You are on page 1of 12

SRL Early Edition


E

The 2016 M w 6.5 Pidie Jaya, Aceh, North


Sumatra, Earthquake: Reactivation of an
Unidentified Sinistral Fault in a Region
of Distributed Deformation
by Muzli Muzli, Muksin Umar, Andri Dian Nugraha, Kyle Edward
Bradley, Sri Widiyantoro, Kemal Erbas, Philippe Jousset, Supriyanto
Rohadi, Irwandi Nurdin, and Shengji Wei

ABSTRACT

On 6 December 2016, an M w 6.5 earthquake occurred in highlighting the need for better understanding of active fault-
Pidie Jaya, Aceh, about 30 km to the north of the Sumatran ing and seismic hazard in this region.
fault (SF) that killed more than 100 people and destroyed
∼3000 buildings. Mainshock focal mechanism inversions
using regional Agency for Meteorology, Climatology, and
Geophysics of Indonesia (BMKG) broadband data and tele- Electronic Supplement: Damage photos, figures of relative main-
seismic waveform data all indicate a strike-slip event with a shock relocation, depth resolution of focal mechanism solution
centroid depth of 11–15 km. The observed macrointensity of the mainshock, focal mechanism solution of the mainshock
data show that most of the damaged buildings are distributed using a cut-and-paste (CAP) method, teleseismic finite-fault
along the coast, approximately perpendicular to the ruptured inversion of the mainshock for the fault plane with a strike of
fault strike instead of parallel with it. The strong shaking and 148°, and the double-difference hypocenter relocation of the
damage sites are primarily located on the coastal sedimentary aftershocks using hypoDD.
soils, highlighting the importance of site conditions in deter-
mining risk. We used one-month data recorded by nine INTRODUCTION
temporal broadband stations to locate aftershocks with grid-
search and double-difference algorithms, thereby resolving a On the early morning of Wednesday, 7 December 2016,
linear trend of seismicity aligned in a northeast–southwest 05:03:33 local time (6 December 22:03:33 UTC), an M w 6.5
direction. The refined aftershock locations indicate a left- earthquake rocked the northern coastal area of north Aceh in
lateral rupture that is in agreement with the preliminary the Pidie Jaya region (hereafter, Pidie Jaya earthquake). The
finite-fault slip inversion as well as geomorphic signatures Agency for Meteorology, Climatology, and Geophysics of In-
of local geological structure. Using a well-located M L 4.2 donesia (BMKG) reported an epicenter location of 5.29° S and
aftershock for path calibration, we relocated the mainshock 96.22° E, with a shallow depth of 15 km. The Pidie Jaya earth-
epicenter with regional P-wave arrivals. The refined epicenter quake occurred at the moment before dawn when many people
falls within the cloud of the well-located aftershocks, whereas were still asleep and many others were in the mosques for
locations from the global and regional catalogs are located morning prayer. The earthquake caused heavy damage within
10–20 km away. Aftershock focal mechanisms determined a radius of ∼35 km of the epicenter. In total, 101 people were
by the first motion reveal similar solutions as the mainshock. killed (mainly caused by buildings collapsing), over 800 people
This earthquake sequence ruptured a previously unidentified were injured, thousands were displaced, and thousands of
fault that is either located at the west of the fault that pro- buildings collapsed or were severely damaged. The estimated
duced the 1967 M w 6.1 earthquake sequence or is actually at economic loss was about $139 million U.S. according to the
the same fault. The Pidie Jaya earthquake and other off-SF National Disaster Management Authority of Indonesia. This
events suggest strong distributed crustal deformation in Aceh, earthquake was well recorded by the national strong-motion

doi: 10.1785/0220180068 Seismological Research Letters Volume XX, Number XX – 2018 1


SRL Early Edition
ence in location up to ∼20 km (Fig. 1). After
the earthquake, both BMKG and USGS re-
leased estimated ground-shaking intensity maps
and the potential impacts (ShakeMap; Wald
et al., 2005). However, these preliminary inten-
sity maps are based on simplified models and
sparse datasets and cannot account for more
localized shaking effects.
Local people’s understanding of earth-
quakes in Aceh has improved since the shaking
and tsunami experience from the M w 9.3 earth-
quake in 2004 (e.g., Lay et al., 2005). Never-
theless, post-earthquake interviews of people on
site indicate that the Pidie Jaya earthquake was
very surprising to them because the shaking was
quite different in nature from their experience
of the 2004 earthquake. The 2016 event pro-
duced very sharp shaking and jerked vertically,
in contrast with the shocks in 2004 that were
relatively slow and moved horizontally. This is
possibly because the 2016 earthquake occurred
much closer to them, and the ground shaking
was dominated by high-frequency body waves,
whereas the more remote 2004 event produced
stronger long-period surface waves. In this ar-
▴ Figure 1. Shallow seismicity (depth < 50 km) and seismic networks in northern ticle, we combine interviews and site investiga-
Sumatra, Indonesia. Green, yellow, black, and blue triangles show broadband, tions to produce an updated intensity map for
strong-motion, temporary networks, and temporary strong-motion stations, respec- the Pidie Jaya earthquake, shown in the Inten-
tively. The transparent, gray, white, and red stars indicate the hypocenter location sity and Ground Motion section.
of the major event from the GEOFON Global Seismic Network (GFZ-Potsdam), the As shown in Figure 1, Pidie Jaya is located
U.S. Geological Survey (USGS), the Agency for Meteorology, Climatology, and Geo- at the northern coast of Aceh, part of northern
physics of Indonesia (BMKG), and this study, respectively. White circles indicate Sumatra, where the Indian Ocean plate sub-
the historical inland earthquakes with magnitudes larger than 5.5 since 1907 ducts beneath the Sunda plate at about
(Pesicek, Thurber, Widiyantoro, et al., 2010; Pesicek, Thurber, Zhang, et al., 2010; 5:2–6:0 cm=yr (McCaffrey, 1992; Sieh and Na-
Nugraha et al., 2018; USGS, BMKG, and International Seismological Centre). Red, tawidjaja, 2000). Although the subduction
yellow, and white circles in the inset showing the historical events on the Suma- process has generated significant uplift all along
tran fault (SF), off-SF, and subducting slab, respectively. Sumatra, the mountainous ridge that traverses
the island is the widest and highest in Aceh,
and broadband networks, with the closest station being where elevations reach ∼3381 m above sea level (Barber et al.,
∼80 km away (Fig. 1). Within 5 min of the origin time, 2005). Oblique subduction produced the right-lateral Suma-
BMKG reported the hypocenter location and magnitude of tran fault (SF) system through a slip partitioning process
the earthquake. A few minutes later, GFZ-Potsdam and U.S. (McCaffrey, 1992, 2009). Fast slip rates of ∼2:7–3:8 cm=yr
Geological Survey (USGS) at international level also issued have previously been proposed for the SF in the Aceh region
similar information (Table 1). All three institutions inferred (Sieh et al., 1994; Genrich et al., 2000). However, a more ac-
a similar location and magnitude of the earthquake, with differ- curate study suggests that the average slip rate along most of the

