ArticlePDF Available

Crustal strain partitioning and the associated earthquake hazard in the eastern Sunda-Banda Arc

Wiley
Geophysical Research Letters
Authors:
  • Badan Riset dan Inovasi Nasional

Abstract and Figures

We use Global Positioning System (GPS) measurements of surface deformation to show that the convergence between the Australian Plate and Sunda Block in eastern Indonesia is partitioned between the megathrust and a continuous zone of back-arc thrusting extending 2000km from east Java to north of Timor. Although deformation in this back-arc region has been reported previously, its extent and the mechanism of convergence partitioning has hitherto been conjectural. GPS observations establish that partitioning occurs via a combination of anticlockwise rotation of an arc segment called the Sumba Block, and left-lateral movement along a major NE-SW strike-slip fault west of Timor. We also identify a westward extension of the back-arc thrust for 300 km onshore into East Java, accommodating slip of ∼6 mm/yr. These results highlight a major new seismic threat for East Java, and draw attention to the pronounced seismic and tsunami threat to Bali, Lombok, Nusa Tenggara and other coasts along the Flores Sea.
Content may be subject to copyright.
Geophysical Research Letters
Crustal strain partitioning and the associated
earthquake hazard in the eastern
Sunda-Banda Arc
A. Koulali
1
, S. Susilo
2
, S. McClusky
1
, I. Meilano
3
, P. Cummins
1
, P. Tregoning
1
, G. Lister
1
,
J. Efendi
2
, and M. A. Syafi’i
2
1
Research School of Earth Sciences, Australian National University, Canberra, Australian Capital Territory, Australia,
2
Bandan Informasi Geospatial, Cibinong, Indonesia,
3
Institute of Technology Bandung, Bandung, Indonesia
Abstract We use Global Positioning System (GPS) measurements of surface deformation to show that
the convergence between the Australian Plate and Sunda Block in eastern Indonesia is partitioned
between the megathrust and a continuous zone of back-arc thrusting extending 2000 km from east Java
to north of Timor. Although deformation in this back-arc region has been reported previously, its extent
and the mechanism of convergence partitioning have hitherto been conjectural. GPS observations
establish that partitioning occurs via a combination of anticlockwise rotation of an arc segment called
the Sumba Block, and left-lateral movement along a major NE-SW strike-slip fault west of Timor. We also
identify a westward extension of the back-arc thrust for 300 km onshore into East Java, accommodating
slip of 6 mm/yr. These results highlight a major new seismic threat for East Java and draw attention
to the pronounced seismic and tsunami threat to Bali, Lombok, Nusa Tenggara, and other coasts along
the Flores Sea.
1. Introduction
Eastern Indonesia encompasses a complex tectonic environment, involving the convergence of four major
tectonic Plates: the Australian, Pacific, Philippine Sea Plates, and the Sunda Block [Hamilton, 1979] (Figure 1).
In this region the Australian Plate subducts northward beneath eastern Java, Nusa Tenggara (114
E125
E),
and the Banda Arc (Figure 1). These three arc segments accommodate a transition in the style of plate con-
vergence from ocean-continent subduction in east Java, to arc-continent collision in Nusa Tenggara and then
to the island arc subduction in the Banda Sea. While historical earthquake observations for this region are
poorly known, at least seven large earthquakes have occurred between 1648 and 1891 [Soloviev and Go, 1974;
Musson, 2012]. Six of these events were associated with macroseismic intensities of IXX and four generated
regional tsunamis in the Flores Sea with estimated runup of 3 m or greater [Soloviev and Go, 1974]. During the
instrumental seismic period, four major events were reported in the area between 112
E and 128
E[Ekström
et al., 2012] (Figure 1). The
M
w
7.9 1992 Flores earthquake was the largest thrust event recorded and gener-
ated a large, destructive tsunami [Beckers and Lay, 1995]. The majority of earthquakes during the last century
are attributed to the back-arc segments of Flores and Wetar and have thrust style focal mechanisms [Ekström
et al., 2012; Beckers and Lay, 1995], suggesting that this fault system is accommodating an important part of
the convergence between the Australian Plate and the Sunda Block.
Marine geophysical surveys [Silver et al., 1983] have revealed evidence for two major back-arc thrusts: the
450 km long Flores thrust north of Sumbawa and western Flores, and the 350 km long Wetar thrust north
of Timor (Figure 1). It has been speculated [Silver et al., 1983] that the thick crust beneath Sumba and Timor,
respectively, facilitates transfer of stress from the fore-arc to the back-arc, while the thinner crust elsewhere
(e.g., Savu basin) enables convergence to be partitioned onto fore-arcand back-arc thrusts and strike-slip faults
that cut the arc at angles oblique to the convergence [McCaffrey, 1988]. However, until now, there has been
no conclusive evidence identifying which if any of these faults are facilitating the transfer of convergence.
Early geodetic investigations [Genrich et al., 1996] concluded that the Timor Trough is inactive and most of the
convergence between Australia, Sundaland, and Eurasia occurs to the north at a rate of 50 mm/yr. In contrast,
later studies [Bock et al., 2003; Nugroho et al., 2009] estimated 15 to 20 mm/yr of motion across the Timor
RESEARCH LETTER
10.1002/2016GL067941
Key Points:
The Sunda-Banda back-arc thrust
system is a large active plate
boundary that extends over 2000 km
Strain is transferred from Java
subduction to the back-arc thrusts
via a left-lateral strike slip
Geodetic strain across the
Sunda-Banda back-arc thrusts
emphasize a high seismic and
tsunami hazard
Supporting Information:
Text S1, Tables S1 and S2,
and Figures S1S9
•DataSetS1
Correspondence to:
A. Koulali,
achraf.koulali@anu.edu.au
Citation:
Koulali, A., S. Susilo, S. McClusky,
I. Meilano, P. Cummins, P. Tregoning,
G. Lister, J. Efendi, and M. A. Syafi’i
(2016), Crustal strain partitioning and
the associated earthquake hazard
in the eastern Sunda-Banda
Arc, Geophys. Res. Lett., 43, 1943–1949,
doi:10.1002/2016GL067941.
Received 24 JAN 2016
Accepted 18 FEB 2016
Accepted article online 19 FEB 2016
Published online 11 MAR 2016
©2016. American Geophysical Union.
All Rights Reserved.
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1943
Geophysical Research Letters 10.1002/2016GL067941
Figure 1. Seismotectonic setting of the Sunda-Banda arc-continent collision, East Indonesia. Major faults
(thick black lines) [Hamilton, 1979]. Topography and bathymetry are from Shuttle Radar Topography Mission
(http://topex.ucsd.edu/www_html/srtm30_plus.html). Focal mechanisms are from the Global Centroid Moment Tensor.
Blue mechanisms correspond to earthquakes with
M
w
>
7
(brown transparent ellipses are the corresponding rupture
areas for Flores 1992 and Alor 2004 earthquakes), while the green focal mechanism shows the highest magnitude
recorded in Sumbawa. Red dots indicate the locations of major historical earthquakes [Musson, 2012].
