You are on page 1of 7

Meilano, I. et al.

Paper:

Slip Rate Estimation of the Lembang Fault West Java


from Geodetic Observation
Irwan Meilano1 , Hasanuddin Z. Abidin1 , Heri Andreas1 , Irwan Gumilar1 ,
Dina Sarsito1 , Rahma Hanifa1,4 , Rino1 , Hery Harjono2 , Teruyuki Kato3 ,
Fumiaki Kimata4 , and Yoichi Fukuda5
1 Geodesy

Research Division, Faculty of Earth Science and Technology,


Institute of Technology Bandung
Ganesa 10, Bandung 40132, Indonesia
E-mail: irwanm@gd.itb.ac.id
2 Indonesian Institute of Sciences (LIPI), Indonesia
3 Earthquake Research Institute, The University of Tokyo
4 Research Center Seismology, Volcanology and Disaster Mitigation, Nagoya University
5 Graduate School of Science, Kyoto University
[Received October 5, 2011; accepted January 10, 2012]

The Sunda arc forms the southern border of the Indonesia Archipelago, where the Indo-Australian plate
is subducted beneath Eurasia. The age of subducting plate increases from Sumatra in the west to Flores
in the east. The increase in age is consistent with an
increase in plate dip along the arc and an increasing
depth of seismic activity. The motion of Australia with
respect to West Java is 68 mm/yr in a direction N11E
orthogonal to the trench. A number of active faults
characterizing this area include Cimandiri fault, Lembang fault and Baribis fault. This research uses campaign and continues GPS data to make a preliminary
estimation of the slip rate of Lembang fault. Our GPS
measurements suggest that Lembang fault has shallow
creeping and deeper locking portion. The estimated
slip rate is 6 mm/yr with fault locking at 3-15 km and
shallow creeping with the same rate. While the results
are preliminary and we need more data for reliable estimations, we point out that these data can contribute
to earthquake risk assessment by constraining earthquake recurrence relationships.
Keywords: Lembang dault, GPS observation, slip rate,
creeping and locking

1. Introduction
Tectonic faults are zone of localized deformation that
accommodates plate motion by creeping aseismically at
depth and by earthquake or episodic creep in the upper
crust (Peltzer at al., 2001 [11]). It is generally assumed
that creep on the deep section of faults and the far-field
plate motion remain steady over long time periods, implying a stable rate of stress loading in the elastic part of the
crust. A comparison between geodetic and long-term geologic slip rate measurements could provide insights into
the causes of earthquake clustering and improve under-

12

standing of seismic hazard (Oskin et al., 2007 [9]).


In recent years GPS techniques have been applied repeatedly in studies on slip rate estimation of active faults
in Indonesia. The first GPS data on active fault in Indonesia found a less than 20 mm/year of slip rate along
the Sorong-Yapen fault (Puntodewo et al., 1994 [12]).
GPS measurements on intermediate points of Palu-Koro
fault transect in Sulawesi Indonesia show that the fault is
locked at a depth estimated around 8-16 km with a left lateral slip rate of 34 mm/yr (Walpersdorf et al., 1988 [17]).
In Jawa Island GPS observations have been done at 14 stations since 1992 covering both Lembang and Cimindiri
Faults. The Cimandiri Fault has a left-lateral strike-slip
motion and takes up N-S compression at 10 mm/yr (Setydji et al., 1997 [14]). This research uses campaign and
continuous GPS data to estimate the slip rate of Lembang
fault. Our GPS measurements suggest that Lembang fault
has shallow creeping and deeper locking portion.