Table 1
Hypocenter Locations of the Mainshock from Three Institutions
Origin Time (UTC)
Number (yyyy/mm/dd hh:mm:ss) Latitude (°) Longitude (°) Magnitude (M w ) Depth (km) Source
1 2016/12/06 22:03:35 5.290 96.220 6.5 15 BMKG
2 2016/12/06 22:03:32 5.281 96.108 6.5 8.2 USGS
3 2016/12/06 22:03:33 5.320 96.070 6.5 10 GFZ
BMKG, Agency for Meteorology, Climatology, and Geophysics of Indonesia; USGS, U.S. Geological Survey.

2 Seismological Research Letters Volume XX, Number XX – 2018


SRL Early Edition
utions differed from that of the SAM-SIP fault
(∼180°), suggesting rupture of a previously
unidentified fault.
The national permanent seismic network
was installed gradually by BMKG after the
2004 M w 9.3 earthquake. It has now reached a
station spacing of ∼100 km in Sumatra, which is
still insufficient to detect and relocate relatively
small earthquakes and aftershock swarms with
enough precision to resolve fault geometries. Pre-
vious temporary networks installed in northern
Sumatra proved to be highly valuable for seismic-
velocity structure inversions and detection of ac-
tive faults (Muksin et al., 2014; U. Muksin, et al.,
unpublished manuscript, 2018; see Data and Re-
sources). Therefore, to better locate aftershocks
and understand fault structure, one week after
▴ Figure 2. The updated intensity map of the Pidie Jaya earthquake. Circles the Pidie Jaya earthquake, BMKG, GFZ-Pots-
indicate the locations of damage observations, colored by the intensity. Note that dam (German Research Center for Geosciences),
we use the earthquake intensity scale of BMKG (Skala Intensitas Gempabumi ITB (Bandung Institute of Technology), and
[SIG]-BMKG). Numbers connected to the circles are the indexes of photos in UNSYIAH (Syiah Kuala University) installed
Ⓔ Figure S1 (available in the electronic supplement to this article). The red star a temporary seismic network composed of nine
shows the relocated mainshock epicenter. Transparent triangles indicate the broadband stations in the ruptured area. The net-
locations of temporary Pijay-Net stations. work was distributed along the coast with an
average station spacing of ∼15 km (Fig. 1). The network was
SF, including in Aceh, is between ∼1:5 and ∼1:6 cm=yr (Brad- named as Pijay-Net and operated for about one month (14 De-
ley et al., 2017). A portion of the total strain produced along cember 2016–15 January 2017), providing valuable waveform
this complex convergent margin has been distributed within dataset for aftershock relocation and focal mechanism studies.
the Sumatran crust and is responsible for the complex modern In this study, we first summarize the macrointensity dis-
topographic pattern, with many active faults occurring away tribution, damage extension, and ground-motion characteris-
from the main strand of the SF. This is particularly true for tics of the mainshock along with site-response analyses. We
the region of the 2016 Pidie Jaya earthquake, in which two then relocate the mainshock, relocate the aftershocks that were
main SF strands (∼5:5° N) appeared to be nearly free of hori- recorded by the Pijay-Net, and determine the focal mechanisms
zontal geodetic strain (Genrich et al., 2000). It was proposed of relatively large aftershocks using P-wave first motion. Finally,
that the strain in this region was likely taking place further to we present a finite-fault inversion that uses the fault geometry
the east on the Lhokseumawe and Samalanga-Sipopok (SAM- constrained by the relocated seismicity.
SIP) faults (Genrich et al., 2000). The 2016 M w 6.5 Pidie Jaya
earthquake was located in the place very close to the
INTENSITY AND GROUND MOTION
northwestern tip of the SAM-SIP fault. Although initial re-
ports (e.g., USGS, BMKG, Global Centroid Moment Tensor Three days after the earthquake, a team from BMKG arrived in
[CMT]) of the earthquake indicated a strike-slip focal mecha- Pidie Jaya to collect mainshock intensity data. We investigated
nism, the preferred strikes (149°/55°) of those fault-plane sol- building damage at 49 sites in the vicinity of the epicenter region
and documented the damage with photographs and interviews.
Table 2 Here, we use the intensity scale defined by BMKG. This new
Earthquake Intensity Scale of BMKG (Skala Intensitas intensity scale, SIG-BMKG (Skala Intensitas Gempabumi-
Gempabumi [SIG]-BMKG) BMKG; Muzli et al., 2016), is a simplified version of the modi-
fied Mercalli intensity (MMI)mscale. It was designed to facilitate
SIG-BMKG Description MMI local understanding of earthquake impacts and implementation
I Not felt except by a very few I–II of hazard preparation and mitigation throughout Indonesia.
under especially conditions SIG-BMKG has been used in routine dissemination of significant
II Felt III–V earthquake information in Indonesia since 2016 (see Data and
III Light damage VI Resources). The SIG-BMKG recognizes only five levels, as shown
IV Moderate damage VII–VIII in Table 2 along with the corresponding MMI classification.
V Heavy damage IX–XII We used interview reports and inspections of building
and infrastructure damage to determine the SIG-BMKG inten-
MMI, modified Mercalli intensity. sity at the 49 sites, shown as circles in Figure 2. In general, the