Trough and 60 mm/yr of shortening across the Flores Sea. The discrepancies in these results reflect the degree
of uncertainty in understanding and assigning slip partitioning, mainly due to the lack of observations in the
vicinity of the back-arc fault system. In addition, the lack of long and adequately sampled GPS time series
makes it difficult to recognize and correct for the effects of postseismic relaxation resulting from both nearby
local and regional earthquakes that have occurred since GPS observations began in the 1990s. Thus, there are
many important but hitherto unanswered questions regarding the southern margin of eastern Indonesia: Is
the present-day back-arc deformation localized on the Flores and Wetar segments? How does the partitioning
of convergence between the subduction megathrust and back-arc vary along the Sunda-Banda Arc transition?
By what mechanism does this partitioning of convergence occur? Answering these questions is essential for
understanding the associated seismic and tsunami hazard. In this study, we use GPS velocities plus earthquake
slip vectors to quantify slip partitioning in eastern Indonesia (110
E to 135
E) and we discuss the implications
of strain distribution for earthquakes hazard in the region.
2. Methods
2.1. GPS Data Processing
The GPS velocity field presented in this study (supporting information Data Set S1) is calculated from obser-
vations at 94 GPS stations located in east Indonesia in combination with a global network of 80 International
Global Navigation Satellite Systems Service tracking sites. Campaign GPS sites have been surveyed irregularly
from 1993 to 2014, while most of the continuous sites operated from 2009 to 2014 (supporting information
Figure S3). The raw data were processed using GAMIT-GLOBK software [Herring et al., 2010], and uncertainties
were estimated following standard procedures described by Reilinger and et al. [2006]. The velocities used
in Figure 2 are with respect to a Sunda Block-fixed reference frame, defined using the velocity of only three
continuous sites: BINT, NTUS, and GETI, for the period prior to the Sumatra-Andaman 2004
M
w
9.2 earth-
quake [Vigny et al., 2005]. This approach eliminates the effects of contamination of the Sunda reference
frame by postseismic effects resulting from the 20042012 Sumatra earthquake sequence Feng et al. [2015].
The weighted root-mean-square for the north and east horizontal velocity components are 0.66 mm/yr and
0.97 mm/yr, respectively, and 1.2 mm/yr for the vertical rate. For our modeling, we use only horizontal veloc-
ities and we do not include any vertical rate estimates since they have large uncertainties due to different
source of systematic errors, making their usage of less importance for our block model inversions.
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1944
Geophysical Research Letters 10.1002/2016GL067941
Figure 2. GPS velocities determined in this study with respect to Sunda Block. Uncertainty ellipses represent 95%
confidence level. The inset figure corresponds to the area of the dashed rectangle in the map. Light blue arrows show
the velocities for East and West Makassar Blocks.
2.2. Kinematic Block Modeling
We model the observed velocities as a sum of block rotations and elastic strain produced by fault locking
[McCaffrey, 2005]. We chose block boundaries based on our qualitative interpretation of the GPS velocities
themselves as well as independent information from earthquakes [Ekström et al., 2012; Shulgin et al., 2011]
and the available seismic and geologic constraints on active faults in the Sunda-Banda Arc [Hamilton, 1979;
McCaffrey, 1988]. We performed a simultaneous inversion of horizontal GPS velocities and earthquake slip
vectors (supporting information Figure S1) to estimate Euler vectors of six blocks, locking depths at major Plate
boundaries and three components of the strain rate tensor for three blocks (Sumba block, Timor block, and
East Makassar block). The flexibility of this approach allows us to verify the significance and the importance
of each plate boundary in the kinematic model. We have investigated the present-day deformation in this
region using four kinematic models with different geometrical configurations of elastic blocks (Figure S5).
Our preferred block model includes two major boundaries that encompass the Sunda-Banda Arc: the Java
Trench and Timor Trough in the southern part and the back-arc thrusts system extending from the Wetar
thrust to the Kendeng thrust east of Java island in the north. We divide the Arc into three blocks: The Timor
Block [McCaffrey, 1988], Sumba Block, and the Eastern Java Block (Figure 3). In the southern part of Sulawesi,
we divide the Makassar Block [Socquet et al., 2006] into eastern and western blocks following the southern
extension of the Walanae Fault and Selayar Trough [Camplin and Hall, 2014]. The boundary we use to separate
the Banda Sea from the Weber Basin is speculative and may have a different geometry, though we do not
currently have observations constraining its precise location. It is not specified as a fault and treated here as
free-slipping boundary.
For the downdip geometry, the faults along the back-arc thrust are assigned a uniform dip angle of 30
in
accordance with seismic reflection profiles [Silver et al., 1983] and discretized with a 5 km downdip interval
from the surface down to 30 km; however, the geometry of Java Trench and Timor Trough was based on the
U.S. Geological Survey slab 1.0 [Hayes et al., 2012] as well as on seismicity cross sections established across the
main thrust for the eastern part were slab 1.0 model is not available. The nodes downdip are placed at depth
every 2 km in the upper 10 km then every 5 km from 10 to 45 km depth and then every 10 km down to 70 km
depth. In order to reduce the number of free parameters, we have estimated uniform locking depths at Wetar,
Flores, Bali-Lombok, and Kendeng thrusts and we have parameterized the locking along the Java Trench as a
function exponentially varying down depth while inverting for the minimum and maximum locking depths
of the transition zone [Wang et al., 2003; McCareyetal., 2007]. The four plausible block model scenarios we
investigate here include different combinations of block boundaries where seismicity or geologic constraints
do not provide a unique solution for the surface expression of an active crustal fault. The assessment of the
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1945
Geophysical Research Letters 10.1002/2016GL067941
Figure 3. Relative slip vectors across block boundaries, derived from our best fit model. Arrows show motion of the
hanging wall (moving block) relative to the footwall (fixed block) with 95% confidence ellipses. The tails of arrows is
located within the “moving” block. Black thick lines show well-defined boundaries we use as active faults in our model
and dashed lines show less well-defined boundaries (green : free-slipping boundaries and black: fixed locked faults) .
Principal axes of the horizontal strain tensor estimated for the SUMB, EMAK, and EJAV are shown in pink. The thick pink
arrow shows the relative motion of Australia with respect to Sunda (AUST/SUND). Abbreviations are Sumba Block
(SUMB), West Makassar Block (WMAK), East Makassar Block (EMAK), East Java Block (EJAV), and Timor Block (TIMO).
The background seismicity is from the International Seismological Centre catalog with magnitudes
5.5 and
depths
< 40 km.
significance of the geometrical complexity of alternate models was based on F test statistics. The detailed
summary of the fit statistics is provided in the supporting information (Table S1).
