You are on page 1of 18

Special Section: Caribbean Tectonics, Seismicity and Earthquake Hazards

The 2022 Seismic Hazard Model for Costa Rica


Diego A. Hidalgo-Leiva*1 , Lepolt Linkimer2 , Ivonne G. Arroyo2 , Mario Arroyo-Solórzano2,3 , Rosey Piedra4 ,
Alvaro Climent4 , Víctor Schmidt Díaz1 , Luis Carlos Esquivel1 , Guillermo E. Alvarado5 , Rolando Castillo6 ,
Marco E. Carranza-Morales7 , Laura Cerdas-Guntanis8 , Jimena Escalante-Meza7 , Sergio Lobo6 ,
María José Rodríguez6 , and Wilfredo Rojas†2

ABSTRACT
Costa Rica is located at the boundary of four tectonic plates where the regularity of destruc-
tive earthquakes highlights the necessity of seismic hazard estimations. This study contains
the most recent Probabilistic Seismic Hazard Assessment (PSHA) for Costa Rica, calculated
with the largest and the most updated earthquake database from both—the Earthquake
Engineering Laboratory and the National Seismological Network of the University of
Costa Rica. For the PSHA, we updated the seismicity parameters for the upper plate, sub-
duction interplate, and intraslab tectonic domains, characterized the upper-plate zones by
percentages of fault types, and used weighted ground-motion models for each of the tec-
tonic domains. The resulted maps of peak ground acceleration (PGA) at return periods of
475 yr (PGA-475) and 2475 yr, as well as the spectral accelerations, show geographic trends
that allow for the division of the country in four seismic hazard levels: (1) extremely high for
the Nicoya, Osa, and Burica peninsulas, situated directly above the subduction interplate,
where the PGA-475 could be 0.55–1.20g; (2) very high for most of the Guanacaste
Province, where the PGA-475 may be 0.55–0.70g; (3) high for most of the country ( ∼41%)
with PGA-475 values of 0.40–0.55g, including Central Costa Rica and the capital city of San
Jose; and (4) moderate for the Talamanca Cordillera and Northern Costa Rica, with PGA-475
up to 0.40g. These ground-motion values are 0.1–0.6g higher than the previous PSHA for the
Pacific peninsulas, Guanacaste, and the southeastern Caribbean. Further, hazard curves, uni-
form hazard spectra, and a hazard disaggregation indicate that the seismic hazard is lower
but more complex in San Jose than in Liberia—the largest city in Guanacaste.

seismic sources are the boundaries of the Panama microplate


KEY POINTS with the Caribbean and Nazca plates (Adamek et al., 1988;
• We conduct a comprehensive Probabilistic Seismic Hazard
Assessment for Costa Rica.
• Four distinct ground-motion levels can be correlated with 1. Earthquake Engineering Laboratory, University of Costa Rica, San José, Costa Rica,
tectonic boundaries. https://orcid.org/0000-0003-3662-4969 (DAH-L); https://orcid.org/0000-0002-
• Ground-motion levels higher than previously estimated 2911-7232 (VSD); https://orcid.org/0000-0003-4208-6482 (LCE); 2. Central
American School of Geology and National Seismological Network (RSN-UCR),
are obtained for the Pacific peninsulas and Guanacaste. University of Costa Rica, San José, Costa Rica, https://orcid.org/0000-0002-1008-
846X (LL); https://orcid.org/0000-0003-3232-9719 (IGA); https://orcid.org/
Supplemental Material 0000-0002-1653-2680 (MA-S); 3. Department of Topography, Geodesy and
Cartography, Polytechnical University of Madrid, Madrid, Spain; 4. National
Seismological Network (RSN-ICE), Costa Rican Institute of Electricity, San José, Costa
Rica, https://orcid.org/0000-0002-4184-4037 (RP); https://orcid.org/0000-0002-
INTRODUCTION 4478-0382 (AC); 5. Department of Seismology and Vulcanology, Costa Rican Institute
of Electricity, San José, Costa Rica, https://orcid.org/0000-0001-7082-4994 (GEA);
Costa Rica is a seismically active country in Central America, 6. School of Civil Engineering and the National Laboratory of Materials and Structural
with a history of at least 68 damaging earthquakes since 1821, Models (LANAMME), https://orcid.org/0000-0002-1456-9770 (RC); https://
orcid.org/0000-0001-7763-9767 (SL); https://orcid.org/0000-0003-2270-0107
including the recent large earthquakes in 1991 (moment mag- RodríguezMaría José; 7. School of Psychology, University of Costa Rica, San José,
nitude) M w 7.7 (Fig. 1), 1992 M w 7.7, and 2012 M w 7.6 and the Costa Rica, https://orcid.org/0000-0001-6437-0586 (MEC-M); https://orcid.org/
1910 M w 6.4 Cartago earthquake, considered the worst disaster 0000-0002-3693-4249 (JE-M); 8. School of Social Work, University of Costa Rica, San
José, Costa Rica, https://orcid.org/0000-0003-4894-9111 (LC-G)
in the Costa Rican history (e.g., Alonso-Henar et al., 2013; *Correspondence author: diego.hidalgo@ucr.ac.cr
Linkimer and Alvarado, 2014). This high seismicity is mainly †Deceased
originated by the subduction of the Cocos plate underneath the Cite this article as Hidalgo-Leiva, D. A., L. Linkimer, I. G. Arroyo, M. Arroyo-
Caribbean plate and the Panama microplate, and by the inter- Solórzano, R. Piedra, A. Climent, V. Schmidt Díaz, L. C. Esquivel, G. E. Alvarado,
R. Castillo, et al. (2022). The 2022 Seismic Hazard Model for Costa Rica, Bull. Seismol.
action between both the overriding plates in Central Costa Rica Soc. Am. XX, 1–18, doi: 10.1785/0120220119
(Fig. 1; Marshall et al., 2000; Montero, 2001). Other relevant © Seismological Society of America

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 1

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
including the Resis II Project in
2008 (Climent et al., 2008;
Benito et al., 2012), which
described the latest seismo-
genic zonation and ground-
motion models (GMM)
applied. There are no recent
PSHA for Costa Rica available,
because in the latest estima-
tions made by the Costa
Rican Seismic Code (CFIA,
2016), the details conducting
to their results (i.e., input
parameters, the GMM, the ini-
tial seismic zonation) were not
published.
The Earthquake Engineering
Laboratory (LIS) and the
National Seismological Network
(RSN) are two centers at the
University of Costa Rica
(UCR) dedicated to monitor
strong motion and seismicity
since 1983 and 1973, respec-
tively. In 2009, a national law
endorsed both centers with
financial support for instrumen-
tation, enabling them to mod-
ernize their networks and
expand countrywide (e.g.,
Linkimer et al. 2018; Moya-
Fernández et al., 2020). The vast
Figure 1. Tectonic framework of Costa Rica. The region contained within the dotted lines represents the Central Costa
Rica Deformed Belt (CCRDB). The dashed line represents the simplified northeast boundary of the Central America forearc amount of new earthquake data
block (CAFAB) along the volcanic arc faults (VAFs). F, Fisher Seamount; NPDB, North Panama deformed belt; PFZ, recorded by both, the LIS and
Panama fracture zone; QP, Quepos plateau; SRB, smooth–rough boundary; and SPDB, south Panama deformed belt. The the RSN during the last decade,
numbered stars represent the largest most recent earthquakes mentioned in the text: (1) 1910 Mw 6.4 Cartago, (2) 1934 and the necessity for an up-to-
Mw 7.6 Armuelles, (3) 1950 Mw 7.7 Nicoya, (4) 1973 Mw 6.5 Tilaran, (5) 1983 Mw 7.4 Golfito, (6) 1990 Mw 7.1 date public PSHA for Costa
Cobano, (7) 1991 Mw 7.7 Limon, (8) 1992 Mw 7.7 Nicaragua, (9) 1999 Mw 6.9 Quepos, and (10) 2012 Mw 7.6 Nicoya.
The gray region represents the Talamanca Cordillera, the orange triangles the active volcanoes, and the black squares the
Rica, motivated a joint project
cities of San Jose and Liberia. The Osa, Nicoya, and Burica (Bur.) peninsulas are labeled. Inset map in the top right corner involving researchers from the
shows the location of Costa Rica. The color version of this figure is available only in the electronic edition. UCR, including the LIS, the
RSN, the National Laboratory
of Materials and Structural
Models (LANAMME), and the
Silver et al., 1990; Kolarsky and Mann, 1995; Westbrook et al., departments of Psychology and Social Work, as well as research-
1995) and the Panama fracture zone (PFZ) (Fig. 1). ers from the Costa Rican Institute of Electricity (ICE). The goal of
Given this very active and complex tectonic setting, seismic this project was to calculate a new seismic hazard model for the
hazard studies are of great importance for the development of country and to contextualize earthquake experiences of rural
the region. The first Probabilistic Seismic Hazard Assessment communities to encourage policies for risk mitigation.
(PSHA) for Costa Rica was estimated in 1977 by the John A. In this article, we present the results for the seismic hazard
Blume Earthquake Engineering Center (Mortgat et al., 1977). model component of the aforementioned joint project. We fol-
Later, different PSHAs were conducted for a larger region of lowed a classical approach in which we selected the seismic
Central America (e.g., Rojas et al., 1993; Climent et al., 2008; sources, built a logic tree with the input parameters and their
Benito et al., 2012; Global Earthquake Model [GEM], 2018a), probabilities, and calculated the hazard for the entire logic tree.

