You are on page 1of 9

EFFECT OF STRUCTURAL CONFIGURATION ON THE SEISMIC FRAGILITY

OF RC SHEAR WALL BUILDINGS

J. Ocampo1, F. Vidales1, D. Feliciano2,6, O. Arroyo2,5,6, J. Carrillo3,5, C. Arteta4,5 and R. Bonett1,5

1 Universidad de Medellin, Medellín, Colombia, Juan José Ocampo, jocampo862@soyudemedellin.edu.co


2 Universidad La Sabana, Bogotá, Colombia.
3 Universidad Militar Nueva Granada, UMNG, Bogotá, Colombia.
4 Universidad del Norte, Barranquilla, Colombia.
5 Colombian Earthquake Engineering Research Network, CEER, Colombia
6 Research center for disaster and climate change, CIDEC, Colombia

Abstract: The reinforced concrete (RC) shear wall system has gained widespread use in high seismicity
countries for housing construction, primarily due to its high stiffness and the advantages of reduced
construction time and cost through industrialization. In countries like Colombia and other Latin American
nations, the structural wall system has evolved to accommodate unconventional plan layouts driven by
architectural requirements. Notably, all architectural partitions serve as gravity load-bearing RC walls, also
functioning as the primary seismic resistance system, irrespective of their geometry or reinforcement detailing.
Furthermore, for over two decades, nonductile electro-welded mesh has been extensively employed as the
primary web reinforcement. However, these buildings have yet to experience high-intensity seismic events,
leaving their actual seismic performance unknown. Consequently, there is a pressing need to advance
numerical studies that quantify the behavior of this structural system. This study presents a seismic fragility
assessment of Thin Lightly-Reinforced Concrete Wall (TLRCW) Buildings, focusing on evaluating the influence
on structural fragility due to wall aspect ratio, wall index, and reinforcement detailing. The research is part of
the Colombian Risk Model, an initiative led by the Colombian Geological Service and the Colombian
Association of Engineering Schools. A total of sixty-eight representative archetypes embodying the prevalent
structural system in Colombia were selected, encompassing heights ranging from four to thirty stories and
representing buildings located in intermediate and high seismic hazard regions. Each archetype underwent
nonlinear response history analyses using OpenSeesPy, employing 3,400 hazard-consistent ground motion
records. The results demonstrate significant variability in structural response among buildings with the same
number of stories but differing architectural configurations, effectively accounting for epistemic uncertainty
when grouping the results by taxonomy. The findings also serve as a valuable benchmark for informing future
structural designs, vulnerability assessments, and risk management practices nationwide by practitioners,
researchers, and professionals.

1. Introduction
The Thin, Lightly-Reinforced Concrete Wall system (TLRCW) in Colombia has evidenced a significant trend
towards the use of thin and slender walls, employing an industrialized process that allows for the construction
of at least one apartment per day. This process involves the use of electro-welded mesh as web reinforcement.
WCEE2024 Ocampo et al.

The strength and stiffness properties of reinforced concrete (RC) walls, combined with the speed achievable
through industrialized construction, have promoted the widespread use of this structural system in Colombia.
Further, since the RC structural walls are also the architectural partitions, construction times are up to 50%
faster, and construction costs are up to 20% lower than those of the traditional frame-wall system. The living
unit areas range between 60 and 150 m2. The thin wall system is also commonly used in Peru, denoted as a
“limited ductility wall system” (Quiroz & Maruyama, 2013). The absence of explicit provisions to control the
minimum thickness and slenderness of walls (i.e., the ratio between their unsupported height and the
thickness, hu/tw) in the Colombian building code NSR-10, has enabled the use of extremely thin and slender
walls in the construction of buildings, primarily aimed at addressing the low-cost housing deficit. The Colombian
TLRCW system has unique features that set it apart from the international context, for example: 1) walls with
very small thicknesses and large slenderness, with thicknesses ranging from 100 mm to 150 mm; 2) walls
controlling drifts having short lengths or high aspect ratio; 3) walls with electro-welded wire meshes (WWM)
as the web reinforcement, which hinders their deformation capacity (Carrillo et al., 2019); and 4) inadequate
or absent confinement at the wall edges.
This work assesses the effect of the main attributes of the structural configuration of RC thin-wall buildings on
their seismic fragility, explicitly focusing on the probability of collapse. The results obtained are part of the
National Seismic Risk Model project for Colombia commissioned by the Colombian Geological Service (SGC,
“Servicio Geológico Colombiano” in Spanish) and the Colombian Association of Engineering Faculties (ACOFI,
“Asociación Colombiana de Facultades de Ingeniería”, in Spanish).