3. GPS Velocity Field
The velocity field with respect to the Sunda Block (Figure 2a) reveals an anticlockwise rotation of the whole of
eastern Sunda-Banda Arc with an increase in the motion of the archipelago from 9 mm/yr in Bali to 82 mm/yr
toward the eastern end of the Timor Trough. This general increase in the northward velocity of the fore-arc is
consistent with the convergence between the Australian Plate and the Sunda Block being progressively trans-
ferred north as the Australian Plate collision transitions from ocean arc in the West to continent-arc collision
in the East. Motion decreases north of the back-arc toward south Sulawesi, reflecting the conclusions of
previous studies that the back-arc is accommodating deformation [McCaffrey, 1988]. However, our new GPS
data along East Java, Madura Island, and Bali Island reveal a significant north-south gradient of velocities
across the Kendeng Basin and the volcanic arc (Figure 2b). This confirms that the boundary marking the transi-
tion from the Kendeng Basin to the Sunda shelf, known as the Kendeng thrust [Smyth et al., 2008], is active and
probably defines the westward onshore extension of the Flores back thrust [Simandjuntak and Barber, 1996].
Our kinematic model results show that the along-arc change in partitioning of plate convergence is caused by
the anticlockwise rotation of the Sunda-Banda Arc with respect to Sunda Block. The relative motion, between
the Asutralian Plate and the Sunda Block, along the Java Trench decreases from 70
±1.0 mm/yr west of
the Lombok Basin (115
E) to 33 ± 0.9 mm/yr south of Rote Island (123
E) (Figure 3). Eastward of Timor Island
(124
E), the relative motion along the Timor Trough gradually diminishes from 32 ± 2.0 mm/yr to an insignif-
icant 1.0 ± 1.7 mm/yr at Tual Island (132
E). In contrast, the relative motion along the back-arc increases
from west to east, from 6 ± 1.0 mm/yr in East Java to 26 ± 1.0 mm/yr at Flores, 28 ± 1.7 mm/yr at Wetar
and 30
± 1.8 mm/yr to the North of Timor, where the direction of vectors changes to more NE implying that
a significant component of shearing must be accommodated by the structures at the back-arc in this area.
This correlates well with the left-lateral strike-slip faulting inferred from the earthquake fault plane solutions
(supporting information Table S2 and Figure S2).
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1946
Geophysical Research Letters 10.1002/2016GL067941
4. Discussion and Implications for Earthquake Hazard
The description of the present-day motion within the eastern Sunda-Banda Arc is a debated question, with
proposed models suggesting that the deformation can be described by several independent crustal blocks
[McCaffrey, 1988; Genrich et al., 1996] and others proposing this region as a wide zone of distributed deforma-
tion with diffuse transitional zones [Nugroho et al., 2009]. Our results from the interpretation of the new GPS
velocity field as well as the kinematic model suggest that the eastern Sunda-Banda is segmented into three
blocks (East Java Block, Sumba Block, and Timor Block) separated by NE transitions of left-lateral faults. The
Semau Fault (SF, Figure 1) is one of a series of NNE-SSW trending left-lateral strike-slip faults west of Timor
[Charlton et al., 1991] and may have been associated with a large earthquake in 1814 [Soloviev and Go, 1974].
Our best fit block model requires a boundary at the Semau Fault connecting the Timor Trough with the Wetar
thrust, thus forming the western boundary of the Timor Block. The Semau Fault is a key component of our
kinematic model as it provides the structural link between the fore-arc and the back-arc.
Our model requires the addition of a boundary approximately perpendicular to the Java Trench, allowing
movement of this arc segment, known as the Sumba Block [McCaffrey, 1988], which is independent of the Java
fore-arc. We chose the location of this fault where marine seismic and gravity modeling studies [Shulgin et al.,
2011] indicate fracturing in the upper 2 km of the oceanic crust and a sharp increase in crustal thickness. This
fault accommodates only 3 to 4 mm/yr of strike-slip motion, less than 5% of the total relative motion, but its
inclusion improves significantly the fit of the data (Table S1).
The northwestern corner of the Sumba Block abuts the offshore extension of the Kendeng thrust, where we
estimate 5
± 0.4 mm/yr of convergence. The presence of mud volcanoes [Istadi et al., 2009] in the eastern
part of the Kendeng Basin is consistent with overpressuring caused by active convergence in this area.
Although some historical earthquakes may have occurred on the Kendeng thrust in the nineteenth century
[Simandjuntak and Barber, 1996], the absence of more recent events raises the question of whether this fault
is slipping aseismically or is fully locked. The current spatial resolution of the GPS network is insufficient to
resolve the heterogeneity in the coupling on the back-arc thrusts. Therefore, we chose to estimate a uniform
locking depth at each segment. On the Kendeng thrust, we estimated a locking depth of 9
± 3 km and found
that the segment north of Sumbawa Island is the deepest locked segment (20 ± 1.8 km) along the back-arc,
with a moment deficit of 2.4 ×10
18
Nm, equivalent approximately to an earthquake of magnitude M
w
= 6.3.
Between 2006 and 2009 a sequence of 3 earthquakes with
M
w
>
6.3
occurred in the eastern part of this seg-
ment at depths ranging between 18 and 20 km. A recent study [Fuchs et al., 2014] detected the occurrence of
triggered nonvolcanic tremors beneath Sumbawa Island, which might increase the recurrence time of major
events (
>
M
w
7) by helping to release strain during the interseismic period.
In contrast to the localization of deformation in the Sunda-Banda back-arc at the Flores and Wetar thrusts
inferred by previous studies, we find that active deformation extends along the back-arc for over
2000 km.
This accounts for a variation from 5% to 40% of the total convergence between the Australian Plate and the
Sunda Block and explains the distribution of both historical and recorded seismic events. Our results elucidate
the role of the left-lateral Semau Fault west of Timor in transferring the shear stress from the trench to the
back-arc (Figure 4), where NE shearing at 20 mm/yr predominates along the transfer boundary with a normal
convergence component of 4 to 9 mm/yr on this fault. This shear zone also marks a transition where the
convergence obliquity along the back-arc changes from normal at Flores (5
N) to a more oblique direction
along east Wetar (17
N), consistent with the difference in shortening [Silver et al., 1983] observed on the Wetar
(
10 km) and Flores (30 km) thrusts. This suggests that the increase of stress due to the obliquity reflects a
recent evolutionary stage of underthrusting across the Wetar back-arc segment migrating eastward.
The concept of slip partitioning in obliquely convergent fault systems is used to explain the accommo-
dation of shear strain resulting from the trench-parallel component of the relative motion [Fitch, 1972].
The classic model requires the megathrust plate boundary to accommodate the trench-normal slip, and a
parallel strike-slip fault in the fore-arc to take up the oblique slip, with both structures isolating a continental
wedge known as a sliver [Fitch, 1972; McCaffrey, 1988]. Along the transition from the eastern Java Trench to
Timor Trough, the direction of the plate convergence becomes progressively oblique to the east, where earth-
quake slip vectors show a complex pattern of deformation typical of highly curved margins as documented
by McCaffrey [1996]. Previous studies demonstrated that the degree of the deformation partitioning is a
function of the convergence obliquity [McCaffrey, 1992; Vernant and Chéry, 2006]. However, they showed that
full partitioning is reached only for high-obliquity values larger than 70
. In our study we predict that only
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1947
Geophysical Research Letters 10.1002/2016GL067941
Figure 4. Fault slip rate components: (a) fault normal (extension positive) and (b) fault parallel (right-lateral positive).