2 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
We used the state-of-the-art open source application Seismic zones
OpenQuake 3.9 (GEM, 2020) and integrated the longest RSN Costa Rica is located on the western margins of the Caribbean
catalog ever compiled (Arroyo-Solórzano and Linkimer, 2021), plate and the Panama microplate (Fig. 1). Beneath both of
an updated LIS strong-motion database (Moya-Fernández et al., them, the Cocos plate subducts at an average velocity of
2020), and a reviewed soil classification for the LIS stations ∼83 mm/yr (DeMets et al., 2010) along the Middle America
(Pinzón et al., 2021). Specific GMMs were used for the three trench (MAT), including several bathymetric features such
main tectonic domains: upper plate, subduction interplate, as the Fisher Seamount, the Quepos plateau, and the Cocos
and subduction intraslab. In the logic tree, 13 spectral points ridge (von Huene et al., 2000; Fig. 1). The Cocos plate subduc-
(i.e., intensity measures) were considered, and specific weights tion has originated a volcanic arc presently located ∼150 km to
for each GMM and intensity measures were assigned. We the northeast of the MAT (Fig. 1).
present the results as seismic hazard curves, spectra, and ground To the southeast, at the Panama triple junction, the MAT
motion maps, and grant special attention to two of the largest ends against the PFZ—a right-lateral strike-slip fault that marks
cities of the country (San Jose and Liberia) in which the disag- the boundary between the Cocos and the Nazca plates (Fig. 1).
gregation of the seismic hazard was performed. Farther east, the limit between the Panama and the Nazca plates
is the southern Panama deformed belt (SPDB), which has been
DATA AND METHODS identified as an oblique subduction zone in western Panama
Earthquake catalog (e.g., Kolarsky and Mann, 1995; Westbrook et al., 1995).
The earthquake catalog and the magnitude–frequency rela- The northern and western boundaries between the
tionship used in this study were taken from Arroyo- Caribbean plate and the Panama microplate have been described
Solórzano and Linkimer (2021). Those authors obtained a joint as broad zones (∼100 km) of crustal deformation. The Central
earthquake catalog by merging the RSN catalog from October Costa Rica Deformed Belt (CCRDB; Fig. 1; Marshall et al., 2000;
1975 to August 2020 with the Resis II project (Climent et al., Montero, 2001) comprises mainly strike-slip faults and the
2008), which was based on Rojas et al. (1993) and covered his- North Panama deformed belt (NPDB)—a sequence of subma-
torical earthquakes from 1522 to 1900. Data from seismologi- rine thrusts and folds (Adamek et al., 1988; Silver et al., 1990;
cal agencies from Nicaragua and Panama were added to Fan et al., 1993). In addition, a succession of strike-slip faults has
improve the completeness of the surrounding regions of been recognized along the volcanic arc in Nicaragua and
Costa Rica. The geographic area chosen to build the joint Northern Costa Rica (volcanic arc fault [VAF], Fig. 1), which
earthquake catalog covers the coordinates 6.0°–12.5° N and allows for the northwestward movement of the Central
81.0°–87.5° W and encompasses the country of Costa Rica America forearc block (CAFAB, e.g., LaFemina et al., 2009;
and portions of Nicaragua and Panama, which were consid- Kobayashi et al., 2014).
ered relevant for the seismic hazard assessment of Costa This complicated tectonic setting has been divided into seis-
Rica. The resulting joint catalog contains 138,424 earthquakes mic zones for the purposes of seismic hazard calculations. This
from 1522 to 2020 with M w from 0.3 to 7.7, and corresponds to work uses the divisions defined by Alvarado et al. (2017), spe-
the largest and more complete catalog ever used for calculating cifically the 28 zones encompassing the territory of Costa Rica
magnitude–frequency relationships in Costa Rica. The details and its vicinity (Fig. 2), which were grouped into three tectonic
on the earthquake selection, merging of the catalogs, and mag- domains: upper plate (18 zones), interplate (5 zones), and
nitude homogenization can be found in Arroyo-Solórzano and intraslab (5 zones). The original nomenclature has been pre-
Linkimer (2021). served, in which each abbreviation starts with the letter C, N, or
Using the classical Stepp (1972) method and the maximum P, referring to the country Costa Rica, Nicaragua, and Panama,
curvature method (e.g., Wiemer and Wyss, 2000), Arroyo- respectively. This first letter is followed by a number for the
Solórzano and Linkimer (2021) determined that the joint upper-plate zones, by “si” and a number for the interplate
earthquake catalog has a magnitude of completeness (M c ) zones, and by “sp” and a number for the intraslab zones.
of 7.0 since 1793 (i.e., all earthquakes with a magnitude 7 The upper-plate zones in the hazard zonation of Alvarado
of greater are included in the catalog since that year). This et al. (2017) contains the regions where shallow seismicity
is possible for a small country like Costa Rica in which mag- (<20 km) is originated by the internal deformation of the over-
nitude 7 earthquakes are felt countrywide, allowing the wit- riding Caribbean and Panama plates and their interaction with
nesses to record their effects in historical documents before other tectonic blocks (Fig. 1). These zones (Fig. 2a) are distrib-
the installation of the first seismograph in San Jose in 1888 uted along the Pacific forearc (N4, C1-N2, C2, C3, and C4), the
(Morales, 1986). The M c decreases through time, as the num- magmatic arc (N6-7, N8, C5, C6, C7, and P5), and the
ber of earthquake detections increases in the region with more Nicaragua Lake and the Caribbean backarc (N9-10, N12,
instrumentation. The most complete portion of the catalog N13-14, C9, and C10). Additional upper-plate zones in reality
has an M c of 3.0 during 1995–2020 (Arroyo-Solórzano and encompass plate boundaries (Fig. 1), such as the PFZ (P1), the
Linkimer, 2021). NPDB (C8-P10), and the SPDB (P2).

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 3

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
The subduction interplate zones are divided into five seg- Figure 2. Seismic zonation for the study area (a) PFZ and the upper-plate
ments that were delineated considering the historical seismicity domain, including arc, backarc, forearc, and the NPDB. (b) Interplate
(depth <40 km) and the bathymetric features in the oceanic domain. (c) Intraslab domain. The color version of this figure is available
only in the electronic edition.
floor of the incoming Cocos plate (Figs. 1 and 2b). These
include the Pacific margin of Nicaragua (zone Nsi16) and
the Nicoya peninsula (Csi11), where a smooth bathymetry atop

4 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
a ∼5–7 km thick oceanic crust (Sallarès et al., 2001) is presently Based on that information, it was considered that 90% of
subducting. The zone off the coast of Central Costa Rica the crustal earthquakes were originated by strike-slip faults
(Csi12) shows an incoming seafloor 40% covered with sea- and 10% by reverse faults. All the strike-slip faults were con-
mounts (von Huene et al., 2000; Barckhausen et al., 2001) sidered vertical, and 60% were classified as right-lateral with a
and an oceanic crust thickness of ∼10 km (Walther, 2003). strike of N50°W and the other 40% as left lateral with a strike
The interplate zone beneath the Osa peninsula (Csi13) corre- of N40°E. Within the reverse faults, 60% were oriented N50°W
sponds to the subduction of the Cocos Ridge (Fig. 1) with a with a dip of 45° to the northeast, and the other 40% were ori-
crust thickness exceeding ∼20 km (Sallarès et al., 2003; ented S50°E and inclined 45° to the southwest. This basic con-
Walther, 2003); and the interplate zone in western Panama figuration scheme was modified for upper-plate zones in which
(Psi9) matches the oblique subduction of the Nazca plate under active faults with different geometries were also documented;
the Panama microplate (Kolarsky and Mann, 1995). for example, in the case of the vertical, right-lateral strike-slip
The subduction intraslab zones describe the seismicity faults oriented N10°E near the political border between Costa
originated within the Cocos slab at intermediate depths from Rica and Panama (zones C4 and C7). The use of these faulting
40 to ∼250 km (Fig. 2c). These include eastern Nicaragua type percentages is a novel approach for a PSHA in Costa Rica.
(Nsp17), where intermediate seismicity dips 65° to the north- The most likely rupture plane parameters for the interplate
east down to ∼250 km (Kyriakopoulos et al., 2015), zones were defined based on the shallower part (<40 km) of
northwestern Costa Rica (Csp14) where the intraslab seismic- the subduction zone revealed by local tomographic studies
ity reaches depths of ∼200 km near the Costa Rica–Nicaragua (e.g., DeShon et al., 2006; Arroyo et al., 2009; Arroyo and
border, Central Costa Rica (Csp15), where the maximum Linkimer, 2021), the slab geometry from Lücke and Arroyo
depths of seismicity shallow to ∼130 km (Lücke and (2015), and the moment tensor solution for the large earthquakes
Arroyo, 2015), and southern Costa Rica (Csp16) and western (M w > 5:5) available in international catalogs and pub-
Panama (Psp11) where the deepest seismicity occurs at lished works.
∼70 km (Lücke and Arroyo, 2015). The most likely rupture plane parameters for the intraslab
zones were entirely based on moment tensors and focal mech-
Seismicity parameters anisms available from the RSN, and international catalogs and
In this study, the upper-plate zones have been modeled as area published works. The quantity of solutions available vary from
sources, and the interplate and intraslab zones as complex dozens in Nicaragua to none for western Panama, in which the
faults. To describe them, for each seismic zone, the following same values as for southern Costa Rica were assumed.
parameters were required in the OpenQuake software: the The magnitude–scaling relationship was determined using
upper and lower seismogenic depths, the most likely hypocen- Wells and Coppersmith (1994). Further, assuming only rectan-
ter, the most likely rupture planes, the magnitude–scaling rela- gular ruptures, an aspect ratio of 1.5 was defined for all zones.
tionship, the rupture aspect ratio, the magnitude–frequency The minimum magnitude for hazard assessment was fixed at a
distribution, the maximum inferred magnitude, and the mini- standard value of M w 4.5.
mum magnitude for hazard calculation (Table S1, available in The a- and b-values of the magnitude–frequency distribu-
the supplemental material to this article). tion for each seismic zone were taken from Arroyo-Solórzano
The upper and lower seismogenic depths, and the most and Linkimer (2021). In that study, a- and b-values were esti-
likely hypocenter for each seismic zone were explored using mated using the classical maximum-likelihood technique of
the earthquakes extracted from the RSN catalog, which were Aki (1965) modified by Weichert (1980), after applying the
recorded between the years of 2011 and 2020, and located with declustering method of Gardner and Knopoff (1974) to remove
the minimum of 12 stations with a root mean square ≥0.5 s. dependent (non-Poissonian) earthquakes (i.e., swarms, after-
The 2011–2020 time period was selected because it has the low- shocks, and foreshocks). The b-value for individual upper-
est location uncertainties and also the best station coverage in plate zones vary from 0.67 (C2 and N4) to 1.12 (C6), with
the RSN history. The minimum depth of 1 km was assumed as an average of 0.90. For intraslab zones, the b-value varies
the upper seismogenic depth for all the upper-plate zones. On between 0.91 (Nsp17) and 1.42 (Csp16), with an average of
the other hand, the lower limit was defined by analyzing seis- 1.14, and for interplate zones, it varies within the smallest
micity cross sections for each zone individually. For the inter- range compared with the upper-plate and intraslab zones: from
plate and intraslab zones, in addition to seismicity profiles, the 0.69 in zone Csi11 to 1.10 in Csi12, with an average of 0.85.
parameters were determined by using the Cocos slab geometry The maximum magnitude (M max or seismic potential) for
proposed by Lücke and Arroyo (2015). each seismic zone was also taken from Arroyo-Solórzano and
The strike, dip, and rake for the most likely rupture plane Linkimer (2021), who calculated it using the cumulative seismic
within each of the upper-plate zones were defined primarily on moment technique of Makropoulos and Burton (1983). The
the compilation of Denyer et al. (2009), and complemented M max results vary from 5.5 to 8.0, and the highest magnitudes
with recent fault maps and earthquake focal mechanisms. estimated within the seismic zones were 8.0 for the NPDB