2. Geological Framework
The West Indonesian island arc results from the interaction of oblique to normal subduction between IndoAustralian plate and Sunda block with velocities up to
6 cm/yr. Moderate to strong earthquakes are common
phenomena, especially in the southern part of West Java.
The 24-km long Lembang fault, the focus of our investigation is located eight kilometers from Bandung city
(Fig. 1). The western part of this fault runs through
densely populated area including Paropong district that
experienced an earthquake on 28 August 2011. Lembang Fault is parallel to Java subduction zone, which lies
300 km off shore to the southern coast of Java Island. To
the west of Lembang fault lies the Cimadiri Fault. The
north facing fault scarp of the Lembang fault has slopes
steeper than 40 degree. The East-West profile shows that
the scrap has a height above the surface of about 300 m

Journal of Disaster Research Vol.7 No.1, 2012

Slip Rate Estimation of the Lembang Fault West Java


from Geodetic Observation

Fig. 1. Simplified geological map of the Bandung Basin, Ql is lake deposit, Qyd is sandy tuff, Qyl is lava, Qyt is tuffaceous
tuff, Qvu is undifferentiated volcanic product and Mtjl is volcanic and marine sediments (a), modified from Silitonga, 1973, Cross
section Geology around Bandung (b).

in Pulasari, 150 m in the vicinity of Maribaya, 40 m in


Maribaya the difference heights gradually decrease westward until in the vicinity of Cisarua district with the outcropping amounting to 15 m.
Van Bemmelen, (1949 [16]) assumed that following the
build up of the Sunda volcano, a gravitational collapse,
due to the loading of enormous amounts of volcanic materials caused thrust faults and diapiric structures in the
epidermal strata of the northern foot slopes. Rifting associated with catastrophic sector failure eruptions destroyed
the volcanic cones, while the depressurization of the main
magma reservoir led to normal faulting and the formation of the Lembang fault (Van Bemmelen, 1949 [16]).
Lembang fault is shown to be terminated by two traverse faults, while a third fault perpendicular to it has
been drawn at the locality of the Cihideung water gap.
The easternmost transverse fault formed the east side of
Mt. Pulusari (Tjia, 1968 [15]).
Tjia (1968 [15]) used topographical, morphological
and structural evidence to indicate that the western part
of the Lembang fault, west of the Cikapundung valley to
be essentially strike slip in nature. Offsets of twelve recognizable streams and valley cuts across the scarp in a
general N-S direction. Out of these ten horizontal off-

Journal of Disaster Research Vol.7 No.1, 2012

sets are sinistral, the remaining two are dextral in nature.


Van Bemmelen (1949 [16]) correlated the presence of the
Lembang scrap with deflection and beheadings of various
lava flows originating from Mt. Tangkuban perahu. The
horizontal displacements range between 75 m and 250 m
with an average displacement of 140 m and annual displacements amounting to at least 3 centimeters. The average ratio of strike slip to dip slip is 2 to 1 (Tjia, 1968 [15]).
The seismicity around Lembang fault is generally very
low due to the record of the earthquake is short and the
sparse seismic station coverage. In July 22, 2011, M2.9
earthquake and in August 28, 2011, M3.3 earthquake
were recorded by local seismic network around Lembang
Fault (Fig. 2). Results of rapid assessment after the August earthquake suggest that most of people in the village
near the earthquake epicenter feel the shaking at the intensity of 3 MMI and 103 non-engineered houses slightly
damages.
Geodetic observations in Lembang fault as estimated
from GPS surveys on June 2006, August 2007 and August 2008 suggest the displacement rates vary spatially
between 0.3 to 1.4 cm/yr (Abidin et al., 2009 [1]). The
displacement due to the Sunda block motion, with the
rate of about 2.7 cm/yr with the southeast direction (Bock

13

Meilano, I. et al.

Fig. 2. Location of the M2.9, July 22, 2011 earthquake and M3.3 August 28, 2011 earthquake.
Black dots (inset) shows distribution of damage houses due to the M3.3 earthquake.

Fig. 3. Observation Network around Lembang fault, cycles show location of continuous GPS point, squares show
location of campaign GPS and triangles show seismic observation.
(a)

et al., 2003 [2]) has been removed. However, the southward horizontal displacements are still apparent. These
horizontal displacements may be associated with the postseismic deformation of the July 2006, Mw7.8 South-Jawa
tsunami earthquake.

3. GPS Observation Around Lembang, West


Java
The GPS measurements in this study consist of campaign mode as well as continuous observations (Fig. 3).
Campaign mode data were collected during 22-24 June
2006, 09-13 Nov. 2006, 16-19 Aug. 2007, 13-16 Aug.
2008 and 29 July-1 Aug. 2009, 25-27 June 2010 and 114

Fig. 4. (a) GPS time series for continuous GPS sites, (b)
GPS time series campaign GPS sites.