37% of the total lateral shear strain is accommodated on Timor Trough and the back-arc thrusts to the north,
suggesting that full partitioning is less likely to occur, consistent with 3-D mechanical modeling predictions
[Vernant and Chéry, 2006]. However, it is unclear how the remaining unaccounted deformation is accommo-
dated and we speculate it is likely to be partitioned further north on the Seram Trough, where the analysis
of focal mechanisms show signs of active deformation. Quantifying precisely where and how this remaining
motion is accommodated is beyond the scope of this paper requiring more dense GPS observations in the
north Banda Sea region.
Our results show a different structural organization, where the convergence itself is transferred to the back-arc.
As with classical slip partitioning, this results in isolation of an arc segment (the Sumba Block), and partitioning
of convergence is achieved through a combination of anticlockwise rotation of the Sumba fore-arc Block and
left-lateral movement along the Semau Fault (Figure 4). A similar organization is observed elsewhere in the
world where back-arc thrusting is active, such as the great Antilles Arc [Mann et al., 2002], the North Panama
Deformed belt [Kobayashi et al., 2014] and the New Hebrides/Vanuatu [Calmant et al., 2003]. These zones of
back-arc thrusting are approximately 500 km, 600 km, and 200 km, respectively, in length. Our results show a
far more extensive zone of active thrusting along a 2000 km section of the eastern Sunda Arc, a potentially
important source of seismic and tsunami hazard.
5. Conclusion
Our results draw a new kinematic framework for active deformation in the eastern Sunda-Banda Arc, high-
lighting the need to reconsider the level of seismic hazard there. Several of the active faults identified here
directly threaten socioeconomic assets vital to Indonesia. The Kendeng thrust passes through the southern
outskirts of Surbaya, Indonesia’s second largest city with a population of over 2.5 million, and traverses a
300 km length of East Java, with a population density of over 800 people per square kilometer. The Semau
Fault skirts the city of Kupang, the main commercial centre of Nusa Tenggara with a population of around
500,000. Finally, earthquakes along the back-arc thrust beneath the sea floor extending 1700 km from eastern
Java to Timor could generate regional tsunamis threatening the coastlines of the Flores Sea. Further studies,
including earthquake, geodetic, and paleoseismic, should be undertaken to better understand these threats.
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1948
Geophysical Research Letters 10.1002/2016GL067941
References
Beckers, J., and T. Lay (1995), Very broadband seismic analysis of the 1992 Flores, Indonesia, earthquake (M
w
= 7.9), J. Geophys. Res., 100(B9),
18,17918,193, doi:10.1029/95JB01689.
Bock, Y., L. Prawirodirdjo, J. F. Genrich, C. W. Stevens, R. McCaffrey, C. Subarya, S. S. O. Puntodewo, and E. Calais (2003), Crustal motion in
Indonesia from Global Positioning System measurements, J. Geophys. Res., 108, 2367, doi:10.1029/2001JB000324.
Calmant, S., B. Pelletier, P. Lebellegard, M. Bevis, F. W. Taylor, and D. A. Phillips (2003), New insights on the tectonics along the New Hebrides
subduction zone based on GPS results, J. Geophys. Res., 108, 2319, doi:10.1029/2001JB000644.
Camplin, D. J., and R. Hall (2014), Neogene history of Bone Gulf, Sulawesi, Indonesia, Mar. Petrol. Geol., 57, 88108.
Charlton, T. R., A. J. Barber, and S. T. Barkham (1991), The structural evolution of the Timor collision complex, eastern Indonesia,
J. Struct. Geol., 13, 489500.
Ekström, G., M. Nettles, and A. M. Dziewonski (2012), The global CMT project 20042010: Centroid-moment tensors for 13,017 earthquakes,
Phys. Earth Planet Inter., 1–9, 200 201.
Feng, L., E. M. Hill, P. Banerjee, I. Hermawan, L. L. H. Tsang, D. H. Natawidjaja, B. W. Suwargadi, and K. Sieh (2015), A unified GPS-based
earthquake catalog for the Sumatran plate boundary between 2002 and 2013, J. Geophys. Res. Solid Earth, 120, 35663598,
doi:10.1002/2014JB011661.
Fitch, T. J. (1972), Plate convergence, transcurrent faults, and internal deformation adjacent to Southeast Asia and the western Pacific,
J. Geophys. Res., 77(23), 44324460, doi:10.1029/JB077i023p04432.
Fuchs, F., M. Lupi, and S. A. Miller (2014), Remotely triggered nonvolcanic tremor in Sumbawa, Indonesia, Geophys. Res. Lett., 41, 41854193,
doi:10.1002/2014GL060312.
Genrich, J. F., Y. Bock, R. McCaffrey, E. Calais, C. W. Stevens, and C. Subarya (1996), Accretion of the southern Banda arc to the Australian plate
margin determined by Global Positioning System measurements, Tectonics, 15(2), 288295, doi:10.1029/95TC03850.
Hamilton, W. B. (1979), Tectonics of the Indonesian region, Tech. Rep. No. 1078, U.S. Govt. Print. O., Wash.
Hayes, G. P., D. J. Wald, and R. L. Johnson (2012), Slab1.0: A three-dimensional model of global subduction zone geometries, J. Geophys. Res.,
117, B01302, doi:10.1029/2011JB008524.
Herring, T. A., R. W. King, and S. C. McClusky (2010), Introduction to GAMIT/GLOBK, Mass. Inst. of Tech., Cambridge.
Istadi, B. P., G. H. Pramono, P. Sumintadireja, and S. Alam (2009), Modeling study of growth and potential geohazard for LUSI mud volcano:
East Java, Indonesia, Mar. Petrol. Geol., 26(9), 1724 1739.
Kobayashi, D., P. LaFemina, H. Geirsson, E. Chichaco, A. A. Abrego, H. Mora, and E. Camacho (2014), Kinematics of the western Caribbean:
Collision of the Cocos Ridge and upper plate deformation, Geochem. Geophys. Geosyst., 15, 1671 1683, doi:10.1002/2014GC005234.
Mann, P., E. Calais, J.-C. Ruegg, C. DeMets, P. E. Jansma, and G. S. Mattioli (2002), Oblique collision in the northeastern Caribbean from GPS
measurements and geological observations, Tectonics, 21(6), 1057, doi:10.1029/2001TC001304.
McCaffrey, R. (1988), Active tectonics of the Eastern Sunda and Banda Arcs, J. Geophys. Res., 93(B12), 1516315182,
doi:10.1029/JB093iB12p15163.
McCaffrey, R. (1992), Oblique plate convergence, slip vectors, and forearc deformation, J. Geophys. Res., 97, 89058915.
McCaffrey, R. (1996), Slip partitioning at convergent plate boundaries of SE Asia, in Tectonic Evolution of SE Asia Symposium, Geol. Soc.
Spec. Publ.,
106, 3 18.
McCaffrey, R. (2005), Block kinematics of the Pacific-North America plate boundary in the southwestern United States from inversion of GPS,
seismological, and geologic data, J. Geophys. Res., 110, B07401, doi:10.1029/2004JB003307.