2. Characterization of TLRCW buildings


The use of the TLRCW system is primarily focused on housing and is employed for constructing buildings
ranging from 5 to 30 stories. Wall length within the system vary in the range 1.0 to 10 m, with an average
around 4.5 m. These features yield significant variations in the wall aspect ratio (height/length), a parameter
directly related to their inelastic rotational capacity. The reinforcement for these walls is usually arranged in a
single layer and, in some cases, in two layers distributed along the length of the wall, with additional
concentrated steel bars at the ends that may or may not be confined. WWMs are widely used as shear and
flexural web reinforcement. In some cases, boundary elements are included; however, their strain capacity is
questionable due to their reduced thickness. Figure 1 shows a typical floor plan of a typical high-rise building.

Figure 1. Example of the floor plan of a 26-story TLRCW building.

These particularities have raised concerns regarding the seismic performance of TLRCW buildings (Gonzales
& López-Almansa, 2012). Studies have particularly focused on evaluating inelastic deformation capacity,
lateral stability, and potential failure mechanisms. In Colombia, a gap has been identified in the current building
code, NSR-10 (AIS, 2010), as this regulation does not include an explicit distinction between the design
requirements for buildings with thick and thin walls (CEER, 2021). For instance, it does not limit the
slenderness of the walls. Additionally, the minimum wall thickness is not explicitly limited.
Numerical and experimental studies reported by Rosso et al. (2022), Blandón & Bonett (2020), Carrillo,
Oyarzo-Vera et al. (2019), Blandón et al. (2018), Rosso et al. (2018), Arroyo et al. (2021), and Arteta (2017)
have demonstrated the potential limitations of this type of structural system. Blandón & Bonett (2020) identified

2
WCEE2024 Ocampo et al.

walls with very limited effective confinement and deficiencies in the reinforcement detailing to restrain buckling
of the longitudinal bars. Carrillo et al. (2019) demonstrated that a significant percentage of the electro-welded
meshes produced in Colombia have limited deformation capacity and should be considered as reinforcement
with very limited ductility. The basic design assumption indicates that when the concrete reaches its maximum
compressive strength, the tensile reinforcement is yielding but not fracturing, which is not the case for flange-
wall reinforced with WWM. Blandón et al. (2018) reported a limited rotation capacity below 0.8% based on a
full-scale experimental program of walls with a thickness of 100 mm, all with T-shaped sections and an M/VLw
ratio equal to 2.0. Rosso et al. (2018) demonstrated a high potential for out-of-plane instability in walls with a
thickness of 80 mm, utilizing reinforcement bars concentrated at the ends of the wall. Table 1 shows the main
attributes of the 67 archetypes analyzed in this study. The first column shows the archetype's taxonomy,
followed by the number of stories (N), total height (H), seismic hazard level, design spectra acceleration (Sa),
wall index (WI), period (T), cracked period (Tcr) mean aspect ratio (Ar mean), and reinforcement type. Figure
2 shows the box and whisker plot for the main parameters analyzed in this study.

Figure 2. Structural parameters of TLRC buildings.

Table 1. Database of TLRCW buildings.

