the allows
 standard the standard
 deviation deviations
 obtained by the bootstr
 to be inferred [36]. Therefore, the standard deviation obtained by the bootstrap
 method not only contains the influence of the original observation error, but also refle method
 not only contains
 the errorthe influence
 caused of the data
 by different original observation
 distribution. error, but
 It is widely used also reflects
 in the the of mo
 estimation
 error caused by different data
 uncertainty [36,37]. distribution. It is widely used in the estimation of model
 uncertainty [36,37].
 To ensure reliable results, the value of B was set to 100 [38], and correspondi
 To ensure reliable results,were
 resampling datasets the value
 createdofinBthe
 was setThese
 case. to 100datasets
 [38], andwere corresponding
 used for inversions w
 resampling the
 datasets were created in the case. These datasets were used
 same model space. Figure 8 shows the estimated uncertainties based onfor inversions with
 the bootstr
 the same model space. Figure 8 shows the estimated uncertainties based on
 method, and the maximum standard deviation is less than 0.22 m. Although the standathe bootstrap
 method, and the maximum
 deviation mainlystandard deviation
 concentrates in theissouthern
 less thanpart
 0.22ofm.the
 Although
 main fault, the standard
 it is small relative
 deviation mainly concentrates in the southern part of the main fault, it
 the peak slip of asperity. Besides, the standard deviation in the aftershock is small relative
 gap is very lo
 to the peakindicating
 slip of asperity. Besides, the standard deviation in the aftershock gap is very
 there is indeed a slip deficit in this region. Overall, the uncertainty of our op
 low, indicating there is indeed
 mal two-fault model ais slip
 verydeficit in this region.
 low, revealing Overall,distribution
 the dislocation the uncertainty of
 is reliable and s
 our optimal two-fault model is very low, revealing the dislocation distribution is reliable
 ble.
 and stable.
PROBING THE FAULT COMPLEXITY OF THE 2017 MS 7.0 JIUZHAIGOU EARTHQUAKE BASED ON THE INSAR DATA - MDPI
Remote Sens. 2021, 13, 1573 9 of 12
Remote Sens. 2021, 13, x FOR PEER REVIEW 10 of 14

 Figure
 Figure 8.
 8. The
 The standard
 standard deviations
 deviations of
 of the
 the (a)
 (a) main
 main fault,
 fault, and
 and (b)
 (b) secondary
 secondary fault, based on
 fault, based on the
 the bootstrap
 bootstrap method.
 method.

 4.2.
 4.2. Fault
 Fault Geometry
 Geometry and and Properties
 Properties
 In
 In our
 ouroptimal
 optimalmodel,
 model,the thesecondary
 secondary fault is atisaatlarge
 fault obtuse
 a large obtuseangleangle
 to thetomain
 the fault,
 main
 termed backward branching. Similar phenomena were also observed
 fault, termed backward branching. Similar phenomena were also observed in the 1992 in the 1992 Landers
 earthquake [39] and the
 Landers earthquake [39]1999
 and Hector
 the 1999 Mine earthquake
 Hector [40]. The mechanism
 Mine earthquake of this back-
 [40]. The mechanism of
 ward branching
 this backward is the rupture
 branching is the stopping or slowing
 rupture stopping down on
 or slowing the main
 down on the fault plane,
 main faultwhich
 plane,
 causes
 which the rupture
 causes to jump to
 the rupture to jump
 a nearbyto afault [39].fault
 nearby This[39].
 phenomenon can also becan
 This phenomenon accommo-
 also be
 dated by a significant
 accommodated shallow slip
 by a significant deficit
 shallow [40].
 slip Apparently,
 deficit no significant
 [40]. Apparently, surface rupture
 no significant surface
 was observed
 rupture in the Jiuzhaigou
 was observed earthquake.