McCaffrey, R., A. I. Qamar, R. W. King, R. Wells, G. Khazaradze, C. A. Williams, C. W. Stevens, J. J. Vollick, and P. C. Zwick (2007), Fault locking,
block rotation and crustal deformation in the Pacific Northwest, Geophys. J. Int., 169, 1315 1340, doi:10.1111/j.1365-246X.2007.03371.x.
Musson, R. M. W. (2012), A provisional catalogue of historical earthquakes in Indonesia, Open Report OR/12/073, pp. 21, Geological Survey,
Edinburgh, British.
Nugroho, H., R. Harris, A. W. Lestariya, and B. Maruf (2009), Plate boundary reorganization in the active Banda Arc-continent collision:
Insights from new GPS measurements, Tectonophysics, 479, 5265, doi:10.1016/j.tecto.2009.01.026.
Reilinger, R., et al. (2006), GPS constraints on continental deformation in the Africa-Arabia-Eurasia continental collision zone and
implications for the dynamics of plate interactions, J. Geophys. Res., 111, B05411, doi:10.1029/2005JB004051.
Silver, E. A., D. Reed, R. McCaffrey, and Y. Joyodiwiryo (1983), Back-arc thrusting in the Eastern Sunda Arc Indonesia: A consequence of
arc-continent collision, J. Geophys. Res., 88(B9), 7429 7448, doi:10.1029/JB088iB09p07429.
Simandjuntak, T. O., and A. J. Barber (1996), Contrasting tectonic styles in the Neogene orogenic belts of Indonesia, Geol. Soc. Spec. Publ.,
106(1), 185 201.
Shulgin, A., H. Kopp, C. Mueller, L. Planert, E. Lueschen, E. R. Flueh, and Y. Djajadihardja (2011), Structural architecture of
oceanic plateau subduction offshore Eastern Java and the potential implications for geohazards, Geophys. J. Int., 184, 1228,
doi:10.1111/j.1365-246X.2010.04834.x.
Smyth, H. R., R. Hall, and G. J. Nichols (2008), Cenozoic volcanic Arc history of east Java, Indonesia: The stratigraphic record of eruptions on
an active continental margin, Geol. Soc. Am. Spec. Pap., 436, 199 222.
Socquet, A., W. Simons, C. Vigny, R. McCaffrey, C. Subarya, D. Sarsito, B. Ambrosius, and W. Spakman (2006), Microblock rotations and
fault coupling in SE Asia triple junction (Sulawesi, Indonesia) from GPS and earthquake slip vector data, J. Geophys. Res., 111, B08409,
doi:10.1029/2005JB003963.
Soloviev, S. L., and Ch. N. Go (1974), A Catalogue of Tsunamis on the Western Shore of the Pacific Ocean, pp 308, Nauka Publishing House,
Moscow.
Vernant, P., and J. Chéry (2006), Mechanical modelling of oblique convergence in the Zagros Iran, Geophys. J. Int., 165, 9911002.
Vigny, C., et al. (2005), Insight into the 2004 Sumatra-Andaman earthquake from GPS measurements in Southeast Asia, Nature, 436,
201206.
Wang, K., R. Wells, S. Mazzotti, R. D. Hyndman, and T. Sagiya (2003), A revised dislocation model of interseismic deformation of the Cascadia
subduction zone, J. Geophys. Res., 108, 2026, doi:10.1029/2001JB001227.
Wessel, P., and W. H. F. Smith (1998), New, improved version of the genericmapping tools released, Eos. Trans. AGU, 79, 579,
doi:10.1029/98EO00426.
Acknowledgments
This research was supported under
the Australian Research Council’s
Linkage Projects funding scheme
(LP110100525). Figures are drawn
with GMT [Wessel and Smith, 1998].
The GPS data were computed on
the Terrawulf II computational facility
at the Research School of Earth
Sciences, a facility supported through
the AuScope initiative. AuScope
Ltd is funded under the National
Collaborative Research Infrastructure
Strategy (NCRIS), an Australian
Commonwealth Government
Program. We are grateful to Nyamadi,
Dodi Sudarmono, Caca Juniarsa,
Budi Parjanto, Heru derajat, Sidik Tri
Wibowo, Munawar Kholil and Putra
Maulida, and all to the personnel
of Badan Informasi Geospatial (BIG),
who participated in GPS surveys
over the past 20 years. We appreciate
constructive comments and
improvements from two
anonymous reviewers.
KOULALI ET AL. CRUSTAL STRAIN IN THE SUNDA-BANDA ARC 1949
... Subsequent analyses by Stevens et al. (2002) and Bock et al. (2003) delineated several distinct crustal blocks within the Indonesian archipelago and constructed velocity fields of crustal motion based on annual GPS field surveys from 1991 to 2001. Additional GPS observations, as exemplified by the work of Koulali et al. (2016), referenced motions with respect to the Sunda reference frame. These observations incorporate data from 94 GPS stations predominantly encompassing the islands of Java, Lombok, Sumbawa, Flores, and Timor, in the western Banda Arc. ...
... Based on the seafloor morphology of these interpreted faults, it is deduced that the current extension of the Aru Trough primarily lies between F2 and F3, that is, Aru Basin. includes velocity data from Stevens et al. (2002), Bock et al. (2003), Koulali et al. (2015Koulali et al. ( , 2016Koulali et al. ( , 2017, and others. However, the coverage of these data sets in the eastern Banda Arc region is sparse. ...
... The primary methods used in this study are: (a) fault mapping based on bathymetry; (b) structural interpretation of seismic profiles; (c) calculating rates and timing of extension from GPS velocities; (d) computation of seismicity (Bock et al., 2003). The gray dashed lines delineate the Bird's Head microplate whose boundaries are defined by the Seram Trough to west, by the Tarera Aiduna strikeslip fault zone and the Lowlands Fault zone to south, by the Manokwari Trough and the New Guinea trench to east, and by the Sorong Fault to north (Bock et al., 2003;Koulali et al., 2016;Stevens et al., 2002;Zhao et al., 2023). (b) GPS velocities around the Aru Trough. ...
Article
Full-text available
The Banda Arc in eastern Indonesia has a complex tectonic history involving oceanic and continental subduction, arc‐continent collision and slab rollback. Among these the subduction rollback that began at 16 Ma shaped the regional tectonic configuration, causing notable upper plate extension evidenced by the Banda Sea and Weber Deep. However, the effects of subduction rollback on the lower plate remain less understood. The Aru Trough, part of subducting Australian continental margin, shows strong deformation and high seismicity, making it ideal for studying these effects. Utilizing multibeam bathymetry, seismic reflection profiles, GPS observations, seismicity and focal mechanisms, we investigate the crustal deformation, driving mechanisms and seismic risks in the Aru Trough under the background of Banda slab rollback. The Aru Trough shaped like an inverted triangle narrows from 160 to 40 km southward. It has a fast 53.8 mm/yr extensional rate at present, when combined with its maximum width suggesting a 3 Ma age. High‐angle faults (>60°) are present at its margins and interior. The trough has a high density of seismicity up to 5.9 events per 25 km², with dominant normal and strike‐slip events, consistent with observed deformational patterns. The low seismic b‐value of 0.82 ± 0.01 suggests a high‐stress state with potential for strong earthquakes (Mw ≥ 6.5). Our findings indicate that lower plate deformation is profoundly influenced by subduction rollback, oblique arc‐continent collision and regional strike‐slip faulting. Understanding deformation in the Aru Trough is crucial for grasping the broader tectonic evolution of the Banda Arc and assessing seismic risks.