N. Seismic WI T Tcr
Archetype H (m) Sa (g) Ar Mean Reinf.
Stories Hazard (%) (s) (s)

0049-MCR-ARM-05P-L 5 12.5 High 0.81 0.11 0.2 0.32 4.83 RB+WWM


0049-MCR-ARM-05P-T 5 12.5 High 0.81 0.18 0.1 0.16 3.07 RB+WWM
0056-MCR-ARM-05P-L 5 12.5 High 0.81 0.27 0.1 0.18 3.44 RB+WWM
0056-MCR-ARM-05P-T 5 12.5 High 0.81 0.41 0.2 0.26 3.75 WWM
0057-MCR-ARM-05P-L 5 12.5 High 0.81 0.28 0.1 0.09 1.76 RB+WWM
0057-MCR-ARM-05P-T 5 12.5 High 0.81 0.42 0.1 0.15 2.59 RB+WWM
0210-MCR-CAS-05P-T 5 12.8 Inter. 0.56 0.29 0.2 0.24 4.43 WWM
0211-MCR-YOP-05P-T 5 12.5 High 0.90 0.29 0.1 0.18 2.59 WWM
0212-MCR-ARM-05P-T 5 12.4 High 0.81 0.31 0.1 0.16 1.77 RB+WWM
0220-MCR-PER-05P-L 5 12.5 High 0.81 0.24 0.2 0.29 4.35 RB+WWM
0220-MCR-PER-05P-T 5 12.5 High 0.81 0.44 0.2 0.25 3.06 RB+WWM
0227-MCR-IBA-05P-L 5 12.5 Inter. 0.70 0.13 0.2 0.27 4.7 RB
0227-MCR-IBA-05P-T 5 12.5 Inter. 0.70 0.54 0.1 0.16 2.43 RB
0038-MCR-BGT-06P-T 6 14.7 Inter. 0.56 0.16 0.2 0.24 2.44 WWM
0041-MCR-BGT-06P-L 6 14.8 Inter. 0.56 0.11 0.4 0.49 4.23 RB+WWM
0041-MCR-BGT-06P-T 6 14.8 Inter. 0.56 0.15 0.2 0.23 3.07 WWM
0042-MCR-BGT-06P-L 6 15.0 Inter. 0.56 0.15 0.2 0.24 2.42 WWM
0042-MCR-BGT-06P-T 6 15.0 Inter. 0.56 0.17 0.3 0.35 4.47 WWM
0209-MCR-YOP-06P-L 6 13.4 High 0.90 0.32 0.2 0.24 4.77 WWM
0209-MCR-YOP-06P-T 6 13.4 High 0.90 0.21 0.3 0.43 5.02 WWM
0208-MCR-MED-07P-L 7 17.5 Inter. 0.56 0.12 0.3 0.4 4.27 RB+WWM
0208-MCR-MED-07P-T 7 17.5 Inter. 0.56 0.43 0.4 0.48 4.89 RB+WWM
0011-MCR-MED-08P-L 8 18.4 Inter. 0.56 0.08 0.3 0.44 5.75 WWM
0011-MCR-MED-08P-T 8 18.4 Inter. 0.56 0.18 0.4 0.5 4.8 WWM
0058-MCR-ARM-08P-L 8 20.0 High 0.81 0.08 0.3 0.46 5.24 WWM
0058-MCR-ARM-08P-T 8 20.0 High 0.81 0.11 0.3 0.41 5.12 WWM
0229-MCR-BGT-09P-L 9 21.6 Inter. 0.56 0.22 0.6 0.77 5.44 RB+WWM
0229-MCR-BGT-09P-T 9 21.6 Inter. 0.56 0.17 0.7 0.88 7.57 RB+WWM
0197-MCR-BUC-11P-L 11 27.7 High 0.77 0.06 0.7 0.99 9.52 RB+WWM
0197-MCR-BUC-11P-T 11 27.7 High 0.81 0.11 0.5 0.72 6.43 RB+WWM
0230-MCR-MED-11P-L 11 26.4 Inter. 0.56 0.06 0.7 0.98 8.47 RB+WWM
0230-MCR-MED-11P-T 11 26.4 Inter. 0.56 0.11 0.5 0.73 7.41 RB+WWM