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 5

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
TABLE 1
Selected Ground-Motion Models (GMM) for the 2022 Seismic Hazard Model for Costa Rica

Seismic Region GMM Region Dependent Variable Component Distance (km) Mw

Upper plate CZ15 (Cauzzi et al., 2015) Worldwide Average horizontal 0–150 4.5–8.0
KA06S (Kanno et al., 2006) Japan Peak square root of the sum of 20–400 5.5–8.0
squares of horizontals
BO14 (Boore et al., 2014) Worldwide Average horizontal 0–400 3.0–7.9
Interplate KA06S (Kanno et al., 2006) Japan Peak square root of sum of squares 20–400 5.5–8.0
of horizontals
ZH06SI (Zhao et al., 2006) Japan Geometric mean 10–300 5.0–8.2
MO17SI (Montalva et al., 2017) Chile Average horizontal 10–300 5.0–9.0
Intraslab AB16SS (Abrahamson et al., 2016) Worldwide Average horizontal 10–300 5.0–8.4
MO17SS (Montalva et al., 2017) Chile Average horizontal 10–300 5.0–8.0
KA06D (Kanno et al., 2006) Japan Peak square root of sum of squares 20–400 5.5–8.0
of horizontals
GA05SS (García et al., 2005) Mexico Quadratic mean 4–400 5.2–7.4
LL08SS (Lin and Lee, 2008) Worldwide Average horizontal 15–630 5.3–8.1

(zones C8-P10, Fig. 2a), 8.0 for the Nicoya peninsula (Csi11, GMM
Fig. 2b), and 7.9 for the southern Nicaragua (Nsi16, Fig. 2b). The choice of the GMM and their weights in the logic tree was
divided into three stages: the review of the GMM available, the
Strong-motion data GMM contrast against the strong-motion dataset, and the
The Costa Rican strong-motion database (Moya-Fernández GMM classification and selection. For the first stage, we
et al., 2020) consists of 2471 triaxial digital records from reviewed the available GMM by applying the exclusion criteria
155 earthquakes recorded from 1998 until 2019 (see the of Cotton et al. (2006) to the list published by Douglas (2021).
Data and Resources section). From this database, 2353 records The 19 GMM initially preselected are included in the
from 139 earthquakes were selected for the current study. OpenQuake Engine 3.9 (GEM, 2020).
These data correspond exclusively to digital accelerometers, For the second stage, the GMM contrast against the strong-
as the older analog records were excluded from the analysis. motion dataset was performed using a strong-motion toolkit
Strong motions from earthquakes lacking source solutions (GEM, 2018b). In the process, we computed the mean total
were also rejected. The magnitudes, locations, and fault-plane residual (mean Res ), the standard deviation of the total residual
solutions were provided mainly by the RSN, but source solu- (σ Res ) obtained from fitting a normal probability function to
tions from the National Earthquake Information Center cata- the residual distribution, and the mean likelihood model (LH,
log of the U.S. Geological Survey (USGS) and the recent Scherbaum et al., 2004; Fig. 4), as well as the log-likelihood
publications were also included (see Table S2). values (LLH, Scherbaum et al., 2009). In most of the GMM
The strong-motion database contains records for hypocen- cases, the mean total residuals are negative for periods larger
tral distances from 7 to 288 km for the upper-plate zones than 0.5 s (Fig. 4), which indicates an overestimation of the
(Fig. 3a), from 10 to 351 km for the interplate earthquakes predicted ground-motion levels. This implies that our results
(Fig. 3b), and from 10 to 379 km for the intraslab zones are conservative, that is, the real ground-motion levels might
(Fig. 3c). The maximum horizontal peak ground acceleration be lower than our estimations.
(PGAh ) varies from 2.3 to 231:0 cm=s2 for the upper-plate Finally, in the GMM classification and selection, we used cri-
zones (Fig. 3d), from 2.0 to 1580 cm=s2 for the interplate zones teria based on Scherbaum et al. (2004) from A for the best qual-
(Fig. 3e), and from 2.0 to 386 cm=s2 for the intraslab zones ity to D for the worst. For the GMM equally classified, the LLH
(Fig. 3f). The database shows a typical distribution, with a scar- was used as a tiebreaker. From the 19 GMM initially considered,
city of low-magnitude events at large distances due to seismic 11 were selected, and some of them were used in more than one
attenuation. The largest PGAh of 1580 cm=s2 (Schmidt-Díaz tectonic domain (Table 1). None of these 11 models (Table 1)
et al., 2014) corresponds to the 2012 M w 7.6 earthquake were specifically developed for Costa Rica or Central America,
recorded in the Nicoya peninsula (Fig. 1). and, even with the use of new earthquake data, the evaluated
The strong-motion database was complemented with a GMM present high dispersion, particularly for higher structural
recent review of the soil classification of the LIS strong-motion periods (Fig. 4). This implies the need to create or modify the
stations (Pinzón et al., 2021), which included the average available models (e.g., Campbell, 2003; Cotton et al., 2006; Kotha
shear-wave velocity up to 30 m depth (V S30 ) and the horizon- et al., 2017). A similar dispersion in the selected GMM was
tal-to-vertical spectral ratio. noticed in the previous studies for Costa Rica (Laporte et al.,

6 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
1994; Benito et al., 2012) that used the GMM from Climent et al. Figure 3. The Costa Rican strong ground motion database (Moya-Fernández
(1994) and Schmidt-Díaz et al. (1997). These GMM were also et al., 2020). Magnitude (Mw ) versus hypocentral distance for (a) upper-
tested in this study with results of low quality (D). plate, (b) interplate, and (c) intraslab domains. Magnitude (Mw ) versus the
maximum horizontal component of the peak ground acceleration (PGAh ) for
(d) upper-plate, (e) interplate, and (f) intraslab domains.
Hazard calculation
As part of the logic tree, we introduced for the first time in a
PSHA of Costa Rica the use of variable weights assigned to the
GMM depending on the 13 intensity measures that were The hazard model was calculated for the continental
defined for the hazard calculation. These levels consist of territory of Costa Rica. A grid with a spacing of 0.02° in both
the PGA and the spectral acceleration (SA) with a damping latitude and longitude for a total of 10,541 points was used to
of 5% for a different period of oscillation (Tn) from 0.1 to calculate the seismic hazard. The ruptures had a mesh spacing
3 s (Table 2). The assigned weights were used in each permu- of 5 km. A constant V S30 of 760 m/s was used, assuming a rock
tation in the logic tree from the maximum of 45 possible paths or rigid soil condition following the most recent seismic code
to the minimum of 18 for the spectral points in which the con- (CFIA, 2016).
sidered weight was null. Hazard curves were also calculated for The truncation level in this work was defined at six standard
the 13 intensity measures, including PGA and SA from 0.1 to deviations. We studied return periods (Trs) of 475 and 2475 yr,
3.0 s for a 5% damping ratio. which corresponds to the minimum annual frequency of

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 7

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
exceedance of 4:04 × 10−4 . Because the truncation level impacts Figure 4. Mean total residual (mean Res ), standard deviation of the total residual
hazard estimations with an annual frequency of exceedance (σRes ), and median likelihood (LH) versus the natural period of oscillation (Tn) for
smaller than 10−6 (Bommer, 2004), no major differences in the selected Ground-Motion Model (see Table 1 for abbreviations) shown as
colored lines in the tectonic domains: (a) upper-plate, (b) interplate, and
our results are expected if a lower truncation level was selected.
(c) intraslab. The color version of this figure is available only in the electronic edition.

