Sei sulla pagina 1di 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/332755848

Fault source of the 2 September 2009 Mw 6.8 Tasikmalaya intraslab


earthquake, Indonesia: Analysis from GPS data inversion, tsunami height
simulation, and stress transfer

Article in Physics of The Earth and Planetary Interiors · April 2019


DOI: 10.1016/j.pepi.2019.04.004

CITATIONS READS

0 91

6 authors, including:

Endra Gunawan Sri Widiyantoro


Bandung Institute of Technology Bandung Institute of Technology
33 PUBLICATIONS 337 CITATIONS 166 PUBLICATIONS 4,363 CITATIONS

SEE PROFILE SEE PROFILE

Gayatri Indah Marliyani Euis Sunarti


Universitas Gadjah Mada Bogor Agricultural University
18 PUBLICATIONS 52CITATIONS 34 PUBLICATIONS 66 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

media gender View project

Earthquake Swarm Monitoring around Mt. Pandan, in Madiun area, East Java Province, Indonesia View project

All content following this page was uploaded by Aditya Riadi Gusman on 01 July 2019.

The user has requested enhancement of the downloaded file.


Ph
ysic
softh
eEa
rthandP
lane
taryIn
terio
rs2
91(2019
)5–
461

Contents lists available at ScienceDirect

Physics of the Earth and Planetary Interiors


journal homepage: www.elsevier.com/locate/pepi

Fault source of the 2 September 2009 Mw 6.8 Tasikmalaya intraslab


earthquake, Indonesia: Analysis from GPS data inversion, tsunami height
simulation, and stress transfer

Endra Gunawana, , Sri Widiyantoroa, Gayatri Indah Marliyanib, Euis Sunartic, Rachmah Idad,
Aditya Riadi Gusmane
a
Global Geophysics Research Group, Faculty of Mining and Petroleum Engineering, Bandung Institute of Technology, Indonesia
b
Geological Engineering Department, Faculty of Engineering, Gadjah Mada University, Indonesia
c
Faculty of Human Ecology, Bogor Agricultural University, Indonesia
d
Faculty of Social Sciences and Political Sciences, Airlangga University, Indonesia
e
GNS Science, Lower Hutt, New Zealand

ARTICLEINFO ABSTRACT

Keywords: We estimate the fault model of the 2 September 2009 Tasikmalaya intraslab earthquake based on the GPS data
Coseismic slip available in western Java, Indonesia. The focal mechanism of the earthquake was used to help construct two
Intraslab earthquake possible fault models: a west-dipping fault with a strike of 160.8° and an east-dipping fault with a strike of 34.0°.
GPS data In this study, vertical information from GPS data is crucial for constructing the top depth of the fault. The
Tsunami simulation subsidence information from GPS data located near the epicenter suggests that the earthquake involved a deeper
Stress transfer
fault model. While the amount of the moment release of the east-dipping fault (Model dipE) is equivalent to Mw
6.9, the moment release of the west-dipping fault (Model dipW) is equivalent to Mw 6.8. The GPS data inversion
indicates that Model dipW produces a better fit than Model dipE. The tsunami simulation indicates that the
tsunami height generated by the east-dipping fault is smaller than that generated by the west-dipping fault,
implying that the maximum tsunami height of the latter is closer to agreement with the reported one. Unlike
Model dipE, the stress transfer analysis of Model dipW indicates that most of the aftershocks were located in the
region where ΔCFF is positive, suggesting positive stress from the ruptured triggered aftershocks. The combined
analysis of GPS data, tsunami simulation, and stress transfer suggests that the fault ruptured during the 2009
earthquake was dipping westward with a steep dip angle.