3
WCEE2024 Ocampo et al.

N. Seismic WI T Tcr
Archetype H (m) Sa (g) Ar Mean Reinf.
Stories Hazard (%) (s) (s)

0045-MCR-ARM-12P-L 12 30.0 High 0.81 0.03 0.6 0.75 3.8 RB+WWM


0045-MCR-ARM-12P-T 12 30.0 High 0.81 0.10 0.6 0.8 5.83 RB+WWM
0195-MCR-BUC-12P-L 12 30.0 High 0.76 0.16 0.8 1.01 6.79 RB+WWM
0195-MCR-BUC-12P-T 12 30.0 High 0.60 0.12 1 1.28 10.17 RB+WWM
0228-MCR-BGT-13P-L 13 31.2 Inter. 0.55 0.15 0.9 1.21 7.37 RB+WWM
0228-MCR-BGT-13P-T 13 31.2 Inter. 0.43 0.12 1.1 1.45 11.01 RB+WWM
0206-MCR-PER-14P-L 14 34.3 High 0.50 0.06 0.9 1.25 12.21 RB
0206-MCR-PER-14P-T 14 34.3 High 0.49 0.14 0.9 1.27 8.65 RB+WWM
0233-MCR-MED-14P-L 14 34.3 Inter. 0.52 0.06 0.9 1.27 12.21 RB
0233-MCR-MED-14P-T 14 34.3 Inter. 0.51 0.14 0.9 1.24 8.66 RB+WWM
0192-MCR-BUC-15P-L 15 36.5 High 0.81 0.07 0.7 0.93 4.38 RB+WWM
0192-MCR-BUC-15P-T 15 36.5 High 0.81 0.08 0.6 0.81 3.07 RB+WWM
0231-MCR-MED-16P-L 16 38.4 Inter. 0.43 0.10 1.1 1.46 9 RB
0231-MCR-MED-16P-T 16 38.4 Inter. 0.43 0.08 1.1 1.48 7.76 RB
0300-MCR-BUC-17P-L 17 45.2 High 0.33 0.09 1.7 2.34 13.63 RB+WWM
0300-MCR-BUC-17P-T 17 45.2 High 0.41 0.08 1.4 1.88 9.03 RB+WWM
0302-MCR-BUC-17P-L 17 42.5 High 0.36 0.07 1.6 2.14 4.17 RB+WWM
0302-MCR-BUC-17P-T 17 42.5 High 0.71 0.06 0.8 1.08 3.25 RB+WWM
0232-MCR-BGT-19P-T 19 45.6 Inter. 0.28 0.08 1.7 2.3 14.22 RB+WWM
0004-MCR-MED-20P-L 20 48.0 Inter. 0.20 0.05 2.4 3.17 13.79 RB+WWM
0004-MCR-MED-20P-T 20 48.0 Inter. 0.33 0.09 1.4 1.94 8.79 RB+WWM
0008-MCR-MED-20P-L 20 48.0 Inter. 0.28 0.08 1.8 2.36 11.24 RB
0008-MCR-MED-20P-T 20 48.0 Inter. 0.27 0.06 1.8 2.38 9.7 RB
0029-MCR-MED-21P-L 21 51.4 Inter. 0.40 0.07 1 1.3 6.53 RB+WWM
0029-MCR-MED-21P-T 21 51.4 Inter. 0.20 0.07 2 2.63 10.48 RB+WWM
0027-MCR-MED-23P-L 23 57.0 Inter. 0.25 0.05 1.6 2.11 8.30 RB+WWM
0027-MCR-MED-23P-T 23 57.0 Inter. 0.21 0.06 1.8 2.48 11.50 RB+WWM
0028-MCR-MED-24P-L 24 59.4 Inter. 0.16 0.04 2.3 3.16 16.12 RB+WWM
0028-MCR-MED-24P-T 24 59.4 Inter. 0.20 0.06 1.9 2.59 9.98 RB+WWM
0308-MCR-MED-25P-L 25 60.3 Inter. 0.26 0.05 1.5 2 10.52 RB
0308-MCR-MED-25P-T 25 60.3 Inter. 0.17 0.08 2.3 3.05 14.43 RB
0317-MCR-MED-26P-L 26 64.8 Inter. 0.29 0.05 1.6 2.2 10.52 RB
0317-MCR-MED-26P-T 26 64.8 Inter. 0.20 0.08 2.4 3.25 14.43 RB
0316-MCR-MED-29P-L 29 69.6 Inter. 0.26 0.07 1.5 2 5.93 RB+WWM
0316-MCR-MED-29P-T 29 69.6 Inter. 0.18 0.02 2.1 2.87 10.47 RB+WWM

