You are on page 1of 7

IOP Conference Series: Earth and Environmental Science

PAPER • OPEN ACCESS You may also like


- Foreshocks and aftershocks of strong
Hypocenter relocation of the Mw 5.9 eastern earthquakes in the light of catastrophe
theory
Manggarai earthquake 2022 and its aftershocks A V Guglielmi

- An active extensional deformation


based on BMKG seismic network example: 19 May 2011 Simav earthquake
(Mw = 5.8), Western Anatolia, Turkey
Alper Demirci, Süha Özden, Tolga Bekler
To cite this article: M Ramdhan et al 2023 IOP Conf. Ser.: Earth Environ. Sci. 1227 012042 et al.

- Omori law for foreshocks and aftershocks


in a realistic earthquake model
O. M. Braun and M. Peyrard

View the article online for updates and enhancements.

This content was downloaded from IP address 114.122.74.156 on 27/09/2023 at 15:09


The 4th Southeast Asian Conference on Geophysics (SEACG 2022) IOP Publishing
IOP Conf. Series: Earth and Environmental Science 1227 (2023) 012042 doi:10.1088/1755-1315/1227/1/012042

Hypocenter relocation of the Mw 5.9 eastern Manggarai


earthquake 2022 and its aftershocks based on BMKG
seismic network

M Ramdhan1,*, Priyobudi2, A Mursitantyo3, K H Palgunadi4,


A L Panjaitan2, J Jatnika2
1
Research Center for Geological Disaster, National Research and Innovation
Agency (BRIN), KST Samaun Samadikun, Jl. Sangkuriang, Bandung, 40135,
Indonesia
2
Indonesian Agency for Meteorology, Climatology, and Geophysics (BMKG),
Jl. Angkasa I No.2, Kemayoran, Jakarta 10720, Indonesia
3
Geophysics Station of Bandung (BMKG-Bandung), Jl. Cemara No.66,
Bandung, 40161, Indonesia
4
Physical Science and Engineering, King Abdullah University of Science
and Technology, Thuwal, Saudi Arabia

E-mail: moha080@brin.go.id

Abstract. The moderate earthquake (Mw 5.9) struck eastern Manggarai and its
surroundings on February 21, 2022, at 12:36:00 UTC. Although this earthquake was
classified as a moderate earthquake, it produced a significant number of aftershocks. The
aftershocks could be observed as the post-seismic activity, which was still well recorded
until March 31, 2022. The hypocenter location of the aftershocks is updated by applying
the double-difference method. The aftershock distribution is striking in west-east
orientation and dipping to the south, emphasizing a geometry of faults. The strike and dip
directions are consistent with the focal mechanism determined by the Indonesian Agency
for Meteorology, Climatology, and Geophysics (BMKG). Most aftershocks after the
relocation are distributed at a depth of < 25 km; the depth of the mainshock is 21.5 km.
The aftershock distribution in this study shows that the earthquake sequence propagates
toward the up-dip direction. The aftershock relocation also reveals that most aftershocks
are situated slightly off the main fault plane in the complex splay faults. Therefore, we
conjecture that the co-seismic activity may trigger the aftershocks. In addition, this study
also shows that the fault system consists of several fault segments. Our study benefits
earthquake disaster mitigation, especially in mapping the segment faults of the Flores
backarc thrust.

1. Introduction
The eastern Manggarai is located on Flores Island and is considered seismologically active. The
earthquake activity on the island is driven by the Flores backarc thrust (FBT) in the northern part