 in the Jiuzhaigou There isThere
 earthquake. a small
 is aslip
 small zoneslipatzone
 the south
 at the of the
 south
 junction between
 of the junction the backward
 between the backward branching
 branchingandandthethe main
 main fault,
 fault,which
 whichmay mayhave
 have been
 caused by a rigid
 caused rigid block
 blockstopping
 stoppingor orslowing
 slowingdowndownthe therupture
 rupture onon thethemain
 main fault. Moreover,
 fault. Moreo-
 the 1992
 ver, Landers
 the 1992 earthquake
 Landers and the
 earthquake and1999
 theHector Mine earthquake
 1999 Hector Mine earthquakeboth occurred within a
 both occurred
 complex
 within a fault
 complexsystem.
 faultThe Jiuzhaigou
 system. earthquakeearthquake
 The Jiuzhaigou appears to be a similar
 appears case,
 to be whose main
 a similar case,
 seismogenic
 whose fault is the hidden
 main seismogenic fault isfaults with nofaults
 the hidden obvious withsurface trace surface
 no obvious [10] or afterslip
 trace [10] [41],
 or
 indicating a young fault [10,41]. This less mature fault system is characterized
 afterslip [41], indicating a young fault [10,41]. This less mature fault system is character- by weak
 faultby
 ized planes
 weakand more
 fault complex
 planes and morerupture processes.
 complex rupture processes.

 4.3. Aftershock
 4.3. Aftershock Distribution
 Distribution
 Based on
 Based onthethedistribution
 distribution of of
 relocated aftershocks,
 relocated therethere
 aftershocks, is an aftershock gap of around
 is an aftershock gap of
 5 km, which may be caused by one or multiple of the following
 around 5 km, which may be caused by one or multiple of the following three factors: three factors: (1) The(1)
 coseismic slip was large and the stress release was sufficient; (2) The strikes of the faults
 The coseismic slip was large and the stress release was sufficient; (2) The strikes of the
 changed at this turning point; (3) There were unbroken asperities [6]. From the InSAR
 faults changed at this turning point; (3) There were unbroken asperities [6]. From the In-
 surface deformation (Figure 3), we find that there are few surface deformations in the
 SAR surface deformation (Figure 3), we find that there are few surface deformations in
 aftershock gap with the indication of no significant coseismic slip below, which is also
 the aftershock gap with the indication of no significant coseismic slip below, which is also
 verified by our preferred slip model. We thus rule out the first possibility. During our finite
 verified by our preferred slip model. We thus rule out the first possibility. During our
 fault tests, the three-fault model suggests that the geometry parameters of the northern
 finite fault tests, the three-fault model suggests that the geometry parameters of the north-
 fault are similar to those of the southern fault, precluding the second possibility. In addition,
 ern fault are similar to those of the southern fault, precluding the second possibility. In
 Li et al. [41] stated that no early afterslip is observed following the Jiuzhaigou earthquake,
 addition, Li et al. [41] stated that no early afterslip is observed following the Jiuzhaigou
 indicating the accumulated strain is unlikely to be released by postseismic deformation in
 earthquake, indicating the accumulated strain is unlikely to be released by postseismic
 the gap. Therefore, we argue that unbroken asperities hindered the coseismic rupturing,
 deformation in the gap. Therefore, we argue that unbroken asperities hindered the coseis-
 resulting in this small seismic gap.
 mic rupturing, resulting in
 Figure 7 illustrates this
 that small
 most seismic gap.
 aftershocks occurred in the down-dip regions of the
 Figurehigh-slip
 coseismic 7 illustrates that most
 asperities, aftershocks
 presenting occurred in the
 a complementary down-dip
 pattern. regions
 Generally, of the
 it can be
 coseismic high-slip asperities, presenting a complementary pattern. Generally,
 interpreted as (1) direct triggering from coseismic stress change [42], or (2) aftershocks it can be
 interpreted as (1) direct triggering from coseismic stress change [42], or
 that are driven by afterslip in the velocity-strengthening zone [30,43]. Therefore, we first (2) aftershocks
 that
 usedarethedriven by afterslip
 numerical in the velocity-strengthening
 code PSGRN/PSCMP [44] followingzone [30,43]. Therefore,
 the strategies of Guo etwe al. first
 [45]
 used the numerical code PSGRN/PSCMP [44] following the strategies
 to estimate the static Coulomb stress changes caused by the coseismic slip. Since of Guo et al. [45]theto
 estimate
 Jiuzhaigou theearthquake
 static Coulomb is a stress changes
 strike-slip caused
 event, we bysetthe
 thecoseismic
 effective slip. Since
 friction the Jiuzhai-
 coefficient to
 gou earthquake is a strike-slip event, we set the effective friction coefficient
 0.4 [24,45–47]. We used the optimal slip model as the driving source. The receiver fault to 0.4 [24,45–
 47].