3. Fragility functions
3.1. Wall modelling
The RC walls were modeled in the OpenSeesPy (Zhu et al. 2018), where a two-dimensional (2D) model
represented each archetype. The walls were simulated as multi-vertical-line-element-models (MVLEM), as this
modeling approach has demonstrated a good balance between simplicity, numerical stability, and accuracy
(Haghi et al., 2020) with reduced computational time (Pozo et al., 2020). The numerical modelling approach
for this study is described in detail in Feliciano et al. (2023).
3.2. Ground motion selection and scaling
The structural response was evaluated through nonlinear dynamic analyses with seismic records consistent
with the seismic hazard of the building’s sites. The National Seismic Risk Model (MNRS, for its acronym in
Spanish) selected a set of ground motions consistent with the seismic hazard in Colombia. The Conditional
Scenario Spectrum (CSS) methodology developed by Arteta & Abrahamson (2019) was used to select the
ground motion records. The CSS effectively discretizes the seismic hazard, as summarized in Table 2. Full
details on the ground motion selection process are reported in Pájaro and Arteta (2022).

Table 2. Characterization of the ground motion records used for nonlinear dynamic analyses.
Return Period, TR (years) [75, 150, 225, 475, 975, 2475, 4975, 9975]
Hazard Level [low; intermediate; high]
Soil Types [rock; soil] = [Vs30 ≥ 760 m/s; 360 m/s ≥ Vs30 ≥ 180 m/s]
Conditional Periods, T* (s) [0.01, 0.1, 0.3, 0.6, 0.7, 1.0, 1.5, 2.0, 3.0]

4
WCEE2024 Ocampo et al.

3.3. Nonlinear dynamic analysis


Each of the archetypes was subjected to more than 3,000 hazard-consistent ground motions records as
uniform excitation at the base. The analyses were performed using an energy-based convergence test at each
step with a tolerance of 10-8, utilizing Newton's method as the solution algorithm and a Newmark integration
scheme. An algorithm for convergence was implemented, adapting both the integration step and the solution
algorithm. The nonlinear response analyses resulted in correlated pairs of Intensity Measures (IM) based on
spectral acceleration and structural response (EDP). The IM of interest corresponds to the spectral
acceleration, 5% damped, at selected structural periods. Figure 3 presents an example with the maximum roof
drift ratio (RDR) results for each seismic record of an 8-story building in Armenia, and intermediate seismic
hazard location.

Figure 3. a) Dispersion of spectral acceleration Sa (T = 0.3s) and the maximum roof drift ratio, b) binned data
for fragility function fitting.

The results of the nonlinear analysis were binned to produce stripes of equal intensity, as presented in Figure
3b. In this case, gray points mark individual EDP (i.e., RDR) values, while continuous lines represent the
median (P50), the fifteenth (P15), and eighty-fifth (P85) percentiles of the bins. With this organization scheme,
it is possible to fit a probability distribution function of seismic demand at each bin, allowing for determining
the probability of exceeding EDP or damage threshold values given IM.
3.4. Collapse definition and fragility curve
Results like those in Figure 3b were used to define the EDP threshold for which the "collapse" of each
archetype is defined. Given that the TLRCW system is not likely to suffer from sudden vertical load-carrying
capacity, the collapse threshold definition comprises evaluating the slopes of the EDP-IM curves at the P50.
Herein, the “collapse” occurs when an apparent break in the RDR-Sa curve is observed and is marked with a
horizontal dashed line in Figure 3b. Henceforth, the collapse damage state is referred to as DS4. Upon
evaluation of the resulting threshold, it is observed, that the change in EDP-IM scaling indicates that at least
one wall of the archetype has reached or is near its deformation capacity. Collapse fragility functions were
obtained for each archetype by fitting a lognormal cumulative distribution function to the results (Baker, 2015).
Figure 4 shows the resulting collapse fragility curve for the 8-story archetype 0058-MCR-ARM-08P. Spectral
accelerations values corresponding to the Design Basis Earthquake (DE) and the Maximum Considered
Earthquake (MCE) are marked with vertical dashed lines. Details of the methodology and results can be found
elsewhere (Bonett et al., 2022).