1. Introduction ~200 km (Fujii and Satake, 2006), with the western region from the
fault identified as coupled and potentially triggered a Mw 8.7 earth-
During the past decade, megathrust earthquakes with magni- quake (Hanifa et al., 2014). At a ~150 km distance to the north from
tudes ≥ 7 have occurred along the Sunda trench in western Sumatra. the epicenter of the 2006 Java tsunami earthquake, occurred a de-
These events were the 2004 Mw 9.2 Sumatra-Andaman and 2005 Mw 8.7 structive earthquake named as the 2 September 2009 Tasikmalaya in-
Nias earthquakes in northern Sumatra (e.g. Banerjee et al., 2007) as traslab earthquake (TIE). The Indonesian National Board for Disaster
well as the 2007 Mw 8.4 Bengkulu and 2010 Mw 7.8 Mentawai earth- Management reported that > 80 people death with another 188,000
quakes in central Sumatra (Gusman et al., 2010). Along the Sunda displaced from their homes due to the 2009 TIE.
trench in southern Java, however, only the 2006 Mw 7.8 Java tsunami The 2009 TIE occurred at ~200 km to the north of the tectonically
earthquake occurred and recorded in the past decade (Ammon et al., active Sunda trench in western Java (Fig. 1), where this subduction
2006; Gunawan et al., 2017). zone represents the tectonic boundary between the Indo-Australian
The 2006 Java tsunami earthquake occurred off the south coast of plate to the south and the Sundaland block to the north (e.g. Schlüter
western Java followed by large tsunami resulted in ~600 deaths and et al., 2002). Using regional broadband seismometer data, Suardi et al.
~75,000 people displaced from their home (Mori et al., 2007). The (2014a) estimated the focal mechanism of the 2009 TIE and suggested a
2006 Java tsunami earthquake involved a fault rupture with length of west-dipping fault model with a strike of 160.8°, a dip of 75.3°, and a


Corresponding author.
E-mail address: endra.gunawan@itb.ac.id (E. Gunawan).

https://doi.org/10.1016/j.pepi.2019.04.004
Received 4 September 2018; Received in revised form 14 April 2019; Accepted 14 April 2019
Av ailableo nlin
e3 0A p ril2019
00 31 -9201/© 2019E lsevie
rB .V.Allrig
h tsre
serv
e d.
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

Fig. 1. Tectonic setting of our study area. Previously identified crustal faults in western Java are shown by solid lines: (1) Cimandiri fault (Marliyani et al., 2016) and
(2) Lembang fault (Afnimar and Rasmid, 2015). Newly identified faults, named as (3) Cipamingkis fault and (4) Jakarta fault, proposed by Gunawan and Widiyantoro
(2019), shown by dashed lines. Focal mechanism of the 2009 TIE was taken from Suardi et al. (2014a). Topography and bathymetry contour every 1000 m derived
from SRTM (Jarvis et al., 2008) and ETOPO 1 (Amante and Eakins, 2009). Inset shows the global map.

rake of 71.3° as well as an east-dipping fault model with a strike of Suardi et al. (2014b) estimated the coseismic slip of the 2009 TIE
34.0°, a dip of 24.0°, and a rake of 140.7°. This suggests that the dip using seismic data by assuming that the second nodal plane, that is the
angle of the westward fault is much higher than that of the plate in- east-dipping fault, is a fault where the earthquake occurred. In this
terface model (Hayes et al., 2012). Moreover, the United States Geo- study, we use the available Global Positioning System (GPS) data in
logical Survey reported that the hypocenter depth is 46 km, which is western Java (Hanifa et al., 2014) to analyze the associated fault source
deeper than the plate interface. This seismological analysis suggests of the 2009 TIE by incorporating information from focal mechanisms.
that the TIE was an intraslab earthquake. In addition, we estimate the tsunami height of the inferred fault models

5
5
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

−6˚ Operating Reference Stations (Ina-CORS; Abidin et al., 2010). Co-


seismic offset at every GPS station associated to the 2009 TIE obtained
by subtracting the mean positions of three days after the mainshock to
three days prior the mainshock (e.g. Banerjee et al., 2005; Gunawan
BAKO et al., 2016b). Horizontal component and vertical component are used
for further analysis.
CLDO Located ~50 km northeast of the 2009 TIE epicenter, the CPMK
CLBG
−7˚ CPTN
station experienced the largest coseismic displacement of ~5 mm, while
the BAKO station at ~200 km northwest of the epicenter experienced
CTVI the smallest displacement, i.e. ~1 mm. At most GPS stations, observa-
CSGT tions were under the level of the uncertainties. For this reason, instead
CUJG
of using fault slip inversion based on a Bayesian information criterion
(Yabuki and Matsu'ura, 1992), we conduct an inversion on a single
CPMK
rectangular fault. We use a priori information for the Green's function
−8˚ calculated in an elastic half-space (Okada, 1992).
Analysis of the focal mechanism of the 2009 TIE (Suardi et al.,
2014a) suggests two possible fault geometries. First is the west-dipping
fault model and second is the east-dipping fault model. In our study, we
estimate a west-dipping fault model (hereinafter referred to as Model
dipW) and an east-dipping fault model (hereinafter referred to as Model
dipE).
−9˚
106˚ 107˚ 108˚ 3. Coseismic slip results