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution
of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
Published under licence by IOP Publishing Ltd 1
The 4th Southeast Asian Conference on Geophysics (SEACG 2022) IOP Publishing
IOP Conf. Series: Earth and Environmental Science 1227 (2023) 012042 doi:10.1088/1755-1315/1227/1/012042

and the Sunda-Banda transition zone in the southern region. The FBT system around the island is
associated with shallow earthquakes, and the second system relates to deep earthquakes. The last
system is a transition zone from the subduction to the collision system. All systems make the
island vulnerable to infrastructure damages that cause casualties and injuries if a big earthquake
hits the island.
The 2022 eastern Manggarai earthquake has a moment magnitude of Mw 5.9. This earthquake
is classified as a moderate earthquake. However, the main shock was later followed by many
aftershocks. The mainshock was felt in Labuan Bajo, Maumere, Ende, Soa, and Ruteng with MMI
III. All locations are parts of Flores island. The earthquake occurred in one of the FBT segments.
The FBT system could potentially generate Mw 7.7 (e.g., in 1992 [1]). The earthquake caused a
tsunami with a submarine landslide, generating wave heights up to 26 m in Maumere. The tsunami
destroyed 90% of buildings, including 800 schools [2]. Therefore, a detailed analysis of the FBT
system, which could reveal an active fault, is highly crucial
This study aims to understand the significant aftershocks driven by the mainshock of the 2022
Mw 5.9 eastern Manggarai earthquake based on the hypocenter relocation of the aftershock. The
study will shed new light on the earthquake characteristics in one of the FBT segments. Knowing
the earthquake characteristics in a particular area is very useful for disaster mitigation, especially
for the construction and development of infrastructures. Hence, if a significant earthquake strikes
in the future, the seismic hazard assessment can reduce casualties and infrastructure damages in
the region.

Figure 1. Tectonic setting and station distribution used in this study. Yellow squares and red
triangles represent a distribution of seismic stations and volcanoes, respectively. The
study area lies inside the black box. The FBT line fault was modified from [3,4].

2. Data and Methods


This study used the Indonesian Agency for Meteorology, Climatology, and Geophysics (BMKG)
catalog at coordinates 119.400-122.50E and 7.00-9.50S from February 21 to March 31, 2022.
Figure 1 represents seismic stations, the tectonic setting, and the study area. The parameters of
the earthquake data were updated using a hypocenter relocation technique, such as updating active
faults and earthquake source hazard maps [5,6]. We applied the double-difference hypocenter
relocation method in Cartesian coordinates using the HypoDD package [7,8]. We applied the 1-

2
The 4th Southeast Asian Conference on Geophysics (SEACG 2022) IOP Publishing
IOP Conf. Series: Earth and Environmental Science 1227 (2023) 012042 doi:10.1088/1755-1315/1227/1/012042

D velocity model from the pre-existing model considering Vp/Vs ratio of 1.73 [9,10]. We only
considered an earthquake with at least 6 phases (both P and S phases or P phase only). The FBT
system is associated with the shallow earthquake, so we neglected more than 30 km of earthquake
depth. The maximum distance for each earthquake to the seismic station is 300 km.

Figure 2. Histograms of relative error locations of hypocenters (X0, Y0, Z0) have a 95%
confidence ellipsoid level.
3. Results and Discussion
The number of earthquake hypocenters successfully relocated, including the mainshock, was 223
out of 232 events from the BMKG catalog. Some earthquakes were not relocated because of a
nonconvergent solution during the inversion process. We assessed the relative error of the
hypocenter location (X0, Y0, Z0) using the bootstrap method [7,11,12]. The results show average
X0 and Y0 errors of less than 0.5 km. While the average error Z0 is less than 0.8 km. The vertical
error is expected to be higher because all stations are on the earth's surface. The statistic
parameters of relative error locations in this study are depicted in Figure 2 and Table 1.

Tabel 1. Statistic parameters for relative error locations of hypocenters (X0, Y0, Z0) using the
bootstrap method.

In general, the strike and dip directions of the FBT are in the west-east and south with a gentle
slope, as shown in Figure 1. The aftershock distribution is consistent with the focal mechanism
released by BMKG. The source focal mechanism has a strike orientation of N94.5 0E and a dip
angle of 23.50. The strike orientation could still be estimated from the BMKG earthquake catalog,
albeit the epicenter distribution is more diverse than the relocated aftershocks, as shown in Figure
3. In contrast to the dip direction, it is challenging to determine the fault dip if the hypocenter is
not relocated, as shown in Figure 4 (c).