 was We used
 set to ourthe optimalfault
 preferred slip model.
 model as the driving
 Results show thatsource.
 mostThe receiver fault
 aftershocks was in
 occurred settheto
 our preferred fault model. Results show that most aftershocks occurred in the Coulomb
Remote Sens.
Remote Sens. 2021,
 2021, 13,
 13, 1573
 x FOR PEER REVIEW 11 of
 10 of 12
 14

 stress shadows (Figure S4). Therefore, we propose that the aftershock distribution may be
 Coulomb stress shadows (Figure S4). Therefore, we propose that the aftershock distribution
 related to aseismic afterslip, indicating the velocity-strengthening region. Further research
 may be related to aseismic afterslip, indicating the velocity-strengthening region. Further
 is needed here. Helmstetter and Shaw [48] suggested that the aftershocks may be related
 research is needed here. Helmstetter and Shaw [48] suggested that the aftershocks may be
 to the tectonic stress on the fault, which could explain the inconsistency between the
 related to the tectonic stress on the fault, which could explain the inconsistency between
 mechanisms of the aftershocks and that of the mainshock. Therefore, we calculated the
 the mechanisms of the aftershocks and that of the mainshock. Therefore, we calculated
 static Coulomb
 the static Coulomb stress changes
 stress on the
 changes on optimally
 the optimallyoriented failure
 oriented planes
 failure for the
 planes for 2017 Jiu-
 the 2017
 zhaigou
 Jiuzhaigou earthquake
 earthquake under thethe
 under tectonic background
 tectonic background stress field.
 stress In this
 field. study,
 In this we set
 study, we the
 set
 maximum principal stress, middle principal stress, and minimum principal
 the maximum principal stress, middle principal stress, and minimum principal stress stress to −10to
 4

 kPa,
 −104−10kPa,kPa,
 3
 −103and
 kPa,0 and
 kPa,0respectively, basedbased
 kPa, respectively, on previous studies
 on previous [23,24].
 studies Although
 [23,24]. Althoughthe
 tectonic stresses may have large uncertainties, it does not prevent us from
 the tectonic stresses may have large uncertainties, it does not prevent us from analyzing analyzing them
 rationally. FigureFigure
 them rationally. 9 reveals the static
 9 reveals theCoulomb stress changes
 static Coulomb at different
 stress changes depths and
 at different the
 depths
 distribution of the aftershocks
 and the distribution within 5within
 of the aftershocks km, from
 5 km,which
 from we
 whichfindwethat most
 find thatof the of
 most after-
 the
 shocks occurred in the region where the Coulomb stress increased. Therefore,
 aftershocks occurred in the region where the Coulomb stress increased. Therefore, the the after-
 shocks of theofJiuzhaigou
 aftershocks earthquake
 the Jiuzhaigou may have
 earthquake maybeenhavecontrolled by the background
 been controlled tectonic
 by the background
 stress field.
 tectonic stress field.

 Figure 9.
 Figure 9. Static
 Static Coulomb
 Coulomb stress changes on
 stress changes on the
 the 3D
 3D optimally
 optimally oriented failure planes
 oriented failure planes created
 created by
 by the
 the Jiuzhaigou
 Jiuzhaigou earthquake
 earthquake
 at different depths and the distribution of the aftershocks within 5 km. Coulomb stress changes at a depth of (a) 5 km, (b)
 at different depths and the distribution of the aftershocks within 5 km. Coulomb stress changes at a depth of (a) 5 km,
 10 km, and (c) 15 km. Black dots represent the relocated aftershocks.