... Kim et al. (2015) Three-dimensional numerical modeling Other findings indicate that the modeling process has an impact on flooding by applying a wave reflection process. Koulali et al. (2016) The 300 km onshore westward extension of the back-arc thrust to East Java, which contains about 6 mm/yr of slip. These findings underline the genuine earthquake and tsunami threat to the Flores Sea shores of Bali, Lombok, Nusa Tenggara, and other islands. ...
... According to Pranantyo & Cummins (2019), only one tide gauge, which is situated in Palopo, Sulawesi, provided proof of the 1992 Flores earthquake and tsunami due to a lack of current technology. The 1992 Flores earthquake and tsunami broke faults that were inclined towards ENE (east-northeast), but the Flores Back Arc Thrust Fault also played a significant role (Beckers & Lay, 1995;Koulali et al., 2016) and is connected to the 2018 Lombok earthquake phenomenon (Felix et al., 2022;Yang et al., 2020). So, according to Koulali et al. (2016), this fault poses a threat to earthquakes and tsunamis in the regions of Bali, Lombok, Nusa Tenggara, and beaches along Flores. ...
... The 1992 Flores earthquake and tsunami broke faults that were inclined towards ENE (east-northeast), but the Flores Back Arc Thrust Fault also played a significant role (Beckers & Lay, 1995;Koulali et al., 2016) and is connected to the 2018 Lombok earthquake phenomenon (Felix et al., 2022;Yang et al., 2020). So, according to Koulali et al. (2016), this fault poses a threat to earthquakes and tsunamis in the regions of Bali, Lombok, Nusa Tenggara, and beaches along Flores. ...
Article
Full-text available
p class="Abstract"> The Flores Sea has experienced devastating earthquakes with magnitudes >7 over the past 30 decades. It can trigger a tsunami and provide important theoretical, experimental, and field information. The seismicity study stated that the island of Flores had experienced tsunamis during the pre-instrumental period (1815, 1818, 1820, and 1836) and the pre-instrumental period in 1992. This study discusses the development of tsunami research in Flores using a literature review approach. The data source comes from the Scopus database, with data analysis using VOSviewer. The search results obtained a total of 22 documents, with the result that the 1992 Flores earthquake became the main research topic and the beginning of the era of modern tsunami science . </p
... Significant earthquakes, including the Mw 7.8 and Mw 7.5 events of the 1992 Flores and 2004 Alor sequences, along with their associated tsunamis (Harris and Major, 2016;Koulali et al., 2016;Cummins, 2017;Coudurier-Curveur et al., 2021), have nucleated on the south-dipping Flores and Wetar thrusts. These thrusts accommodate a substantial portion of the convergence between the Australian and Sunda plates (e.g. ...
... These thrusts accommodate a substantial portion of the convergence between the Australian and Sunda plates (e.g. Koulali et al., 2016;Coudurier-Curveur et al., 2021). Our analysis of earthquake focal mechanisms suggests two steep northdipping fault planes for earthquake events of Mw > 7.0. ...
... The Kendeng Fault, which comprises multiple distinct segments with a thrust or fold system, extends 300 km westward from the Flores back arc fault to the East Java mainland [2] [3]. Although the Kendeng Fault itself hasn't triggered significant earthquakes to date, the Flores thrust, which is part of the same back arc, has recorded several seismic events [4]. ...
... In the event of an earthquake impacting the velocity vector's direction, it would signify coseismic or postseismic activity, leading to velocity adjustments towards the earthquake source [6]. Our study highlights northeast-directed movement in the GPS velocity ( Figure 5), despite the Kendeng Fault primarily being of thrust-type with an additional shear component [2] that forms the crux of our investigation ( Figure 4). Consequently, we solely consider the horizontal component to estimate the Kendeng Fault's horizontal motion. ...
Conference Paper
Full-text available
The Kendeng Fault is recognized as an extension of the Flores back-arc thrust, primarily exhibiting thrust fault behavior. However, prior research has indicated the presence of a significant right-lateral shear component along this fault. While no significant seismic events have been documented in recent years, historical records suggest that the region experienced significant seismic events prior to the 20th century, necessitating a careful consideration of the earthquake hazard associated with this fault. The question arises whether the absence of recent seismic activity is a consequence of energy accumulation or gradual energy release. Our investigation utilizes GPS data collected over a four-year period starting in 2016 to determine the slip rate of the Kendeng Fault. Our findings indicate that the slip rate along the Kendeng Fault is approximately 0.45 ± 0.07 cm/year. These results provide valuable insights into the behavior of the Kendeng Fault and contribute to a better understanding of its seismic potential and associated risks.
... b. Location map of the Singaraja-Sukasada area in Buleleng Regency, where the Gejer Bali disaster occurred With its neighboring islands of Lombok and Sumbawa, Bali is part of the volcanic arc of the Lesser Sunda Islands, which was created during the Miocene subduction of the Indo-Australian plate (Hall 2002;Koulali et al. 2016). Thus, earthquakes in southern Bali are distributed along the Java trench, like the 1917 event of M6.6, which killed 1,500 people due to landslides principally (Soloviev and Go 1974). ...
Article
Full-text available
In November 1815, the deadliest “natural” disaster in Balinese history was caused by the exceptional combination of multiple natural hazards that occurred simultaneously and cascaded in the present-day province of Buleleng. This major disaster, which is thought to have claimed more than 10,000 lives, has never been scientifically analyzed. The study conducts an in-depth analysis of this cascading disaster, from the root causes and chronology of natural hazards to their environmental and societal effects, by thoroughly examining all available written sources about this event, whether colonial or Indonesian. Seven months after the Tambora eruption, a magnitude 7.3 earthquake, which occurred in the Bali Sea off the northern coast of the island, triggered a very large landslide on the northern flank of the Buyan-Bratan caldera. The initial mass movement evolved into a cohesive debris flow that reached the sea after traveling up to twenty kilometers through Banyumala River Valley and Singaraja City downstream. According to historical accounts, fifteen villages were buried or devastated by the debris flow. The large volume of sediment entering the sea triggered a local tsunami along Buleleng’s coast. This geohistorical approach offers a comprehensive overview of various sources describing Singaraja’s situation before the crisis, the hazard succession, the cascading hazard intensities, and the short- to long-term impacts on Buleleng. Based on the written sources, Bali took around fifteen years to recover from the 1815 disasters.
... The northward movement of the Australian Plate and the resistance to subduction led to the arc-continent collision. This shift is evidenced by global positioning system observations that show a deceleration from 70 mm/year at the southern Timor Trough to 30 mm/year at Seram (Figure 1; Koulali et al., 2016). The collision between the arc-islands and the Australian continental margin causes the transition from an accretionary wedge to fold and thrust belts (R. Harris, 2011) and contributes to the formation of the spoon-shaped Banda slab and the 180°bend of the Banda arc (Spakman & Hall, 2010). ...