Seismological Research Letters Volume XX, Number XX – 2018 3


SRL Early Edition

▴ Figure 3. Postevent survey photographs of closely adjacent sites showing drastically different responses of buildings to similar
ground-shaking intensity. (a) Site 31: complete collapse of a building, Kec. Meurah Dua at 5.233° N, 96.268 ° E; and (b) site 32: minimal
damage of an office at the same subdistrict at 5.235° N, 96.274° E.

intensities derived from interviews have larger uncertainties, a few damage reports from inland areas to the south of the main-
because they are subjective; on the other hand, inspections of shock epicenter. The direction of the building collapse or move-
buildings resulted in more accurate intensity estimations. Photo- ment agreed well with a left-lateral strike-slip focal mechanism
graphs of damage at these sites (numbered in Fig. 2) can be (e.g., sites 10, 11, 27, and 40 in Ⓔ Fig. S1), consistent with the
found in Ⓔ Figure S1 (available in the electronic supplement fault geometry defined by the relocated aftershocks. These pho-
to this article), with representative damage shown in Figure 3. tos provide independent evidence for the ruptured fault plane,
The number of fatalities and damaged buildings are shown in preferring the one with a strike of 055°. The aftershock distri-
Table 3. The maximum inferred intensity is V in SIG-BMKG or bution also implicates the fault plane with a strike of 055°, rather
IX in MMI, corresponding with total building collapse (e.g., than the orthogonal plane with a strike of ∼149°, which is more
Fig. 3). It is likely that the complete building collapse was caused parallel to the coast and to the SF.
by a combination of poor structural design, strong shaking, and We therefore assume a relocated hypocenter of 96.225°
site effect. At several sites, neighbored structures to collapsed E/5.235° N/15 km, M w of 6.5, and a strike-slip focal mechanism
buildings exhibited only light damage, indicating very different of the mainshock and predict the distribution of peak ground
building design and quality. For instance, sites 31 (Fig. 3a) and velocity (PGV) and peak ground acceleration (PGA) based on
the ground-motion prediction equation (GMPE) in Newmark
32 (Fig. 3b) are directly adjacent, but the degrees of damage are
and Hall (1982) and Zhao et al. (2006). We then convert the
substantially different. The most severe building damage was
PGV into intensity, with V S30 (the average shear-wave velocity
found near the relocated mainshock epicenter (see the details
down to 30 m) taken into account (Fig. 2; Wald et al., 2005).
of relocation in the Seismicity Relocation and Aftershock Focal
The conversion of ground motion to intensity is based on the
Mechanism Inversion section). The damage zone spreads in a empirical formula of the ground-motion-to-intensity conversion
nearly east–west direction, parallel to the coastline, with only equation as described by Worden et al. (2012). It shows that the
maximum predicted intensity is VIII in MMI at the relocated
Table 3 epicenter location. This predicted intensity is slightly smaller
Casualties and Building Damages Produced by the than that implied by the observed damage (see Fig. 2 and Ⓔ
Earthquake (BNPB) Fig. S1; e.g., indexes 18, 38, and 40), probably due to poor build-
ing quality. The predicted intensity map also shows an elongated
Number Description Total distribution of high intensity along the coast. The east–west-ori-
1 Deaths 101 ented damage zone in the residential area was likely strongly re-
2 Injured 857 lated to site effects rather than to the finite rupture process,
3 House damaged because the soft soil (very low V S30 ) zone is distributed along
Heavy 2992 the coast instead of along the strike of the fault (Fig. 4). Further
Moderate 94 inland to the south, the population is relatively low, the area is
Light 8582 hilly, and the soil is harder and thinner than in the coastal area.
Thus, the damage reported in this area is much less.
BNPB, National Disaster Management Authority of The thickness of sediment in the coastal region is ∼4:0 km
Indonesia. according to CRUST1.0 (Laske et al., 2013). However,

4 Seismological Research Letters Volume XX, Number XX – 2018


SRL Early Edition
CRUST1.0 is a low-resolution model, and a
more accurate sedimentary structure in this re-
gion awaits further investigation. Other data that
can indicate the thickness of the soil layer and
amplification effect are the horizontal-to-vertical
spectral ratios (HVSRs) from ambient seismic
noise data (Nakamura, 1989, 2008). Many stud-
ies (e.g., Parolai et al., 2002; Gosar and Lenart,
2010) have shown that there is a strong correla-
tion between the resonance frequency from
HVSR and soil thickness. The results of HVSR
from ambient noise data at Pijay-Net stations are
shown in Figure 5, in which peak resonance fre-
quency ranges from 0.7 to 8.0 Hz, indicating var-
iable soil thickness beneath the stations. The
lowest resonance frequency is observed at the
A22 station (0.7 Hz), corresponding with a
▴ Figure 4. Intensity data collected during the postevent survey (colored circles) 188-m soil thickness based on an empirical for-
overlain on an empirical V S30 map (Wald and Allen, 2007; Allen and Wald, 2009).
mula in Parolai et al. (2002). This soil layer has a
See Figure 2 for the intensity definition of the circles. The National Earthquake
significant effect on the local amplification of
Hazards Reduction Program (NEHRP) site classification (A, B, C, D, and E repre-
ground motion and resonance. Based on HVSR
senting the hard rock, rock, soft rock, stiff soil, and soft soil, respectively) is based
(table 2 in Zhao et al., 2006), we also estimated
on V S30 . The red star shows the relocated mainshock epicenter. Colored triangles
the site classification at the eight Pijay-Net sta-
indicate the locations of temporary Pijay-Net stations, and the colors represent site
tions. We found that, in general, the HVSR-
classification based on horizontal-to-vertical spectral ratio (HVSR; see Fig. 5).
derived site classifications agree well with that
derived from V S30 (Fig. 4); however, there are

▴ Figure 5. HVSR at Pijay-Net stations, solid and dashed lines showing the average HVSR and its 95% confidence level, respectively. The
variable frequency response patterns and different peak magnitudes of HVSR illustrate the importance of local site conditions for shallow
amplification of seismic waves. Site classifications based on HVSR (see table 2 in Zhao et al., 2006) are shown on each subfigure below
the station names.