3
The 4th Southeast Asian Conference on Geophysics (SEACG 2022) IOP Publishing
IOP Conf. Series: Earth and Environmental Science 1227 (2023) 012042 doi:10.1088/1755-1315/1227/1/012042

After the hypocenter relocation, the fixed depths (10 km) no longer exist; therefore, the
aftershock distribution reveals a fault geometry, as represented in Figures 4 (b) and (d). The results
from this study emphasize the significance of relocating the earthquake catalog before tectonic or
geological interpretation for various scales, as Enghdal et al. have updated the global earthquake
catalog [13]. In comparison, this study updates the earthquake catalog locally at Cartesian
coordinates.

Figure 3. seismicity of the BMKG catalog (a) and the seismicity of the relocated hypocenter (b).
Lines A-A' and B-B' are vertical sections that pass through the strike and dip
directions. The hypocenter distance (D) to the two lines is -50 km≤D≤50 km. Yellow
stars show on both maps indicating the epicenter of the mainshock. The topographic
map utilizes SRTM Non-Void Filled [14].

Figure 4. Vertical section of BMKG catalog (a) and after hypocenter relocation (b) along the
strike direction. For (c) and (d), same as (a) and (b) but along the dip direction. The
yellow star represents the mainshock.

4
The 4th Southeast Asian Conference on Geophysics (SEACG 2022) IOP Publishing
IOP Conf. Series: Earth and Environmental Science 1227 (2023) 012042 doi:10.1088/1755-1315/1227/1/012042

Figure 5. Reconstructed model of the FBT fault plane (red dotted line) and the complex fault
zone (curved red lines) based on hypocenter relocation results. The depth of the source
mechanism is generated from the updated hypocenter.

From the results of the hypocenter relocation and the source mechanism, we could explain the
characteristics of the 2022 Mw 5.9 Eastern Manggarai earthquake based on a vertical section
across the dip direction, as shown in Figure 5. Studies of several earthquakes with Mw 6.2-6.6 in
2002-2009 in the north of Sumbawa relating to the FBT occurred at depths between 15-20 km.
These earthquakes are caused by the presence of a barrier that separates the FBT fault plane so
that it forms short segments [15]. If all of these segments release energy simultaneously, it could
potentially generate an earthquake of Mw 7.7, as occurred in 1992 [1]. The seismicity map in
Figure 3 (b) shows that the aftershocks are divided into two clusters, namely clusters located in
the west and east. A narrow gap between two clusters extends to the dip direction, considered a
barrier between fault segments.
The aftershock distribution occurs at a depth of less than 17 km with a 20 km width at a
distance of 42-62 km (Figure 5). Multibeam data shows a complicated thrust fault zone in northern
Flores in the latitude range of 7.750-8.000S or a width of 18 km extending west-east between 1200-
121,500E [16]. In this study, the distribution of aftershocks occurred in the latitude range of 7.90 0-
8.200S, or slightly south, noting the down-dip part of these faults. The aftershock distribution
spans 20 km, corresponding to the 18 km wide fault zone within the complicated thrust fault zone.
The main shock on the faults may generate these aftershock zones in the splay faults through
static or dynamic stress transfer. However, assessing whether the main fault causes significant
aftershocks requires further detailed study.
The main fault plane of the FBT is reconstructed from one of the nodal planes, which has a
dip to the south (see Figure 5). The northern FBT reconstruction line (up-dip) corresponds to the
FBT line from the bathymetric data interpretation north of the Flores basin, as shown in Figure 5
[14]. The reconstruction results show that the aftershock distribution is not exactly on the FBT
fault plane. The mainshock's hypocenters were outside the aftershocks' distribution zone. This
phenomenon occurs due to a seismic gap between the mainshock and the aftershocks. This gap
usually indicates a rupture or asperity zone during a co-seismic event. It can explain why there
are no aftershocks in the gap zone because the area's stress has most likely been released.
However, studies of the Coulomb stress change and slip inversion need to be conducted to confirm
the provisional conjecture. All those studies will describe a comprehensive knowledge of the
characteristics of the earthquake. The aftershock zone may have occurred in the unstable zone or
is critically stressed; therefore, a slight stress increase could trigger a co-seismic slip on the splay
faults. A moderate earthquake (Mw < 6.0) which releases many aftershocks, has become the main
attention of an earthquake event. A moderate earthquake followed by many aftershocks also