Fig. 2. Distribution of GPS stations used in this study. Location of each station We conduct series of analysis by calculating fault model with
is shown by red triangles. Dashed line denotes slab contour for every 20-km shallow top depth, i.e. 1 km, up to deeper depth of 45 km, shown in
depth from Slab1.0 (Hayes et al., 2012). (For interpretation of the references to
Fig. 3. Vertical information from the GPS data at the CPMK station
color in this figure legend, the reader is referred to the web version of this
crucially explains the top depth of the fault model. We find that model
article.)
with deeper fault depth generates lower misfit than the shallower one.
Meanwhile, model with different fault length generate relatively similar
misfit for the same fault depth. Our preference for fault Model dipW is a
and the stress transfer of the mainshock related to 3-month aftershock fault with length of 85 km along strike and top depth of 30 km, while
data. the best fit fault Model dipE is a fault with length of 60 km along strike
and top depth of 30 km. In those two models, the earthquake hypo-
2. Data and method center of a 46 km depth is located within our fault geometry.
The coseismic slip result of Model dipW is shown in Fig. 4. The fault
Our study employs daily time series data at eight GPS stations lo- slip of Model dipW is 17.1 cm with a rake direction of 58.1°. Misfit
cated in western Java (BAKO, CLBG, CLDO, CPMK, CPTN, CSGT, CTVI, between the displacement data and model is calculated using
CUJG; see Fig. 2) available from early 2008 to the end of 2010 (Hanifa χ2 = (∑i=1, n(datai − modeli)2)/n, and we find that χ 2
= 3 mm. For
dipW
et al., 2014; Gunawan et al., 2016a). These GPS stations are maintained each station, the worst fit is estimated at the CLBG station with the best
by the Indonesian Geospatial Information Agency (BIG) as part of the fit estimated at BAKO. The seismic moment of Model dipW corresponds
nationwide GPS network named the Indonesian Continuously to 2.21 × 1019 N m (equal to Mw 6.8), calculated using a value of

Fig. 3. Misfit between displacement data and model for (a) fault Model dipW, and (b) fault Model dipE.

5
6
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

Fig. 4. Slip result of Model dipW, with comparison displacements between data (black vectors) and model (gray vectors) for (a) horizontal and (b) vertical com-
ponents. Solid black line represents the fault top and yellow star indicates the epicenter of the 2009 TIE. Error ellipses are shown at the 68% confidence level. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 5. Slip result of Model dipE, with comparison displacements between data (black vectors) and model (gray vectors) for (a) horizo ntal and (b) vertical com-
ponents. (See Fig. 4 for a detailed description of the figure legend.)

50 GPa for rigidity (e.g. Ohta et al., 2011; Kita and Katsumata, 2015). of 133.1°. We find a larger misfit of Model dipE, that is χdipE2 = 4 mm.
Fig. 5 shows the coseismic slip result of the second model, i.e. Model For each station, the worst fit occurs at CSGT with BAKO experiencing
dipE. In this model, estimated fault slip is 9.5 cm with a rake direction the best fit. We estimate a seismic moment of Model dipE that is

Table 1
Estimated fault parameters.

Model Longitudea Latitudea Top depthb Widthb Lengthb Strikec Dipc Rakec Slipd

dipW 107.1689 −7.7921 30 30 85 161 75.3 58.1 171


dipE 106.7582 −8.2639 30 100 60 34 24.0 133.1 95

a
Upper-left corner of a rectangular fault plane.
b
In km.
c
In degree.
d
In mm.

5
7
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

Fig. 6. Simulated maximum tsunami amplitude distribution from Model dipW. Green circle indicates the location of Pamengpeuk area. Upper and right panels are
the maximum tsunami amplitude along the coast across the longitude and latitude, respectively. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