Seismological Research Letters Volume XX, Number XX – 2018 5


SRL Early Edition

▴ Figure 6. The strong-motion record at Station JAKO of the


m b 4.9 aftershock event on 11 December, with the peak ground
▴ Figure 7. PGA as a function of source distance for the major
acceleration (PGA) ∼67 gals. The hypocentral distance to the sta-
event and aftershocks with different magnitudes and source dis-
tion is ∼19 km.
tances up to 600 km. Solid lines show predicted PGA based on the
ground-motion prediction equation from Zhao et al. (2006) for
some discrepancies; for instance, the site classification of A21 is crustal events with the magnitudes 2, 3, 4, 5, 6, 6.5, and 7, from
A + B and D from HVSR and V S30 , respectively. This difference lower to higher, respectively. The labels on the green circles in-
is not unexpected, because V S30 is derived from topography and dicate the station names of the temporary network.
has large uncertainties (e.g., Lemoine et al., 2012).
As part of the Pijay-Net, we also deployed one strong- shock Focal Mechanism Inversion section, the strike of the
motion station (JAKO) at 5.2393° N and 96.2499° E near the fault defined by the aftershocks and focal mechanisms does
major event epicenter two days after the earthquake (see the not align with the strike of these surface deformations. Because
station location in Fig. 1) that has recorded the aftershocks the largest asperity in the preliminary finite rupture model (see
between 9 December 2016 and 11 December 2016. An after- the Finite-Fault Slip Model section) and the majority of after-
shock with an mb of 4.9 was recorded by this station; the three- shocks (see the Seismicity Relocation and Aftershock Focal
component waveform of the earthquake recorded is shown in Mechanism Inversion section) were also located at the
Figure 6. Because the epicenter was located only 19 km away, 8–20-km depth range, the observed surface cracks were likely
the recorded PGA was ∼67 gals. Using data from the national caused by secondary ruptures, postseismic deformation, or
strong-motion network (BMKG), we assembled all strong-mo- near-surface effects.
tion records for the earthquake sequence. We plotted all the
PGA against the hypocenter distance in Figure 7, along with
the PGA predictions from GMPE described in Zhao et al. SEISMICITY RELOCATION AND AFTERSHOCK
(2006). We used the GMPE from Zhao et al. (2006) because FOCAL MECHANISM INVERSION
there is still no GMPE available in Sumatra, and the region of
Japan, which was studied in Zhao et al. (2006), is probably one To better understand the fault geometry and associated tecton-
of the most similar places to Sumatra in terms of tectonics. ics, we used the arrival-time picks on the waveform data
Both Sumatra and Japan are located on the overriding plates recorded by Pijay-Net to locate the aftershocks that occurred
in the subduction-zone systems, where the significant ground between 14 December 2016 and 15 January 2017. The
motion comes from the megathrust on the plate interface and regional broadband data, that is the closest at 80 km away from
earthquakes on the inland faults. The predicted PGA and PGV the aftershocks, were not included for the small-events reloca-
of the mainshock at the epicenter location (5.235° N; tion. During this period, Pijay-Net recorded more than 300
96.225° E) are 284 gals and 23:8 cm=s, respectively. Overall, events with magnitudes larger than 0.50. Among these events,
the ground motions predicted by GMPE are comparable with we have 260 earthquakes with M L > 1:0, for which we can
our observations for the mainshock (see Fig. 7), although the pick P-wave arrivals with high accuracy. Because the stations
prediction underestimates the PGA at the closest sites and are very close to the event locations, the averaged accuracy of
overestimates it in the far field for the M 4.9 and 4.2. arrival picks is as high as ∼0:1 s. These picks were then used to
During the postearthquake survey, we did not locate clear locate the earthquakes through a grid-search manner (e.g., Wei
evidence of surface rupture along the strike of the fault. et al., 2015). To calculate the theoretical arrival time, we use
However, the earthquake did produce some localized surface the CRUST1.0 model (Laske et al., 2013) and a grid spacing of
deformations along the coastal area (Ⓔ Fig. S1; e.g., indexes 0.5 km. After the initial grid search, we further relocate the
22 and 28). As shown in the Seismicity Relocation and After- events with a double-difference algorithm implemented in

6 Seismological Research Letters Volume XX, Number XX – 2018


SRL Early Edition

▴ Figure 8. Aftershock distribution and focal mechanism solution of the 2016 Pidie Jaya earthquake after relocation using grid-search
and hypoDD methods. Numbers on the focal mechanism solution are the event ID in the Ⓔ electronic supplement. Green squares indicate
the three events with their waveforms recorded at Station A22 (gray triangle) as shown in Figure 9. Blue circles indicate the sequence of
an historically significant M 6.1 earthquake in 1967.

hypoDD (Waldhauser and Ellsworth, 2000; Waldhauser, (shown as blue circles). The surface projection of the relocated
2001). The relocation results in map view and vertical profiles seismicity shows a clear lineation that has a fault-length dimen-
are shown in Figure 8. We also plotted USGS locations for the sion of ∼20 km and strikes toward an azimuth of ∼N042°,
historical 1967 M 6.1 earthquake sequence on the same figure agreeing well with the strike of one of the fault-plane solutions