 (b) 10 km, and (c) 15 km. Black dots represent the relocated aftershocks.
 5.
 5. Conclusions
 Conclusions
 In
 In this
 this study, we tested
 study, we tested four
 four different
 different finite
 finite fault
 fault models
 models based
 based on
 on the
 the InSAR
 InSAR data,
 data,
 and
 and determined
 determined that
 that the
 the seismogenic fault of
 seismogenic fault of the
 the 2017
 2017 MwMw 6.5
 6.5 Jiuzhaigou earthquake was
 Jiuzhaigou earthquake was
 a brand-new
 brand-new two-fault
 two-faultmodel.
 model.The Thehigh-slip
 high-slip regions
 regionsin in
 thethe
 main fault
 main are are
 fault distributed
 distributedbe-
 tween
 between 2–82–8
 km, with
 km, witha peak slipslip
 a peak of 1.51 m. m.
 of 1.51 There are are
 There twotwo
 asperities thatthat
 asperities are are
 perfectly de-
 perfectly
 marcated
 demarcated by by
 thethe
 aftershock
 aftershock gap, which
 gap, may
 which mayindicate a barrier.
 indicate TheThe
 a barrier. secondary faultfault
 secondary is atisa
 large
 at a large obtuse angle to the main fault, whose slip extends to 11 km, with a peak 0.85
 obtuse angle to the main fault, whose slip extends to 11 km, with a peak slip at slip
 m. The m.
 at 0.85 mainThefault
 mainoffault
 the Jiuzhaigou earthquake
 of the Jiuzhaigou is the NW-trending
 earthquake is the NW-trendinghiddenhidden
 fault offault
 the
 Huya
 of the fault,
 Huyawhich
 fault, is one ofisthe
 which onebranches of the East
 of the branches of Kunlun
 the Eastfault.
 Kunlun There were
 fault. a fewwere
 There after-
 a
 shocks at both ends
 few aftershocks of the
 at both main
 ends fault
 of the where
 main faultthe Coulomb
 where stress increased,
 the Coulomb which deserve
 stress increased, which
 deserve
 our our particular
 particular attention. attention.

 Supplementary Materials:
 Materials: The
 Thefollowing
 followingare
 are available
 available online
 online at www.mdpi.com/xxx/s1, Figure
 at https://www.mdpi.com/article/10
 S1—Search results of the
 .3390/rs13081573/s1, one-fault
 Figure model;
 S1—Search Figure
 results ofS2—Search results
 the one-fault of Figure
 model; the two-fault model;
 S2—Search Figure
 results of
 S3—Search results
 the two-fault model;of Figure
 the three-fault
 S3—Search model; Figure
 results S4—Static
 of the Coulomb
 three-fault model;stress changes
 Figure caused
 S4—Static by the
 Coulomb
 Jiuzhaigou earthquake
 stress changes caused byat the
 different depthsearthquake
 Jiuzhaigou and the distribution ofdepths
 at different the aftershocks within 5 km.of the
 and the distribution
 aftershocks within 5 km.
 Author Contributions: Conceptualization, R.G. and X.T.; methodology, J.X.; software, R.G.; valida-
 tion,
 AuthorJ.X.,Contributions:
 R.G., H.S., X.C.Conceptualization,
 and J.Z.; formal analysis,
 R.G. andR.G.; writing—original
 X.T.; methodology, J.X.; draft preparation,
 software, R.G.;X.T.;
 val-
 writing—review and editing, X.T., J.X., R.G., H.S., X.C. and J.Z.; visualization, X.T. and
 idation, J.X., R.G., H.S., X.C. and J.Z.; formal analysis, R.G.; writing—original draft preparation, R.G.; super-
 vision, J.X., H.S., X.C. and
 X.T.; writing—review and J.Z.; project
 editing, administration,
 X.T., J.X., R.G., H.S.,X.C.