RESULTS
Ground-motion maps
We present the classical ground-motion maps for engineering the Osa and Burica peninsulas and the inner part of
purposes (Fig. 5), including the PGA for Tr of 475 yr (PGA- Guanacaste, where the city of Liberia is located. The intermedi-
475) and 2475 yr (PGA-2475), as well as the SA for Tn of 0.2 ate levels correspond to most of the countries, including
and 1 s and for Tr of 475 and 2475 yr, abbreviated in this study Central Costa Rica, where the capital city of San Jose is situ-
as: SA(0.2)-475, SA(1.0)-475, SA(0.2)-2475, and SA(1.0)-2475. ated, and regions in the Southeast Pacific and the Caribbean.
All these maps display similar trends: The highest ground- The lowest ground-motion levels are present in the Talamanca
motion levels occur in the Nicoya peninsula, followed by both Cordillera and Northern Costa Rica (Fig. 5).

8 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
TABLE 2
Weights Assigned to Each Ground-Motion Model (GMM, See Table 1 for Abbreviations) and the Intensity Measures in the Logic
Tree Assessment

Intensity Measures

Seismic SA SA SA SA SA SA SA SA SA SA SA SA
Domains GMM PGA (0.1) (0.2) (0.3) (0.4) (0.5) (0.6) (0.7) (0.8) (0.9) (1.0) (2.0) (3.0)

Upper-plate CZ15 0.4 0.4 0.4 0.4 0.45 0.45 0.45 0.45 0.5 0.5 0.5 0.6 0.6
KA06S 0.4 0.4 0.4 0.4 0.45 0.45 0.45 0.45 0.3 0.3 0.3 0.2 0.2
BO14 0.2 0.2 0.2 0.2 0.1 0.1 0.1 0.1 0.2 0.2 0.2 0.2 0.2
Interplate KA06S 0.5 0.5 0.35 0.35 0.3 0.3 0.3 0.3 0.6 0.6 0.6 0.6 0.6
ZH06SI 0.5 0.5 0.55 0.55 0.4 0.3 0.3 0.3 0.15 0.15 0.15 0.3 0.3
MO17SI 0.0 0.0 0.1 0.1 0.3 0.4 0.4 0.4 0.25 0.25 0.25 0.1 0.1
Intraslab AB15SS 0.4 0.4 0.4 0.4 0.3 0.3 0.3 0.3 0.1 0.1 0.1 0.2 0.2
MO17SS 0.0 0.0 0.0 0.0 0.3 0.3 0.3 0.3 0.6 0.6 0.6 0.6 0.6
KA06D 0.2 0.2 0.2 0.2 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.0 0.0
GA05SS 0.4 0.4 0.0 0.0 0.0 0.0 0.0 0.0 0.1 0.1 0.1 0.1 0.1
LL08SS 0.0 0.0 0.4 0.4 0.3 0.3 0.3 0.3 0.1 0.1 0.1 0.1 0.1

PGA, peak ground acceleration; and SA, spectral acceleration.

Our PGA-475 results for a 10% of probability of exceedance Costa Rica, from 0.60 to 1.60g. The rest of the country has values
in 50 yr (Fig. 5a) correspond to values of 0.70–1.20g for the of 1.60–2.00g.
Nicoya and Burica peninsulas and the southeastern part of The trends of the estimated SA(1.0) are analogous to those
the Osa peninsula. PGA-475 levels of 0.55–0.70g are expected calculated for a period of 0.2 s (Fig. 5e,f). Within a Tr of 475 yr
in inner Guanacaste, which includes the city of Liberia, and (Fig. 5e), the Nicoya peninsula displays ground-motion values
parts of the Osa peninsula. For most (48%) of the country of 0.70–1.20g, and the inner Guanacaste and the Southern
we observe PGA-475 values of 0.40–0.55g, including the capital Peninsulas displays ground-motion values of 0.40–0.70g.
San Jose in central Costa Rica. Finally, PGA-475 values of less Acceleration values of 0.25–0.40g are estimated for most of
than 0.40g are calculated for Talamanca and Northern the country, and the lowest values for Northern Costa Rica
Costa Rica. and Talamanca are inferior to 0.25g.
The PGA-2475 maps for a 2% of probability of exceedance For a Tr of 2475 yr, the highest SA(1.0) reach values of
in 50 yr (Fig. 5b) display a similar pattern. The highest PGA- 1.60–3.00g in the Nicoya peninsula (Fig. 5f). The Southern
2475 values of 1.20–1.80g are predicted for the Pacific penin- Peninsulas and inner Guanacaste display accelerations of
sulas. The PGA-2475 levels correspond to 1.00–1.20g for inner 0.80–1.60g, Talamanca and Northern Costa Rica of less than
Guanacaste and some areas of southeastern Costa Rica. Middle 0.20g, and the rest of the country, between 0.50 and 0.80g.
values of 0.75–1.00g are found in central and southeastern Estimations of the dynamic amplification factor (i.e., the
Costa Rica, both in the Pacific and the Caribbean sides. ratio between SA and PGA) display significant variations over
Acceleration values lower than 0.75g occur in Talamanca the country. For an SA(0.2 s)-475, it varies from 1.98 to 2.75,
and Northern Costa Rica. with the greatest values in the Pacific peninsulas and the lower
The results for the SA(0.2)-475 (Fig. 5c) distribute in a pat- ones in Northern Costa Rica. The same pattern is obtained for
tern comparable to PGA. In this case, accelerations of 2.10–3.00g Tr of 2475 yr, with numbers ranging from 1.98 to 3.03. Our
are expected in the Nicoya peninsula, whereas in the Southern results contrast with the current practice in the country, in
Peninsulas and Guanacaste the values are 1.30–2.10g, including which a unique dynamic amplification factor of 2.5 has been
1.70g for Liberia. Levels between 0.90 and 1.30g are anticipated used in the seismic code (CFIA, 2016).
for most of the country, including the capital city, with the Likewise, for SA(1.0s)-475 the dynamic amplification factor
exception of Talamanca and Northern Costa Rica, where the varies from the lowest values of 0.56–0.60 for the Southern
estimated accelerations are smaller than 0.90g. Caribbean up to the highest of 0.98–1.09 in the Nicoya pen-
The calculated SA(0.2)-2475 yr (Fig. 5d) are the highest val- insula. Considering a Tr of 2475 yr, the amplification factor
ues obtained in this work. The uppermost values from 3.00 to ranges from 0.63 to 0.73 in Talamanca and up to 1.12–1.23
5.50g are observed in the Nicoya peninsula, the Southern in the Nicoya peninsula. Again, the seismic code uses a con-
Peninsulas, and the western part of inner Guanacaste, encom- stant amplification factor of 1.0, expecting the same accelera-
passing Liberia. Zones of inner Guanacaste and the southeastern tion as the PGA. Therefore, the use of constant amplification
Pacific display accelerations of 2.00–3.00g, whereas the lowest factors fails to reflect the expected variation of acceleration lev-
values are calculated once more for Talamanca and Northern els between different sites and Trs in Costa Rica.

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 9

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
Figure 5. Results of the 2022 Costa Rican Seismic Hazard Model for selected represent the cities of San Jose (S) and Liberia (L). (a) PGA for a Tr of 475 yr,
intensity measures of the peak ground acceleration (PGA) and the spectral (b) PGA for a Tr of 2475 yr, (c) SA(0.2 s) for a Tr of 475 yr, (d) SA(0.2 s) for a
acceleration (SA) in fraction of g for a specified return period (Tr). In each Tr of 2475 yr, (e) SA(1.0 s) for a Tr of 475 yr, and (f) SA(1.0 s) for a Tr of
map, the intensity levels are specified by colors, and specific contours are 2475 yr. The Tr of 475 and 2475 yr correspond to a 10% and 2%
labelled. Main tectonic boundaries as in Figure 1. The black squares exceedance probability in 50 yr, respectively.