5
WCEE2024 Ocampo et al.

Figure 4. Fragility curves for the 8-story archetype in Armenia.

4. Effect of the reinforcement type


The electro-welded wire mesh has been widely used in Colombia as the web reinforcement of Thin Lightly
Reinforced Concrete Walls. The ultimate strain capacity of the electro-welded meshes manufactured in
Colombia is less than 2%, with an average ultimate strain εsu = 1.24% as reported by Carrillo et al. (2019). The
use of this type of reinforcement in areas where incursion in the plastic range of response is anticipated is not
advisable due to its limited ductility, as evidenced by experimental results reported by Blandon et al. (2018)
and Blandón and Bonett (2020). The effect on the seismic fragility of using an electro-welded mesh with and
without ductile reinforcement at the wall edges is evaluated using 30 low-rise archetypes. The probability of
collapse is estimated for each building at two intensity levels, DE and MCE. Figure 5 compares the dispersion
of these probabilities for buildings reinforced only with electro-welded wire meshes (WWM) and those
reinforced with ductile bars concentrated at the ends and WWM as web reinforcement (RB+WWM).

Figure 5. Exceedance Probability vs. Type of Reinforcement for the state of Collapse for: a) DE y b) MCE in
low-rise buildings.

The results show higher probability of collapse in buildings reinforced solely with WWM. The expected seismic
performance of a building designed according to a seismic design code for a 475-year earthquake (i.e., for the
DE level) does not anticipate collapse, hence the probability should be very low or negligible. For the MCE,
Figure 5 shows a median probability of collapse of the WWM model exceeding 10%, with some archetypes
exceeding 20%, which does not represent satisfactory seismic performance according to the FEMA P-695
(ATC, 2009). In contrast, the median probability for buildings detailed with ductile bars and WWM ranged

6
WCEE2024 Ocampo et al.

between 8% to 12%, with a median close to 10%. Notably, the results dispersion for either intensity level is
greater for the WWM models.

5. Effect of Aspect Ratio


Figure 6 shows the dispersion in probability of collapse as a function of the mean aspect ratio of the TLRCW
building database in Table 1. Buildings were organized into three categories: a) Ar ≤ 4, b) 4 < Ar ≤ 10, and c)
Ar > 10. Overall, an increase in the probability of collapse is observed as the aspect ratio increases. At the
MCE level, archetypes with a mean aspect ratio less than 4 have a median probability of collapse below 10%,
which is within the recommended limits by FEMA P-695. Conversely, for archetypes with an average aspect
ratio greater than 10, the median collapse probability is close to 20% and it reaches 50% for some archetypes,
exceeding the limits proposed in FEMA-P695.

Figure 6. Exceedance Probability vs. Mean Aspect ratio for the state of Collapse for: a) DE and b) MCE in
the buildings of database.

6. Effect of Wall Index, WI.


The Wall Area Index (WI) is defined as the ratio between the area of the web of the walls sustaining the base
shear in one direction, and the area of all the floors above it. WI is a parameter related to the lateral stiffness
of buildings. Figure 7 shows the dispersion of the probability of collapse in terms of WI the database in Table
1. As WI increases, a decrease in the collapse exceedance probability is observed. For the DE level, the
probability of collapse is below 10%, reaching values under 5% for WI>0.30 (Figure 7a). For WI ≤ 0.15, the
median collapse probability is 20% for the MCE level and can even reach values over 35%. For a wall index
greater than or equal to 0.30, a low dispersion is observed, and the exceedance probability is close to 10%,
with no archetypes exceeding 20%, meeting the limits proposed by FEMA-695 (Figure 7b).