2.85 × 1019 N m, equivalent to Mw 6.9 (using 50 GPa for rigidity). The data for the numerical simulation. The time interval for the numerical
estimated fault parameters of Model dipW and Model dipE are listed in tsunami simulation is 1 s to satisfy the CFL (Courant–Friedrichs–Lewy)
Table 1. condition. The initial sea surface displacement is calculated by the
Okada (1992) formula using the earthquake source parameters
4. Discussion (Table 1). Our tsunami height estimation of Model dipW and Model
dipE is shown in Figs. 6 and 7.
Comparing the estimated displacement between our two models, In general, our simulation suggested that the tsunami height gen-
Model dipW produced a better fit than Model dipE by 1 mm. This dif- erated by Model dipE is smaller than that from Model dipW. The tsu-
ference of 1 mm is slightly smaller than the displacement error at the nami simulation from both models agree that the Pamengpeuk area (at
GPS stations. This suggests that we need additional information to 107.7° E, −7.7° S) experienced a higher tsunami than the others. While
support our findings. For this reason, we conducted a tsunami height the tsunami simulation from Model dipW indicated a maximum height
simulation based on our fault models. of ~19 cm estimated in the Pamengpeuk area, the simulation from
The Indonesian Geospatial Information Agency (previously named Model dipE suggested a height of ~12 cm. From the tsunami simulation
as BAKOSURTANAL) reported that a tsunami with a peak amplitude of result, Model dipW is our preferred model due to the higher tsunami
the first wave of ~40 cm was observed at the Pamengpeuk tide gauge height compared with Model dipE, even though the maximum tsunami
station (Schöne et al., 2011). To compare the tsunami height estimated height of Model dipW is ~21 cm lower than the reported one in the
by our fault models with the reported one, we simulate tsunamis by Pamengpeuk area. One of the possible reasons for the height difference
solving the linear shallow water equations using a finite difference is most likely due to the use of 30 arc-sec GEBCO gridded bathymetry
method with a staggered grid scheme in a spherical coordinate system data in our analysis, as it is not capable of resolving small-scale coastal
(Gusman et al., 2010). We use 30 arc-sec GEBCO gridded bathymetry features such as harbors (e.g. Heidarzadeh et al., 2019).

5
8
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

Fig. 7. Simulated maximum tsunami amplitude distribution from Model dipE. Green circle indicates the location of Pamengpeuk area. Upper and right panels are the
maximum tsunami amplitude along the coast across the longitude and latitude, respectively. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

We also checked the possibility of stress transfer from the main- horizontal and vertical displacement of Model dipW fits the observed
shock triggering aftershocks. Nugraha et al. (2018) reported a relocated data better than the one generated by the east-dipping fault, i.e. Model
earthquake along the Sunda arc region, Indonesia. We use their in- dipE. Further, Model dipW produces a tsunami height closer to agree-
formation to view the location of the 3-month aftershocks after the ment with the reported one than Model dipE. Finally, Model dipW is
2009 TIE. The relocated aftershocks are concentrated at a shallower capable of explaining the triggered aftershock by the mainshock based
than 25 km depth, which is shallower than the mainshock (Fig. 8). on the positive ΔCFF. Therefore, we propose that the 2009 TIE ruptured
We calculate the Coulomb failure stress changes ΔCFF by using a at a west-dipping fault, i.e. Model dipW, with a slip of 17.1 cm and rake
coefficient of friction of 0.8 (Gunawan et al., 2018). The positive (ne- of 58.1°.
gative) value of ΔCFF indicates that stress promotes (opposes) a slip.
The stress transfer analysis of Model dipE indicates that aftershocks are
located in the negative stress. This indicates that stress transfer from Acknowledgments
Model dipE failed to explain the aftershocks. On the contrary, the stress
transfer analysis of Model dipW indicates that most of the aftershocks Our work benefited from discussion with Nuraini Rahma Hanifa,
are located in the region where ΔCFF is positive (Fig. 8). This finding is Iman Suardi, and Daryono on an earlier draft of the manuscript. We
in agreement with the idea of positive stress from the mainshock trig- thank the Indonesian Agency for Meteorological, Climatological, and
gering the aftershocks (Lin and Stein, 2004; Toda et al., 2005). Geophysics and the Indonesian Geospatial Information Agency for
maintaining the seismic network and GPS data. We also thank the two
5. Conclusion anonymous reviewers and the editor, Vernon Cormier, for their thor-
ough overview and constructive comments. This study was supported
We estimated the coseismic slip of the 2009 TIE on two possible by the 2018 World Class University Research Fund of Bandung Institute
fault plane models using GPS data inversion. The amount of moment of Technology for National Collaborative Research. Figures were pro-
release of Model dipW is equivalent to Mw 6.8. The calculated duced using generic mapping tools (Wessel and Smith, 1998).