 X.C.and
 and J.Z.; funding
 J.Z.; acquisition,
 visualization, X.T. J.X., H.S.,
 and R.G.;
 X.C. and J.Z. J.X.,
 supervision, All authors have
 H.S., X.C. andread
 J.Z.;and agreed
 project to the published
 administration, X.C.version offunding
 and J.Z.; the manuscript.
 acquisition, J.X.,
 H.S., X.C. This
 Funding: and J.Z. All authors
 research have read
 was funded andB-type
 by the agreed to the published
 Strategic version of
 Priority Program ofthe
 themanuscript.
 Chinese Acad-
 emy of Sciences, grant number XDB41000000, and the National Natural Science Foundation of
 China, grant numbers 41974023, 41874094, 41674083, 41874026.
Remote Sens. 2021, 13, 1573 11 of 12

 Funding: This research was funded by the B-type Strategic Priority Program of the Chinese Academy
 of Sciences, grant number XDB41000000, and the National Natural Science Foundation of China,
 grant numbers 41974023, 41874094, 41674083, 41874026.
 Institutional Review Board Statement: Not applicable.
 Informed Consent Statement: Not applicable.
 Data Availability Statement: Not applicable.
 Acknowledgments: We thank the National Earth System Science Data Center, National Science &
 Technology Infrastructure of China (http://www.geodata.cn, accessed on 14 April 2021) for the
 data provided. We are grateful to Lingyun Ji of the Second Monitoring Center, China Earthquake
 Administration for providing the InSAR data. We used the Generic Mapping Tools (GMT) open-
 source collection of computer software tools to create the figures, which is developed and maintained
 by Paul Wessel and Walter H. F. Smith.
 Conflicts of Interest: The authors declare no conflict of interest.

References
1. Li, Q.; Tan, K.; Wang, D.Z.; Zhao, B.; Zhang, R.; Li, Y.; Qi, Y.J. Joint inversion of GNSS and teleseismic data for the rupture process
 of the 2017 Mw6.5 Jiuzhaigou, China, earthquake. J. Seism. 2018, 22, 805–814. [CrossRef]
2. Zheng, A.; Yu, X.; Xu, W.; Chen, X.; Zhang, W. A hybrid source mechanism of the 2017 Mw 6.5 Jiuzhaigou earthquake revealed by
 the joint inversion of strong-motion, teleseismic and InSAR data. Tectonophysics 2020, 789, 228538. [CrossRef]
3. Ji, L.; Liu, C.; Xu, J.; Liu, L.; Long, F.; Zhang, Z. InSAR observation and inversion of the seismogenic fault for the 2017 Jiuzhaigou
 MS7.0 earthquake in China. Chin. J. Geophys. 2017, 60, 4069–4082.
4. Deng, Q.; Cheng, S.; Ma, J.; Du, P. Seismic activities and earthquake potential in the Tibetan Plateau. Chin. J. Geophys. 2014, 57,
 2025–2042.
5. Xu, X.; Chen, G.; Wang, Q.; Chen, L.; Ren, Z.; Xu, C.; Wei, Z.; Lu, R.; Tan, X.; Dong, S.; et al. Discussion on seismogenic structure
 of Jiuzhaigou earthquake and its implication for current strain state in the southeastern Qinghai-Tibet Plateau. Chin. J. Geophys.
 2017, 60, 4018–4026.