10 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
Seismic hazard for the cities of San Jose and Liberia is 28% for San Jose (Fig. 7b). Because the trends in the disag-
In this article, we select two key centers in terms of population gregation for SA(0.2) are similar to PGA, we do not display
and economic significance to present hazard curves for three those results.
intensity measures (Fig. 6a): the capital San Jose, in central
Costa Rica, and Liberia—the biggest city in northwestern DISCUSSION
Costa Rica. The hazard curves show that the maximum ampli- The ground motions from the 2022 seismic hazard model for
tudes occur at the spectral points of 0.2 s, whereas the mini- Costa Rica display, as expected, a strong connection with the
mum values are achieved at 1.0 s. The uniform hazard spectra tectonic boundaries. The previous studies have shown particu-
computed from the hazard curves illustrate the variations of larly the influence of the interplate zone of the subduction of
the intensity measures for different Tr (Fig. 6b). It can be the Cocos plate, resulting in a higher seismic hazard along the
clearly noticed from these curves and the strong-motion maps Pacific coast of the country (e.g., Mortgat et al., 1977; Benito
(Fig. 5) that San Jose has lower levels of seismic hazard (e.g., et al., 2012; CFIA, 2016). In general, in these studies the seismic
PGA-475 of 0.44g and PGA-2475 of 0.77g) when compared hazard also decreases as the distance increases from the Pacific
with Liberia (0.70 and 1.19g). interplate zone.
In order to study the contribution of different seismic A comparison between the PGA-475 calculated in this study
sources to the seismic hazard of these cities, we present the and that from the project Resis II (Benito et al., 2012) shows
disaggregation of the seismic hazard as a function of distance marked differences (Fig. 8a). In that study, the highest PGA
and magnitude for PGA and SA(1.0 s) at Tr of 475 and 2475 yr values of 0.50–0.60g occur in areas along the Pacific coast
(Fig. 7). For San Jose, three main sources are clearly identified: and the lowest values of 0.20–0.25g occur in areas along
the first one is the upper plate (zone C6) with historical inter- Northern Costa Rica. In our study, there is a slight decrease
mediate-magnitude earthquakes (5:0 < M w < 6:4) that are of 0.10–0.20g in the accelerations estimated for central Costa
relevant for PGA (Figs. 1 and 7a), for example, the 1910 Rica. On the other hand, there is an increase of 0.10–0.20g in
M w 6.4 Cartago earthquake located ∼20 km southeast of most of Guanacaste, the Osa peninsula, and the southern
San Jose (Fig. 1). The second source is the interplate zone Caribbean, and a significant increase of 0.20–0.60g in the
(Csi12), with historical events like the 1990 M w 7.1 Cobano Nicoya peninsula (Fig. 8a).
and the 1999 M w 6.9 Quepos earthquakes, located at around We also compare our results with the seismic code (CFIA,
90–95 km from the city (Fig. 1). Finally, the third source com- 2016). That study shows three seismic zones for PGA-475
bines the interplate seismic zones Csi11 and Csi13, more than (Fig. 7b) of 0.25, 0.38, and 0.50g, which, in general, corresponds
100 km away from San Jose that could potentially affect long to northwest–southeast-oriented zones named II, III, and IV
period structures (Figs. 1 and 7b). The most recent largest his- corresponding to northern, central, and the Pacific of Costa
torical earthquakes in these interplate zones have been the Rica, respectively. Similarly, CFIA (2016) shows the three seis-
1983 M w 7.4 Golfito earthquake and the 2012 M w 7.6 mic zones for the PGA-2475 but with higher values.
Nicoya earthquake (Fig. 1). The PGA-475 values estimated in our study (Fig. 5a) are
The seismic hazard of the city of Liberia is mainly controlled generally higher than those of the CFIA (2016; Fig. 8b).
by two seismic sources. The interplate zone Csi11 has a Remarkable differences are observed for the Nicoya penin-
predominant effect for both the PGA (Fig. 7c) and a natural sula and the rest of Guanacaste, in which our results are
period of 1.0 s (Fig. 7d). The largest historical earthquakes 0.20–0.60g higher, as well as in the Osa peninsula, in which
in this seismic zone are the 1950 M w 7.7 and 2012 M w 7.6 there is an increase of 0.10–0.30g. For the PGA-2475, our
Nicoya earthquakes, located 60–100 km southwest of Liberia results are 0.19–0.44g higher for Central Costa Rica and
(Fig. 1). The second contributing source is the upper-plate the Southern Caribbean, and 0.45–1.05g for the North
zone C5 with historical intermediate-magnitude earthquakes Pacific. As it can be appreciated in our results, because there
(5:0 < M w < 6:5) like the 1973 M w 6.5 Tilaran earthquake is a variability among the regions, there is no single factor for
located ∼62 km to the southeast of the city (Fig. 1). converting the PGA-475 values to PGA-2475 values. These
The seismic hazard disaggregation shows that even though results cast doubt on the suitability of applying a single factor
the seismic hazard is lower in San Jose than in Liberia, there is a to convert intensity values from one Tr to another across a
larger complexity in the contributing seismic sources for San territory.
Jose. The capital city is affected by earthquakes occurring in a The higher seismic hazard values obtained from our results
larger range of distances (up to 150 km) than Liberia, and from in certain areas, when compared with the previous works, can
more sources, consequently, the hazard assessment is more be explained by the use of new earthquake data, updated seis-
complex for San Jose. The hazard disaggregation also shows micity parameters, and an improved depth geometry for the
that the contribution of the interplate sources is distinctly interplate regions. The ∼35,000 new well-located earthquakes
higher for Liberia, with a contribution of 58% (Fig. 7d), that the RSN acquired during the last decade allowed for a bet-
whereas the maximum contribution from the interplate source ter determination of the magnitude–frequency distribution

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 11

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
shown by Arroyo-Solórzano and Linkimer (2021). This new Figure 6. (a) Seismic hazard curves and (b) uniform hazard spectrum (UHS)
combination of the a- and b-values used in our study implies calculated for the cities San Jose and Liberia. Intensity measures and Trs are
a higher probability of occurrence of large earthquakes in the color coded as indicated. The color version of this figure is available only in
the electronic edition.
Nicoya peninsula and Guanacaste compared with the values
used by Benito et al. (2012), leading to a higher hazard esti-
mation. In addition, in the Resis II project (Benito et al.,
2012), the interplate zone Csi11 was defined from 26 to seismometers (e.g., DeShon et al., 2006; Arroyo et al., 2009;
63 km depth. But based on the latest model of the top of Arroyo and Linkimer, 2021), and rupture and slip models
the Cocos slab (Lücke and Arroyo, 2015), earthquake locations of the 2012 M w 7.6 Nicoya earthquake (e.g., Yue et al.,
and tomographic models obtained from ocean-bottom 2013; Protti et al., 2014; Liu et al., 2015), in our study we

12 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
latest largest (M w > 7:4) earth-
quakes occurred in the Nicoya
peninsula in 1950 and 2012
(Güendel, 1986; Protti et al.,
2014) and in the Osa peninsula
in 1983 (Fig. 1). The extremely
high seismic hazard zone has
PGA-475 values of 0.55–1.20g
and PGA-2475 of 1.00–1.80g.
The highest PGA instrumen-
tally recorded by the LIS net-
work has been of 1.60g at
25 km from the epicenter of
the 2012 M w 7.6 Nicoya earth-
quake (Schmidt-Díaz et al.,
2014), demonstrating the
extremely high intensities
reached during large earth-
quakes in this peninsula.
The region with a very high
seismic hazard corresponds to
most of the inner part of
Guanacaste (Fig. 8c), which
represents the ∼21% of the
country. This region is
bounded by the VAF to the
Figure 7. Seismic hazard disaggregation for the cities of San Jose and Liberia. (a) PGA for a Tr of 475 yr in San Jose, northeast and by the CCRDB
(b) SA (1.0 s) for a Tr of 2475 in San Jose, (c) PGA for a Tr of 475 yr in Liberia, and (d) SA (1.0 s) for a Tr of 475 yr.
to the southeast. Together
Annotations with arrows show the seismic zones contributing to the seismic hazard for each city (Fig. 1).
with the Nicoya peninsula, this
region is moving northwest as
part of the CAFAB (LaFemina
defined interplate seismogenic depths from 12 to 35 km. This et al., 2009; Kobayashi et al., 2014; Fig. 1). Destructive shallow
locates the interplate zone directly beneath the Nicoya penin- earthquakes are common in Guanacaste, including the 1973
sula, leading again to much higher hazard values than the pre- M w 6.5 Tilaran earthquake (Fig. 1), which is the largest
vious works, especially along the coastline. upper-plate earthquake on record along the Costa Rican vol-
Considering the observed trends in the ground-motion lev- canic arc (e.g., Güendel, 1986). In this region, the calculated
els and the active tectonic features, four zones with distinct values for PGA-475 are 0.55–0.70g and for PGA-2475 are
seismic-hazard levels can be identified (Fig. 8c), and we 1.00–1.20g. The SA (0.2) and SA (1.0) results also characterize
described them in the categories of moderate, high, very high, this region very distinctly (Fig. 5).
and extremely high seismic hazard. The geometry of these A high level of seismic hazard encompasses the majority
zones was drawn using both the PGA and SA contours, and (∼41%) of Costa Rica and its population, including Central
the general location of tectonic boundaries. Costa Rica, the Southern Caribbean, and parts of the Southeast
The region with an extremely high seismic hazard contains Pacific (Fig. 8c). This zone is affected mostly by the CCRDB and
the three main peninsulas in the Pacific: Nicoya, Osa, and the inland continuation of the NPDB. The deadliest (∼600 fatal-
Burica (Fig. 8c), and corresponds to ∼16% of the Costa Rican ities) disaster in Costa Rica history—the 1910 M w 6.4 Cartago
territory. Even though the ground-motion values expected for earthquake (Fig. 1) is just one of many destructive earthquakes
the Nicoya peninsula are higher than Osa and Burica, the val- located within the high seismic hazard zone (e.g., Alonso-Henar
ues in the latter are generally higher than the rest of the coun- et al., 2013). The calculated PGA-475 ranges mostly from 0.40 to
try, and the tectonic context and seismicity favor its description 0.55g and the PGA-2475 from 0.75 to 1.00g.
as a unit. The three peninsulas override part of the megathrust Finally, a moderate seismic hazard level is the lowest
seismogenic zone between the subducting Cocos and the upper category for Costa Rica, spanning ∼22% of the continental
Caribbean and Panama plates, and are the regions where territory, and including the Talamanca Cordillera and
M w > 6:5 earthquakes are more frequent in Costa Rica. The Northern Costa Rica. Although the latter belongs to a backarc