5
9
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

Fig. 8. Calculated Coulomb stress change for (a) Model dipW, and (b) Model dipE. Open circles indicate the location of aftershocks.

References Molucca Sea, Indonesia, earthquake determined from GPS measurements. Phys. Earth
Planet. Inter. 259, 29–33.
Gunawan, E., Meilano, I., Hanifa, N.R., Widiyantoro, S., 2017. Effect of coseismic and post-
Abidin, H.Z., Subarya, C., Muslim, B., Adiyanto, F.H., Meilano, I., Andreas, H., Gumilar, I., seismic deformation on homogeneous and layered half-space and spherical analysis: model
2010. The Applications of GPS CORS in Indonesia –– Status, Prospect and Limitation, in FIG simulation of the 2006 Java, Indonesia, tsunami earthquake. Journal of Applied Geodesy
(International Federation of Surveyors) Congress 2010 –– Facing the Challenges –– 11 (4), 207–214.
Building the Capacity. Gunawan, E., Widiyantoro, S., Rosalia, S., Daryono, M.R., Meilano, I., Supendi, P., Ito, T., Tabei,
Afnimar, Yulianto, Rasmid, E., 2015. Geological and tectonic implications obtained from first T., Kimata, F., Ohta, Y., Ismail, N., 2018. Coseismic slip distribution of the July 2, 2013 Mw
seismic activity investigation around Lembang fault. Geoscience Letters 2 (1), 4. 6.2 Aceh, Indonesia, earthquake and its tectonic implication estimated from aftershock
Amante, C. and Eakins, B. W., 2009. ETOPO1 1 arc-minute global relief model: procedures, data data and joint inversion of GPS and geological offsets. Bull. Seismol. Soc. Am. https://doi.
sources and analysis (p. 19). US Department of Commerce, National Oceanic and org/10.1785/0120180035.
Atmospheric Administration, National Environmental Satellite, Data, and Information Gusman, A.R., Tanioka, Y., Kobayashi, T., Latief, H., Pandoe, W., 2010. Slip distribution of the
Service, National Geophysical Data Center, Marine Geology and Geophysics Division. 2007 Bengkulu earthquake inferred from tsunami waveforms and InSAR data. Journal of
Ammon, C.J., Kanamori, H., Lay, T., Velasco, A.A., 2006. The 17 July 2006 Java tsunami Geophysical Research: Solid Earth 115 (B12).
earthquake. Geophys. Res. Lett. 33 (24). Hanifa, N.R., Sagiya, T., Kimata, F., Efendi, J., Abidin, H.Z., Meilano, I., 2014. Interplate
Banerjee, P., Pollitz, F.F., Bürgmann, R., 2005. The size and duration of the Sumatra-Andaman coupling model off the southwestern coast of Java, Indonesia, based on continuous GPS
earthquake from far-field static offsets. Science 308 (5729), 1769–1772. data in 2008–2010. Earth Planet. Sci. Lett. 401, 159–171.
Banerjee, P., Pollitz, F., Nagarajan, B., Bürgmann, R., 2007. Coseismic slip distributions of the Hayes, G.P., Wald, D.J., Johnson, R.L., 2012. Slab1.0: a three-dimensional model of global
26 December 2004 Sumatra–Andaman and 28 March 2005 Nias earthquakes from GPS subduction zone geometries. Journal of Geophysical Research: Solid Earth (B1), 117.
static offsets. Bull. Seismol. Soc. Am. 97 (1A), S86–S102. Heidarzadeh, M., Tappin, D.R., Ishibe, T., 2019. Modeling the large runup along a narrow
Fujii, Y., Satake, K., 2006. Source of the July 2006 West Java tsunami estimated from tide gauge segment of the Kaikoura coast, New Zealand following the November 2016 tsunami from a
records. Geophys. Res. Lett. 33 (24). potential landslide. Ocean Eng. 175, 113–121.
Gunawan, E., Widiyantoro, S., 2019. Active tectonic deformation in Java, Indonesia inferred Jarvis, A., Reuter, H. I., Nelson, A. and Guevara, E., 2008. Hole-filled SRTM for the globe
from a GPS-derived strain rate. J. Geodyn. 123, 49–54. https://doi.org/10.1016/j.jog. Version 4. available from the CGIAR-CSI SRTM 90m Database (http://srtm.csi.cgiar.org).
2019.01.004. Kita, S., Katsumata, K., 2015. Stress drops for intermediate-depth intraslab earthquakes beneath
Gunawan, E., Meilano, I., Abidin, H.Z., Hanifa, N.R., 2016a. Investigation of the best coseismic Hokkaido, northern Japan: differences between the subducting oceanic crust and mantle
fault model of the 2006 Java tsunami earthquake based on mechanisms of postseismic events. Geochem. Geophys. Geosyst. 16 (2), 552–562.
deformation. J. Asian Earth Sci. 117, 64–72. https://doi.org/10.1016/j.jseaes.2015.12. Lin, J., Stein, R.S., 2004. Stress triggering in thrust and subduction earthquakes and stress
003. interaction between the southern San Andreas and nearby thrust and strike-slip faults.
Gunawan, E., Kholil, M., Meilano, I., 2016b. Splay-fault rupture during the 2014 Mw7. 1 Journal of Geophysical Research: Solid Earth 109 (B2).