6. Fang, L.; Wu, J.; Su, J.; Wang, M.; Jiang, C.; Fan, L.; Wang, W.; Wang, C.; Tan, X. Relocation of mainshock and aftershock sequence
 of the Ms7.0 Sichuan Jiuzhaigou earthquake. Chin. Sci. Bull. 2018, 63, 649–662. [CrossRef]
7. Zebker, H.A.; Rosen, P.A.; Goldstein, R.M.; Gabriel, A.; Werner, C.L. On the derivation of coseismic displacement fields using
 differential radar interferometry: The Landers earthquake. J. Geophys. Res. Space Phys. 1994, 99, 19617–19634. [CrossRef]
8. Hussain, E.; Wright, T.J.; Walters, R.J.; Bekaert, D.P.S.; Lloyd, R.; Hooper, A. Constant strain accumulation rate between major
 earthquakes on the North Anatolian Fault. Nat. Commun. 2018, 9, 1–9. [CrossRef]
9. Nie, Z.; Wang, D.-J.; Jia, Z.; Yu, P.; Li, L. Fault model of the 2017 Jiuzhaigou Mw 6.5 earthquake estimated from coseismic
 deformation observed using Global Positioning System and Interferometric Synthetic Aperture Radar data. Earthplanets Space
 2018, 70, 55. [CrossRef]
10. Sun, J.; Yue, H.; Shen, Z.; Fang, L.; Zhan, Y.; Sun, X. The 2017 Jiuzhaigou Earthquake: A Complicated Event Occurred in a Young
 Fault System. Geophys. Res. Lett. 2018, 45, 2230–2240. [CrossRef]
11. Zhao, D.; Qu, C.; Bürgmann, R.; Gong, W.; Shan, X. Relaxation of Tibetan Lower Crust and Afterslip Driven by the 2001 Mw7.8
 Kokoxili, China, Earthquake Constrained by a Decade of Geodetic Measurements. J. Geophys. Res. Solid Earth 2021, 126. [CrossRef]
12. Malinowska, A.A.; Witkowski, W.T.; Guzy, A.; Hejmanowski, R. Mapping ground movements caused by mining-induced
 earthquakes applying satellite radar interferometry. Eng. Geol. 2018, 246, 402–411. [CrossRef]
13. Sopata, P.; Stoch, T.; Wójcik, A.; Mrocheń, D. Land Surface Subsidence Due to Mining-Induced Tremors in the Upper Silesian
 Coal Basin (Poland)—Case Study. Remote Sens. 2020, 12, 3923. [CrossRef]
14. Shan, X.; Qu, C.; Gong, W.; Zhao, D.; Zhang, G. Coseismic deformation field of the Jiuzhaigou MS7.0 earthquake from Sentinel-1A
 InSAR data and fault slip inversion. Chin. J. Geophys. 2017, 60, 4527–4536.
15. Hong, S.; Zhou, X.; Zhang, K.; Meng, G.; Dong, Y.; Su, X.; Zhang, L.; Li, S.; Ding, K. Source Model and Stress Disturbance of the
 2017 Jiuzhaigou Mw 6.5 Earthquake Constrained by InSAR and GPS Measurements. Remote Sens. 2018, 10, 1400. [CrossRef]
16. Zhao, D.; Qu, C.; Shan, X.; Gong, W.; Zhang, Y.; Zhang, G. InSAR and GPS derived coseismic deformation and fault model of the
 2017 Ms7.0 Jiuzhaigou earthquake in the Northeast Bayanhar block. Tectonophyicsics 2018, 726, 86–99. [CrossRef]
17. Chen, W.; Qiao, X.J.; Liu, G.; Xiong, W.; Jia, Z.G.; Li, Y.; Wang, Y.B.; You, Z.L.; Long, F. Study on the coseismic slip model and
 Coulomb stress of the 2017 Jiuzhaigou MS7.0 earthquake constrained by GNSS and InSAR measurements. Chin. J. Geophys. 2018,
 61, 2122–2132.
18. Shen, W.; Li, Y.S.; Jiao, Q.; Xie, Q.; Zhang, J. Joint inversion of strong motion and InSAR/GPS data for fault slip distribution of the
 Jiuzhaigou 7.0 earthquake and its application in seismology. Chin. J. Geophys. 2019, 61, 115–129.