3 April 2011. Data from continuous GPS stations span


2008-2010 (Fig. 3). The observation times vary from
8 hrs to 24 hours with 30 seconds time interval.
Dual-frequency carrier phase and pseudorange observation were processed using Bernese GPS processing
software version 5.0 (Hugentobler et al., 2004 [6]), developed at the University of Bern. The International GNSS
service (IGS) precise ephemeris, earth rotation parameters
and satellite clock coefficients were used and the coordinate reference system was established by connecting to
nearby IGS stations. Integer biases were fixed with QuasiJournal of Disaster Research Vol.7 No.1, 2012

Slip Rate Estimation of the Lembang Fault West Java


from Geodetic Observation

(b)
Fig. 4. Continued.

Journal of Disaster Research Vol.7 No.1, 2012

15

Meilano, I. et al.

Fig. 5. GPS vector for GPS point around the Lembang fault. Black arrow show result from campaign GPS, while blue from continuous GPS, ellipse corresponding 95%
confidence interval.

Ionosphere Free (QIF). All relevant geodynamic reductions were applied in order to enable a careful determination of crustal deformation. The ocean tidal loading was
considered using the GOT00 model (Bos and Scherneck,
2004 [3]). The elevation cut-off angle used in the analyses
was set to 10 . For each station troposphere zenith path
delays were estimated for a time interval of 2 hrs introducing the condition that the troposphere delay changes
only linear with the time for each interval. Antenna phase
center variations were taken into consideration using consistent, absolute models of both receiver and satellite antenna phase center (Schmid and Rothacher, 2003 [13]).
The daily solutions are combined into a free network
solution. Residual tropospheric zenith delays are estimated simultaneously with the station coordinated by
least squares adjustment. The BAKO continuous GPS
is used as a reference for the time series. To assess the
quality of the processing, we estimated station positions
for each day and for each campaign by fitting in the least
square sense. Daily and long-term position time series
(Fig. 4a) were examined to detect outlier. Fig. 4 shows
position time series for ITB that is located inside campus
of Institut Teknologi Bandung and CLBG that is located
very close to Lembang Fault. Velocities determined at the
campaign sites are based on three to four measurement
and the uncertainties are 2-4 higher to continuous GPS
and should be interpreted carefully (Fig. 4b).
The general pattern shows GPS velocity vector oriented
roughly toward the south with amplitudes increasing for
the station close to the trench. The velocity for the station
located close to the Lembang fault is 3-10 mm/yr. While
the GPS point close to the south coast shows southward
vectors with velocities 10-22 mm/yr (Figs. 5 and 6). This
pattern suggests the transient postseismic signal following
the 2006 South Java tsunami earthquake still observed in
GPS time series. Therefore we modeled the postseismic
deformation and removed the effect from the observed
16

Fig. 6. GPS displacement vector, red arrow shows GPS


point located close to the south coast.

GPS vectors. After this correction, the slip rates of the


Lembang fault were estimated.

4. Postseismic Deformation and Model Estimation


The Mw7.8 July 17th 2006 South Java earthquake occurred on Java trench about 220 km south of Java Island.
The earthquake involved shallow-low angle thrust faulting in the Java trench excited a tsunami of 3-8 m in height
that inundated the southern coasts of Java. The coseismic
displacement observed at BAKO GPS station 350 km to
the north of the epicenter is 4 mm to the southward (Hanifa et al., 2007 [5]). The source of postseismic slip area
was assumed to be located on the lower portion of the coseismic slip. Aftershocks data were used to constrain the
geometry of the subducting interface. Fig. 7 shows the
Journal of Disaster Research Vol.7 No.1, 2012

Slip Rate Estimation of the Lembang Fault West Java


from Geodetic Observation

Fig. 7. Cross-section of seismicity (USGS) and focal mechanism (Global CMT). The postseismic area is located bellow
the coseismic region.