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 13

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
subduction environment, the Talamanca Cordillera is a rapidly Figure 8. Difference between our results and the previous studies for PGA as
uplifting range, where extrusive volcanic activity ended a fraction of g for a Tr of 475 yr. The black squares represent the cities of San
∼3.5 Ma (MacMillan et al., 2004). There are no historical dam- Jose (S) and Liberia (L). Tectonic boundaries as in Figure 1. (a) The Resis II
project (Benito et al., 2012). (b) The Costa Rican Seismic Code (CFIA, 2016).
aging earthquakes with epicenters in the interior of this region;
Peculiar shapes for the geographic regions are preserved from the political
however, earthquakes in the proximities may still induce division of districts used by CFIA (2016) to display the PGA values.
strong shaking as shown by our results: PGA-475 could be (c) Seismic hazard zonation based on both our results and the location of
as high as 0.40g and PGA-2475 as high as 0.75g. tectonic boundaries. These zones were qualitatively described in the text as
Our results show that ∼78% of the continental territory of regions with moderate, high, very high, and extremely high seismic hazard.
Costa Rica could be described with a high, very high, and The color version of this figure is available only in the electronic edition.
extremely high seismic hazard. In other subduction zones of
the world, several PSHA have shown ground-motion levels
comparable to those obtained here for the Pacific peninsulas.
In Ecuador, the highest PGA values predicted are also found 1.37g for a Tr of 2475 yr (Parra et al., 2016). Further high inten-
along the coastline, especially in three peninsulas, which are sity levels of 0.40g for a Tr of 475 yr and of 0.65g for a Tr of
the areas closest to the interplate seismogenic zone. There, 2475 yr are expected close to north Andean block limit. Along
PGA values of 0.76g are expected for a Tr of 475 yr and of the Pacific coast of Colombia, the calculated PGA reach levels

14 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
of 0.60–0.90g for a Tr of 475 yr in the south, surpassing 0.90g The seismic hazard estimated for the cities of San Jose—the
for a Tr of 2475 yr, mainly associated with the subduction capital located in central Costa Rica, and Liberia—the biggest
interplate sources (Arcila et al., 2020). PGA values of city in Northern Costa Rica, show definite differences. In
0.75–1.00g have been modeled in north and central Chile Liberia, the seismic hazard is larger than for San Jose and is
for Tr of 475 yr, increasing to 1.50g for a Tr of 2475 yr mainly controlled by the interplate zone. The hazard for
(Medina et al., 2017). A PSHA for Taiwan, located also in a San Jose is more complex and multisource from distances
very complicated tectonic setting involving two subduction up to 150 km, although lower than in Liberia.
zones, have estimated accelerations up to 1.40–1.60g for a Regular updates of the seismic hazard for Costa Rica are a
Tr of 475 yr (Chan et al., 2020). necessary task to incorporate the latest understanding of the
In addition, our SA(0.2) values are, in general, comparable seismic sources, identify further research needs, contribute
or higher than those calculated in other subduction zones. For to the seismic code, and provide a more robust and realistic
a Trof 475 yr, for example, the highest values in parts of Chile framework for seismic risk mitigation.
are 1.00–2.00g (Medina et al., 2017), whereas in Costa Rica the
highest values are 2.10–3.00g in the Nicoya peninsula, which is DATA AND RESOURCES
sitting above the locked portion of the interplate seismogenic The National Seismological Network (RSN) catalog was obtained
zone. For a Tr of 2475 yr in the western coasts of Canada from the National Seismological Network (RSN; http://www.rsn.ucr
(Adams et al., 2019) and the U.S. (USGS, 2022), western .ac.cr, doi: 10.15517/TC). The Earthquake Engineering Laboratory
(LIS) catalog is available at https://www.crsmd.lis.ucr.ac.cr/. The
Sumatra (Sengara et al., 2020), and Taiwan (Wang et al.,
intensity measures and statistical summary can be found on the
2016), the estimated values are 2.40–4.00g. Similarly, values
website http://crsmd.lis.ucr.ac.cr/crsmdb.zip. Some plots were made
of 2.40–4.00g have been calculated in the Honshu Island of using Quantum Geographic Information System (QGIS) Geographic
Japan (Fujiwara et al., 2012) and in New Zealand (Stirling et al., Information System version 3.16.3 (Open Source Geospatial
2012), even reaching 7.00g in the Shizuoka Prefecture and the Foundation project: http://www.qgis.org). The supplemental material
Canterbury and West Coast regions, respectively. In Costa is composed of two tables. Table S1 has the seismicity parameters
Rica, the peak values are 3.00–5.50g. We believe that these high established in the OpenQuake software. Table S2 has the magnitudes,
values along the western coastlines in the Pacific peninsulas of locations, and fault-plane solutions of the selected earthquakes. All
Costa Rica express the extreme hazard of these regions located websites were last accessed in June 2022.
at only 10–15 km on top of the interplate seismogenic zone
(e.g., DeShon et al., 2006; Arroyo et al., 2009; Arroyo and DECLARATION OF COMPETING INTERESTS
Linkimer, 2021). The authors acknowledge that there are no conflicts of interest
recorded.
CONCLUSIONS
The 2022 Seismic Hazard Model for Costa Rica displays four ACKNOWLEDGMENTS
regions with distinct ground-motion levels, which correlate This research was partially funded by the National Emergency and
Risk Prevention Law N°8488 of the Republic of Costa Rica and the
with tectonic boundaries. These levels are: (1) extremely high
UCREA funds from the University of Costa Rica through the project
for ∼16% of the territory involving the Nicoya, Osa, and Burica
referenced as B9-780 as well as the research projects B5-704 and B9-
peninsulas, situated directly above the subduction interplate, in 911. The authors thank Marco Pagani and Luis Mixco for a number of
which the PGA-475 could be 0.55–1.20g; (2) very high for most helpful comments and discussions regarding OpenQuake, and two
of the Guanacaste Province (∼21% of the country), in which anonymous reviewers whose comments greatly improve the article.
the PGA-475 may be 0.55–0.70g; (3) high for ∼41% of Costa
Rica, including the central region and the capital city of San REFERENCES
Jose, with PGA-475 values from 0.40 to 0.55g; and (4) moderate Abrahamson, N., N. Gregor, and K. Addo (2016). BC hydro ground
for the Talamanca Cordillera and Northern Costa Rica, with motion prediction equations for subduction earthquakes, Earthq.
PGA-475 up to 0.40g and corresponding to ∼22% of the Spectra 32, no. 1, 23–44, doi: 10.1193/051712EQS188MR.
continental territory. Adamek, S., C. Frohlich, and W. Pennington (1988). Seismicity of the
The expected ground motions are, in general, higher than Caribbean-Nazca boundary: Constraints on microplate tectonics
of the Panama Region, J. Geophys. Res. 93, 2053–2075, doi:
obtained in the previous studies, except for the central part of
10.1029/JB093iB03p02053.
the country, in which values are 0.10–0.20g smaller for PGA-
Adams, J., T. Allen, S. Halchuk, and M. Kolaj (2019). Canada’s 6th
475. The Nicoya, Osa, and Burica peninsulas present the larg-
Generation Seismic Hazard Model, as Prepared for the 2020
est increment (0.1–0.6g) in PGA-475 when compared with the National Building Code of Canada, 12th Canadian Conf. on
past works. In addition, we estimated a relative wide range of Earthquake Engineering, Quebec City.
dynamic amplification factors from 1.98 to 3.03, which chal- Aki, K. (1965). Maximum likelihood estimated of b in the formula log
lenge the use of a single value—a practice routinely performed N = A – b*M and its confidence limits, Rev. Earthquakes Res. Inst.
in Costa Rica. Tokyo Univ. 43, 237–239.