6
0
E. Gunawan, et al. P
hysicsofth
eEa
rtha
ndP
lan
eta
ryIn
teriors29
1(2
019
)5–
461

Marliyani, G.I., Arrowsmith, J.R., Whipple, K.X., 2016. Characterization of slow slip rate faults Schöne, T., Illigner, J., Manurung, P., Subarya, C., Zech, C., Galas, R., 2011. GPS-controlled tide
in humid areas: Cimandiri fault zone. Indonesia. Journal of Geophysical Research: Earth gauges in Indonesia–a German contribution to Indonesia's Tsunami Early Warning System.
Surface 121 (12), 2287–2308. Nat. Hazards Earth Syst. Sci. 11 (3), 731–740.
Mori, J., Mooney, W.D., Kurniawan, S., Anaya, A.I., Widiyantoro, S., 2007. The 17 July 2006 Suardi, I., Widiyantoro, S., Yagi, Y., 2014a. Moment tensor analysis of the September 2, 2009
tsunami earthquake in west Java, Indonesia. Seismol. Res. Lett. 78 (2), 201–207. Tasikmalaya, West Java earthquake using the waveform inversion method of near field
Nugraha, A.D., Shiddiqi, H.A., Widiyantoro, S., Thurber, C.H., Pesicek, J.D., Zhang, H., Wiyono, data. International Journal of Tomography & Simulation™ 25 (1), 63–74.
S.H., Ramdhan, M., Irsyam, M., 2018. Hypocenter relocation along the Sunda Arc in Suardi, I., Yagi, Y., Widiyantoro, S., Meilano, I., 2014b. Analysis of source rupture process of the
Indonesia, using a 3D seismic-velocity model. Seismol. Res. Lett. 89 (2A), 603–612. September 2, 2009 Tasikmalaya earthquake by using the joint inversion method of near
Ohta, Y., Miura, S., Ohzono, M., Kita, S., Iinuma, T., Demachi, T., Tachibana, K., Nakayama, T., field and Teleseismic data. International Journal of Tomography & Simulation™ 27 (3),
Hirahara, S., Suzuki, S., Sato, T., 2011. Large intraslab earthquake (2011 April 7, M 7.1) 1–22.
after the 2011 off the Pacific Coast of Tohoku earthquake (M 9.0): Coseismic fault model Toda, S., Stein, R.S., Richards-Dinger, K., Bozkurt, S.B., 2005. Forecasting the evolution of
based on the dense GPS network data. Earth, planets and space 63 (12), 1207–1211. seismicity in southern California: animations built on earthquake stress transfer. Journal of
Okada, Y., 1992. Internal deformation due to shear and tensile faults in a half-space. Bull. Geophysical Research: Solid Earth (B5), 110.
Seismol. Soc. Am. 82 (2), 1018–1040. Wessel, P., Smith, W.H., 1998. New, improved version of generic mapping tools released. Eos.
Schlüter, H.U., Gaedicke, C., Roeser, H.A., Schreckenberger, B., Meyer, H., Reichert, C., Transactions American Geophysical Union 79 (47), 579.
Djajadihardja, Y., Prexl, A., 2002. Tectonic features of the southern Sumatra-western Java Yabuki, T., Matsu'ura, M., 1992. Geodetic data inversion using a Bayesian information criterion
forearc of Indonesia. Tectonics 21 (5). for spatial distribution of fault slip. Geophys. J. Int. 109 (2), 363–375.

6
1

View publication stats

Potrebbero piacerti anche