Seismological Research Letters Volume XX, Number XX – 2018 7


SRL Early Edition
into consideration in the future seismic hazard
estimation in this region. Notably, the preferred
strike of ∼055° is inconsistent with the orien-
tation of north–south strike-slip faults that ap-
pear on maps but have no clear topographic
expression (such as the Samalanga-Sipopok
fault), as well as with the north–south strikes
of inverted basin-bounding normal faults of
the well-known north Aceh basin just to the
east of the study area (Barber et al., 2005).
Among the 260 aftershocks, there are 23
events having magnitudes larger than 3.0. The
largest magnitude (M L 4.2) recorded by Pijay-
Net occurred on 17 December 2016 was also
well recorded by the regional network. How-
ever, the location of this earthquake from the
USGS epicenter report is 15.5 km away from
our relocated epicenter. Because our location is
much more accurate than the global and
▴ Figure 9. Waveforms of the deepest event (28 km) and other two events (21 and regional catalogs, due to our proximal and dense
11 km) at Station A22 (see the event locations in Fig. 8). The S (T2) to P (T1) differ- network, we can calibrate the paths to the
ential arrival times increase with event depth (h). BMKG and GFZ or Incorporated Research In-
stitutions for Seismology (IRIS) broadband sta-
tions and refine the epicenter of the mainshock. To do that, we
of the mainshock. Interestingly, a vertical cross section along
calculate the theoretical arrival times using the CRUST1.0
the direction perpendicular to this strike reveals more complex-
velocity model at regional broadband stations (Fig. 1) for
ity in the aftershock distribution. The majority of the seismic-
the M L 4.2 aftershock and compare them with high-quality
ity defines a fault plane that dips ∼63° toward southeast, again
handpicked arrivals. The residuals are then used to correct
consistent with one of the mainshock fault-plane solutions. the theoretical arrival time for the mainshock. With this cor-
However, there are at least two vertical branches at depths rection, we refine the mainshock epicenter in a grid-search
of 12 and 18 km (Fig. 8), indicating possible vertical segmen- manner. Our relocation of the mainshock epicenter
tation of the fault. The lineup of the seismicity at one of the (5.235° N; 96.225° E) is consistent with our relocated after-
branches dips ∼25° to the northwest. This part of the seismic- shocks (Fig. 8 and Ⓔ Fig. S2). The path calibration for these
ity is very localized, as indicated in the map view, and could regional stations can be used to refine the location of future
represent a small fault that either ruptured during or was sub- earthquakes in northern Sumatra.
sequently activated by the mainshock. A planar extrapolation We then estimate focal mechanisms of the large after-
of the main seismicity along a dip angle of 63° intersects the shocks (M L > 3:0) using an inversion of the P-wave first-
surface about 5 km to the west of the seismicity cloud. This motion data from the Pijay-Net, with an approach proposed
surface projection coincides with linear topographic ridges that by Suetsugu (1997). Waveform records of M L > 3:0 events
align with the strike of the fault and that mark anticlines and show clear P-wave first-motion polarities, due to their relatively
synclines deforming Late Miocene and Pliocene sediments and large magnitude. All together, we obtained focal mechanisms of
overlying Pleistocene arc volcanics (Bennett et al., 1980; Fig. 8). 17 events that are shown in Figure 8, both in map view and
We name this structure the Panteraja fault. The depths of depth profiles. The majority of these events have a mechanism
aftershocks are primarily distributed between 8 and 23 km, similar to that of the mainshock; such consistency further veri-
with the deepest event located at a depth of 28 km (Fig. 8), fies the fault geometry determined by our refined seismicity.
indicating that the lower crust is seismogenic because the We also invert the double-couple solution of the mainshock by
crustal thickness in this region is only 30 km, according to using broadband waveforms recorded by the regional BMKG
the CRUST1.0 model. The depth of the seismicity is further stations that have a good azimuthal coverage for the earth-
verified by inspecting the relative arrival time between P and S quake. After removing the instrument response from the raw
waves at the closest station, A22 (Fig. 9). It is clear that the data, we applied the cut-and-paste method (Zhu and Helm-
deepest event has the largest S-to-P differential arrival time. berger, 1996) to invert for strike/dip/rake/M w /depth of the
A deep potential seismogenic depth should therefore be con- earthquake in a grid-search manner. A 1D velocity model
sidered in the future seismic hazard estimation in this region. A extracted from Collings et al. (2012) was used to calculate
deeper seismogenic zone can physically facilitate a longer rup- the Green’s function library for the inversion. The depth res-
ture dimension along strike and therefore generate larger earth- olution and best waveform fits are shown in Ⓔ Figures S3 and
quakes (Weng and Yang, 2017). This factor should be taken S4, respectively. Both data and synthetics show the largest am-

8 Seismological Research Letters Volume XX, Number XX – 2018


SRL Early Edition

▴ Figure 10. The teleseismic finite-fault inversion result of the mainshock. (a) The teleseismic displacement waveform fits where the
actual data are shown in black and the synthetic in red. The station names are indicated at the beginning of each waveform pair along
with the azimuth (upper) and epicenter distance (lower) in degree. The peak amplitude in the data in micrometers is displayed at the end
of each waveform. (b) The depth profile of the finite-fault slip model in which the contours indicate the rupture time and the arrows
indicate the slip direction. (c) The moment rate function of the finite-fault model.

plitude on the tangential components, which is typical for a FINITE-FAULT SLIP MODEL
strike-slip focal mechanism. As shown, the centroid depth
of the earthquake is ∼11 km, which is in agreement with the Although the magnitude of the Pidie Jaya earthquake is rela-
finite-slip distribution derived from teleseismic inversion, as tively small (M w 6.5) compared with other damaging events,
shown in the Finite-Fault Slip Model section. (The relocated we attempted a teleseismic finite-fault inversion to resolve the
aftershocks and focal mechanism can be found in the Ⓔ elec- fault plane of the earthquake. We downloaded broadband
tronic supplement.) teleseismic P waves recorded by the Global Seismic Network