19. Xie, Z.; Zheng, Y.; Yao, H.; Fang, L.; Liu, C.; Wang, M.; Shan, B.; Zhang, H.; Ren, J.; Ji, L.; et al. Preliminary analysis on the source
 properties and seismogenic structure of the 2017 Ms7.0 Jiuzhaigou earthquake. Sci. China Earth Sci. 2018, 61, 339–352. [CrossRef]
Remote Sens. 2021, 13, 1573 12 of 12

20. Gui-Xi, Y.; Feng, L.; Ming-Jian, L.; Hui-Ping, Z.; Min, Z.; You-Qing, Y.; Zhi-Wei, Z.; Yu-Ping, Q.; Si-Wei, W.; Yue, G.; et al. Focal
 mechanism solutions and seismogenic structure of the 8 August 2017 M7.0 Jiuzhaigou earthquake and its aftershocks, northern
 Sichuan. Chin. J. Geophys. 2017, 60, 4083–4097.
21. Hu, X.; Sheng, S.; Wang, Y.; Liang, S. Fault Plane Parameters of 2017 Jiuzhaigou Ms7. 0 Earthquake Determined by Aftershock
 Distribution. J. Seismol. Res. 2019, 42, 366–371.
22. Liu, G.; Xiong, W.; Wang, Q.; Qiao, X.; Ding, K.; Li, X.; Yang, S. Source Characteristics of the 2017 Ms 7.0 Jiuzhaigou, China,
 Earthquake and Implications for Recent Seismicity in Eastern Tibet. J. Geophys. Res. Solid Earth 2019, 124, 4895–4915. [CrossRef]
23. Toda, S.; Stein, R.S.; Richards-Dinger, K.; Bozkurt, S.B. Forecasting the evolution of seismicity in southern California: Animations
 built on earthquake stress transfer. J. Geophys. Res. Space Phys. 2005, 110, 1–17. [CrossRef]
24. Wang, J.; Xu, C. Coseismic Coulomb stress changes associated with the 2017 MW6.5 Jiuzhaigou earthquake (China) and its
 impacts on surrounding major faults. Chin. J. Geophys. 2017, 60, 4398–4420.
25. Massonnet, D.; Feigl, K.L. Radar interferometry and its application to changes in the Earth’s surface. Rev. Geophys. 1998, 4,
 441–500. [CrossRef]
26. Goldstein, R.M.; Werner, C.L. Radar interferogram filtering for geophysical applications. Geophys. Res. Lett. 1998, 25, 4035–4038.
 [CrossRef]
27. Rosen, P.A.; Hensley, S.; Zebker, H.A.; Webb, F.H.; Fielding, E.J. Surface deformation and coherence measurements of Kilauea
 Volcano, Hawaii, from SIR-C radar interferometry. J. Geophys. Res. Space Phys. 1996, 101, 23109–23125. [CrossRef]
28. Okada, Y. Surface deformation due to shear and tensile faults in a half-space. B Seismol. Soc. Am. 1985, 75, 1135–1154.
29. Guo, R.; Zheng, Y.; Diao, F.; Xu, J. Rupture model of the 2013 M W 6.6 Lushan (China) earthquake constrained by a new GPS data
 set and its effects on potential seismic hazard. Earthq. Sci. 2018, 31, 117–125.
30. Guo, R.; Zheng, Y.; Xu, J.; Jiang, Z. Seismic and Aseismic Fault Slip Associated with the 2017 M w 8.2 Chiapas, Mexico, Earthquake
 Sequence. Seismol. Res. Lett. 2019, 90, 1111–1120. [CrossRef]
31. Wang, R.; Diao, F.; Hoechner, A. SDM—A geodetic inversion code incorporating with layered crust structure and curved fault
 geometry. In Proceedings of the EGU General Assembly 2013, Vienna, Austria, 7–12 April 2013.
32. Wang, R.; Martín, F.L.; Roth, F. Computation of deformation induced by earthquakes in a multi-layered elastic crust—FORTRAN
 programs EDGRN/EDCMP. Comput. Geosci. UK 2003, 29, 195–207. [CrossRef]
33. Laske, G.; Masters, G.; Ma, Z.; Pasyanos, M.E. CRUST1.0: An updated global model of Earth’s crust. In Proceedings of the EGU
 General Assembly 2012, Vienna, Austria, 22–27 April 2012.
34. Wang, Y.; Zhao, T.; Li, C.X.; Liu, C. Relocations and focal mechanism solutions of the 2017 Jiuzhaigou, Sichuan MS 7.0 earthquake
 sequence. Process Geophys. 2019, 34, 469–478.