Article
Full-text available
The Banda arc‐continent collision zone signifies one of the most seismically active and tectonically intricate zones. The high convergence rate across the region, coupled with the exceptionally arcuate arc and subducted slab, makes it an ideal locale for investigating interactions between plate (slab) kinematics and plastic flow in the asthenosphere, which can be diagnosed by seismic anisotropy from shear wave splitting analyses. In total, 206 pairs of splitting measurements using teleseismic SKS, SKKS, and PKS, along with 43 pairs using local S phases, are obtained by utilizing broadband seismic data from five permanent seismic stations. To reduce the ambiguity in determining the origin of anisotropy leading to the teleseismic splittings, which lack vertical resolution, crustal anisotropy is constrained according to the sinusoidal moveout of converted S phases at the Moho using receiver functions. A layered anisotropic structure based on joint analyses of the anisotropy measurements characterizing different depth layers suggests the presence of trench‐parallel flow both in the mantle wedge and the sub‐slab region. The northeastward motion of the slab, entrained by the fast‐moving Australian Plate, deflects asthenospheric materials. The modulation results in trench‐parallel plastic mantle flows and leads to the steepening of the southern portion of the asymmetric spoon‐shaped Banda slab. In the shallower part of the sub‐slab region, the northeastward Australian Plate motion produces simple shear in the transitional layer between the rigid lithosphere and the viscous asthenosphere. The shear deformation induces seismic anisotropy with resulting fast orientations in accordance with the plate motion direction.
... Most significantly, the prevalence of the motif of the sinking island seems to suggest that this was a common enough narrative in the late 1500s and early 1600s for the legend to reach Eredia, who was stationed in Malacca. In fact, while some have argued that the body of observations in Eastern Indonesia is poorly known and consists of sparse reports (Koulali et al. 2016;Pranantyo and Cummins 2019), there are good reasons to believe in the existence of oral accounts that could predate European descriptions (Liu and Harris 2013). For example, there is a large compendium of myths and legends in this part of the archipelago relating events of island uplift and submergence, and the destruction of island masses by earthquakes, partial flooding, and powerful tsunamis (Table 1). ...
Article
Full-text available
There is a large corpus of myths and legends about sea creatures in the maritime world, a record that portrays incredible and wondrous feats, wrecks, calamities, and disasters. In this article, I present an account of the mythological cosmology of the Endenese, a group of fearless seafarers that scoured the Eastern Indonesian seas for over four centuries. By discussing the legend of Kota Djogo, an island that disappeared into the sea in time immemorial, I reconstruct Endenese explanations for luck and uncertainty in a world plagued by volcanic eruptions, earthquakes, tsunamis, and typhoons. Going beyond the legend’s veracity, I build on the study of the narrative to show how symbolic accounts of environmental events can provide important clues to the understanding of ecological disasters. I argue that the indigenous rationalizations of uncertainty present in oral legends and myths can function as coping mechanisms that reconcile communities with the unpredictable and the ambiguous.
Article
The Lombok earthquake in August 2018 triggered a sequential rupture with doublet earthquake up to Mw 6.9. This tectonic activity occurred near the main earthquake due to the decay of residual energy from one event to another. This activity is suspected to be a post-seismic deformation process such as afterslip and viscoelastic. In this paper, we conducted a study to determine the deformation pattern. Each of these processes can be investigated by extracting InSAR observational data. Time series from Sentinel-1 SAR is processed using LiCSBAS as data observation and then compare with the model based on exponential and logarithmic functions. The results of combined logarithmic and exponential fitting suggest the Lombok multi-event earthquakes were influenced by seismic activity from dual-releasing residual energy comprises of afterslip and viscoelastic as a dual mechanism with long duration rather than single mechanism.
Article
Full-text available
Earth Physics Research Group (EPRG) is one of three groups of research running at Physics Study Program, the State University of Surabaya, Indonesia, where a number of research projects with corresponding topics have been conducted (and some are in progress) by the group members and associated students having final projects in the field of earth physics since 2018. Whereas the research roadmap of the group has been presented in association with definitive research projects for 25 years long starting from 2011, the specific goal of this paper is to shortly summarise all academic achievement in terms of research performance made by the group members during the last five years. The majority of the recent works was mainly based on computational work, where some were completed in collaboration with researchers from other universities and a national agency and others were performed by the group members and selected students. The topics were spread across disciplines in earth physics that included tectonic earthquakes, tsunami generation and propagation of seismic and non-seismic origin, volcanic eruptions and an integrated disaster mitigation study. A small portion of the projects were performed using a chosen method of applied geophysics. These studies have ended up with publications in recent years, where the saline points of the key findings are here presented. Future studies focusing on vulnerability to earthquake hazards in the northern areas of Java and on volcanic and meteo-tsunamis are also discussed in the context of possible tsunamis induced by seismic sources or volcanic processes.
Article
Full-text available
We have compiled the first self-consistent GPS-based earthquake catalog for the Sumatran plate boundary. Using continuous daily position time series from the Sumatran GPS Array (SuGAr), we document 30 earthquakes which occurred within or outside the SuGAr network from August 2002 through the end of 2013, and we provide estimates of both vertical and horizontal coseismic offsets associated with 1 M9.2, 3 M8, 6 M7, 19 M6, and 1 M5.9 earthquakes, as well as postseismic decay amplitudes and times associated with 9 M > 7 earthquakes and 1 M6.7 earthquake. For most of the previously studied earthquakes, our geodetic catalog provides more complete coseismic displacements than those published, showing consistent patterns of motion across a large range of distances. For many of the moderate to large earthquakes, we publish their coseismic displacements for the first time, providing new constraints on their locations and slip distributions. For the postseismic time series, we have tackled the challenge of separating the signals for individual events from the overlapping effects of many other earthquakes. As a result, we have obtained either new or much longer time series than previously published. Based on our long time series, we find logarithmic decay times ranging from several days to more than 20 years, and sometimes a second decay time is needed, suggesting that when studying large to great Sumatran earthquakes, we need to consider multiple postseismic mechanisms. Our geodetic catalog provides rich spatial and temporal Sumatran earthquake cycle information for future studies of the physics and dynamics of the Sumatran plate boundary.
Article
Full-text available
We present, for the first time, evidence for triggered tremor beneath the island of Sumbawa, Indonesia. We show triggered tremor in response to three teleseismic earthquakes; the Mw 9.0 2011 Tohoku earthquake, and two oceanic strike slip earthquakes (Mw 8.6 and Mw 8.2) offshore of Sumatra in 2012. We constrain an apparent triggering threshold of 1 mm/s ground velocity that corresponds to about 8 kPa dynamic stress. Peak tremor amplitudes of about 180 nm/s are observed, and scale with the ground velocity induced by the remote earthquakes. Triggered tremor responds to 45–65 s period surface waves and predominantly correlates with Rayleigh waves, even though the 2012 oceanic events have stronger Love wave amplitudes. We could not locate the tremor because of minimal station coverage, but data indicates several potential source volumes including the Flores Thrust, the Java subduction zone or Tambora volcano.