Seismological Research Letters Volume XX, Number XX – 2018 9


SRL Early Edition
from IRIS (see Data and Resources). We removed the instru- the SF. Thus far, earthquakes on off-SF structures in Aceh have
ment response and converted the waveform data into displace- been strike-slip events, but comparison with other strike-slip
ment at 2.0 Hz and a lower frequency. We used a hypocenter of margins and the occurrence of very high topographic relief in
15 km, as determined by BMKG, and a 1D velocity model the region suggests that thrust or normal-mechanism events
extracted from CRUST1.0 in the inversion. Adopting an in- could also occur. Assessment of earthquake hazard and risk typ-
version scheme proposed by Ji et al. (2002), we performed fi- ically relies on geological mapping of active faults; however, the
nite-fault inversions using the two fault planes in the Global 2016 Pidie Jaya and 2013 Bener Meriah earthquakes clearly
CMT solution, allowing the slip to vary from 0 to 3 m and the show that unmapped structures of unexpected orientation
rupture speed to change from 2.0 to 3:0 km=s. Interestingly, as may rupture even in areas where basement fault systems have
shown in Figure 10 and Ⓔ Figure S5, it is clear that the in- already been mapped. Geophysical monitoring may therefore
version on a fault plane with a strike of 057° and a dip of 63° prove to be a more reliable method for discovering active faults
fits the data much better than the auxiliary fault plane, which is in this region prior to damaging ruptures.
again consistent with the relocated seismicity and surface geol- To better prepare for future damaging earthquakes in
ogy. The slip distribution of the finite-fault inversion Sumatra, it is also critical to collect the seismic ground-shaking
(∼12 km) also agrees well with the centroid depth in the mo-
data from local to regional scale, which can be used for a series
ment tensor solution obtained by regional waveform inversions
of studies that include site effects and attenuations, as partly
(Ⓔ Figs. S3 and S4).
covered in this study. Temporary network and national
strong-motion and broadband networks provide the funda-
DISCUSSION mental dataset for such analyses. With more data collected and
analyzed, much better GMPEs in Sumatra or Aceh can be
Despite the rapidly increasing population of Aceh, the location, derived. In addition to off-SF earthquakes, such as the Pidie
geometry, and slip rates of many secondary but still hazardous Jaya earthquake in this study, the Sumatran region experiences
faults remain uncertain. These faults are more active in Aceh gigantic megathrust earthquakes that can generate strong
than are the other secondary faults in Sumatra, as shown by ground shaking at both long periods and high frequencies
background seismicity (Fig. 1, inset). Over the last 100 yrs, (e.g., Jousset and Douglas, 2007). The civil engineering com-
off-SF faults in Aceh have produced 17 M > 5:5 shallow munity usually ignores long-period ground shaking; however,
events, compared with only three such events in the rest of the long-period shaking produced by the 2011 Tohoku-Oki
Sumatra. These earthquakes include an M w 6.1 earthquake earthquake swayed the top of a 52-floor building in Osaka
that took place near the SAM-SIP fault in 1967 about 12 km bay more than 1.3 m horizontally, despite the building being
away from the recent Pidie Jaya earthquake (USGS; Pesicek, located more than 800 km away from the epicenter (Kanamori,
Thurber, Widiyantoro, et al., 2010; Pesicek, Thurber, Zhang, 2014). Such 3D structural complexity as well as building re-
et al., 2010) and the M w 6.1 Bener Meriah earthquake in 2013 sponses should also be taken into consideration for future seis-
that occurred about 50 km away (BMKG). A Global Position- mic hazard preparation, in particular for Jakarta, capital city of
ing System (GPS) velocity profile across the SF near the 2016
Indonesia.
Pidie Jaya earthquake shows a smooth transition in fault-
parallel velocity across the SF trace, indicating that the sur-
rounding crust is stressed by fault locking (Ito et al., 2012).
CONCLUSIONS
Transmission of stress from the subduction interface and
across the SF is presumably responsible for earthquakes such as
The results of aftershock analysis from Pijay-Net and main-
the Pidie Jaya event. However, because the slip rates of secon-
shock relocation show that the ruptured fault plane has a strike
dary faults are presumably very low, it is challenging to identify
of 42°, a dip of 63°, and a fault dimension of ∼20 km along
them, even with decadal GPS data. Although the national
strike and as deep as 25 km. The depths of the aftershocks are
broadband network in Sumatra region is coarse (∼100 km
station spacing), temporary seismic networks can be installed primarily distributed between 8 and 23 km, with maximum
to densify the station coverage and provide near-field depth of 28 km. We name this structure the Panteraja fault
data to precisely locate microseismicity or aftershocks (e.g., and tentatively correlate it with near-surface deformational fea-
U. Muksin, et al., unpublished manuscript, 2018; see Data tures that parallel its strike. Focal mechanisms of aftershocks
and Resources). We therefore suggest installing a more tempo- with a magnitude greater than 3 also indicate a fault geometry
rary short-period seismic network in the region to complement consistent with relocated seismicity and the focal mechanism
the national broadband network, particularly after significant of the mainshock. Damage caused by the earthquake was dis-
earthquakes. tributed primarily in the east–west direction, instead of along
The geological context of the moderate magnitude (M 6+) the strike of the rupture fault plane, due to the effect of
earthquakes in Aceh remains unclear and could be improved ground-motion amplification caused by greater sediment thick-
through more detailed field and geophysical observations of ness along the coastline. To better understand the neotectonics
active faulting. Certainly, the Pidie Jaya earthquake highlights and prepare for the future seismic hazard, we suggest improving
the hazard posed by auxiliary faults oriented at a high angle to the seismic monitoring of the entire Aceh region.