35. Yi, S.-J.; Wu, C.-H.; Li, Y.-S.; Huang, C. Source tectonic dynamics features of Jiuzhaigou Ms 7.0 earthquake in Sichuan Province,
 China. J. Mt. Sci. 2018, 15, 2266–2275. [CrossRef]
36. Schnaidt, S.; Heinson, G. Bootstrap resampling as a tool for uncertainty analysis in 2-D magnetotelluric inversion modelling.
 Geophys. J. Int. 2015, 203, 92–106. [CrossRef]
37. McLaughlin, K.L. Maximum-likelihood event magnitude estimation with bootstrapping for uncertainty estimation. B Seismol.
 Soc. Am. 1988, 78, 855–862.
38. Efron, B.; Tibshirani, R. Bootstrap Methods for Standard Errors, Confidence Intervals, and Other Measures of Statistical Accuracy.
 Stat. Sci. 1986, 1, 54–75. [CrossRef]
39. Fliss, S.; Bhat, H.S.; Dmowska, R.; Rice, J.R. Fault branching and rupture directivity. J. Geophys. Res. Space Phys. 2005, 110, 110.
 [CrossRef]
40. Oglesby, D.D.; Day, S.M.; Li, Y.; Vidale, J.E. The 1999 Hector Mine Earthquake: The Dynamics of a Branched Fault System. B.
 Seismol. Soc. Am. 2003, 93, 2459–2476. [CrossRef]
41. Li, Y.; Bürgmann, R.; Zhao, B. Evidence of Fault Immaturity from Shallow Slip Deficit and Lack of Postseismic Deformation of the
 2017 Mw 6.5 Jiuzhaigou Earthquake. Bull. Seism. Soc. Am. 2020, 110, 154–165. [CrossRef]
42. Stein, R.S.; King, G.C.P.; Lin, J. Stress Triggering of the 1994 M = 6.7 Northridge, California, Earthquake by Its Predecessors.
 Science 1994, 265, 1432–1435. [CrossRef]
43. Guo, R.; Zheng, Y.; An, C.; Xu, J.; Jiang, Z.; Zhang, L.; Riaz, M.S.; Xie, J.; Dai, K.; Wen, Y. The 2018 Mw 7.9 Offshore Kodiak, Alaska,
 Earthquake: An Unusual Outer Rise Strike-Slip Earthquake. J. Geophys. Res. Solid Earth 2020, 125, e2019JB019267.
44. Wang, R.; Lorenzo-Martín, F.; Roth, F. PSGRN/PSCMP—a new code for calculating co- and post-seismic deformation, geoid and
 gravity changes based on the viscoelastic-gravitational dislocation theory. Comput. Geosci. 2006, 32, 527–541. [CrossRef]
45. Guo, R.; Zheng, Y.; Xu, J. Stress modulation of the seismic gap between the 2008 M s 8.0 Wenchuan earthquake and the 2013 M s
 7.0 Lushan earthquake and implications for seismic hazard. Geophys. J. Int. 2020, 221, 2113–2125. [CrossRef]
46. Lin, X.; Chu, R.; Zeng, X. Rupture processes and Coulomb stress changes of the 2017 Mw 6.5 Jiuzhaigou and 2013 Mw 6.6 Lushan
 earthquakes. Earthplanets Space 2019, 71, 1–15. [CrossRef]
47. Shan, B.; Zheng, Y.; Liu, C.; Xie, Z.; Kong, J. Coseismic Coulomb failure stress changes caused by the 2017 M7.0 Jiuzhaigou
 earthquake, and its relationship with the 2008 Wenchuan earthquake. Sci. China Earth Sci. 2017, 60, 2181–2189. [CrossRef]
48. Helmstetter, A.; Shaw, B.E. Relation between stress heterogeneity and aftershock rate in the rate-and-state model. J. Geophys. Res.
 Space Phys. 2006, 111. [CrossRef]
You can also read