Article
Full-text available
Version 3.1 of the Generic Mapping Tools (GMT) has been released. More than 6000 scientists worldwide are currently using this free, public domain collection of UNIX tools that contains programs serving a variety of research functions. GMT allows users to manipulate (x,y) and (x,y,z) data, and generate PostScript illustrations, including simple x-y diagrams, contour maps, color images, and artificially illuminated, perspective, and/or shaded-relief plots using a variety of map projections (see Wessel and Smith [1991] and Wessel and Smith [1995], for details.). GMT has been installed under UNIX on most types of workstations and both IBM-compatible and Macintosh personal computers.
Article
Full-text available
Subduction of the Cocos plate and collision of the Cocos Ridge have profound effects on the kinematics of the western Caribbean, including crustal shortening, segmentation of the overriding plate, and tectonic escape of the Central American fore arc (CAFA). Tectonic models of the Panama Region (PR) have ranged from a rigid block to a deforming plate boundary zone. Recent expansion of GPS networks in Panama, Costa Rica, and Colombia makes it possible to constrain the kinematics of the PR. We present an improved kinematic block model for the western Caribbean, using this improved GPS network to test a suite of tectonic models describing the kinematics of this region. The best fit model predicts an Euler vector for the counterclockwise rotation of the CAFA relative to the Caribbean plate at 89.10°W, 7.74°N, 1.193° Ma-1, which is expressed as northwest-directed relative block rates of 11.3±1.0 - 16.5±1.1 mm a-1 from northern Costa Rica to Guatemala. This model also predicts high coupling along the Nicoya and Osa segments of the Middle American subduction zone. Our models demonstrate that the PR acts as a single tectonic block, the Panama block, with a predicted Euler vector of 107.65°W, 26.50°N, 0.133° Ma-1. This rotation manifests as northeast migration of the Panama block at rates of 6.9±4.0 - 7.8±4.8 mm a-1 from southern Costa Rica to eastern Panama. We interpret this motion as tectonic escape from Cocos Ridge collision, redirected by collision with the North Andes block, which migrates to the northwest at 12.2±1.2 mm a-1.
Article
Full-text available
The region offshore Eastern Java represents one of the few places where the early stage of oceanic plateau subduction is occurring. We study the little investigated Roo Rise oceanic plateau on the Indian plate, subducting beneath Eurasia. The presence of the abnormal bathymetric features entering the trench has a strong effect on the evolution of the subduction system, and causes additional challenges on the assessment of geohazard risks. We present integrated results of a refraction/wide-angle reflection tomography, gravity modelling, and multichannel reflection seismic imaging using data acquired in 2006 south of Java near 113°E. The composite structural model reveals the previously unresolved deep geometry of the oceanic plateau and the subduction zone. The oceanic plateau crust is on average 15 km thick and covers an area of about 100 000 km2. Within our profile the Roo Rise crustal thickness ranges between 18 and 12 km. The upper oceanic crust shows high degree of fracturing, suggesting heavy faulting. The forearc crust has an average thickness of 14 km, with a sharp increase to 33 km towards Java, as revealed by gravity modelling. The complex geometry of the backstop suggests two possible models for the structural formation within this segment of the margin: either accumulation of the Roo Rise crustal fragments above the backstop or alternatively uplift of the backstop caused by basal accumulation of crustal fragments. The subducting plateau is affecting the stress field within the accretionary complex and the backstop edge, which favours the initiation of large, potentially tsunamogenic earthquakes such as the 1994 Mw= 7.8 tsunamogenic event.
Article
Bone Gulf is one of the inter-arm basins of the unusual K-shaped island of Sulawesi. Its age, character and origin are disputed. This study is based on recently acquired 2D seismic lines, seabed multibeam mapping and limited well data, and is linked to stratigraphy on land. The gulf is probably underlain by pre-Neogene volcanogenic, sedimentary, metamorphic and ultramafic rocks, and includes crust of Australian origin. We favour basin initiation in the Miocene rather than Eocene, by extension associated with strike-slip deformation. The main basin trends N-S and is divided into several sub-basins and highs. The highs segment the gulf and their WNW-ESE orientations reflect pre-Neogene basement structures. They are interpreted as strike-slip fault zones active at different times in the Neogene. A southern high was active relatively early, whereas further north there is evidence of young displacements during the Late Neogene. These are visible on the seabed above a high linked to the Kolaka Fault on land. Early basin-bounding faults are oriented NNW-SSE and record extension and strike-slip movements, like the sub-parallel Walanae Fault of South Sulawesi which can be traced offshore into extensional faults bounding the young and narrow Selayar Trough. Sediment in the basins came mainly from the north with contributions from both west and east. Carbonate deposits formed at the margins while deeper marine sediments were deposited in the axial parts of the gulf. An Early Pliocene unconformity can be mapped across the study area marking major uplift of Sulawesi and subsidence of Bone Gulf. This regional event caused major influx of clastic sediments from the north, development of a southward-flowing canyon system, and back-stepping and drowning of carbonates at the basin margins. Hydrocarbons are indicated by seeps, and Bone Gulf has potential sources, reservoirs and seals, but the complex faulting history is a risk.
Chapter
The stratigraphic record of volcanic arcs provides insights into their eruptive history, the formation of associated basins, and the character of the deep crust beneath them. Indian Ocean lithosphere was subducted continuously beneath Java from ca. 45 Ma, resulting in formation of a volcanic arc, although volcanic activity was not continuous for all of this period. The lower Cenozoic stratigraphic record on land in East Java provides an excellent opportunity to examine the complete eruptive history of a young, well-preserved volcanic arc from initiation to termination. The Southern Mountains Arc in Java was active from the middle Eocene (ca. 45 Ma) to the early Miocene (ca. 20 Ma), and its activity included significant acidic volcanism that was overlooked in previous studies of the area. In particular, quartz sandstones, previously considered to be terrigenous clastic sedimentary rocks derived from continental crust, are now known to be of volcanic origin. These deposits form part of the fill of the Kendeng Basin, a deep flexural basin that formed in the backarc area, north of the arc. Dating of zircons in the arc rocks indicates that the acidic character of the volcanism can be related to contamination of magmas by a fragment of Archean to Cambrian continental crust that lay beneath the arc. Activity in the Southern Mountains Arc ended in the early Miocene (ca. 20 Ma) with a phase of intense eruptions, including the Semilir event, which distributed ash over a wide area. Following the cessation of the early Cenozoic arc volcanism, there followed a period of volcanic quiescence. Subsequently arc volcanism resumed in the late Miocene (ca. 12–10 Ma) in the modern Sunda Arc, the axis of which lies 50 km north of the older arc.
Article
Earthquake moment tensors reflecting seven years of global seismic activity (2004–2010) are presented. The results are the product of the global centroid-moment-tensor (GCMT) project, which maintains and extends a catalog of global seismic moment tensors beginning with earthquakes in 1976. Starting with earthquakes in 2004, the GCMT analysis takes advantage of advances in the mapping of propagation characteristics of intermediate-period surface waves, and includes these waves in the moment-tensor inversions. This modification of the CMT algorithm makes possible the globally uniform determination of moment tensors for earthquakes as small as MW = 5.0. For the period 2004–2010, 13,017 new centroid-moment tensors are reported.