10 Seismological Research Letters Volume XX, Number XX – 2018


SRL Early Edition
DATA AND RESOURCES Barber, A. J., M. J. Crow, and J. Milsom (Editors) (2005). Sumatra: Geol-
ogy, Resources and Tectonic Evolution, Geological Society of London,
London, United Kingdom.
Seismograms of the aftershocks used in this study were col- Bennett, J. D., D. M. Bridge, N. R. Cameron, A. Djunuddin, S. A. Gha-
lected by the temporary network of Pijay-Net. These data will zali, D. H. Jeffery, W. Keats, N. M. S. Rock, S. J. Thompson, and R.
be released to public after 5 yrs (November 2022) through the Whandoyo (1980). Geologic Map of the Banda Aceh Quadrangle,
data center of webdc.eu, the Incorporated Research Institutions North Sumatra, Bandung, Indonesia, Geological Research and
for Seismology (IRIS), or a data center at the Earth Observa- Development Centre, Bandung, Indonesia.
Bradley, K. E., L. Feng, E. M. Hill, D. H. Natawidjaja, and K. Sieh
tory of Singapore. Some aftershock and mainshock seismo- (2017). Implications of the diffuse deformation of the Indian
grams were collected from the Agency for Meteorology, Ocean lithosphere for slip partitioning of oblique plate convergence
Climatology, and Geophysics (BMKG), the GEOFON Global in Sumatra, J. Geophys. Res. 122, 572–591, doi: 10.1002/
Seismic Network (GFZ-Potsdam), and the Global Seismic 2016JB013549.
Collings, R., D. Lange, A. Rietbrock, F. Tilmann, D. H. Natawidjaja, B.
Network (GSN) from IRIS (iris.edu). Part of the historical
Suwargadi, M. Miller, and J. Saul (2012). Structure and seismogenic
catalog data were obtained from the U.S. Geological Survey properties of the Mentawai segment of the Sumatra subduction
(USGS) website and the International Seismological Centre zone revealed by local earthquake traveltime tomography, J. Geophys.
(1900–1959). Damage photos were collected during the field Res. 117, no. B1, doi: 10.1029/2011JB008469.
survey. An example of earthquake information based on the Genrich, J. F., Y. Bock, R. McCaffrey, L. Prawirodirdjo, C. W. Stevens, S.
S. O. Puntodewo, C. Subarya, and S. Wdowinski (2000). Distribu-
intensity scale of Skala Intensitas Gempabumi (SIG)-BMKG tion of slip at the northern Sumatran fault system, J. Geophys. Res.
is on the BMKG website: www.bmkg.go.id/berita/?p=english- 105, no. B12, 28,327–28,341.
magnitude-6-5-earthquake-hits-aceh&lang=EN. Some plots were Gosar, A., and A. Lenart (2010). Mapping the thickness of sediments in
made using the Generic Mapping Tools v.4.5.6 (www.soest the Ljubljana Moor basin (Slovenia) using microtremors, Bull.
.hawaii.edu/gmt; Wessel and Smith, 1998; Wessel et al., 2013). Earthq. Eng. 8, no. 3, 501–518.
Ito, T., E. Gunawan, F. Kimata, T. Tabei, M. Simons, I. Meilano, Y. Ohta,
All websites were last accessed on May 2018. All other data and I. Nurdin, and D. Sugiyanto (2012). Isolating along-strike variations
resources used in this article came from published sources listed in the depth extent of shallow creep and fault locking on the
in the references. The unpublished mauscript by U. Muksin, northern Great Sumatran fault, J. Geophys. Res. 117, no. B6,
K. Bauer, M. Muzli, T. Ryberg, Irwandi, Masturyono, and doi: 10.1029/2011JB008940.
M. Weber (2018), “AcehSeis project provides insights into the Ji, C., D. J. Wald, and D. V. Helmberger (2002). Source description of the
1999 Hector Mine, California, earthquake, part I: Wavelet domain
detailed seismicity distribution and relation to fault structures inversion theory and resolution analysis, Bull. Seismol. Soc. Am. 92,
in Central Aceh, Northern Sumatra,” submitted to J. Asian no. 4, 1192–1207.
Earth Sci. Jousset, P., and J. Douglas (2007). Long-period earthquake ground
displacements recorded on Guadeloupe (French Antilles), Earthq.
Eng. Struct. Dynam. 36, no. 7, 949–963.
Kanamori, H. (2014). The diversity of large earthquakes and its implica-
ACKNOWLEDGMENTS tions for hazard mitigation, Annu. Rev. Earth Planet. Sci. 42, 7–26.
Laske, G., G. Masters, Z. Ma, and M. Pasyanos (2013). Update on
The authors are grateful to the team from the Agency for CRUST1.0—A 1-degree global model of Earth’s crust, EGU
Meteorology, Climatology, and Geophysics of Indonesia General Assembly, Vienna, Austria, 7–12 April, EGU2013–2658.
(BMKG), the ITB (Bandung Institute of Technology), Lay, T., H. Kanamori, C. J. Ammon, M. Nettles, S. N. Ward, R. C. Aster,
S. L. Beck, S. L. Bilek, M. R. Brudzinski, R. Butler, et al. (2005). The
UNSYIAH (Syiah Kuala University), and Profil Pusat Studi great Sumatra–Andaman earthquake of 26 December 2004, Science
Gempa Bumi Nasional (PUSGEN) for their great support 308, no. 5725, 1127–1133.
in the fieldwork, especially Zulfakriza, Kadek Hendrawan Pal- Lemoine, A., J. Douglas, and F. Cotton (2012). Testing the applicability
gunadi, Rizkiray Ryannugroho, Angga Setiyo Prayogo, Rasmit, of correlations between topographic slope and V S30 for Europe,
and Tommy Ardiansyah. The authors also thank local citizens Bull. Seismol. Soc. Am. 102, no. 6, 2585–2599.
McCaffrey, R. (1992). Oblique plate convergence, slip vectors, and
for their help in completing the interviews. This work is sup- forearc deformation, J. Geophys. Res. 97, no. B6, 8905–8915.
ported by the Earth Observatory of Singapore fund McCaffrey, R. (2009). The tectonic framework of the Sumatran subduc-
(M443B50240) and partly funded by the BMKG, ITB, and tion zone, Annu. Rev. Earth Planet. Sci. 37, 345–366.
UNSYIAH. The authors used stations from the Geophysical Muksin, U., C. Haberland, M. Nukman, K. Bauer, and M. Weber (2014).
Instrumental Pool of Potsdam for the temporary network of Detailed fault structure of the Tarutung pull-apart basin in Sumatra,
Indonesia, derived from local earthquake data, J. Asian Earth Sci.
broadband stations. They would like to thank Paul Tapponnier 96, 123–131.
for insightful discussions. They thank three anonymous re- Muzli, M., M. Masturyono, J. Murjaya, and M. Riyadi (2016). Studi Awal
viewers and the editors who provided valuable comments that Penyusunan Skala Intensitas Gempabumi Badan Meteorologi
improved the original article. Klimatologi dan Geofisika (Preliminary Study on Earthquake In-
tensity Scale of BMKG), J. Meteorol. Geofis. 17, no. 2, 89–99
(in Indonesian).
Nakamura, Y. (1989). A method for dynamic characteristics estimation of
REFERENCES subsurface using microtremor on the ground surface, Q. Rep.
Railway Tech. Res. Inst. 30, no. 1, 25–33.
Allen, T. I., and D. J. Wald (2009). On the use of high-resolution Nakamura, Y. (2008). On the H/V spectrum, The 14th World Conf. on
topographic data as a proxy for seismic site conditions (V S30 ), Bull. Earthquake Engineering, Beijing, China, 12–17 October 2008.
Seismol. Soc. Am. 99, no. 2A, 935–943.