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 15

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
Alonso-Henar, J., W. Montero, J. Martínez-Díaz, J. Álvarez-Gómez, J. Chan, C. H., K. F. Ma, J. B. H. Shyu, Y. T. Lee, Y. J. Wang, J. C. Gao, Y.
Insua-Arévalo, and W. Rojas (2013). The Aguacaliente fault, T. Yen, and R. J. Rau (2020). Probabilistic seismic hazard assess-
source of the Cartago 1910 destructive earthquake (Costa Rica), ment for Taiwan: TEM PSHA2020, Earthq. Spectra 36, 137–159,
Terra Nova 25, no. 5, 368–373, doi: 10.1111/ter.12045. doi: 10.1177/8755293020951587.
Alvarado, G. E., B. Benito, A. Staller, Á. Climent, E. Camacho, W. Climent, Á., W. Rojas, G. E. Alvarado, and M. B. Benito (2008).
Rojas, G. Marroquín, E. Molina, J. E. Talavera, S. Martínez- Proyecto Resis II Evaluación de la amenaza sísmica (in Spanish).
Cuevas, et al. (2017). The new central American seismic hazard Climent, Á., W. Taylor, M. Ciudad-Real, W. Strauch, M. Villagran, A.
zonation: Mutual consensus based on up to day seismotectonic Dahle, and H. Bungum (1994). NORSAR technical report: Spectral
framework, Tectonophysics 721, 462–476, doi: 10.1016/ strong motion attenuation in central America, NORSAR Technical
j.tecto.2017.10.013. Report No. 2-17, , 63 pp.
Arcila, M., J. García, J. S. Montejo Espitia, J. F. Eraso, J. Valcárcel, M. Cotton, F., F. Scherbaum, J. J. Bommer, and H. Bungum (2006).
G. Mora Cuevas, D. Vigano, M. Pagani, and F. J. Díaz Parra (2020). Criteria for selecting and adjusting ground-motion models for spe-
Modelo nacional de amenaza sísmica para Colombia, doi: cific target regions: Application to central Europe and rock sites, J.
10.32685/9789585279469 (in Spanish). Seismol. 10, no. 2, 137–156, doi: 10.1007/s10950-005-9006-7.
Arroyo, I. G., and L. Linkimer (2021). Modelo unidimensional de DeMets, C., R. G. Gordon, and D. F. Argus (2010). Geologically cur-
velocidades sísmicas y características tomográficas tridimensio- rent plate motions, Geophys. J. Int. 181, 1–80, doi: 10.1111/j.1365-
nales del Caribe Sur de Costa Rica, Revista Geológica de 246X.2009.04491.x.
América Central 65, 1–15, doi: 10.15517/Rgac.V0i65.46696 (in Denyer, P., W. Montero, and G. E. Alvarado (2009). Atlas tectónico de
Spanish). Costa Rica, Editorial Universidad de Costa Rica, San José, Costa
Arroyo-Solórzano, M., and L. Linkimer (2021). Spatial variability of Rica, 64 pp. (in Spanish).
the b-value and seismic potential in Costa Rica, Tectonophysics doi: DeShon, H., S. Schwartz, A. V. Newman, V. González, M. Protti, L. M.
10.1016/j.tecto.2021.228951. Dorman, T. H. Dixon, D. E. Sampson, and E. R. Flueh (2006).
Arroyo, I. G., S. Husen, E. R. Flueh, J. Gossler, E. Kissling, and G. E. Seismogenic zone structure beneath the Nicoya peninsula,
Alvarado (2009). Three-dimensional P-wave velocity structure on Costa Rica, from three-dimensional local earthquake P and S-wave
the shallow part of the central Costa Rican Pacific margin from tomography, Geophys. J. Int. 164, no. 1, 109–124, doi: 10.1111/
local earthquake tomography using off- and onshore networks, j.1365-246X.2005.02809.x.
Geophys. J. Int. 179, no. 2, 827–849, doi: 10.1111/j.1365- Douglas, J. (2021). Ground motion prediction equations 1964-2021,
246X.2009.04342.x. available at http://www.gmpe.org.uk/ (last accessed May 2021).
Barckhausen, U., C. R. Ranero, R. von Huene, S. C. Cande, and H. A. Fan, G., S. Beck, and T. Wallace (1993). The seismic source param-
Roeser (2001). Revised tectonic boundaries in the Cocos plate off eters of the 1991 Costa Rica aftershock sequence: Evidence for a
Costa Rica: Implications for the segmentation of the convergent transcurrent plate boundary, J. Geophys. Res. 98, 15,759–15,778.
margin and for plate tectonic models, J. Geophys. Res. 106, Fujiwara, H., N. Morikawa, T. Okumura, Y. Ishikawa, and N. Nojima
no. B9, 207–220, doi: 10.1029/2001JB000238. (2012). Revision of probabilistic seismic hazard assessment for
Benito, M. B., C. Lindholm, E. Camacho, Á. Climent, G. Marroquin, E. Japan after the 2011 Tohoku-Oki mega-thrust earthquake
Molina, W. Rojas, J. J. Escobar, E. Talavera, G. E. Alvarado, et al. (M9.0), 15th World Conf. on Earthquake Engineering, Lisbon,
(2012). A new evaluation of seismic hazard for the central America Portugal, 24–28 September 2012.
region, Bull. Seismol. Soc. Am. 102, no. 2, 504–523, doi: 10.1785/ García, D., S. K. Singh, M. Herraíz, M. Ordaz, and J. F. Pacheco
0120110015. (2005). Inslab earthquakes of central Mexico: Peak ground-motion
Bommer, J. J. (2004). The challenge of defining upper bounds on parameters and response spectra, Bull. Seismol. Soc. Am. 95, no. 6,
earthquake ground motions, Seismol. Res. Lett. 75, no. 1, 82–95, 2272–2282, doi: 10.1785/0120050072.
doi: 10.1785/gssrl.75.1.82. Gardner, J., and L. Knopoff (1974). Is the sequence of earthquakes in
Boore, D. M., J. P. Stewart, E. Seyhan, and G. M. Atkinson (2014). southern California, with aftershocks removed, Poissonian? Bull.
NGA-West2 equations for predicting PGA, PGV, and 5% damped Seismol. Soc. Am. 64, no. 5, 1363–1367.
PSA for shallow crustal earthquakes, Earthq. Spectra 30, no. 3, Global Earthquake Model (GEM) (2018a). CCARA project:
1057–1085, doi: 10.1193/070113EQS184M. Caribbean and central America earthquake risk assessment, avail-
Campbell, K. W. (2003). Prediction of strong ground motion using the able at https://www.globalquakemodel.org/project/ccara (last
hybrid empirical method and its use in the development of accessed April 2022).
ground-motion (attenuation) relations in eastern North Global Earthquake Model (GEM) (2018b). gmpe-smtk, available at
America, Bull. Seismol. Soc. Am. 93, no. 3, 1012–1033, doi: https://github.com/GEMScienceTools/gmpe-smtk (last accessed
10.1785/0120020002. October 2020).
Cauzzi, C., E. Faccioli, M. Vanini, and A. Bianchini (2015). Updated Global Earthquake Model (GEM) (2020). The OpenQuake-
predictive equations for broadband (0.01–10 s) horizontal engine user manual, Global Earthquake Model (GEM)
response spectra and peak ground motions, based on a global data- OpenQuake Manual for Engine version 3.9.0 doi: 10.13117/
set of digital acceleration records, Bull. Earthq. Eng. 13, no. 6, GEM.OPENQUAKE.MAN.ENGINE.3.9.0.
1587–1612, doi: 10.1007/s10518-014-9685-y. Güendel, F. (1986). Seismotectonics of Costa Rica: An analytical view
CFIA (2016). Código Sísmico de Costa Rica 2010 (Revisión 2014), Fifth of the southern terminus of the middle America trench, University
Ed., Editorial Tecnológica de Costa Rica, Cartago (in Spanish). of California at Santa Cruz, Ph.D. Thesis, 157 pp.