7. Final remarks
The results obtained in the development of the National Seismic Risk Model allowed for the calculation of
seismic fragility functions for thin-walled reinforced concrete buildings representative of the construction
practices in Colombia. The buildings range from 5 to 30 levels and exhibit variability in their structural
configuration, particularly in terms of the wall index, average aspect ratio of the walls, and reinforcement
detailing type. The following findings are drawn from the results obtained:
 The use of welded wire mesh results in collapse probabilities that exceed the limits recommended by the
FEMA P695 for the Maximum Considered Earthquake (MCE).
 The use of walls with an aspect ratio greater than 10 to control the seismic behavior of buildings increases
the collapse probability up to unacceptable levels.
 Increasing the wall index helps in controlling and reducing the collapse exceedance probability.

7
WCEE2024 Ocampo et al.

Figure 7. Exceedance Probability of collapse vs. Wall index: a) Design Earthquake (DE) and b)
Maximum Considered Earthquake (MCE).

8. Acknowledgements
The results presented in this work were obtained during the execution of the national research project titled:
“National Seismic Risk Model for Colombia” (MNRS in Spanish), led by the Colombian Geological Service
(SGC) and the Colombian Association of Engineering Faculties (ACOFI). The authors wish to express their
gratitude to all the institutions and universities that have actively participated in this project. A special
acknowledgment to all the students from Colombian universities who actively participated in the project, to the
professional associations in Colombia, and to the consulting companies that provided the structural plans to
consolidate the TLRCW database.

9. References
AIS. (2010). ASOCIACIÓN COLOMBIANA DE INGENIERÍA SÍSMICA. Reglamento Colombiano de
Construcción Sismo-Resistente, NSR-10.
Almeida, J., Prodan, O., Rosso, A., & Beyer, K. (2017). Tests on Thin Reinforced Concrete Walls Subjected
to In-Plane and Out-of-Plane Cyclic Loading. Earthquake Spectra, 33(1), 323–345.
https://doi.org/10.1193/101915EQS154DP
Applied Technology Council. (2009). Quantification of building seismic performance factors. US Department
of Homeland Security, FEMA.
Arroyo, O., Feliciano, D., Carrillo, J., & Hube, M. A. (2021). Seismic performance of mid-rise thin concrete
wall buildings lightly reinforced with deformed bars or welded wire mesh. Engineering Structures, 241,
112455.Arteta, C. (2017). Simple mechanics of reinforced concrete thinwall,design considerations for
Colombia – CEER. VII Congreso Nacional de Ingeniería Sísmica. http://ceer.co/mecanica-simple-de-
muros-delgados-con-aleta-aspectos-a-considerar-para-su-diseno-en-colombia-
2/?login=success&lang=en
Arteta, C. A., & Abrahamson, N. A. (2019). Conditional Scenario Spectra (CSS) for Hazard-Consistent
Analysis of Engineering Systems. Earthquake Spectra, 35(2), 737–757.
https://doi.org/10.1193/102116EQS176M
Baker, J. W. (2015). Efficient Analytical Fragility Function Fitting Using Dynamic Structural Analysis.
Https://Doi.Org/10.1193/021113EQS025M, 31(1), 579–599. https://doi.org/10.1193/021113EQS025M
Blandon, C. A., Arteta, C. A., Bonett, R. L., Carrillo, J., Beyer, K., & Almeida, J. P. (2018). Response of thin
lightly-reinforced concrete walls under cyclic loading. Engineering Structures, 176, 175–187.
https://doi.org/10.1016/J.ENGSTRUCT.2018.08.089
Blandón, C., & Bonett, R. (2020). Thin slender concrete rectangular walls in moderate seismic regions with a
single reinforcement layer. Journal of Building Engineering, 28, 101035.
https://doi.org/10.1016/J.JOBE.2019.101035