Seismological Research Letters Volume XX, Number XX – 2018 11


SRL Early Edition
Newmark, N. M., and W. J. Hall (1982). Earthquake spectra and design, eters and modified Mercalli intensity in California, Bull. Seismol.
Geotechnique 25, no. 2, 139–160. Soc. Am. 102, no. 1, 204–221.
Nugraha, A. D., H. A. Shiddiqi, S. Widiyantoro, C. H. Thurber, J. D. Zhao, J. X., J. Zhang, A. Asano, Y. Ohno, T. Oouchi, T. Takahashi, H.
Pesicek, H. Zhang, S. H. Wiyono, M. Ramdhan, and M. Irsyam Ogawa, K. Irikura, H. K. Thio, P. G. Somerville, et al. (2006).
(2018). Hypocenter relocation along the Sunda Arc in Indonesia, Attenuation relations of strong ground motion in Japan using site
using a 3D seismic velocity model, Seismol. Res. Lett. 89, no. 2A, classification based on predominant period, Bull. Seismol. Soc. Am.
603–612. 96, no. 3, 898–913.
Parolai, S., P. Bormann, and C. Milkereit (2002). New relationships Zhu, L., and D. V. Helmberger (1996). Advancement in source estima-
between V S , thickness of sediments, and resonance frequency tion techniques using broadband regional seismograms, Bull. Seis-
calculated by the H/V ratio of seismic noise for the Cologne area mol. Soc. Am. 86, no. 5, 1634–1641.
(Germany), Bull. Seismol. Soc. Am. 92, no. 6, 2521–2527.
Pesicek, J. D., C. H. Thurber, S. Widiyantoro, H. Zhang, H. R. DeShon, Muzli Muzli1
and E. R. Engdahl (2010). Sharpening the tomographic image of the
subducting slab below Sumatra, the Andaman Islands and Burma, Kyle Edward Bradley
Geophys. J. Int. 182, no. 1, 433–453. Shengji Wei2
Pesicek, J. D., C. H. Thurber, H. Zhang, H. R. DeShon, E. R. Engdahl, Earth Observatory of Singapore
and S. Widiyantoro (2010). Teleseismic double-difference reloca- Nanyang Technological University
tion of earthquakes along the SumatraAndaman subduction zone Singapore
using a 3-D model, J. Geophys. Res. 115, no. B10, doi: 10.1029/
2010JB007443. shjwei@ntu.edu.sg
Sieh, K., and D. H. Natawidjaja (2000). Neotectonics of the Sumatran
fault, Indonesia, J. Geophys. Res. 105, no. B12, 28,295–28,326. Muksin Umar
Sieh, K., J. Zachariasen, Y. Bock, L. Edwards, F. Taylor, and P. Gans Irwandi Nurdin
(1994). Active tectonics of Sumatra, Geol. Soc. Am. Abstr. Progr. Department of Physics
26, 382.
Suetsugu, D. (1997). Source mechanism practice, Lecture Notes, Syiah Kuala University
International Institute of Seismology and Earthquake Engineering, Banda Aceh
Tsukuba, Japan. Indonesia
Wald, D. J., and T. I. Allen (2007). Topographic slope as a proxy for
seismic site conditions and amplification, Bull. Seismol. Soc. Am. Andri Dian Nugraha
97, no. 5, 1379–1395.
Wald, D. J., B. C. Worden, V. Quitoriano, and K. L. Pankow (2005). Sri Widiyantoro
ShakeMap Manual: Technical Manual, User’s Guide, and Software Faculty of Mining and Petroleum Engineering
Guide (No. 12-A1), available at https://pubs.usgs.gov/tm/2005/ Bandung Institute of Technology
12A01/ (last accessed November 2017). Indonesia
Waldhauser, F. (2001). hypoDD—A program to compute double-differ-
ence hypocenter locations (hypoDD version 1.0-03/2001), U.S.
Geol. Surv. Open-File Rept. 01, 113 pp. Kemal Erbas
Waldhauser, F., and W. L. Ellsworth (2000). A double-difference earth- Philippe Jousset
quake location algorithm: Method and application to the northern Helmholtz Centre Potsdam GFZ German Research Centre for
Hayward fault, California, Bull. Seismol. Soc. Am. 90, no. 6, 1353– Geosciences
1368. Germany
Wei, S., S. Barbot, R. Graves, J. J. Lienkaemper, T. Wang, K. Hudnut, Y.
Fu, and D. Helmberger (2015). The 2014 M w 6.1 South Napa
earthquake: A unilateral rupture with shallow asperity and rapid Supriyanto Rohadi
afterslip, Seismol. Res. Lett. 86, no. 2A, 344–354. Agency for Meteorology, Climatology, and Geophysics
Weng, H., and H. Yang (2017). Seismogenic width controls aspect ratios (BMKG)
of earthquake ruptures, Geophys. Res. Lett. 44, no. 6, 2725–2732, Indonesia
doi: 10.1002/2016GL072168.
Wessel, P., and W. H. Smith (1998). New, improved version of Generic
Mapping Tools released, Eos Trans. AGU 79, no. 47, 579. Published Online 25 July 2018
Wessel, P., W. H. Smith, R. Scharroo, J. Luis, and F. Wobbe (2013).
Generic Mapping Tools: Improved version released, Eos Trans. 1
Also at Agency for Meteorology, Climatology, and Geophysics
AGU 94, no. 45, 409–410. (BMKG), Indonesia.
Worden, C. B., M. C. Gerstenberger, D. A. Rhoades, and D. J. Wald 2
Also at Asian School of the Environment, Nanyang Technological Uni-
(2012). Probabilistic relationships between ground-motion param- versity, Singapore.

12 Seismological Research Letters Volume XX, Number XX – 2018

You might also like