16 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
Kanno, T., A. Narita, N. Morikawa, H. Fujiwara, and Y. Fukushima Medina, F., S. C. Harmsen, and S. E. Barrientos (2017). Probabilistic
(2006). A new attenuation relation for strong ground motion in seismic hazard analysis for Chile, 16th World Conf. on Earthquake
Japan based on recorded data, Bull. Seismol. Soc. Am. 96, no. 3, Engineering, Santiago, Chile, 9–13 January 2017.
879–897, doi: 10.1785/0120050138. Montalva, G. A., N. Bastías, and A. Rodriguez-Marek (2017). Ground-
Kobayashi, D., P. LaFemina, H. Geirsson, E. Chichaco, A. A. Abrego, motion prediction equation for the Chilean subduction zone, Bull.
H. Mora, and E. Camacho (2014). Kinematics of the western Seismol. Soc. Am. 107, no. 2, 901–911, doi: 10.1785/0120160221.
Caribbean: Collision of the Cocos Ridge and upper plate deforma- Montero, W. (2001). Neotectónica de la región central de Costa Rica:
tion, Geochem. Geophys. Geosys. 15, doi: 10.1002/ 2014GC005234. frontera oeste de la microplaca de Panamá, Revista Geológica de
Kolarsky, R. A., and P. Mann (1995). Structure and neotectonics of an América Central 24, 29–56 (in Spanish).
oblique subduction margin, southwestern Panama, in Geologic and Morales, L. D. (1986). Historia de la sismología en Costa Rica, Revista
Tectonic Development of the Caribbean Plate Boundary in Southern de Filosofía Universidad de Costa Rica 24, no. 59, 93–104.
Central America, P. Mann (Editor), Geological Society of America, Mortgat, C. P., T. C. Zsutty, H. C. Shah, and L. Lubetkin (1977). A
Special Paper, Colorado, U.S.A., 295–349 pp. Study of Seismic Risk for Costa Rica, John A. Blume Earthquake
Kotha, S. R., D. Bindi, and F. Cotton (2017). From ergodic to region- Engineering Center, Stanford.
and site-specific probabilistic seismic hazard assessment: Method Moya-Fernández, A., L. A. Pinzón, V. Schmidt-Díaz, D. A. Hidalgo-
development and application at European and middle eastern sites, Leiva, and L. G. Pujades (2020). A strong-motion database of Costa
Earthq. Spectra 33, no. 4, 1433–1453, doi: 10.1193/081016EQS130M. Rica: 20 yr of digital records, Seismol. Res. Lett. 91, no. 6, 3407–
Kyriakopoulos, C., A. V. Newman, A. M. Thomas, M. Moore-Driskell, 3416, doi: 10.1785/0220200036.
and G. T. Farmer (2015). A new seismically constrained subduc- Parra, H., M. B. Benito, and J. M. Gaspar-Escribano (2016). Seismic
tion interface model for central America, J. Geophys. Res. 120, hazard assessment in continental Ecuador, Bull. Earthq. Eng. 14,
5535–5548, doi: 10.1002/2014JB011859. 2129–2159, doi: 10.1007/s10518-016-9906-7.
LaFemina, P., T. H. Dixon, R. Govers, E. Norabuena, H. Turner, A. Pinzón, L. A., D. A. Hidalgo-Leiva, A. Moya-Fernández, V. Schmidt-
Saballos, G. Mattioli, M. Protti, and W. Strauch (2009). Forearc Díaz, and L. G. Pujades (2021). Seismic site classification of the
motion and Cocos Ridge collision in central America, Geochem. Costa Rican strong-motion network based on Vs30 measurements
Geophys. Geosys. 10, doi: 10.1029/2008GC002181. and site fundamental period, Earth Sci. Res. J. 25, no. 4, 383–389,
Laporte, M., C. Lindholm, H. Bungunr, and A. Dahle (1994). NORSAR doi: 10.15446/esrj.v25n4.93927.
technical report: Seismic hazard for Costa Rica, NORSAR Technical Protti, M., V. González, and A. Newman (2014). Nicoya earthquake
Report No. 2-14, 73 pp. rupture anticipated by geodetic measurement of the locked plate
Lin, P. S., and C. T. Lee (2008). Ground-motion attenuation relation- interface, Nature Geosci. 7, 117–121, doi: 10.1038/ngeo2038.
ships for subduction-zone earthquakes in northeastern Taiwan, Rojas, W., H. Cowan, C. Lindholm, A. Dahle, and H. Bungum (1993).
Bull. Seismol. Soc. Am. 98, no. 1, 220–240, doi: 10.1785/0120060002. NORSAR technical report: Regional seismic zonation for central
Linkimer, L., and G. E. Alvarado (2014). Distribución espacio-temporal de América a preliminary modelNORSAR report under the RESIS I
la sismicidad sentida en Costa Rica (1976-2013) en el marco histórico regional seismology project, 40 pp.
del 30 aniversario (1982-2012) de la Red Sismológica Nacional Sallarès, V., P. Charvis, E. R. Flueh, and J. Bialas (2003). Seismic struc-
(RSN: UCR-ICE), Rev. Geol. Amér. Central, Spec. 30, aniversario ture of Cocos and Malpelo volcanic ridges and implications for hot
(1984-2014), 45–71, doi: 10.15517/rgac.v0i0.16569 (in Spanish). spot-ridge interaction, J. Geophys. Res. 108, no. B12, 2564, doi:
Linkimer, L., I. G. Arroyo, G. Alvarado, M. Arroyo, and H. Bakkar 10.1029/2003JB002431.
(2018). The National Seismological network of Costa Rica (RSN): Sallarès, V., J. J. Dañobeitia, and E. R. Flueh (2001), Lithospheric
An overview and recent developments, Seismol. Res. Lett. 89, structure of the Costa Rican Isthmus: Effects of subduction zone
no. 2A, 392–398. magmatism on an oceanic plateau, J. Geophys. Res. 106, no. B1,
Liu, C., X. Xiong, Y. Zheng, A. López, R. Wang, and J. Li (2015). 621–643, doi: 10.1029/2000JB900245.
Rupture processes of the 2012 September 5 Mw 7.6 Nicoya, Sengara, I. W., M. Irsyam, I. D. Sidi, A. Mulia, M. Asrurifak, D.
Costa Rica earthquake constrained by improved geodetic and seis- Hutabarat, and W. Partono (2020), New 2019 risk-targeted ground
mological observations, Geophys. J. Int. 203, no. 1, 175–183, doi: motions for spectral design criteria in Indonesian seismic building
10.1093/gji/ggv295. code, E3S Web Conf. 156, 03010, doi: 10.1051/e3sconf/
Lücke, O. H., and I. G. Arroyo (2015). Density structure and geometry of 202015603010.
the Costa Rican subduction zone from 3-D gravity modeling and local Scherbaum, F., F. Cotton, and P. Smit (2004). On the use of response
earthquake data, Solid Earth 6, 1169–1183, doi: 10.5194/se-6-1169-2015. spectral-reference data for the selection and ranking of ground-
Makropoulos, K., and P. Burton (1983). Seismic risk of circum-pacific motion models for seismic-hazard analysis in regions of moderate
earthquakes I. Strain energy release, Pure Appl. Geophys. 121, seismicity: The case of rock motion, Bull. Seismol. Soc. Am. 94,
no. 2, 247–266. no. 6, 2164–2185, doi: 10.1785/0120030147.
MacMillan, I., P. B. Gans, and G. E. Alvarado (2004). Middle Miocene Scherbaum, F., E. Delavaud, and C. Riggelsen (2009). Model selection
to present plate tectonic history of the southern central American in seismic hazard analysis: An information-theoretic perspective,
volcanic arc, Tectonophysics 392, 325–348. Bull. Seismol. Soc. Am. 99, no. 6, 3234–3247, doi: 10.1785/
Marshall, J., D. Fisher, and T. Gardner (2000). Central Costa Rica 0120080347.
deformed belt: Kinematics of diffuse faulting across the western Schmidt-Díaz, V., A. Dahle, and H. Bungum (1997). NORSAR tech-
Panamá block, Tectonics 19, 468–492. nical report: Costa Rican spectral strong motion attenuation,

Volume XX Number XX – 2023 www.bssaonline.org Bulletin of the Seismological Society of America • 17

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user
NORSAR report under the RESIS I regional seismology project, Weichert, D. (1980). Estimation of the earthquake recurrence param-
45 pp. eters for unequal observation periods for different magnitudes,
Schmidt-Díaz, V., D. A. Hidalgo-Leiva, A. L. Acuña, A. Moya- Bull. Seismol. Soc. Am. 70, no. 4, 1337–1346.
Fernández, E. Cordero, C. Segura, and E. López (2014). Wells, D. L., and K. J. Coppersmith (1994). New empirical relation-
Aceleraciones del terremoto de Sámara del 05 de setiembre del ship between magnitude, rupture length, rupture width, rupture
2012, Revista En Torno a La Prevención 12, no. 1, 38–47 (in Spanish). area, and surface displacement, Bull. Seismol. Soc. Am. 84,
Silver, E. A., D. L. Reed, J. E. Tagudin, and D. J. Heil (1990). no. 4, 974–1002.
Implications of the north and south Panama thrust belts for Westbrook, G. K., N. C. Hardt, and R. Heath (1995). Structure and
the origin of the Panama Orocline, Tectonics 9, no. 2, 261–281, tectonics of the Panama–Nazca boundary, in Geologic and
doi: 10.1029/TC009i002p00261. Tectonic Development of the Caribbean Plate Boundary in
Stepp, J. (1972). Analysis of completeness of the earthquake sample in Southern Central America, P. Mann, (Editor), Geological
the Puget Sound area and its effect on statistical estimates of Society of America, Special Paper, Colorado, U.S.A., 295, 349 pp.
earthquake hazard, Proc. of the Microzonation Conference, Wiemer, S., and M. Wyss (2000). Minimum magnitude of completeness
Vol. 2, 897–909. in earthquake catalogs: Examples from Alaska, the western United
Stirling, M., G. McVerry, M. Gerstenberger, N. Litchfield, R. Van States, and Japan, Bull. Seismol. Soc. Am. 90, no. 4, 859–869.
Dissen, K. Berryman, P. Barnes, L. Wallace, P. Villamor, R. Yue, H., T. Lay, S. Y. Schwartz, L. Rivera, M. Protti, T. H. Dixon, S.
Langridge, et al. (2012). National seismic hazard model for Owen, and A. V. Newman (2013). The 5 September 2012 Nicoya,
New Zealand: 2010 update, Bull. Seismol. Soc. Am. 102, no. 4, Costa Rica Mw 7.6 earthquake rupture process from joint inver-
1514–1542, doi: 10.1785/0120110170. sion of high-rate GPS, strong-motion, and teleseismic P wave
U.S. Geological Survey (USGS) (2022). Available at https://www.usgs data and its relationship to adjacent plate boundary interface
.gov/programs/earthquake-hazards/science/national-seismic- properties, J. Geophys. Res. 118, 5453–5466, doi: 10.1002/
hazard-maps (last accessed June 2022). jgrb.50379.
Von Huene, R., C. R. Ranero, W. Weinrebe, and K. Hinz (2000). Zhao, J. X., J. Zhang, A. Asano, Y. Ohno, T. Oouchi, T. Takahashi, H.
Quaternary convergent margin tectonics of Costa Rica, segmenta- Ogawa, K. Irikura, H. K. Thio, P. G. Somerville, et al. (2006).
tion of the Cocos plate, and central American volcanism, Tectonics Attenuation relations of strong ground motion in Japan using site
19, no. 2, 314–334, doi: 10.1029/1999tc001143. classification based on predominant period, Bull. Seismol. Soc. Am.
Walther, C. H. E. (2003). The crustal structure of Cocos Ridge off Costa 96, no. 3, 898–913, doi: 10.1785/0120050122.
Rica, J. Geophys. Res. 108, no. B3, 2136, doi: 10.1029/2001JB000888.
Wang, Y.-J., Y.-T. Lee, C.-H. Chan, and K.-F. Ma (2016). An investi-
gation of the reliability of the Taiwan earthquake model PSHA2015, Manuscript received 22 June 2022
Seismol. Res. Lett. 87, no. 6, 1287–1298, doi: 10.1785/0220160085. Published online 2 December 2022

18 • Bulletin of the Seismological Society of America www.bssaonline.org Volume XX Number XX – 2023

Downloaded from http://pubs.geoscienceworld.org/ssa/bssa/article-pdf/doi/10.1785/0120220119/5743748/bssa-2022119.1.pdf


by Universidad de Costa Rica user

You might also like