8
WCEE2024 Ocampo et al.

Bonett, R., Arroyo, O., Carrillo, J., Feliciano, D., Ocampo, J., Vidales, F. (2022). Funciones de fragilidad y
vulnerabilidad sísmica para la tipología constructiva de muros en concreto reforzado: Reporte MNRS
No. 004-2022. Bogotá: Preparado por la Asociación Colombiana de Facultades de Ingeniería (ACOFI)
para el Servicio Geológico Colombiano (SGC).
CAMACOL. (2019). Número de construcciones en Colombia. https://camacol.co/informacion-economica
Carrillo, J., Diaz, C., & Arteta, C. A. (2019). Tensile mechanical properties of the electro-welded wire meshes
available in Bogotá, Colombia. Construction and Building Materials, 195, 352–362.
https://doi.org/10.1016/J.CONBUILDMAT.2018.11.096
Carrillo, J., Oyarzo-Vera, C., & Blandón, C. (2019). Damage assessment of squat, thin and lightly-reinforced
concrete walls by the Park & Ang damage index. Journal of Building Engineering, 26, 100921.
https://doi.org/10.1016/J.JOBE.2019.100921
Feliciano, D., Arroyo, O., Bonett, R., Carrillo, J., Arteta, C, Vidales, F. and Ocampo, J. (2023). The 2D-MVLEM
formulation as a tool for assessing RC wall buildings with non-planar walls: case study in Colombia. 18th
World Conference on Earthquake Engineering.
Haghi, N., Epackachi, S., & Taghi Kazemi, M. (2020). Macro modeling of steel-concrete composite shear
walls. Structures, 23, 383–406. https://doi.org/10.1016/J.ISTRUC.2019.10.018
Hassan, A. F., & Sozen, M. A. (1997). Seismic vulnerability assessment of low-rise buildings in regions with
infrequent earthquakes. ACI Structural Journal, 94(1), 31-39.
Kolozvari, K., Arteta, C., Fischinger, M., Gavridou, S., Hube, M., Isaković, T., Lowes, L., Orakcal, K.,
Vásquez, J., & Wallace, J. (2018). Comparative Study of State-of-the-Art Macroscopic Models for
Planar Reinforced Concrete Walls. Structural Journal, 115(6), 1637–1657.
https://doi.org/10.14359/51710835
Pájaro, C. and Arteta, C., (2022). Selección de acelerogramas: Reporte MNRS No. 004-2022. Bogotá:
Prepared by the Colombian Association of Engineering Schools (ACOFI) for the Colombian Geological
Survey (SGC).
Quiroz, L. G., & Maruyama, Y. (2013). Comparison of numerical fragility curves for thin RC walls used in
Lima, Peru considering variations of ground motion datasets.
https://cris.ulima.edu.pe/en/publications/comparison-of-numerical-fragility-curves-for-thin-rc-walls-used-i
Rosso, A., Jiménez-Roa, L. A., Almeida, J. P. De, & Beyer, K. (2022). Instability of Thin Concrete Walls with
a Single Layer of Reinforcement under Cyclic Loading: Numerical Simulation and Improved Equivalent
Boundary Element Model for Assessment. Journal Of Earthquake Engineering, 26(1), 493–524.
https://doi.org/10.1080/13632469.2019.1691679
Rosso, A., Jiménez-Roa, L. A., De Almeida, J. P., Zuniga, A. P. G., Blandón, C. A., Bonett, R. L., & Beyer, K.
(2018). Cyclic tensile-compressive tests on thin concrete boundary elements with a single layer of
reinforcement prone to out-of-plane instability. Bulletin of Earthquake Engineering, 16(2), 859–887.
https://doi.org/10.1007/S10518-017-0228-1/TABLES/4
Zhu, M., McKenna, F., & Scott, M. H. (2018). OpenSeesPy: Python library for the OpenSees finite element
framework. SoftwareX, 7, 6-11

You might also like