cross section and geometry of the postseismic area. The


strike, dip, rake angles and depth of the postseismic t slip
are 289 , 30 , 90 and 50 km respectively.
Dislocation theory was used in a semi-infinite homogeneous perfect elastic body to calculate the displacement
at each GPS station from slip on the postseismic plane. A
grid search technique was used to find the best solution.
The best estimated postseismic slip rate is 11 cm/year.
The horizontal crustal deformation rate measurements
were used to constrain a model of the slip rate of Lembang fault. The horizontal surface displacement rate was
modeled to be the result of the strike-slip motion on vertical dislocations representing the Lembang fault. We use
grid search method for the interseismic fault slip rates using the surface displacement field and the Greens function relation for surface displacement due to dislocations
in a homogeneous, isotropic, elastic half-space (Okada,
1985 [8]). The fault was divided into three vertical depth
ranges: a creep region from 0-3 km. a seismogenic region from 3-15 km, and a deep slip zone (> 15 km). A
uniform vertical dip angle was assumed since the coverage of the GPS network cannot solve complex model of
the source. The evidence of shallow creep region can be
detected from displacement GPS point very close to the
fault (CLBG). It was found that the best estimate of shallow creeping and deep creeping rate is 6 mm/yr. The estimated model cannot explain all the deformation at CITR
and TNKP (Fig. 8) due to poor quality of campaign observation.

5. Conclusion
The aim of this research was to quantify the active deformation of the Lembang fault. The preliminary determined slip rate is 6 mm/yr with fault locking at 3-15 km.
Although the true geometry of the Lembang fault is undoubtedly far more complex than our model results, it
seems inevitable that there are locked region on this fault.
Journal of Disaster Research Vol.7 No.1, 2012

Fig. 8. Fault parallel component of observed and calculated


displacement.

This research also identified a shallow creeping rate of


6 mm/yr. The relatively low fault slip rates are consistent
with the lower seismicity in Lembang fault compared to
the active fault in Sumatra.
This study indicates that fault Lembang fault has a
strike-slip (left lateral) component but this is based on
horizontal GPS observations. Due to the low quality of
the GPS observations on the vertical component the possibility of other mechanisms such as normal slip (Tjia,
1968 [15]) can not be detected. Longer continuous GPS
observation data is required to detect the vertical movement of the Lembang fault.
Acknowledgements
This work is partially supported by Ministry of National Education of Indonesia, Japan Science and Technology Agency - Japan
International Cooperation Agency (JST-JICA) and AustraliaIndonesia Facility for Disaster Reduction (AIFDR). The authors
would like to acknowledge National Coordinating Agency for
Surveys and Mapping Indonesia (BAKOSURTANAL) for providing continuous GPS data.

References:
[1] A. Z. Hasanuddin, H. Andreas, T. Kato, T. Ito, I. Meilano, F. Kimata, D. H. Natawidjaya, and H. Harjono, Crustal Deformation
Studies In Java (Indonesia) Using GPS, Journal of Earthquake and
Tsunami, Vol.3, No.2, pp. 77-88, 2009.
[2] Y. Bock, L. Prawirodirdjo, J. F. Genrich, C. W. Stevens, R. McCaffrey, C. Subarya, S. S. O. Puntodewo, and E. Calais, Crustal
motion in Indonesia from global positioning system measurements, Journal of Geophysical Research, Vol.108, No.B8, p. 2367,
doi:10.1029/2001JB000324, 2003.
[3] M. S. Bos and H. G. Scherneck, Free ocean tide loading provider,
2004, http://www.oso.chalmers.se/loading/ [Accessed at 5 August
2011]
[4] M. A. C. Dam, The Late Quaternary evolution of the Bandung
basin, West-Java, Indonesia Thesis, Vrije Universiteit, Amsterdam,
p. 252, 1994.
[5] H. N. Rahma, M. Irwan, T. Sagiya, F. Kimata, and H. Z. Abidin,
Numerical Modelling of the 2006 Java Tsunami Earthquake, Advance in Geoscience, Vol.13, Solid Earth, 2007.
[6] U. Hugentobler, R. Dach, and P. Fridez (Eds.), Bernese GPS Software, Version 5.0, University of Bern, 2004.