5
The 4th Southeast Asian Conference on Geophysics (SEACG 2022) IOP Publishing
IOP Conf. Series: Earth and Environmental Science 1227 (2023) 012042 doi:10.1088/1755-1315/1227/1/012042

occurred in Plampang, Sumbawa, on June 13, 2020 [17,18]. The studies in this area indicate a
very brittle zone along Sumbawa to Flores island.

4. Conclusions
The characteristics study of an earthquake will be helpful for earthquake disaster mitigation,
especially for infrastructure development. Our study successfully reconstructed the fault plane of
the FBT and revealed the complex fault zone affected by co-seismic activity, which had never
been considered before in this area. This study also shows that the FBT system consists of some
fault segments. The seismic hazard study needed to be conducted after this step because it is
proven to minimize fatalities and injuries if a significant earthquake hits a particular area.

Acknowledgment
We are very grateful to the Center for Earthquake and Tsunami (PGT), Indonesian Agency for
Meteorology, Climatology, and Geophysics (BMKG) for providing the earthquake catalog for
this research.

References
[1] Ekström G, Nettles M and Dziewoński A M 2012 Physics of the Earth and Planetary
Interiors 200–201 1–9
[2] Harris R and Major J 2016 Geological Society, London, Special Publications, 441(1), 9–46
[3] Hamilton W B 1979 (US Govt. Print. Off.)
[4] Koulali A, Susilo S, McClusky S, Meilano I, Cummins P, Tregoning P, Lister G, Efendi J
and Syafi'i M A 2016 Geophysical Research Letters 43 1943–9
[5] Priyobudi P and Ramdhan M 2020 Jurnal Lingkungan dan Bencana Geologi 11 1–9
[6] PUSGEN 2017 (Kementerian Pekerjaan Umum dan Perumahan Rakyat)
[7] Waldhauser F and Ellsworth W L 2000 Bulletin of the Seismological Society of America 90
1353–68
[8] Waldhauser F 2001 US Geol. Surv. Openfile report 01–113
[9] Koulakov I, Bohm M, Asch G, Lühr B-G, Manzanares A, Brotopuspito K S, Fauzi P,
Purbawinata M A, Puspito N T, Ratdomopurbo A, Kopp H, Rabbel W and Shevkunova E
2007 Journal of Geophysical Research: Solid Earth 112
[10] Wadati K 1933 Geophys. Mag 7 101–11
[11] Billings S D 1994 Geophysical Journal International 118 680–92
[12] Efron B 1982 (Society for Industrial and Applied Mathematics)
[13] Engdahl E R, Di Giacomo D, Sakarya B, Gkarlaouni C G, Harris J and Storchak D A 2020
Earth and Space Science 7 e2019EA000897
[14] Earth Resources Observation And Science (EROS) Center 2017 Shuttle Radar Topography
Mission (SRTM) Non-Void Filled
[15] Sianipar D, Huang B-S, Ma K-F, Hsieh M-C, Chen P-F and Daryono D 2022 Journal of
Asian Earth Sciences 229 105167
[16] Silver E A, Breen N A, Prasetyo H and Hussong D M 1986 Journal of Geophysical
Research: Solid Earth 91 3489–500
[17] Ramdhan M, Priyobudi, Mursityanto A, Palgunadi K H, and Daryono 2021 IOP Conf.
Ser.: Earth Environ. Sci. 873 012070
[18] Priyobudi P and Ramdhan M 2021 EKSPLORIUM 42 111–8

You might also like