17

Meilano, I. et al.

[7] B. J. Meade and B. H. Hager, Block models of crustal motion


in southern California constrained by GPS measurements, J. Geophys. Res., Vol.110, B03403, doi:10.1029/2004JB003209, 2005.
[8] Y. Okada, Surface deformation due to shear and tensile faults in a
half space, Bull. Seismol. Soc., Vol.75, No.4, pp. 1135-1154, 1985.
[9] M. Oskin, L. Peng, D. Blumentritt, S. Mukhopadhyay, and A.
Iriondo, Slip rate of the Calico fault: Implications for geologic
versus geodetic rate discrepancy in the eastern California shear
zone, Journal of Geophysical Research, Vol.112, B03402, doi:
10.1029/2006JB004451, 2007.
[10] B. C. Papazachos, E. M. Scordilis, D. G. Panagiotopoulos, C. B. Papazachos, and G. F. Karakaisis, Global Relations Between Seismic
Fault Parameters and Moment Magnitude of Earthquakes, Bulletin
of Geological Society of Greece, Vol.XXVI, pp. 1482-1489, 2004.
[11] G. Peltzer, F. Crampe, S. Hensley, and P. Rosen, Transient strain
accumulation and fault interaction in the Eastern California Shear
Zone, Geology, Vol.29, pp. 975-978, 2001.
[12] S. S. O. Puntodewo, et al., GPS measurements of crustal deformation within the Pacific Australia plate boundary zone in Irian
Jaya, Tectonophysics, Vol.237, pp. 141-153, 1994.
[13] R. Schmid and M. Rothacher, Estimation of elevation-dependent
satel- lite antenna phase center variations of GPS satellites, Journal
of Geodesy, Vol.77, pp. 440-446, doi: 10.1007/s00190-003-03390., 2003.
[14] B. Setydji, I. Murata, J. Kahar, S. Suparka, and T. Tanaka, Analysis
of GPS measurement in West-Java, Indonesia, Ann. Disas. Prev.
Res. Inst. Kyoto Univ., Vol.40, No.B-1, pp. 27-33, 1997.
[15] H. D. Tjia, The Lembang Fault, West Java, Geologie En Mijnbouw, Vol.47, No.2, pp. 126-130, 1968.
[16] R. W. Van Bemmellen, The Geology of Indonesia, Vol.IA General
Geology, The Hague, 1949.
[17] A. Walpersdorf, C. Rangin, and C. Vigny, GPS compared to longterm geologic motion of the north arm of Sulawesi, Earth planet.
Sci. Lett., Vol.159, pp. 47-55, 1988.
[18] L. D. Wells and J. K. Coppersmith, New empirical relationships
among magnitude, rupture length, rupture width, rupture area, andsurface displacement, Bull. seism. Soc. Am., Vol.84, pp. 974-1002,
1994.

Name:
Irwan Meilano

Affiliation:
Geodesy Research Division, Faculty of Earth
Science and Technology, Institute of Technology
Bandung

Address:
Ganesha 10, Bandung 40132, Indonesia

Brief Career:
2006-2008 Postdoctoral Researcher, Nagoya University
2008-2011 Lecturer, Geodesy Research Group, Institute of Technology
Bandung (ITB)
2011 Associate Professor, Geodesy Research Group, Institute of
Technology Bandung (ITB)

Selected Publications:

M. Irwan, F. Kimata, and N. Fujii, Time Dependent of Magma Intrusion


Model During the Early Stage of The 2000 Miyakejima Activity,
JVGR(150), pp. 202-212, 2006.
M. Irwan, F. Kimata, K. Hirahara, T. Sagiya, and A. Yamagiwa,
Measuring Ground Deformation with 1-hz GPS Data : the 2003
Tokachi-oki Earthquake (Preliminary Report), Earth Planets and Space,
Vol.56, pp. 389-393, 2004.

Academic Societies & Scientific Organizations:


American Geophysical Union (AGU)
Indonesian Association of Geophysics (HAGI)

18

Journal of Disaster Research Vol.7 No.1, 2012

You might also like